You are on page 1of 229

Study on Ballistic Performance of Hybrid

Soft Body Armour

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

Doctor of Philosophy

in the Faculty of Engineering and Physical Sciences

2015

YANFEI YANG

SCHOOL OF MATERIALS
TABLE OF CONTENTS
LIST OF FIGURES ............................................................................................ 6
LIST OF TABLES ............................................................................................ 11
ABBREVIATIONS ........................................................................................... 12
ABSTRACT ....................................................................................................... 13
DECLARATION ............................................................................................... 14
COPYRIGHT STATEMENT .......................................................................... 15
ACKNOWLEDGEMENT ................................................................................ 16
PUBLICATIONS .............................................................................................. 17
Chapter 1 Introduction ..................................................................................... 18
1.1 Background.................................................................................................. 18
1.2 Research aim and objectives ...................................................................... 19
1.3 Thesis Layout............................................................................................... 21
Chapter 2 Lite rature review............................................................................. 23
2.1 Background of the ballistic body armour ................................................ 23
2.1.1 Introduction of the ballistic soft body armour ...................................... 23
2.1.2 Ballistic performance evaluation ............................................................ 26
2.1.2.1 Evaluation criteria................................................................................. 26
2.1.2.2 Selection of cartridge and fragment simulator................................... 28
2.1.3 Materials applied for ballistic resistance ............................................... 29
2.1.3.1 Aramid fibres......................................................................................... 30
2.1.3.2 UHMWPE fibres ................................................................................... 32
2.1.3.3 Influence on material properties.......................................................... 33
2.2 Factors influencing on ballistic performance ........................................... 37
2.2.1 Impact response of woven fabric ............................................................ 37
2.2.1.1 A single yarn unde r impact .................................................................. 37
2.2.1.2 A fabric under impact........................................................................... 38
2.2.2 Fabric structures ...................................................................................... 40
2.2.2.1 Fabric weave structure ......................................................................... 40
2.2.2.2 Inte r-yarn friction ................................................................................. 42
2.2.2.3 Uni-directional laminates ..................................................................... 43
2.2.3 Impact conditions ..................................................................................... 45
2.2.3.1 Striking velocity and projectile geometry ........................................... 45
2.2.3.2 Panel size and boundary condition ...................................................... 47
2.3 Ballistic response of the soft armour panel ............................................... 48
2.3.1 Multi-layer panel with uniform laye rs ................................................... 48
2.3.1.1 Inte raction between layers.................................................................... 48

1
2.3.1.2 Ballistic performance of the multi-layer armour panel ................. 49
2.3.2 Hybrid armour panel ............................................................................... 50
2.3.2.1 Hybrid materials and hybrid structures ......................................... 50
2.3.2.2 Mechanisms responding to the hybridisation effect....................... 51
2.4 Failure mechanisms of armour panels unde r ballistic impact ................... 52
2.4.1 Fracture of fibres in ballistic fabric........................................................ 52
2.4.1.1 Fracture mechanis ms of polymer fibres........................................................ 52
2.4.1.2 Fracture mechanis ms of ballistic fibres ......................................................... 53
2.4.2 Failure modes of armour panels ............................................................. 55
2.4.2.1 Woven panel.................................................................................................................... 55
2.4.2.2 UHMWPE UD panel ................................................................................................. 56
2.5 Analysis techniques ...................................................................................................................... 57
2.5.1 Experime ntal techniques........................................................... 57
2.5.2 Finite Ele ment Analysis (FEA) techniques ............................................ 58
2.5.3 Analytical techniques ............................................................... 60
2.6 Concluding re marks .................................................................................................................... 60
CHAPTER 3 Experimental studies of ballistic armour panels .................................. 63
3.1 Materials specifications and mechanical prope rties ............................................... 63
3.1.1 Materials specifications............................................................................ 64
3.1.1.1 Twaron yarns ................................................................................................................. 64
3.1.1.2 Twaron fabrics .............................................................................................................. 65
3.1.1.3 Dyneema UD ................................................................................................................... 70
3.1.2 Mechanical properties .............................................................. 71
3.1.2.1 Tensile test........................................................................................................................ 71
3.1.2.2 Shear frame test............................................................................................................ 73
3.2 Ballistic panels preparation .................................................................................................... 75
3.2.1 Woven panels ......................................................................... 75
3.2.1.1 Panels with a single layer fabric ......................................................................... 75
3.2.1.2 Perforated multi-layer woven panels .............................................................. 76
3.2.1.3 Non-perforated woven panels .............................................................................. 76
3.2.2 UD panels .............................................................................. 77
3.2.3 Hybrid panels ......................................................................... 78
3.2.3.1 Different layering sequences of components .............................................. 78
3.2.3.2 Varied amount of each component ................................................................... 80
3.2.3.3 Designed hybrid panels ............................................................................................ 81
3.3 Ballistic test ....................................................................................................................................... 82
3.3.1 Ballistic test equipme nt and procedure ........................................ 82

2
3.3.2 Data acquisition ...................................................................... 83
3.3.2.1 Striking velocity and residual velocity ........................................................... 83
3.3.2.2 Energy absorption of the panel ........................................................................... 84
3.3.2.3 Ballistic resistance tests............................................................................................ 84
3.3.2.4 Quantification of the indentation in the clay .............................................. 85
3.3.2.5 Backface Signature (BFS) ...................................................................................... 87
3.3.2.6 Indentation volume ..................................................................................................... 88
3.4 Photographic observations ...................................................................................................... 88
3.4.1 Impact process observations ...................................................... 88
3.4.2 Post-impact panels observations ................................................. 89
3.5 Test data .............................................................................................................................................. 90
3.6 Summary ............................................................................................................................................ 92
CHAPTER 4 Numerical modelling of armour panels under ballistic impact ..... 94
4.1 Introduction of nume rical modelling ................................................................................ 94
4.2 FE modelling of armour panels under ballistic impact ......................................... 95
4.2.1 FE model of a single layer fabric ................................................ 95
4.2.1.1 Geometry of yarns ....................................................................................................... 96
4.2.1.2 FE modelling of fabric under impact .............................................................. 98
4.2.2 FE model of perforate d multi-layer panels ................................. 100
4.2.3 FE model of non-perforated multi-layer panels ........................... 102
4.3 Material prope rties ....................................................................................................................103
4.3.1 Introduction of material assignment ......................................... 103
4.3.2 Projectile ............................................................................. 106
4.3.3 Twaron yarns and fabric ........................................................ 106
4.3.4 Dyneema UD laminate ............................................................ 110
4.3.5 Clay .................................................................................... 111
4.4 FE results and validation........................................................................................................112
4.4.1 Single layer fabric under impact........................................................... 112
4.4.2 Perforated multi-layer panels ............................................................... 118
4.4.3 Non-perforated multi-layer panels ....................................................... 121
4.5 Summary ..........................................................................................................................................122
CHAPTER 5 Exploration of the ballistic responses of the multi-layer system.124
5.1 The perforated woven panel .................................................................................................124
5.1.1 Fracture time .......................................................................................... 124
5.1.2 Transverse deformation......................................................................... 126
5.1.3 Stress distribution .................................................................................. 129
5.1.4 Energy absorption mechanis ms ................................................ 134

3
5.1.4.1 Energy absorption distribution ...................................................... 134
5.1.4.2 Energy absorption efficiency of each laye r ................................... 137
5.2 The non-perforated woven panel................................................................. 139
5.2.1. Transverse deflection ............................................................ 139
5.2.2 Stress distribution ................................................................................. 140
5.2.3 Energy absorption ................................................................. 142
5.2.3.1 Energy absorption distribution and efficiency ............................. 142
5.2.3.2 Influence of the number of layers .................................................. 144
5.2.3.3 Effect of the striking velocity .......................................................... 146
5.2.4 The number of broken layers ................................................... 148
5.2.5 Backface Signature BFS and the indentation volume ........................ 150
5.3 Summary........................................................................................................ 151
CHAPTER 6 INVESTIGATION ON FAILURE MODES OF ARMOUR
PANELS UNDER IMPACT .................................................................................. 154
6.1 Twaron woven panel ..................................................................................... 154
6.1.1 Failure characte ristics of woven panel ....................................... 154
6.1.1.1 A single layer woven fabric............................................................. 154
6.1.1.2 The perforated woven panel ........................................................... 157
6.1.1.3 The non-perforated woven panel ................................................... 162
6.1.2 Failure modes of the woven panel ............................................. 167
6.1.2.1 Failure modes ................................................................................... 167
6.1.2.2 Transverse deformation and stress distribution........................... 168
6.1.2.3 Energy absorption ........................................................................... 169
6.2 Dyneema UD panel ....................................................................................... 170
6.2.1 Failure modes of Dyneema UD ................................................. 171
6.2.2 Mechanisms of thermal damage ............................................... 176
6.3 Summary........................................................................................................ 177
CHAPTER 7 DESIGN OF HYBRID ARMOUR PANELS ............................... 179
7.1 Selection of components for hybridisation ................................................. 179
7.1.1 Ballistic performance of fabric with different weave structures ..... 179
7.1.1.1 Effect of areal density...................................................................... 180
7.1.1.2 Weave structures in a single layer fabric ...................................... 181
7.1.1.3 Construction of the multi-layer panel............................................ 185
7.1.2 Ballistic performance of Dyneema UD ....................................... 188
7.1.2.1 Energy absorption capacity ............................................................ 188
7.1.2.2 Advantage of small BFS .................................................................. 189
7.2 Layering sequence of hybridisation ............................................................ 191
7.2.1 Hybrid structures .................................................................. 191
4
7.2.1.1 Energy absorption ........................................................................... 192
7.2.1.2 Stress distribution............................................................................ 193
7.2.2 Hybrid materials ................................................................... 195
7.2.2.1 Energy absorption ........................................................................... 195
7.2.2.2 Backface Signature (BFS) ............................................................... 197
7.3 Design of hybrid armour panels .................................................................. 197
7.3.1 Design principle of the hybrid armour panels ............................. 198
7.3.2 The function and the range of each group in a panel .................... 200
7.3.2.1Heat resistant layers ......................................................................... 200
7.3.2.2 Energy absorption layers ................................................................ 203
7.3.2.3 Backface Signature (BFS) constraining layers ............................. 206
7.4 Ballistic performance evaluation of the hybrid panels .............................. 208
7.4.1 Hybrid panels manufacture ..................................................... 208
7.4.2 Ballistic performance evaluation of the hybrid panels .................. 209
7.4.2.1 Ballistic resistance tests ................................................................... 209
7.4.2.2 Backface Signature (BFS) tests ...................................................... 210
7.5 Summary........................................................................................................ 210
CHAPTER 8 CONCLUSIONS AND FUTURE WORK .................................... 212
8.1 Conclusions.................................................................................................... 212
8.2. Recommendations for future research work............................................. 215
REFERENCES ....................................................................................................... 217
APPENDIX 1 THE THICKNESS OF SPECIMEN ............................................ 227
APPENDIX 2 FRACTURED YARNS IN TWARON WOVEN PANEL 8F9 ... 228

5
LIST OF FIGURES

Fig. 2.1 Ballistic body armour vest.......................................................................... 24


Fig. 2.2 Energy dissipated in soft armour panel .................................................... 24
Fig. 2.3 Interceptor® body armour ........................................................................ 25
Fig. 2.4 Ballistic limit plot ........................................................................................ 26
Fig. 2.5 BFS measure ment NIJ standard-0101.06 ................................................. 28
Fig. 2.6 Cartridges .................................................................................................... 28
Fig. 2.7 Fragment Simulation Projectiles with 2-, 4-, 16-, and 64 grain size....... 29
Fig. 2.8 Tensile properties of different materials................................................... 30
Fig. 2.9 Molecular structures of PPTA ................................................................... 31
Fig. 2.10 Polyethylene unit and orientation of polyethylene ................................. 32
Fig. 2.11 Tensile behaviour at different strain rates (a) Kevlar fibres (b)
UHMWPE fibres................................................................................................ 34
Fig. 2.12 Stress-strain response of Twaron® CT716 (110tex, 12.2×12.2ends/cm,
280g/m2 ) at different strain rates...................................................................... 35
Fig. 2.13 Stress-strain response of 0/90°Dyneema HB26 laminate (SK76 fibres
with polyurethane (PU) matrix) at different strain rates............................... 35
Fig. 2.14 Breaking force of Dyneema® UD SB21 at ageing process at 700ºC and
humidity of 50% ................................................................................................ 36
Fig. 2.15 Projectile impact into body armour ........................................................ 37
Fig. 2.16 Transverse impact on a single ply of fabric (a) side view, (b) top view,
(c) bottom vie w ................................................................................................... 39
Fig. 2.17 Sche matic diagrams of fabrics upon ballistic impact ............................ 41
Fig. 2.18 Wedge through under ballistic impact.................................................... 41
Fig. 2.19 Ballistic impact load on yarns .................................................................. 42
Fig. 2.20 Sketch of the processing steps in the manufacture of the Dyneema ® HB
26 laminate material. ......................................................................................... 44
Fig. 2.21 Sche matic of Dyneema® UD ..................................................................... 44
Fig. 2.22 Cross-section of Dyneema® HB26............................................................ 44
Fig. 2.23 Three regimes of energy absorption for conical projectiles at different
impact velocity ................................................................................................... 46
Fig. 2.24 Different projectile geometries ................................................................ 46
Fig. 2.25 Energy absorption of fabric with diffe rent boundary conditions ........ 47
Fig. 2.26 Multi-layer systems (a) the layered system (b) the spaced system ...... 48
Fig. 2.27 Axial splitting of PPTA fibres .................................................................. 54
Fig. 2.28 Thermal damage of Dyneema fibres ....................................................... 54
Fig. 2.29 Two forces applied on yarns under ballistic impact .............................. 55

6
Fig. 2.30 FE model of fabric under impact (a) 3D solid FE mode (b) Pin-joint FE
model,(c) Shell FE model ............................................................................... 59
Fig. 3.1 Tensile curve of Twaron yarn (93tex) ...................................................... 65
Fig. 3.2 The cross-section of yarns in Twaron woven fabric 8F........................... 66
Fig. 3.3 Gaps in five Twaron fabrics (a) 8F, (b) 9F, (c) 10F, (d) 12F, (e) 13F ..... 69
Fig. 3.4 The specimen of Dyneema UD SB71 for tension test............................... 71
Fig. 3.5 Tensile test .................................................................................................. 72
Fig. 3.6 Tensile stress-strain curve of Dyneema UD .............................................. 72
Fig. 3.7 Pure shear deformation .............................................................................. 73
Fig. 3.8 Shear deformation of the specimen (a) The shear frame (b) Shear
deformation ........................................................................................................ 73
Fig. 3.9 Shear frame test .......................................................................................... 74
Fig. 3.10 Shear tensile curve of Dyneema UD ........................................................ 74
Fig. 3.11 Sche matic of experime ntal ballistic range .............................................. 82
Fig. 3.12 Ballistic test set-up .................................................................................... 82
Fig. 3.13 The indentations in the clay behind (a) Twaron woven fabric panel; (b)
Dyneema UD panel ............................................................................................ 85
Fig. 3.14 Wax mould of the indentation behind Twaron woven panel 8F24 (left)
and Dyneema 7U20 (right) ................................................................................. 86
Fig. 3.15 3D surface of indentations behind the (a) Twaron woven panel 8F24 (b)
Dyneema UD panel 7U20 .................................................................................... 86
Fig. 3.16 BFS measure ment ..................................................................................... 87
Fig. 3.17 Measurement of the indentation volume ................................................ 88
Fig. 4.1 Fibres packing in yarns (a) square packing, (b) hexagonal packing ...... 96
Fig. 4.2 Sche matic of yarn cross-section ................................................................. 97
Fig. 4.3 FE model of Twaron fabric (a) a single yarn, (b) the fabric and
projectile ............................................................................................................. 99
Fig. 4.4.The mesh of FE model (a) fabric and projectile, (b) yarn..................... 100
Fig. 4.5 The FE model of a multi-layer panel....................................................... 100
Fig. 4.6 Hybrid mesh (a) the fine mesh of the primary yarn, (b) the coarse mesh
of the secondary yarn, (c) hybrid mesh in fabric .......................................... 101
Fig. 4.7 Quarter-symmetry FE model of the non-perforated panel under impact
........................................................................................................................... 102
Fig. 4.8 Typical stress-strain curve of elastic-plastic material ........................... 104
Fig. 4.9 linear-elastic response of high pe rformance fibres ................................ 104
Fig. 4.10 FE results of a single layer Twaron fabric............................................ 113
Fig. 4.11 Fracture time of a single layer Twaron fabric under impact ............. 113
Fig. 4.12 Fracture d yarns in fabric 8F (a) FE model; (b) post-impact fabric... 114
Fig. 4.13 FE simulation of Twaron fabric 8F under ballistic impact................. 115

7
Fig. 4.14 Stress contours of Twaron fabric 8F at diffe rent impact time ........... 116
Fig. 4.15 Stress and strain distribution of Twaron fabric 8F ............................. 117
Fig. 4.16 The transverse deformation of fabric 8F at 5.6µs ................................ 118
Fig. 4.17 Energy absorption in three sample sizes.............................................. 119
Fig. 4.18 Energy absorption of Twaron panels with the small model size......... 119
Fig. 4.19 Energy absorption of the FE model with different mesh .................... 120
Fig. 4.20 Energy absorption of Twaron panels with the hybrid mesh............... 120
Fig. 4.21 The clay behind the Twaron panel 11F24 .............................................. 121
Fig. 4.22 FE model of the clay behind the Twaron panel 11F24 ......................... 121
Fig. 4.23 The indentation produced by the jacket ............................................... 122
Fig. 5.1 The residual velocity history of Twaron panel....................................... 125
Fig. 5.2 The maximum transverse deflection of each layer in Twaron woven
fabric panels (a) 8F, (b) 8F3 , (c) 8F6 and (d) 8F9 ........................................... 128
Fig. 5.3 Stress distribution on the first layer of the panel 8F3 ............................ 130
Fig. 5.4 Stress distribution on the second layer of the panel 8F3 ........................ 131
Fig. 5.5 Stress distribution on the third layer of the panel 8F3 .......................... 132
Fig. 5.6 Stress gradient on panels of (a) 8F6 , and (b) 8F9 at 0.3µs under impact
........................................................................................................................... 133
Fig. 5.7 The energy absorption of woven fabric .................................................. 134
Fig. 5.8 The specific energy absorption (SEA) of woven fabric ......................... 134
Fig. 5.9 The proportion of the dominant forms of energy absorption in fabric 8F
........................................................................................................................... 135
Fig. 5.10 Energy absorption of each layer in panels of (a) 8F3 , (b) 8F6 and (c) 8F9
........................................................................................................................... 136
Fig. 5.11 Energy absorption efficiency of each layer in panels 8F3 , 8F6 and 8F9
........................................................................................................................... 138
Fig. 5.12 The maximum transverse deflection of each layer in Twaron panel
11F24 .................................................................................................................. 139
Fig. 5.13 The transverse deformation of fabric layers in Twaron panel 11F24 . 140
Fig. 5.14 The stress distribution on each layer in Twaron woven panel 11F24 (a)
front layers, (b) middle layers and (c) back layers ....................................... 141
Fig. 5.15 Energy absorption of each layer in Twaron woven panel 11F24 ........ 142
Fig. 5.16 Energy absorption efficiency of Twaron woven panel 11F24 .............. 143
Fig. 5.17 Energy absorption of each layer in Twaron woven panel 11F36 ........ 144
Fig. 5.18 Energy absorption of each layer in Twaron woven panel 11F48 ......... 144
Fig. 5.19 Energy absorption of Twaron panels .................................................... 146
Fig. 5.20 Energy absorption distribution at different striking velocities .......... 147
Fig. 5.21 The number of broken layers in non-perforated panels ..................... 148
Fig. 5.22 The number of broken layers in non-perforated panels at different
striking velocities ............................................................................................. 149
8
Fig. 5.23 BFS of non-perforated panels with different areal density................. 150
Fig. 5.24 Indentation volume in the clay behind non-perforated panels .......... 151
Fig. 6.1 Impact process of a single layer Twaron fabric ..................................... 155
Fig. 6.2 The perforated Twaron woven fabric ..................................................... 155
Fig. 6.3 Failure morphology of Twaron fibres .................................................... 156
Fig. 6.4 Shear failure appearance of broken Twaron fibres.............................. 157
Fig. 6.5 Impact process of Twaron woven panel 11F9 ......................................... 157
Fig. 6.6 Perforated laye rs in Twaron woven panel 11F9 ..................................... 158
Fig. 6.7 Perforated holes on fabric layers of Twaron panel 11F9 ....................... 159
Fig. 6.8 Fractured yarns in different layers in Twaron woven fabric panel 11F9
........................................................................................................................... 160
Fig. 6.9 Broken fibres in fabric layers of Twaron panel 11F9 ............................ 162
Fig. 6.10 Impact process of the non-perforated Twaron woven panel 11F24 .... 162
Fig. 6.11 The non-perforated Twaron panel 11F24 (a) post-impact Twaron panel
and (b) the indentation in the clay.................................................................. 163
Fig. 6.12 Different extent of crease on fabric layers in Twaron panel 11F24..... 164
Fig. 6.13 The impact site of fabric layers in Twaron panel 11F24 ...................... 165
Fig. 6.14 Debris of broken yarns in the Twaron panel 11F24 ............................. 165
Fig. 6.15 Broken fibres in the non-perforated panel 11F24 (a) ply-1 (b) ply-10 166
Fig. 6.16 Sche matic of the stress propagation ...................................................... 167
Fig. 6.17 Sche matic of the fibre splitting development ....................................... 168
Fig. 6.18 Impact process of a single Dyneema UD layer ..................................... 171
Fig. 6.19 Impact process of non-perforated Dyneema UD panel 11F24 ............. 171
Fig. 6.20 Perforated Dyneema UD ........................................................................ 172
Fig. 6.21 Ply splitting on the exit face of the post-impact Dyneema UD 7U9 : (a)
Ply splitting, (b) Region 1, and (c) Region 2 .................................................. 173
Fig. 6.22 Fractured fibres on Dyneema UD panel 7U9 at different magnification
times: (a)×750 times, (b) ×850 times, (c) ×1800 times, and (d) ×3200 times
........................................................................................................................... 174
Fig. 6.23 Fractured fibres of ply-9 in the post-impact Dyneema UD panel 7U24 at
different magnification times: (a)×25 times, (b)×500 times, (c)×800 times,
and (d)×1000 times........................................................................................... 175
Fig. 6.24 Broken fibres with axial splitting in Dyneema UD panel 7U24 ........... 175
Fig. 6.25 Broken fibres with kink bind in Dyneema UD panel 7U24 .................. 176
Fig. 6.26 The interaction between the projectile and UD layer under impact.. 176
Fig. 7.1 Energy absorption of woven fabric with varie d areal density .............. 180
Fig. 7.2 Effect of areal density on energy absorption of woven fabric .............. 181
Fig. 7.3 Energy absorption of 10F (10.4ends/cm) and 10M (10.7ends/cm) ....... 182
Fig. 7.4 Energy absorption of woven fabrics for a given yarn count of 93tex .. 182

9
Fig. 7.5 Energy absorption of woven fabrics with varied crimp ratios ............ 183
Fig. 7.6 The maximum transverse deformation of 8F and 11F .......................... 183
Fig. 7.7 Stress distribution on a primary yarn of fabric (a) 8F and (b) 11F ..... 184
Fig. 7.8 Stress wave velocity on two fabrics 8F and 11F ..................................... 185
Fig. 7.9 Effect of weave density on energy absorption of woven panels 8F4 and
11F3 .................................................................................................................... 186
Fig. 7.10 Effect of yarn count on energy absorption of woven panels 11F4 and
10M3 .................................................................................................................. 186
Fig. 7.11 BFS of woven panels and Dyneema UD panel...................................... 187
Fig. 7.12 The specific energy absorption of Twaron and Dyneema UD panel.. 188
Fig. 7.13 Stress distribution contours at 3µs ........................................................ 189
Fig. 7.14 Stress distribution in the plane of Twaron panel and Dyneema UD
panel .................................................................................................................. 190
Fig. 7.15 Stress distribution perpendicular to the plane of Twaron panel and
Dyneema UD panel .......................................................................................... 190
Fig. 7.16 The maximum transverse deformation area of the last layer in Twaron
woven panel and Dyneema UD panel............................................................. 191
Fig. 7.17 Experime nt results of energy absorption of two hybrid panels .......... 192
Fig. 7.18 FE results of energy absorption of each laye r in two hybrid panels .. 193
Fig. 7.19 The stress distribution on each layer of 8F/11F ................................... 194
Fig. 7.20 The stress distribution on each layer of 11F/8F ................................... 194
Fig. 7.21 Specific energy absorption of hybrid material panels ......................... 196
Fig. 7.22 Specific energy absorption of six-laye r hybrid material panels ........ 196
Fig. 7.23 Effect of the layering up sequence on the BFS of panels ..................... 197
Fig. 7.24 Sche matic of a panel division ................................................................. 199
Fig. 7.25 BFS of panels with different components in the first group ............... 201
Fig. 7.26 Effect of diffe rent weave structure in the first group .......................... 202
Fig. 7.27 Effect of the range of the first group on specific depth ....................... 202
Fig. 7.28 The number of pe rforated layers in the multi-layer system ............... 203
Fig. 7.29 The probability of the number of perforated layers in the panel 11F24
........................................................................................................................... 205
Fig. 7.30 Division of three groups in the non-perforated panel 11F24 ............... 206
Fig. 7.31 The specific depth of different hybrid panels....................................... 207
Fig. 7.32 The specific BFS of hybrid panel composed of 11F and Dyneema UD
........................................................................................................................... 207

10
LIST OF TABLES
Table 2.1 Properties of several high performance fibres ...................................... 30
Table 2.2 Mechanical properties of UHMWPE fibre bundles ............................. 36
Table 3.1 Mechanical properties of Twaron® yarns............................................. 64
Table 3.2 Specifications of Twaron woven fabric .................................................. 65
Table 3.3 Weave structures in different fabrics .................................................... 69
Table 3.4 Specifications of Dyneema UD panels .................................................... 70
Table 3.5 Material properties of Twaron 11F and Dyneema UD SB71 .............. 75
Table 3.6 Perforated panels with varied numbe r of laye rs .................................. 76
Table 3.7 Woven panels with different constructions ........................................... 76
Table 3.8 Non-perforated woven panels with increasing layers........................... 77
Table 3.9 Perforated Dyneema UD panels ............................................................. 77
Table 3.10 Non-perforated Dyneema UD panels ................................................... 78
Table 3.11 Specification of each component for hybridisation ............................ 78
Table 3.12 Perforated hybrid panels with different layering sequences ............. 79
Table 3.13 Non-perforated hybrid panels with different layering sequences ..... 79
Table 3.14 Hybrid panels with different amount of UD layers ............................ 80
Table 3.15 Hybrid structure panels ........................................................................ 81
Table 3.16 Two hybrid panels ................................................................................. 81
Table 3.17 Ballistic performance of a single layer panel ...................................... 90
Table 3.18 Ballistic performance of the perforated panels .................................. 91
Table 3.19 Ballistic performance of the non-perforated panels .......................... 91
Table 3.20 Ballistic performance of the perforated hybrid panels ..................... 92
Table 3.21 Ballistic performance of the non-perforated hybrid panels ............. 92
Table 4.1 Yarns geometry in FE models of Twaron fabrics ................................. 98
Table 4.2 Material properties of the projectile .................................................... 106
Table 4.3 Material properties of Twaron yarn and fabric ................................. 107
Table 4.4 Material properties used in woven and UD models............................ 109
Table 4.5 Yield stress ratios of Twaron fabric and Dyneema UD...................... 110
Table 4.6 Material properties of clay.................................................................... 111
Table 4.7 Hardening properties ............................................................................ 112
Table 5.1 Fracture time of each layer in perforated Twaron panels ................. 126
Table 5.2 Energy absorption of Twaron woven panels....................................... 145
Table 5.3 Non-perforated Twaron woven panels ................................................ 149
Table 7.1 Division of the reference panel 11F24 ................................................... 205
Table 7.2 Ballistic test results of the hybrid panel and woven panel ................. 209
Table 7.3 Ballistic test results of the hybrid panel and woven panel ................. 210

11
ABBREVIATIONS

Ar : Areal density
BFS: Backface Signature
EA: Energy Absorption
FE: Finite Element
FEA: Finite Element Analysis
FEM: Finite Element Modelling
FSP: Fragment Simulation Projectile
NIJ: National Institute of Justice
PPTA: P-phenylene Terephthalamides
RCC: Right Circular Cylinder
SD: Specific Depth
SEA: Specific Energy Absorption
SEM: Scanning Electron Microscope
STD: Standard Deviation
UD: Uni-directional
UHMWPE: Ultra High Molecular Weight Polyethylene
Vi : indentation volume
v s: striking velocity
v r: residual velocity

12
ABSTRACT

Soft body armour is usually constructed by layering numerous layers of the same
fabric. Such a construction, however, may not be the most efficient in providing the
required protection due to different ballistic resistant efficiency of each layer. This
research aims to optimise the construction of the panels for soft body armour by
hybridisation in order to achieve the improvement of ballistic performance a nd
reductions in weight. Twaron woven fabrics with different weave structures and
Dyneema uni-directional (UD) laminates were used as components for the hybrid
design of panels. Two complementary research approaches were employed in this
study, namely the empirical method and the Finite Element (FE) analysis.
The first part of this research systematically revealed the different ballistic
characteristics of each layer in different positions of an armour panel and the way of
energy absorption in the panel. The fabric layers in the front, middle and back of the
panel exhibited different extent of transverse deformation and stress distribution. The
energy absorption increases from front layer and reaches to the maximum value in
the last perforated layer and then decreases gradually in the following back layers.
Such pattern of energy absorption was not affected by either the striking velocity or
the total number of layers in the panel, but the position, in the thickness, of the peak
value in energy absorption was shifted more towards the back of the panel when the
striking velocity increases. Such findings contribute to the understanding of different
ballistic responses in different positions of an armour panel under ballistic impact.
The second part of this research put forward a new hybrid design concept. According
to above theoretical understandings of different ballistic characteristics in different
positions of an armour panel, the fabric layers in the panel were discretely divided
into three groups. In addition to the performance of different components for the
panel and the influences of the laying sequence, a procedure for constructing hybrid
armour panels has been established. The first group was composed of the first few
layers on the striking face. The heavyweight fabrics as heat resistant layers were used
in this group to resist the heat generated on the striking face. The second group
contained some middle layers close to the last perforated layers. The lightweight
fabric was combined in this group due to the higher energy absorption capacity. All
back layers were classified into the third group. Dyneema UD laminates were placed
in this group to constrain the large transverse deflection of the lightweight fabric and
to minimize BFS of the panel.
Two hybrid panels were designed and evaluated. In the perforation ballistic tests, the
hybrid panel was more likely to stop the projectile compared to Twaron woven
panels with the same areal density. In the non-perforation ballistic tests, the hybrid
panel exhibited significantly lower BFS and achieved the reductions in weight. Such
hybrid design makes best use of different available materials to achieve the
improvement of ballistic performance and lightweight of a panel. It has a practical
significance for the soft armour panel design.

13
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.

14
COPYRIGHT STATEMENT

1) The author of this thesis (including any appendices and/or schedules to this thesis)
owns certain copyright or related right in it (the „Copyright‟) and he has given The
University of Manchester certain rights to use Copyright, including for
administrative purpose.

2) Copies of this thesis, either in full or in extracts and whether in hard or electronic
copy, may be 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.

3) The ownership of certain Copyright patents, designs, trademarks and any and all
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 Rights 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 Rights and/or Reproductions.

4) Further information on the conditions under which disclosure, publication and


commercialisation of this thesis, the Copyright and any Intellectual Property and/or
Reproductions described in it may take place is available in the University IP Policy
(seehttp://www.campus.manchester.ac.uk/medialibrary/policies/intellectualproperty.
pdf), in any relevant Thesis restriction declarations deposited in the University
Library, The University Library‟s regulations (see http://www.manchester.ac.uk
/library/ aboutus/regulations) and in The University‟s po licy on presentation of
Theses.

15
ACKNOWLEDGEMENT

I would like to thank my supervisor Dr Xiaogang Chen for his keen support,
professional advices and guidance for this work.

I also appreciate the financial support provided by the Zhongyuan University of


Technology to cover the tuition fees. Thanks for the 8- month support of the
maintenance to this research from University of Manchester. I also acknowledge the
provision of Twaron® yarns from Teijin® and Dyneema® laminate from DSM for this
research.

I am thankful to Mr Tomas Kerr and Mr Mark Chadwick for their assistance in


weaving fabrics, to Dr Xiangli Zhong for her training and guide in SEM observation,
and to Mr Adrian Handley and Ms Alison Harvey for their technical support in
testing mechanical properties of fabrics.

Many sincere thanks to Dr Ying Wang, who helped, encouraged and inspired me
during my studies. I also wish to express my appreciation to every member in this
group, including Yi Zhou, Ms Yanyan Chu, Miss Shengnan Min, Miss Nan Wang
Miss Yue Xue, and Mr Zishun Yuan, for their friendship and support.

Special thanks go to my dearest family for their endless love, patience,


encouragement, and huge supports during my study.

16
PUBLICATIONS

1. Y Yang, X Chen. Energy absorption among layers in the multiply system under
ballistic impact, Proceedings of the 89th Textile Institute World Conference. Wuhan,
China.2014.

17
Chapter 1 Introduction

1.1 Background

Body armour has been used in personnel protection for centuries since the Roman era
(145 B.C.) [1]. Although ballistic body armour experienced a rapid evolution during
the modern warfare, the primary functions still remain the same: to stop the
penetration of weapon into the human body and to dissipate the impact energy [2].
Nowadays, ballistic body armour is widely used by armed services, law enforcement
and security personnel. With the development of military operations, weapons and
ammunition, the body armour is required not only to possess superior ballistic
protection but also to be lightweight and flexible to ensure the wearer‟s mobility and
comfort. How to balance these two conflicting demands is always the key issue of
the body armour design.

During the 20th century, innovations in materials and manufacturing technology led
to the advent of high performance synthetic fibres (such as nylon, aramid, and high
modulus polyethylene). The application of high performance fibres in ballistic
protection has greatly pushed the development of soft body armour. All of these high
performance fibres have very high moduli, high tensile strength to weight ratios and
low density. These superior material properties contribute to the remarkable ballistic
resistant capacity and reduced weights.

It is recognized that further improvements of ballistic performance of soft body


armour will be increasingly difficult to achieve, due to the increasing financial costs
of developing new ballistic materials and the long term to market commercialization
[3]. Before new ballistic materials emerge onto the market, soft body armour still
needs further development to meet the increasing threat levels from sophisticated
weaponry and the invention of new firearms, modern military operations and
armoured vehicles. Therefore, optimising the construction of armour panels and
making full use of the current available materials will be an efficient and feasible
option in the current situation.

Traditionally, soft body armours was manufactured by layering numerous layers of


the same woven fabrics. Usually a ballistic proof vest varies from 10 to 50 layers
with weight of around 3-5kg [4]. The ballistic resistant efficiency of each layer is
18
supposed not to be uniform due to varied positions of each layer in armour panels.
As a result, combining the same fabric layers can‟t be the most efficient in providing
the best ballistic performance. In addition, there is no one material that can provide
all the required properties for ballistic protection. To make best use of materials
properties, different ballistic materials can be combined together to make the hybrid
armour panel. This is referred to as hybridisation. The study of the hybrid panel
mainly focuses on the combination of different available materials, which is feasible
to be carried out. Therefore, hybrid panels for ballistic body armour have become an
increasingly interesting topic to researchers for many years.

Many patents were put forward and some commercial hybrid products have been
proved to be very efficient in providing superior ballistic performances with reduced
weights. [63,92,93,94 ]. However, the mechanisms responding to the hybridisation
effect have not been fully understood, such as the ballistic resistant efficiency of each
panel component, the layering sequence, the proportion of each component in a
hybrid panel and performance of different hybridisation manners. To provide the
effective guide for the soft body armour design, the ballistic response of the hybrid
panel and the related hybridisation mechanisms still need to be further investigated
and thoroughly understood.

1.2 Research aim and objectives

This research aims to optimise the construction of panels for soft body armour by
hybridisation in order to achieve the improvement of ballistic performance and
reduction in weight. To achieve this aim, the research objectives are organized as
follows.

(1) Understanding the ballistic response of multi- layer armour panel under impact.
This includes different ballistic characteristics of each layer in a panel, energy
absorption distribution in the multi- layer system, related BFS, as well as failure
modes of armour panels under ballistic impact.

a. Different ballistic characteristics of each layer will be analysed experimentally and


numerically, such as fracture time, transverse deflection, stress distribution, energy
absorption, and BFS. This investigation will reflect the different ballistic resistant

19
efficiency of each layer in the panel, through which the mechanisms of energy
absorption of the multi- layer system can be identified.

b. Failure modes are important aspects to reflect ballistic response of an armour


panel under impact. It will be systematically investigated through the photographic
observations of post- impact panels layer by layer. This work will provide a wealth of
physical evidences to reflect different ballistic characteristics of each layer in a
ballistic armour panel.

(2) Optimisation of the ballistic panel design. The construction of the multi- layer
panel will be optimised by combining different components of ballistic materials in
the most effective positions. To design the hybrid armour panel, the mechanisms of
hybridisation will be analysed and identified before the design principle is put
forward.

a. Selection of components for hybridisation will be studied. In this research, Twaron


woven fabric with different structural features and Dyneema UD laminate, two of the
most popular ballistic materials, will be used as components for hybridisation.
Ballistic performance of these two proposed components will be investigated on their
advantages and disadvantages for ballistic protection.

b. The hybridisation effect will be investigated by analysing the ballistic performance


of hybrid structures and hybrid materials. Layering sequences of different
components in an armour panel will be explored to determine the bettercombination
manner of materials.

c. A procedure for constructing a hybrid armour panel will be established. An armour


panel may be divided into several groups according to different ballistic
characteristics of each layer. Ballistic functions and the range of each group will be
discussed and identified. This will provide a guide for establishment of hybridisation
procedure. The ballistic performance of hybrid panels will be evaluated
experimentally to confirm the positive hybridisation effect.

This study will be carried out using two research approaches, namely the empirical
method and the Finite Element (FE) analysis. The empirical studies will be centred
on ballistic tests and photographic observations of post- impact panels. FE numerical

20
simulation will be used to explore the ballistic response of each layer in the multi-
layer system where the empirical method is unable to acquire data.

1.3 Thesis Layout

Following the chapter of introduction, Chapter 2 reviews some fundamental


mechanisms responding to ballistic impact on armour panels in previous studies.
This literature review mainly introduces: (1) general knowledge of soft body armour;
(2) fundamental mechanisms of ballistic impact; (3) the related damage mechanisms
of armour panels under ballistic impact; (4) and analysis techniques widely used in
ballistic impact research.

The research methodology contains experimental studies and Finite Element (FE)
numerical studies, which is introduced in Chapter 3 and Chapter 4 respectively.
Chapter 3 introduces experiment studies, including material properties tests,
preparation of armour panel specimen, ballistic tests and results, and photographic
observations of post-impact panels. Chapter 4 presents Finite Element (FE)
numerical modelling of armour panels under ballistic impact at perforation and non-
perforation case. This includes geometrical modelling, material properties
assignment, and simulation results validation.

Results analysis is developed in Chapter 5, Chapter 6, and Chapter 7. Chapter 5


explores ballistic response of the multi- layer armour panels at perforation and non-
perforation cases. The ballistic characteristics of each layer are analysed, including
the transverse deformation, stress distribution, energy absorption and BFS. Energy
absorption distribution in the multi- layer panel and energy absorption efficiency of
each layer are discussed and identified.

Chapter 6 investigates the failure modes of the multi- layer armour panel under
ballistic impact. Different failure characteristics of each layer in post- impact Twaron
woven panels and UD panels are observed and analysed to reflect different ballistic
characteristics of each layer.

Chapter 7 explores the design of hybrid amour panel. This includes the selection of
components for hybridisation, the investigation of the hybridisation effect and

21
establishment of a procedure of hybrid armour panels design. Finally the ballistic
performance of the hybrid armour panels is evaluated at last.

Chapter 8 ends the thesis with conclusions and recommendations for future work.

22
Chapter 2 Literature review

In the modern society, sophisticated weaponry increases threat effectiveness level.


To develop more effective soft body armour, numerous ballistic research studies
have been carried out since the last century. Although all ballistic responses of soft
body armour have not been thoroughly and quantitatively understood, some
mechanisms responding to ballistic impact has been identified through experimental
and theoretical investigation. This chapter presents a comprehensive literature review
of previous studies, which provides a guide for this study to develop the hybrid soft
body armour.

In this chapter, the background of ballistic body armour is introduced in section 2.1.
The ballistic responses of armour panels are reviewed in section 2.2. Section 2.3
analyse the failure modes of armour panels under ballistic impact. In the section 2.4,
some widely used analysis techniques in ballistic researches are described. A
summary is presented in the last section.

2.1 Background of the ballistic body armour

2.1.1 Introduction of the ballistic soft body armour

Ballistic body armour is a class of protective clothing which protects human body
from the bullets and steel fragments from handheld weapons and exploding
munitions [5], which is commonly used in the military and law enforcement forces
for personnel protection. It must be worn to be effective. The basic requirement is to
provide extraordinary ballistic protection levels at a reasonable weight to ensure the
wearer‟s necessary mobility and comfort. It means that the ballistic body armour
must conform to the user‟s body, properly distribute their weight over the body to
minimize the user‟s fatigue, provide sufficient breathability for extended use, and
must not interfere with or restrict the user‟s mobility [3]. Due to modern military
systems becoming faster, more agile, and more mobile [6], an increased development
emphasis is being placed on the reductions in weight of ballistic body armour.
However, the weight reduction is always related to the compromise of ballistic
performance due to the reduced material. In ballistic researches, how to balance the
two conflict requirements is always the key issue needed to be solved.

23
Depending on application characteristics, ballistic body amour can be broadly
divided into soft body armour and hard body armour. So ft body armour is made from
woven fabrics, nonwoven, and compliant laminates (armour grade composites with
only~20% resin), which are typically flexible, strong and lightweight. It can protect
wearers against most common low and medium impact energies with striking
velocities up to 500 m/s [7]. To increase the ballistic protection level of the body-
armour vests at high risk attack, with respect to 0.30 caliber or small threats, ceramic
insert strike-plates are commonly used to cover vital organs as shown in Fig.
2.1(a).The hard body armour is made from rigid materials, such as polymer
composite, ceramic, metal, silicon carbide (SiC) and boron carbide (B4 C) plates [4].
It is rigid and thicker, which is widely used in the lightweight vehicle armour,
combat shields, and blast protection for infrastructure [8].

Ceramic plate

Vest Cover Soft armour panel

(a) (b) (c)


Fig. 2.1 Ballistic body armour vest [3]

Traditionally, soft armour panels are manufactured by layering up multi- layer woven
fabrics stitched together as shown in Fig. 2.1 (b). Depending on the protection level
and the yarn count, the number of fabric layers of a ballistic proof vest varies from
10 to 50 with weight around 3-10 kg [5].

Impact energy Transmitted energy

Blunt trauma

Energy absorption in a panel

Fig. 2.2 Energy dissipated in soft armour panel [3]

24
Soft body armour dissipates the kinetic energy of a projectile mainly by fabric
deformation, yarns stretching, and yarns fracture. Some impact energy is absorbed in
the armour panel. The transmitted energy through the panel can cause potential
injury to the wearer due to blunt trauma as shown in Fig. 2.2. To achieve better
ballistic performance, the impact energy is expected to be absorbed in soft body
armour as much as possible.

With the invention of new firearms, modern military operations and complicated
weaponry, ballistic body armour has experienced a rapid evolution. The first ballistic
armour made of 30 layers of cotton, which was used in battles by US Navy in 1871.
It can only resist against low- velocity bullets (around 100m/s) [9]. After then, several
types of body armour, such as the chrome nickel steel „Brewster Body Shield‟
(around 18 kg) during World War I and Nylon „flak jackets‟ (around 10 kg)in the
early stages of World War II, have also been quickly out of date due to clumsy,
expensive and inefficient. With the advent of high performance fibres such as Kevlar
in the 1960‟s, today's modern generation of concealable body armour has achieved
significant improvement of ballistic resistant capacity and weight reduction.
Technologies such as the „Personnel Armour System‟ (PASGT) for Ground
Troop (from the mid-1980s until the mid-2000s), and „Interceptor body
armour‟ (IBA) (from the late 1990s up till now) have been widely applied in the
military field.

Fig. 2.3 Interceptor® body armour [10]


25
The newer Interceptor® armour system issued by United States military can protect
form 9 mm sub- machine gun fire. The armour vest is only 3.8kg. When two ceramic
plates are inserted front and back (around 7.4 kg) as shown in Fig. 2.3, it can resist
threats up to NATO (the North Atlantic Treaty Organization) rifle rounds. Their
application has greatly reduced casualty and death in the Iraq and Afghanistan war
[6]. Since 1990‟s, the vests with combining stab and ballistic protection have been
developed to provide multiple functions used by the law enforcement [10].

2.1.2 Ballistic performance evaluation

Different countries and regions have their own body armour standards to ensure
proper performance and reliability levels against ballistic and fragments threats.
Several most widely used standards in the world are US National Institute of Justice
(NIJ Standard), North Atlantic Treaty Organization (NATO), UK Home Office
Scientific Development Branch (HOSDB), US National Bureau of Standards (NBS),
US Personnel Protection Armour Association (PPAA) and so on. The National
Institute of Justice (NIJ Standard) is the most representative one, which was broadly
accepted by many researches and producers of soft body armour. It establishes
minimum performance requirements and methods of testing for the ballistic
resistance of police body armour intended to protect the torso against gunfire [3].

2.1.2.1 Evaluation criteria

According to NIJ standards-0101.06 [7], the ballistic performance of soft body


armour is often characterized by the ballistic limit. The ballistic limit velocity is the
striking velocity below which the projectile will fail to perforate the armour panel.
Energy Absorbed by Fabric

Ballistic limit (BL)


No Penetration
Penetration

Initial Projectile Velocity Vs

Fig. 2.4 Ballistic limit plot [3]


26
For different armour panels, a higher the ballistic limit velocity indicates the better
the ballistic resistant capacity. However, the variability in ballistic tests reduces the
predictive power of the ballistic limit. The variability may be come from the armour,
test backing materials, bullet, casing, powder, primer and the gun barrel. So a variety
of ballistic limit velocities including V0, V50 and V100 are defined with a specified
statistical significance. According to NIJ standard-0101.06, V50 is typically denoted
as the ballistic limit [7]. It is the velocity at which the bullet is expected to perforate
the armour at 50% chance for a given bullet or projectile type. The ballistic limit is
often plotted to the specific energy absorption as shown in Fig. 2.4. At the ballistic
limit, fabric can achieve the highest energy absorption.

Energy absorption in the panel is another popular criterion of ballistic performance in


ballistic tests. It directly reflects the energy absorption capacity of panels. In the
perforation case, it can be easily obtained according to the striking velocity and
exiting velocity of the projectile. In the non-perforation case, the energy absorption
in the panel is hard to be identified. Karahana [11] proposed that it can be calculated
by removing the transmitted energy in the clay from the kinetic energy of projectile.
According to the tested energy absorption per unit indentation volume (8.721×10-4
J/mm3 ), the amount of transmitted energy can be calculated b y the volume of the
indentation in the clay. However, the indentation volume is also significantly
affected by the striking velocity [12]. Due to the variability of the striking velocity,
the transmitted energy is hard to be accurately identified just by the indentation
volume.

Even if a bullet or projectile is stopped by the armour panel, the „behind-armour


blunt trauma‟ (BABT) can cause a potential injury to the wearer. This will result in
internal organs damage or even death in extreme cases. Therefore, Backface
Signature (BFS) is used to specify the impact deformation limits of soft body armour
in the backing material caused by non-perforation impact. It is the perpendicular
distance from the reference plane of the backing material surface to the deepest point
of indentation as shown in Fig. 2.5. The backing materials is usually Roma Plastilina
No. 1 (RP #1) oil based modelling clay, which has the similar mass properties to
those of human body. The clay does not spring back after impact, so it can memorize
the shape of dent. According to NIJ standard-0101.06, BFS behind body armour
cannot be greater than 44 mm in US [7].
27
Fig. 2.5 BFS measurement NIJ standard-0101.06 [7]

2.1.2.2 Selection of cartridge and fragment simulator

Personnel protection armour is designed to resist penetration from small arms


ammunition. Small arms ammunition also known as cartridges are used in rifles,
handguns and machine guns as shown in Fig. 2.6. The range in sizes is from 0.22
calibre cartridges through to 30 millimetre cartridges as shown in Fig. 2.6 [13]. It
should be noted that all armour systems have a limit of protection level for a certain
type of impact.

Projecti l e

Casing

Propellant

Rim

Primer

Fig. 2.6 Cartridges [14]

Based on the large body of casualty data after the Vietnam War, the greatest threat to
soldiers is fragments rather than the bullet in a combat situation [15]. It was reported
28
that in Operation Iraqi Freedom II, the numbers of injuries were classified into blast
(41%), ballistic impact (38%), and other (21%) [16].

Fragments are generated when a bomb, grenade or artillery shell explodes.


According to the study of velocity distributions of fragments, warhead explosives
have blast speed of 6,100 to 9,100 m/s. This results in very high speed of ejected
fragments over 1000m/s [17]. 95% of all fragments from a bomb blast under 4 grains
(0.26g) have a velocity of 910m/s or less.

For fragment test, Right Circular Cylinder (RCC) is specified as one of most
common Fragment Simulation Projectiles (FSPs) used in fragment tests according to
NATO standards. The test series most often includes 2 grain (0.13g), 4 grain
(0.263g), 16 grain (1.0g), and 64 grain (4.2g) mass as shown in Fig. 2.7. In contrast
to the deformable lead bullets, steel fragments do not change shape, which are steel
and cannot be deformed by textile materials. The fragment tests are always used
exclusively for screening tests of body armour materials, acceptance testing of body
armour systems, and quality assurance testing.

Fig. 2.7 Fragment Simulation Projectiles with 2-, 4-, 16-, and 64 grain size[3]

2.1.3 Materials applied for ballistic resistance

The performance of soft body armour is highly dependent on material properties [18].
With the advent of Kevlar fibres in the 1960‟s, the ballistic resistant capacity of soft
body armour has been greatly improved. All high performance fibres used for
ballistic protection, such as aramid and UHMWPE fibres, possess superior
mechanical properties of high strength, high moduli and low strain as shown in Fig.
2.8.

29
Dyneema

Tenacity (cN/tex)
Aramid

Polyester

Steel wire Nylon

Elongation (%)
Fig. 2.8 Tensile properties of different materials [19]
Table 2.1 Properties of several high performance fibres [3]

Elastic Tensile Fibre


Density Strain to
Material Fibre Modulus strength diameter
(g/cm3 ) failure (%)
(GPa) (GPa) (µm)
Polyamide Nylon 66 1.14 6 1.16 19 25
Twaron 1.45 121 3.1 2 10
Kevlar 29 1.44 70-83 3.6 4 12
Kevlar 129 1.44 96 3.39 3.5 12
Aramid
Kevlar 49 1.44 113 3.6-4.1 2.6 12
Kevlar 149 1.47 186 3.3 2 12
Kevlar M2 1.44 70 3.3 4 12
Dyneema
0.97 109-132 3.30-3.90 3.5-4.0 21
SK75,76,78
Dyneema
0.97 65-100 2.40-3.30 3.5 21
SK60,62,65
Dyneema
0.97 52 2.2 3.5 21
UHMWPE SK25
Spectra 900 0.97 73 2.4 2.8 28
Spectra
0.97 103 2.83 2.8 38
1000
Spectra
0.97 124 3.34 3 38
2000

2.1.3.1 Aramid fibres

Aramid fibre is aromatic polyamides, which was first introduced by DuPont in the
early 1960s. Due to excellent tensile properties, the para-aramid, also known as p-
phenylene terephthalamides (PPTA) fibre, is widely applied in aerospace, military
protective clothing, and ballistic composites. Kevlar® (DuPont) and Twaron® (Teijin)
30
are two of the most popular commercial products. All PPTA fibres possess the high
tensile strength (with 5 times than steel), high tensile modulus and high toughness.
These superior material properties are derived from the special molecular structures
as shown in Fig. 2.9.

Fig. 2.9 Molecular structures of PPTA [19]

PPTA filaments consist of long molecular chains, which are constructed with
aromatic structures and amide groups as shown in Fig. 2.9. Any one structure of
aromatic and amide is very strong. These molecular chains are highly oriented
parallel with many inter-chain bonding (hydrogen-bonding) as shown in Fig. 2.10
[20]. In addition, the high crystallinity of PPTA filament is above 90%. All of such
characteristics of molecular structure contribute to the high strength of PPTA
filaments. Due to the rigid flat rod molecules and the extrusion processes, PPTA
filament exhibits anisotropic properties. The filament is stronger and stiffer in the
axial than that of in the transverse direction.

The aromatic rings in molecular chains assure the stable thermal properties of PPTA
fibres with excellent heat resistance. Such fibres do not melt or support combustion.
The decomposing temperature is at about 427°C to 482ºC [6]. The thermal stability
ensures integrity of the ballistic structure at relatively high environment temperature.

Kevlar® is one of the PPTA products from DuPont, which replaced nylon in the
modern generation of lightweight body armours in the 1960s. Some most common
products for ballistic protection includes Kevlar 29, Kevlar 49, Kevlar 129, and
Kevlar KM2 [6]. The specifications of these material properties are listed in Table
2.1. Among them, Kevlar 29 was used for manufacturing the PASGT® bullet-proof
vests. Kevlar KM2 has been used for the Interceptor® vests to achieve the higher
ballistic protection with casualty reduction.

31
Twaron® is another product of the para-aramid filament developed by Teijin®
Aramid Japan, which was introduced into the market in the last 20 years. Twaron
fibres have a higher number of fine microfilaments. Many properties were improved
with the developments of finer filaments combined with higher tensile properties. It
was well adapted to the specific demands of hard and soft ballistic applications [9].
The detailed specification of Kevlar® and Twaron® fibres were listed in Table 2.1.

2.1.3.2 UHMWPE fibres

The Ultra High Molecular Weight Polyethylene (UHMWPE) fibres are


®
commercialized in the late 1970s by the Dutch chemical company DSM . Due to the
extreme high tensile properties and the lowest density among all fibres (0.97 g/cm3 ),
it is reported that this kind of fibre has superior energy absorption capacity with four
times of aramid fibre [21]. Therefore, UHMWPE fibre has been quickly applied in
the field of ballistic protection products.
High-performance Regular
polyethylene polyethylene

orientation>95% orientation low;


crystallinity<85% crystallinity<60%

(a) Polyethylene unit [21] (b) orientation of polyethylene [19]


Fig. 2.10 Polyethylene unit and orientation of polyethylene

UHMWPE fibre is made up of extremely long chains of polyethylene with a large


molecular mass numbering in the millions [22]. The chemical formula for the
monomer of polyethylene fibres is shown in Fig. 2.10(a). The high tensile strength is
largely derived from the long individual molecule chain with the primary covalent
bonds, which transfers load more effectively, in addition to the accumulated effect of
Van der Waals bonds between chains [23]. Polyethylene is amorphous in the gel
state. The gel-spun fibres can attain a high degree of chain extension with parallel
orientation greater than 95% and a high level of crystallinity up to 85% by drawing

32
process as shown in Fig. 2.10 (b) [19]. Due to the higher orientation of the molecular
structure, UHMWPE fibres have fairly lower friction co-efficient (0.2) compared to
the aramid fibres (0.32) [24]. In addition, UHMWPE fibre does not contain any
chemical groups that can react with common chemicals. Hence the fibre is inertness
and stability [19]. It‟s very resistant to water, moisture, most chemicals, and UV
radiation. Due to the weak bonding between olefin molecules, UHMWPE fibres have
poorer heat resistance with relatively low melting point (130-145°C) [25]. This has
been some concerns regarding the thermal damage effect on ballistic performance of
soft body armour.

Dyneema® from DSM and Spectra® from Honeywell, are two of the most popular
commercial UHMWPE products. The linear density of multifilament in a yarn varies
from about 0.44dtex to 11dtex. The tenacity may well be over 5N/tex, and modulus
can be over 120N/tex [19]. The mechanical properties of two products are listed in
Table 2.1.

2.1.3.3 Influence on material properties

Under ballistic impact, the body armour experiences very high strain rates. It is
estimated to be around 1000s-1 based on interaction length of 2cm at the impact
velocity of ~500m/s [5, 26]. The influence of strain rate on material properties has
been widely studied.

In many studies, the mechanical properties of ballistic fibres have been found to be
sensitive to strain rates [20, 27-29]. Wang and Xia tested Kevlar 49 fibre bundles at
varied strain rates from 10-4 s-1 to 103 s-1 . They observed the initial elastic modulus,
tensile strength and failure strain increased with increasing strain rates as shown in
Fig. 2.11 (a) [27]. However, someone argued that this may be due to the slack and
variability in test [20, 30, 31]. For UHWMPE fibres, the maximum strength also
exhibit some strain-rate dependency as shown in Fig. 2.11 (b) [20].

33
Strain rate ( s-1)
1350
440
140
0.01

Stress (GPa)
0.0001

-6
Strain (10 )
(a) Kevlar fibres [27]

-1
800 s
Engineering Stress (GPa)

530 s -1
340 s -1

Engineering Strain
(b) UHMWPE fibres[20]
Fig. 2.11 Tensile behaviour at different strain rates (a) Kevlar fibres [27] (b)
UHMWPE fibres[20]

For fabric or laminate made of these ballistic fibres, the dynamic mechanical
properties at different strain rates have also been investigated by many researchers.
Shim [32] observed the tensile stress and elastic modulus of Twaron® fabric
increased with the increasing strain rate, while the failure strain decreased
remarkably as shown in Fig 2.12. They analysed that the amount of energy
absorption of fabric reduced at high strain rates due to the ductile-to-brittle failure
transition. Such characteristics have also been observed by Chocron et al [33].
Russell studied the effect of strain rate on UHMWPE fibres, yarns and Dyneema
laminates [34]. The stress-strain response of 0/90 Dyneema laminate was shown in
Fig 2.13. At the lowest strain rate of 10-4 s-1 , an initial non- linear behaviour is
observed. As the strain rate is increased, the response becomes more linear, and the
strain to failure decreases. The peak strength of 725 MPa is independent of strain rate
34
over the range tested, which is obviously higher than that of Twaron fabric (507MPa).
Koh and Shim [35] tested Spectra® Shield laminate and found there were an increase
in stress and stiffness at failure and a decrease in strain with an increase in strain rate.
At strain rates greater than 400s-1 , this trend reversed, reducing failure stresses and
stiffness and an increase in failure strain. The study cites SEM images of broken
filaments that undergo ductile and shear failure, as opposed to the br ittle failures seen
below 400s-1 (and above quasi-static strain rates).

495s -1
238s -1
1s -1
-1
0.1s
Stress (MPa)

0.01s -1

Strain

Fig. 2.12 Stress-strain response of Twaron® CT716 (110tex, 12.2×12.2ends/cm,


280g/m2 ) at different strain rates [32]

10-3s-1
10-2s-1

10-4s-1
Stress (MPa)

Strain
Fig. 2.13 Stress-strain response of 0/90°Dyneema HB26 laminate (SK76 fibres
with polyurethane (PU) matrix) at different strain rates [34]

In the daily use, the soft body armour is required to withstand different
environmental temperatures. The elevated temperature may cause the degradation of
material properties [36]. For soft body armour made of aramid fibres, the thermal

35
stability of aramid fibres ensures the integrity of the ballistic structure at relatively
high environment temperature. DuPont Company tested 79 bulletproof vests made of
para-aramid fabrics that were used from 2 to 10 years. Only 16 of them showed only
a reduction of 8-18% in the ballistic protection limit V50 [37].

Due to the lower melting point, UHMWPE fibres exhibited more sensitivity to the
elevated temperature than that of PPTA fibres. The ageing problem during usage has
always been a concern since UHMWPE products used in ballistic armour panels.
Some studies proved that the material properties of UHMWPE products has been
obviously influenced by the elevated temperature as shown in Table 2.2 and Fig.
2.14 [38]. In the laboratory ageing tests, ballistic performance degradation of ballistic
vests were also observed. The chemical structure of macromolecules and crystallinity
rate had changed with the time of ageing [37]. Therefore, UHMWPE fibres should
not be exposed to temperature exceeding 80-100ºC for long periods of time [19].

Table 2.2 Mechanical properties of UHMWPE fibre bundles [38]

E(Gpa) бb(Gpa) εb(%)

25℃ 80 2.55 6.52


300/s
70℃ 61 2.47 7.54
25℃ 82 2.55 6.26
700/s
70℃ 68 2.48 6.57
Breaking force, lengthewise [N]

Accelerated ageing time [days]


Fig. 2.14 Breaking force of Dyneema® UD SB21 at ageing process at 700ºC and
humidity of 50%[37]

36
However, some other tests results showed that the material properties of UHMWPE
fibres are not changed at the elevated environment temperature and the ballistic
performance can be retained during the usage [20, 38, and 39]. Such controversial
conclusions indicated that the ageing problem of UHMWPE ballistic products still
needs further investigation.

2.2 Factors influencing on ballistic performance

Soft body armour is constructed from numerous hierarchies including the fibre, yarn
and fabric level. Under ballistic impact, mechanisms influencing on the performance
of soft body armour is a combination of many simultaneous factors. Ballistic
performance is not only dependent on the material properties of the fibre, but is also
was affected by fabric structures, construction of the panel and impact conditions. It
is difficult to isolate each mechanism related to the individual factors [39].
Correspondingly, ballistic behaviour of soft armour panel becomes very complex.

2.2.1 Impact response of woven fabric

2.2.1.1 A single yarn under impact

As a starting point, the transverse impact into a fibre and fabric was described firstly.
When a projectile strikes a fibre, the longitudinal wave and transverse wave begin to
propagate from the point of impact. The longitudinal tensile wave travels along the
fibre axis at the sound speed of the material. As the tensile wave propagates away
from the impact point, the material behind the wave front flows toward the impact
point to produce the transverse movement of the yarn, which is the transverse wave.
The transverse wave is propagated at a lower velocity than that of the longitudinal
velocity of the material [39].

Fig. 2.15 Projectile impact into body armour [39]


37
Materials which possess a high modulus and low volume density have high stress
wave velocities as can be seen from the equation (2-1). Such material can dissipate
the stress wave rapidly away from the impact point [40]. This reduces the high stress
concentration and distributes the impact energy over wider area. Correspondingly,
more materials are in strain and involved in energy absorption. Therefore, the
materials possessing the high stress wave velocity have the higher energy absorption
capacity [5, 41]. Dingenen stated that the tenacity and failure elongation of yarns
determine the amount of energy absorption [19].

Although mechanical properties of material play a large role in ballistic performance,


such as tensile strength, modulus and failure strain, each individual property cannot
*
determine it [39]. Cunniff has derived a dimensionless fibre property U to assess
the performance of fibres as shown in equation (2-3). It took into account of the
specific fibre toughness and its strain wave velocity [42].

√ (2-1)

√ (2-2)

√ (2-3)

where c is longitudinal wave speed; E is the yarn tensile elastic modulus, ρ is


volumetric density. u is transverse wave speed , σ is the tensile stress and ε is the
strain.

During ballistic impact, fibre straining is the primary mechanisms of energy


absorption. Due to the impact force, fabric produces the transverse deflection. Some
yarns are engaged in strain and stretched. As a result, the strain energy was produced
in fabric before fracture.

2.2.1.2 A fabric under impact

When a projectile impacts fabric, there is a similar impact response as in a single


yarn. The longitudinal stress waves generate in primary yarns which directly contact
with the projectile and propagates down the axis of the yarns as shown in Fig. 2.16.
Additionally, secondary yarns, defined as yarns that intersect with primary yarns, are
38
then pulled out of the original fabric plane by the primary yarns. These secondary
yarns undergo transverse deformation and develop a stress wave like those observed
in primary yarns. Analogously, these orthogonal yarns then drive yarns with which
they intersect. These yarn-yarn interaction produce bowing, the misalignment of the
orthogonal yarns toward the impact point. Thus, the stress wave is propagated over
wider area on fabric. The transverse deflection proceeds until the strain at the impact
point reaches a breaking straining of material [39].

As woven fabric is orthotropic material, the stress wave can‟t propagate in all
directions around the impact point. It can only propagate along primary yarns at first.
This can be seen from the higher stress on the primary yarns under impact as shown
in Fig 2.17. As a result, a pyramid shape of the deformed fabric was observed upon
impact. Roylance analysed that the majority of the ballistic impact energy is
deposited in primary yarns, while other yarns are essentially ineffective [41].

Fig. 2.16 Transverse impact on a single ply of fabric (a) side view, (b) top view,
(c) bottom view [39]

Smith [43] discussed that due to the wave reflection from crossovers between warp
and weft, there was stress attenuation in fabric along its yarns. Cunniff [44]
concluded that the stress wave attenuation on fabric produced a strain gradient on
each yarn. The maximum value of the stress and strain was at the point of impact and
decreased along the length of the yarn. Roylance [45] proposed that the wave
velocity in fabric was a fixed fraction of the wave velocity in a single fibre taking
account of the effective double of linear density of a fibre caused by crossovers.

39
(2-4)

where c‟ is stress wave speed on the fabric, c is the stress wave speed on the fibre

Under ballistic impact, the lost impact energy is mainly transferred in forms of yarn
kinetic energy, yarn strain energy, frictional energy dissipation, if heat dissipation,
acoustic energy dissipation and any rotational kinetic energy of the projectile are
neglected for simplification [3, 46]. The impact energy dissipation in fabric is
derived from transverse deflection of fabric and inward movement of yarn material
towards the impact point. Fibre straining due to the transverse movement is the
primary mechanisms of energy absorption. Lee et al. [47] found that in the fabric
composites, the number of broken yarns exhibited good correlation with energy
absorption. A proportion of the impact energy is dissipated through friction between
yarns and between projectile and yarns [48].

2.2.2 Fabric structures

2.2.2.1 Fabric weave structure

Woven fabric is the most commonly used traditional structure for soft body armour.
In addition to material properties, fabric weave structure plays a key role in ballistic
performance, which includes a series of weave parameters.

Ballistic fabric usually incorporates yarns with little twists to ensure the efficient
conversion of fibres strength. Increasing twist level will reduce tensile strength and
modulus of fibres. Rao studied the optimal twist angle of yarn is around 7 o [49]. A
twist factor of 1.1 was found to be suitable for Kevlar yarns [50].

The yarn count also has an influence on ballistic performance of the fabric. Fabrics
made from fine aramid yarns of 220-440dtex were observed to exhibit better
performances than those made from the coarser yarns. However, due to high costs of
the finer yarns and weaving, 1100dtex is usually used [51].

Plain weave and basket weave are two typical weave patterns applied in ballistic
protection fabric. The most crossovers in fabric contribute to better yarn gripping and
the balanced stress distribution. By contrast, the unbalanced weave fabrics have
inferior ballistic performance. Cunniff [44] observed the asymmetric transverse

40
deflection in the unbalanced fabric. As a result, less material was in strain and
involved into energy absorption. To spread the impact load more evenly and widely
in fabric, the triaxial fabric was investigated on ballistic performance. However,
Hearle observed the inferior V50 performance of triaxial fabric compared to the
biaxial fabric due to the inherent openness of the weave as shown in Fig.2.17 [52].

(a) Biaxial fabric [28] (b) Triaxial fabric[52]


Fig. 2.17 Schematic diagrams of fabrics upon ballistic impact

Weave density known as „cover factor‟ is one of key parameters determining the
fabric structure. If the weave is too tight, the transverse deflection will be restricted.
Due to high stress concentration at the impact point, fabric will quickly fail. If fabric
is too loose or has low inter- yarn friction, yarns in fabric are easily to be pushed
aside by the projectile. Such „wedge through‟ phenomenon on loose woven fabrics
was observed by many researchers as shown in Fig. 2.18 [47, 53-55]. This will
produce too large transverse deflection, which is related to blunt trauma to the wearer
[5]. Chitrangad [56] analysed the cover factors of fabric should be from 0.6 to 0.95
for the effective ballistic application.

Fig. 2.18 Wedge through under ballistic impact [39]

To prevent yarn slippage and increase yarn gripping in the fabric under impact,
narrow fabric with two-sided selvages was designed by Cork [57]. Ballistic tests
41
results confirmed that energy absorption in narrow fabrics, particular for the closer
weaves, increased than that of the wide fabric panels. However, narrow fabrics
possessed lines of weakness between strips. Some researches employed leno
structures in ballistic fabric to increase the yarn gripping [58, 59].

Yarn crimp is the undulation of yarns due to the interlacing between yarns in fabric.
Due to the crimp effect in fabric, impact loading on fabric induces transverse loads at
crossover with yarns decrimping as shown in Fig. 2.19. This reduces the translation
of fibre strength to fabric strength [60]. Tan [61] analysed that yarns decrimping was
the initial stage of fabric deformation. The fabric only started to resist the projectile
when the yarns had straightened and began to stretch. Taking account of different
crimp ratios between warp and weft yarns in the plain fabric, Chitrangad [56]
designed a fabric with balanced crimp using weft yarn having larger elongation to
break to achieve the improved ballistic limit velocity.

Ballistic
Impact
Fx

Stress along
fibre axis
Fy
Ballistic
Impact

Fig. 2.19 Ballistic impact load on yarns [62]

2.2.2.2 Inter-yarn friction

In previous researches, the inter-yarn friction in fabric was demonstrated to have


influence on energy absorption of fabric under ballistic impact. Duan [46] analysed
that although the frictional energy dissipation only accounts for a very small portion
of the total energy absorption, the inter-yarn friction hindered the movement of yarns
thus engaged more yarns in strain. As a result, both the yarn strain energy and yarn
kinetic energy increased with the friction existing. The frictional characteristics of
fabric are often determined by yarn pull-out test, which is conducted by extracting
single yarns from woven fabric and monitoring the force-displacement response.

42
Such quasi-static tests results was believed to be quantitatively correlated with yarn
pull-out during ballistic impact [63, 64]. Zeng et al. [65] investigated the optimum
friction coefficient range was from 0.1 to 0.6. If the friction coefficient between
fibres is too high, fibres will cut each another during the penetration. While if it is
very low, the material will not offer any resistance to the penetration. Bazhenov [66]
experimentally observed about 40% reduction of ballistic protection of wet armour
panel, due to the water as a lubricant decreased friction between the bullet and yarn.

Ballistic performance fibres used for soft body armour usually have lower friction
coefficient due to high drawing during manufacture process. This results in the easy
yarns slippage under impact and reduces energy absorption of fabric. Therefore,
many efforts have been put to increase the friction coefficie nt of fibres by chemical
or mechanical manners to improve energy absorption of fabric. Some studies
modified the friction coefficient of fibre by coating [67, 68], impregnated with shear
thickness fluids (STF) [69] or plasma etching [70, 71]. Lee et al [72] reported
compared with 14 layers of neat Kevlar fabric, the impregnated STF Kevlar fabric
composites can dissipate the same amount of energy and provide higher impact
resistant at equivalent weight. Hogenboom [73] designed a hybrid yarn to increase
the inter-yarn friction by combining the filaments of high frictional coefficient with
high performance fibres. These two types of fibres were twisted or core spun
together as a yarn to obtain improvement of ballistic performance.

2.2.2.3 Uni-directional laminates

Uni-directional (UD) laminates is a kind of complaint composite. Unlike traditional


structural composites, UD laminates only contain less than 20% weight fraction
matrix. In the manufacture process as shown in Fig. 2.20, all fibres are laid in parallel
bonded with thermoplastic resin to form a sheet. Several plies of sheets are stacked in
the direction of 0/90° and hot pressed between two films to produce the final
consolidated laminate as shown in Fig. 2.21 and Fig. 2.22. The configuration of the
cross-section of UD showed the fibre diameter is unchanged by the hot-pressing
operation [34]. Commercial UD laminates used for ballistic protection include
Spectra Shield® (UHMWPE fibres), Gold Shield® (Kevlar fibres), K-Flex® (Kevlar
®
fibres) from Honeywell company and Dyneema ® UD (UHMWPE fibres) from
DSM® company.

43
UHMWPE in solvent

Step I Gel-spinning + hot drawing


Polymer Fibre
Oven
solution
Fibre (yarn) Cooling bath

Step II Ply fabrication


Yarn
Resin application Drying
Ply precursor
Ply precusor

Step III Consolidation

Ply precusor Heat +pressure Laminate [0/90]48


laminate

Fig. 2.20 Sketch of the processing steps in the manufacture of the Dyneema® HB
26 laminate material. [34]

Fig. 2.21 Schematic of Dyneema® UD [34]

Fig. 2.22 Cross-section of Dyneema® HB26[34]

Due to no cross-over between fibres in UD laminate, the stress wave on UD


laminates is believed to propagate along fibres more freely and quickly than the
conventional woven fabric. Although the composite has around 20% content in
matrix, the wave propagation velocity on Dyneema® UD strips was found not to be
44
affected by the matrix at all, which is very similar to that of dry fibres [74].
Therefore, the UD laminate is believed to have a higher energy absorption capacity
than that of woven fabric. According to ballistic test results, Karahan found that K-
Flex® panels absorbed 12.5–16.5% more energy at unit weight compared to Twaron®
woven fabric panels [62].

2.2.3 Impact conditions

It should be noted that the protection level of all body armour systems is only
meaningful for a certain impact condition. The ballistic performance of an armour
panel is greatly influenced by a series of impact conditions, such as the striking
velocity, types of projectile, panel size and the boundary condition.

2.2.3.1 Striking velocity and projectile geometry

During ballistic impact, Gray [75] analysed three regimes of the striking velocities.
In the regime-I at low striking velocities, material fails in brittle materials by linear
elastic fracture mechanisms. In the regime-II, the striking velocity approaches sonic
velocity (~500m/s) with the strain rates of 10 3 /s. The impact loading is associated
with the stress wave propagation. In the regime-III, the increasing striking velocity
would eventually result in the melting and fusion of the material at the impact region.
The related three principle failure modes to the striking velocity have been identified,
namely plastic deformation, brittle fracture, and melting [5].

At different striking velocities, the same fabric exhibited varied energy absorption
capacity. Generally, energy absorption in fabric usually reaches the maximum at the
ballistic limit and then sharply decreases with the continually increasing projectile
velocity. Shim [53] and Tan [48] concluded the mechanisms of energy absorption at
three distinct regimes of striking velocities as shown in Fig. 2.23. At the low striking
velocity, fabric has enough time to produce transverse deflection, thus allowing more
energy absorption in fabric. At the high striking velocity, fabric failed quickly, which
results in the localised damage and limited energy absorption in fabric.

45
Energy Absorbed by Fabric (J)

Impact Velocity (m/s)


Fig. 2.23 Three regimes of energy absorption for conical projectiles at different
impact velocity [48]

The geometry of a projectile has an influence on its pe netrating ability. Tan [48] and
Lim [55] studied the influence of different heads of the projectile (as shown in
Fig.2.24) on performance of Twaron fabric. They found perforation mechanisms are
highly dependent on the shape of the projectile. The conical and ogival projectiles
perforated fabric by slipping through the weave, which resulted in the lowest V50 of
58 and 76m/s. The flat head projectile, sheared yarns across the thickness with V50
of 100 m/s, while the hemispherical projectile produced the most yarn pull-out and
has the highest V50 of 159 m/s. Cunniff [44] observed that ballistic performance of
armour materials against Right Circular Cylinder (RCC) was significantly different
from that of the same systems against chisel- nosed fragment simulators. He
concluded that this could be due to an increase in shear stress in the region of the
impact point. However, with the increasing number of layers in a panel, the effect of
the projectile geometry becomes less significant [54, 55, 76].

Fig. 2.24 Different projectile geometries [48]


46
2.2.3.2 Panel size and boundary condition

In ballistic tests, the sample size and boundary condition are also important. Cunniff
[44] tested the ballistic limit of Kevlar and Spectra fabric with four types of sample
sizes (1-, 2-, 4-, and 8-inch) and found the ballistic limit velocity of fabric was highly
dependent on the sample size. The smaller sample size was related to the lower
ballistic limit. He analysed that was due to the transverse deflection of the fabric was
constrained in small size. However, the effect of sample size decreased rapidly when
the striking velocity exceeded the ballistic limit. Zhang [77] analysed by FE model
that fabric with smaller sample size had less energy absorption due to less material
involved. Sidney Chocron [78] observed numerically that the rarefaction wave in the
full-size (30.5cm) panel took longer time to reflect from the boundary compared to
the small sample (15.24cm). Mamivand [79] analysed that a certain value of the
sample size has very significant influence on ballistic limit. This value was varied
with different striking velocities for the same panel. He also found that the
performance of the panel can be underestimated with the reducing sample size in the
multi- layer system.

In ballistic perforation tests, the panel was clamped in a frame. The performance of
fabric was sensitive to boundary conditions in ballistic tests. Energy absorption in
fabric with two edges clamped was observed to be more than that of fabric with four
edges clamped as shown in Fig. 2.25. It was observed that the stress on yarns
increased rapidly when yarns ends were clamped. As a result, the fabric quickly
failed and less energy was absorbed [46,80].

Two clamped edges

Four clamped edges at 45 degrees

Four clamped edges at 0 degrees


Energy absorption (J)

Impact velocity (m/s)


Fig. 2.25 Energy absorption of fabric with different boundary conditions [80]
47
However, with the increasing striking velocity above a certain value (>300m/s), the
effect of the boundary conditions on the energy absorption was not obvio us. Energy
absorption of fabric was found to be a function of the clamping pressure. The
insufficient clamping pressure can result in fabric slippage from clamps. This can
contribute 4.5 times greater energy absorption than that of non-slippage case [47].

2.3 Ballistic response of the soft armour panel

In previous studies, some ballistic responses of a single layer fabric or UD laminate


have been identified. However, due to the interactions between layers in armour
panels, the ballistic responses of the multi- layer system became more complex. As a
result, the ballistic performances of the multi- layer system are not necessarily scale
with the performance of the individual layer.

2.3.1 Multi-layer panel with uniform layers

2.3.1.1 Interaction between layers

Cunniff experimentally observed that the layered system absorbed less ener gy than
spaced single layer system as shown in Fig. 2.26. He analysed that such deleterious
system effects of the panel was due to possible constraint of the transverse deflection
by subsequent layers on the front layers. As a result, the interference between layers
may prevent the stacked layers from achieving their individual energy absorption
capacities [44].

(a) the layered system (b) the spaced system


Fig. 2.26 Multi-layer systems (a) the layered system (b) the spaced system [81]

Mamivand [79] found that with the increasing layer gap in spaced system, the
percentage of perforated layers increased in the FE model of a panel. He analysed
that layers in the panel with large gap gradually came into contact with the projectile.
This resulted in the reduction of the reacting force. However, Lim [55] investigated
the double- layer system of Twarons CT716 and got different results. The energy
absorption of layered system was better than spaced systems for sharp-nosed
48
projectile. The reverse was true for hemispherical- nosed projectiles, and no
distinction difference for flat-nosed projectile. So he concluded that the ratio of
energy absorbed in the double-ply system to that of the single-ply system varies with
the striking velocity and projectile geometry.

2.3.1.2 Ballistic performance of the multi-layer armour panel

In general, in a single layer fabric, energy absorption was roughly proportional to the
fabric areal density [82]. In the multi- layer system, energy absorption of a panel will
definitely increase with more layers added. However, energy absorption of the panel
is less than the total amount of energy absorption of a single fabric layer with the
same number of layers [83]. Gogineni et.al observed that the growth rate of strain
energy absorption in fabric decreased with the increasing number of layers. [84].
Such results indicated that the contribution of each layer to total energy absorption of
a panel has been demonstrated not to be uniform.

Cunniff [85] observed that the material near the strike face had little influence on
ballistic performance due to quickly failure. Karahan [62] supposed most of total
energy absorption in armour panels was absorbed by the first layers of a pa nel in the
non-perforation case. Novotny et al. [83] concluded that with the increasing areal
density, the strain along the thickness of the panel decreased. As a result, less energy
absorption in back layers at the beginning of impact.

Joo and Kang [86] [87] employed a FE model of 4- layer woven panel at different
striking velocity to analyse ballistic behaviour of each layer. In the non-perforation
case, energy absorption of the first layer was the highest and followed by the
subsequent layers. This was due to more significant deformation of the first layer
than back layers. When perforation occurs, the sequence was reversed. In addition,
they found in the non-perforation case, yarn slippage and viscous deformation were
dominant mechanisms responding to energy absorption. In the perforation case,
inelastic collision was relatively significant.

Prosser [88] experimental analysed ballistic behaviour of nylon cloth system


impacted by fragment-simulating projectiles (FSP). He founded that the work of
penetration through each interior layer was essentially constant, which were different
from the strike layer and back layer. Zohdi [89] simulated the number of ballistic

49
fabric sheets needed to stop an incoming projectile. He found that energy absorption
in per sheet was controlled by the mode of deformation, which dictated the amount
of energy kinetically and elastic strain energy by the sheet. According to previous
studies, energy absorption of each layer was identified not to be uniform. However,
energy distribution in the multi- layer armour panel still needs to be further
investigated and qualified.

Ballistic resistance performance of armour panel in the non-perforation case has also
been studied by many researches. In Shen‟s study [90], test and simulation results
revealed three distinct loading phases during impacting. In the first phase up to about
60µs, the bullet transferred a major portion of its energy into the body armour with
significant deceleration. In the second phase, the armour kept on absorbing energy by
yarn stretching till about 100µs when energy absorption reached its peak. In the third
phase, the deformation of the armour panel continued to grow and energies inside
both armour and targets redistributed through wave propagation. The results
indicated that armour deformation and energy absorption in the second and third
phases were significantly affected by the material properties and geometrical
characteristics of the targets. To assess the transmitted energy in the clay, Park [12]
found that the indentation volume in the clay was proportional to the mass of the
impactor, while the energy absorption per unit indentation volume increased linearly
with the striking velocity. Lee et al. [72] also identified an empirical linear
relationship between the striking velocity (within a relatively narrow range) and BFS.
Such results indicated that BFS of the non-perforated panel was mainly determined
by the impact energy and energy absorption in the panel.

2.3.2 Hybrid armour panel

2.3.2.1 Hybrid materials and hybrid structures

Hybrid body armour as an innovation concept has attracted researchers for many
years. Many patents and commercial hybrid products have been proved to be very
efficient in providing superior ballistic performances and reductions in weight.
Karahan [62] experimentally observed that hybrid panels combining para-aramid
woven fabrics and K-Flex ® UD laminate can achieve around 4.5% reduction in BFS
and 8.5% improvement in energy absorption per unit weight compared to 100%
woven fabric panels. In Chabba‟s patent [91], a multi- layered panel was composed of
50
at least two sub-stacks of trauma reducing layers (such as UD) separated by at least
one fibrous layer (such as woven fabric or knitting, braiding, non-woven). The
decreasing BFS of the hybrid panel can reach 9.52% compared to the Dyneema ® UD
panel.

The components of different materials and the construction of hybrid panels played
important roles in ballistic performance. The Honeywell Company [92] designed 21-
layers hybrid armour panel with a sandwich structure by combining three Spectra
Shields®, two Gold Flex®, seven Kevlar®, four Spectra Shield® and five Gold Flex®.
It was claimed to provide the lighter weight, a greater range of threat protection, and
improved comfort, flexibility. The desired effect can be achieved by manipulat ing
the order of layering up and the number of layers. The similar sandwich design was
found in the patent of Zufle [93]. The panel combined ten Kevlar® layers between
two groups of eight Spectra Shield ® layers, and a single Kevlar® outer layer on the
top and beneath respectively. In addition, a reinforcing panel of 1.5mm made from
Lexan polycarbonate is placed against the last Kevlar layer. As a result, the back- face
deformation was reduced to almost a half compared with the normal armour.

2.3.2.2 Mechanisms responding to the hybridisation effect

Although many patents and products of the hybrid soft armour panel demonstrated
the positive effect of the hybridisation on ballistic performance of the multi- layer
system, the mechanisms responding to the hybridisation effect have not been
identified. Cunniff [44] experimentally studied the effect of layering sequence of a
two- layer system combining Kevlar® (low modulus) and Spectra® (high modulus)
woven fabric. The measured V50 of two panels with reverse layering sequence
showed some difference. He explained that the material with higher modulus had a
wider transverse deflection than that of the material with lower modulus. When the
high modulus material was placed on the striking face, the transverse deflection
would have been constrained by the second layer with more narrow transverse
deflection. Therefore, the fabric panel with Spectra ® (with high modulus) on the
striking face had a reduction of energy absorption by over 80%. Porwal and Phoenix
[94] developed a theoretical and numerical model to analyse above hybridisation
effect. The results showed that the interference between two layers had a significant

51
influence on the strain evolution in layers, particularly near the edge of the projectile
where failure initiates.

Park [95] observed the hybrid panel with components in the order of decreasing
modulus enhanced the penetration resistance, due to the less constraint force from the
back layers. While the hybrid panel with components in the order of increasing
modulus can obtain the enhanced trauma resistance. This was expla ined the stronger
yarns in back layers with high modulus can sustain longer, which resulted in more
impact energy dissipation. Such layering sequence of the hybridisation contributes to
the decreased BFS. Rahman [96] compared the contact force of a hybrid panel
combining with glass fabric and glass fabric laminates to panels with each
component. The numerical results showed that the contact force of the woven panel
was higher than that of the rigid panel. This results in more energy absorption in the
woven panel.

Most of these studies on hybrid armour panel just focus on the hybridisation effect.
The mechanisms relating to the hybridisation has not been fully understood. In
addition, the hybrid soft armour has not theoretical design concept.

2.4 Failure mechanisms of armour panels under ballistic impact

Failure modes are important aspects of ballistic response of an armour panel under
impact. Different materials and structures are related to different failure
characteristics. These have been investigated both at the microscopic level and
macroscopic level in many researches. The microscopic studies concentrate on the
microstructures of the polymer, which provides theoretica l interpretation for failure
modes at the macroscopic level. The macroscopic observation focuses on damage
characteristics or failure modes of yarns and fabric under ballistic impact.

2.4.1 Fracture of fibres in ballistic fabric

2.4.1.1 Fracture mechanisms of polymer fibres

At the microscopic level, polymer molecules are combined by primary bonds and
secondary bonds. The primary bond is the internal covalent bond holding individual
atoms together to form the polymer chains. The secondary bond refers to Van der
Waals forces and hydrogen bonds, which holds groups of polymer chains together to
52
form the molecular structure. [97] A bond breaks when it becomes excited beyond its
activation energy, which is dependent on circumstances (e.g. temperature) and
conditions (e.g. applied force).

The brittle fracture and the ductile fracture are two dominant failure modes for
polymer fibres. In brittle fracture, both primary and secondary bonds break, which
across the fracture surface perpendicular to the fibre axis. In this process, primary
bond breakage is usually more pronounced and there is little or no plastic flow. In
ductile fracture, the intermolecular slip is the dominant mechanism resulting in the
secondary bonds broken. This kind of failure modes usually requires the sufficient
mobility of the polymer chains to produce the plastic flow. Primary bonds breakage
needs more fracture energy compared to the intermolecular slip. In most cases,
mixed failure modes can be observed [98].

According to time-temperature equivalence principle, at high strain rates or below


the glass transition temperature of polymer, most polymers undergo brittle fracture.
In this case, both primary and secondary bonds are found to rupture. At low strain
rates (0.01s-1 ) and above the glass transition temperature, plastic deformation and
intermolecular slippage involving the rupture of secondary bonds occur in preference
to primary bond breakage [99]. Such failure mechanisms provide a good
interpretation for the changed mechanical properties of ballistic material with strain
rates. Due to insufficient time for plastic deformation and intermolecular slip, the
brittle fracture reduced the amount of energy absorption in the material at high strain
rates [32, 100].

2.4.1.2 Fracture mechanisms of ballistic fibres

Most high performance fibres applied in ballistic protection, such as aramid,


UHMWPE fibres, have tough molecular chains. Under tensile stress along fibre axis,
the filament fracture does not easily occur within the strong ordered chains at high
strain rate. Therefore, an alternative load transfer occurs through the amorphous
regions instead of the crystalline phases [101]. Hearle observed that, due to the shear
stress along the fibre axis, the breakage of intermolecular bonds resulted in the axial
splitting occurring on fibres as shown in Fig.2.27 [25]. The axial splitting is also
named fibrillation. It is the dominant morphology of high performance fibres with
high strength, high modulus and high toughness.
53
Fig. 2.27 Axial splitting of PPTA fibres [20]

Fig. 2.28 Thermal damage of Dyneema fibres [102]

Under ballistic impact, the thermal damage of UHMWPE fibres induced by the
generated heat in armour panels has been observed since the 1950s [39]. The
morphologies including fibre fusion, polymer bridge, and fibre melting have been
observed in many studies [54, 102, 103]. Most fracture UHMWPE fibres ends
exhibited a smooth, partly globular appearance as shown in Fig. 2.28, which
suggested the material underwent a phase of softening and plasticity flow [25]. It was
reported that the high performance polyethylene fibres would undergo a 100- fold
contraction when exposed to temperatures exceeding the melting point [104]. Prosser
observed the crystallinity changes on the impacted Spectra-1000 yarns by the
polarized light micrographs [54].

Languerand [20] investigated fracture of PPTA and UHMWPE fibre bundles at


varied strain rates. For PPTA fibres, fibrillation was always the main fracture
mechanisms at different strain-rates. While for UHMWPE fibres, crazing in
amorphous regions was more common at high strain-rate, which was more important

54
energy absorption mechanism than at quasi-static rates. As a result, UHMWPE fibres
appear to be able to dissipate more strain energy than PPTA fibres at high strain rate.
Koh [35] observed that the ductile failure of ballistic fibres, which displayed uneven
fracture surfaces and necking of tapered fibre ends at a low strain rate. With the
increasing strain rate, many broken filaments exhibited clean surfaces due to brittle
fracture.

2.4.2 Failure modes of armour panels

2.4.2.1 Woven panel

On the post-impact fabric, different damage characteristics of yarns can be observed,


such as yarns fracture, yarn pull-out and yarn bowing. Yarns fracture is one of the
dominant mechanisms of impact energy dissipation. Under ballistic impact, two
causes of yarns rupture are stretching of yarns along the length and shearing
perpendicular to the length as shown in Fig. 2.29 [55]. Tan [48] experimentally
investigated that the energy dissipated through the cutting action was much lower
than that of yarn stretching to breakage.
Yarn shearing

Yarn stretching

Fig. 2.29 Two forces applied on yarns under ballistic impact

Due to the inter-yarn frictional force, some yarns are pulled out of fabric without
fibres fracture. The yarn decrimping and subsequent yarn translations are two main
mechanisms responding to impact energy dissipation [63]. In particular, at low
striking velocity, in a small panel with free boundary conditions, the energy
dissipation due to yarn pull-out was found more significant [105].

Yarn bowing refers to the phenomenon where warp yarns become non-orthogonal to
weft yarns. It can be caused by two mechanisms. First, the projectile pushes yarns
aside during the passage through the fabric. Second, the stress wave propagation
causes stretched yarns to displace from crossovers of the weave [48]. Lim et al.
observed that bowing was more obvious in the back layer of the double-ply system
due to the quickly decelerated striking velocity [55]. At high striking velocities,
bowing was observed to be more localised for all types of projectile [48].
55
Under ballistic impact, fabric layers in an armour panel experienced different failure
mechanisms due to varied impact force in different positions. Prosser concluded the
cutting or shearing action was the prime mode of penetration in multi- layer panels by
an FSP or RCC projectile [54, 88, 106]. Due to the force required for shearing failure
of yarns was less than that of by tensile failure, the projectile will choose the easiest
mechanism of penetration. Chen and Zhu [107] used an analytical model of ballistic
impact on woven panel to analyse the transition of damage mechanisms from shear
to tensile failure in layers from front to back.

2.4.2.2 UHMWPE UD panel

Due to the special structure, failure modes of UD laminate are different from woven
fabric and traditional structural composites. In the thin UD laminate, the observed
failure modes included fibre stretching, fibre pull-out, sequential delamination, cut-
out of a plug, and laminate deformation.

When UD laminate is under impact, fibres on the striking face were driven into the
underlying plies before they fail. Scott [47] reported that fibre stretching was the
dominant mode of energy absorption in the UD laminate, even in the first few plies
within compliant laminates. Beneath the projectile, the material was compressed and
absorbed the remaining energy through fibre elongation and fibre pull out [108, 109].
Some studies also found that the projectile penetrated the thin laminate by moving
fibres laterally or through fibre pull-out [110, 100]. The sequential delamination and
cut-out of a plug was the typical damage modes of the composite, which were also
found in the UD laminate. Such failure modes induced by the shear perpendicular to
the thickness of the laminate combined with the mode of tensile failure of fibres
[100].

Some studies discussed the global response of UHMWPE composites during ballistic
penetration. Such work has identified the dominance of shear failure on the striking
face of a panel and membrane loading with tensile failed fibres towards the exiting
face [39, 108, 111]. Critescu [110] concluded there was a similar failure mechanism
for Dyneema UD panel under impact as the traditional structure composite. The
fibres experienced the initial impact failed in shear or cutting at the edge of the
projectile. While in the back layers of the panel, the mode closely resembled tensile
failure.
56
Due to the low melting point of UHMWPE fibres, investigations of failure modes of
UD panel under ballistic impact have mainly focused on the thermal damage. The
generated heat in a local impact region was identified in many studies, although the
exact temperature of the heat generated on the fabric was hard to be measured. It was
estimated that the temperature in the local region of a fibre can exceed the melting
point of UHMWPE [74, 112]. Prevorsek analysed Spectra composite under impact
and concluded that the temperature can reach 330°C, but only in a very small region
around the interface, due to the short interaction time and the low thermal
conductivity of Spectra fibre [104]. Koh and Shim observed that the generated heat
was not evenly distributed on a fibre, because individual filaments were not equally
stretched at high strain-rate. As a result, the localised temperature in “hot spots” can
increase sufficient to cause a strength decrease of the fibre [35]. The reason of the
generated heat was commonly analysed as the result of the friction between fibres in
armour panel and projectile [54, 104].

2.5 Analysis techniques

In previous studies, some different techniques have been used to analyse the ballistic
response of soft armour panel, mainly including experimental, numerical analytical
methods and so on. These techniques were complementary to provide a thorough
investigation of ballistic response on the armour panel.

2.5.1 Experimental techniques

Experimental studies include experimental work and analysis of data obtained


through experiment tests. With advancements in photography, microscopy,
measurement, and material testing systems, the complex phenomena during the
impact process can be captured [64]. Experimental studies not only directly reflect
some ballistic responses and performance of armour panels, but also provide an
important source of data for theoretical analysis.

In previous researches, ballistic tests were conducted to reflect ballistic performance


of armour panels [44, 48, 52, 62, 95]. By the high-speed digital cameras, the dynamic
ballistic response of armour panels can be observed and quantified [28, 74, 62, 113].

57
In addition, mechanical properties were measured and failure modes of materials
were observed. Mechanical properties of fibres, yarns and fabric are most important
material properties for ballistic protection, especially at high strain rates. A split
Hopkinson bar apparatus and impact tensile testing apparatus are most commonly
used to determine these properties [27, 28, 32, 35]. Yarn pull-out testing was often
performed to determine the fabric‟s frictional characteristics, which can be conducted
under quasi-static conditions [63]. Photographic investigation was usually adopted to
observe failure modes of post- impact panels. This has been conducted by
microscopic, Scanning Electronic Microscopic (SEM), Atomic Force Microscopy
(AFM) [20, 102, 114].

Up till now, a great deal of experimental studies contributes to understanding the


ballistic behaviour of armour panels. Such experimental programs are critical for
ensuring the utility and effectiveness of the armour systems. However, they are
generally expensive, time-consuming, and involve destructive testing [115].

2.5.2 Finite Element Models techniques

To take insight into the ballistic impact behaviour of the armour panel and reduce the
dependent on the experimental tests, the computation-based engineering analyses and
simulations were employed in the ballistic researches. Numerous FE models have
been employed in previous studies.

The three-dimensional continuum Finite Element Model (FEM) has been


demonstrated to be powerful for capturing and elucidating in detail dynamic ballistic
responses of single- layer fabrics [116]. FE model can be created at yarn level as
shown in Fig.2.30 (a), which was used in many studies [77, 46, 117,118]. A single
yarn is simulated as a solid continuum rod with a certain crimp wave and some shape
of cross-section. These 3D FE models can take account for the interaction between
yarns during impact. However, this approach is limited by the computing power.
When the armour systems contain more fabric layers, such as 30-50, these FE models
were found to be computationally very expensive and became unstable with the
increasing number of layers [64].

To reduce the computation expense, the pin-joined model and the shell or membrane
model are widely used for modelling the multiply system. In the pin-joint model, the

58
fabric is assumed to be two-dimensional assembly of pin- jointed orthogonal bars at
the crossover of the weave as shown in Fig. 2.30 (b). Multi- layer fabrics can be
modelled by stacking layers of masses and strings. The pin-jointed FE model has
proved to be very efficient in simulating the dynamic behaviour of woven fabric. But
the discrete nature of the yarn model is associated with inherent oversimplifications,
which cannot be accounted for the weave structure, yarn pullout, yarn crimp and
inter- yarn friction. This work includes Rolance [45], Shim [53], Ting [119], Zeng
[80], Novotny and Cepus [83], Joo and Kang [87, 89, 113].

In membrane/shell FE model, the fabric is discretized using membrane/shell


elements to represent the dynamic response of the fabric under impact as shown in
Fig.2.30 (c) [120, 121, 115]. Compared to a solid FE model and a membrane model,
the computational efficiency of the membrane FE model of a single-ply plain-woven
fabric was found to increase significantly [115]. However, it cannot adequately
capture the impact response along the fabric thickness.

(a) 3D solid FE mode [46] (b) Pin-joint FE model [82]

(c) Shell FE model [114]


Fig. 2.30 FE model of fabric under impact (a) 3D solid FE mode[46], (b) Pin-
joint FE model [83],(c) Shell FE model [115]

59
No one FE model proposed so far can account for all complex phenomena involved
in penetration mechanisms during ballistic impact. There is always some
compromise in terms of the modelling accuracy. Therefore, all FE models are the
balance between the performance characteristics and material properties [5].

2.5.3 Analytical techniques

Analytical methods make use of general continuum mechanics equations and laws,
such as the conservation of energy and momentum. Governing equations are set up
using various parameters involved in impact process. This includes work by Parga-
Landa [122], Gu [28], Phoenix and Porwal [94, 121], Zohid [89], Zhu and Chen
[107], and Naik et al. [21] .

Analytical models in the mechanics of high-speed penetration can be classified as


empirical, semi- or quasi-empirical, engineering, simplified, analytical, semi-
analytical and approximate [123]. The simplified models in high speed penetration
mechanics was used widely by the number of studies. Approximate models are also
important because they can stimulate new engineering so lutions and indicate
directions for further theoretical and experimental investigations.

Analytical methods are useful to handle simple physical phenomena due to their
direct applicability. The calculations based on such models do not require large
computer resources. However, the analytical models will become increasingly
complicated as the phenomena become more complex and involve many variables.

2.6 Concluding remarks

Since the introduction of high performance fibres in the 1960s, considerable


researches into the ballistic impact of soft armour panel has grown rapidly. Up till
now, although a thorough, quantitative understanding of all mechanisms does not yet
exist, some fundamental understanding of impact response of woven fabric under
impact has been identified. It has been realised that the performance of armour
panels to be a function of many factors in a coupled manner. It is difficult to isolate
each mechanism. The advanced of photographic observations equipment improved
the understanding of ballistic behaviour of armour panels and related failure
characteristics. With the development of numerical simulation techniques, the

60
internal nature of ballistic impact, such as the stress wave propagation, transverse
deflection, and different forms of energy dissipation, can be reflected and quantified.
This makes it possible to analyse the complex interaction between yarns and fabric
layers in armour panels.

According to previous studies, some complicated ballistic responses and the related
mechanisms of the multi- layer system under ballistic impact still need to be further
investigated. In previous studies, it has been demonstrated the ballistic resistance of
each layer in an armour panel is not to be the same. However, the contribution of
each layer to energy absorption has not been quantified. The way of energy
absorption in the multi- layer system needs to be exactly identified. Such fundamental
understanding would provide a guide for optimising the construction of traditional
armour panels.

It has already been identified that the ballistic performance of soft armour panel is
greatly dependent on material properties. However, there is no material that can
provide all the required properties for ballistic resistance. To make the best use of
advantages and avoid weak points of materials properties, different current available
materials should be utilized to obtain the most efficient performance of armour
panels.

Currently, hybridisation is one of the most popular innovative approaches to improve


the ballistic performance of soft body armour. Although the improved ballistic
performance of hybrid soft armour panel has been proved in many products, the
related mechanisms have never been identified. Most of these studies just focused on
the hybridisation effect. The mechanisms responding to hybridisation still need to be
thoroughly investigated. This can provide an effective guide for the hybrid design of
armour panels.

Many studies analysed failure mechanisms of soft body armour under ballistic
impact. Most of these studies only focused on local failure modes of fibres and yarns
in some individual layer in a panel. None of these studies have gone into depth as to
the reationship between failure modes of each layer and the relative contribution to
energy absorption. Understanding the partition of energy between failure modes is a
key to develop simple and effective ballistic design algorithms [42]. The global

61
failure modes of the multi- layer system still need to be thoroughly investigated by
identifying the difference of failure modes between layers.

According to previous studies, in order to achieve further improvement of ballistic


performance with reduced weights with current available materials, optimising the
construction of armour panels is an efficient and feasible option. Different materials
can be combined into the most effective positions of an armour panel. In this way,
the hybrid panel can make full use of different materials properties. Therefore,
understanding mechanisms responding to ballistic impact on the multi- layer system
is the key issue to develop the effective construction of armour panels.

62
CHAPTER 3 Experimental studies of ballistic armour
panels

To thoroughly investigate the ballistic response of the multi- layer system, two
complementary research approaches were adopted in this study, namely the empirical
method and the Finite Element (FE) analysis. This chapter describes experimental
method, which mainly focuses on ballistic tests and photographic observations of
post-impacted panels. This method directly reflects ballistic performance and failure
modes of armour panels under impact. The test results are discussed in great details
in Chapter 5, Chapter 6, and Chapter 7.

In this study, Twaron fabric and Dyneema UD laminates, two of the most popular
ballistic materials are used as components for the hybrid panel. The mechanical
properties of fibres, fabric and UD laminate are measured, which are described in
Section 3.1 at first. And then, to identify the ballistic performance of different panels
composed of these two components, a series of samples for ballistic tests are
prepared. The specifications of samples are described in Section 3.2. Ballistic tests
including perforation and non-perforation cases are conducted at the ballistic
laboratory in University of Manchester. The test equipment and test procedure are
introduced in Section 3.3. And then, section 3.4 describes photographic observations
of the impact process and post- impact panels. A summary is presented in the last
section 3.5.

3.1 Materials specifications and mechanical properties

Twaron® woven fabric and Dyneema® UD laminate are two of the most popular
ballistic materials for soft body armour. Due to their superior material properties with
high strength and high moduli, both of them possess excellent ballistic resistant
capacities.

Twaron woven fabric is one of the typical traditional ballistic materials, which has
been commonly used for the armour vest in many modern types of ballistic body
armours. It dissipates impact energy over a large area of an armour panel by fabric
deformation and yarn stretching. Therefore, Twaron woven fabric exhibits high
energy absorption during ballistic impact. In addition, woven fabric has advantages

63
of lighter weight and better flexibility compared with the metal and composite, which
can ensure the wearer‟s mobility and comfort.

Dyneema UD, a kind of compliant composite, is another new type of ballistic


material. The structure of UD laminate is different from conventional woven fabrics.
The typical laminate structure is characterized by placing Dyneema fibres parallel at
0/90°in each cross ply with little resin fraction. Without interlacing between fibres,
the stress wave can be quickly dissipated widely. So the impact energy can be
dissipated by the effective fibre stretching. In addition, Dyneema fibres have the
lowest density of all fibres, which is beneficial in providing lighter weights for
ballistic armour panel. Therefore, it has been increasingly used in ballistic protection
in recent years.

In this study, these two materials both were used as components of hybrid panels.
Several types of Twaron woven fabrics with different structures were designed and
engineered. In addition, two types of Dyneema UD laminates were used. Materials
specifications and mechanical properties were investigated firstly.

3.1.1 Materials specifications

3.1.1.1 Twaron yarns

To investigate the effect of yarn count, three types of Twaron® yarns were adopted,
including the fine yarn of 93tex, middle yarn of 110tex and thick yarn of 168tex.
These yarns were provided by Teijin® Company. Tensile properties of yarns were
provided to product properties of Teijin®, which were listed in Table 3.1. Taking
Twaron yarn 93tex as an example, the tensile c urve indicated that Twaron yarns
possess the linear elastic material properties of as shown in Fig. 3.1.

Table 3.1 Mechanical properties of Twaron® yarns

Yarn Strength Tenacity Young’s


Elongation
count Filament Twaron type at break at break Modulus
(%)
(tex) (N) (N/tex) (GPa)
2040
93 1000 225 2.35 3.45 89
microfilament
2040
110 1000 267 2.35 3.45 91
microfilament
168 1000 2000 or 2042 385 2.35 3.45 91

64
Fig. 3.1 Tensile curve of Twaron yarn (93tex)

3.1.1.2 Twaron fabrics

In this study, nine types of Twaron plain woven fabric were used to investigate the
ballistic performance. The specifications were listed in Table 3.2. The sample label
contains one number and a letter. The number represented the weave density. The
letter means the yarn count. „F‟ is the fine yarn of 93tex. „M‟ is the middle yarn
count of 110tex. „T‟ is the thick yarn of 168tex.

Table 3.2 Specifications of Twaron woven fabric

Warp/weft Warp/weft Crimp Areal


Sample Thickness
Material yarn count weave density ratio density
label (mm)
(tex) (ends/cm) (%) (g/m2 )
8F 93 8.3 0.20 0.41 155.35
9F 93 9.3 0.22 0.83 186.42
10F 93 10.4 0.25 1.36 202.08
11F 93 10.9 0.26 1.52 196.85
Twaron
woven 12F 93 11.6 0.30 1.71 227.95
fabric
13F 93 12.6 0.32 2.16 251.76
7M 110 6.5 0.20 1.07 153.25
10M 110 10.7 0.34 1.79 259.28
7T 168 7.0 0.30 0.77 260.41

65
Among them, two kinds of commercial ballistic fabric 11F and 7M were
manufactured by Teijin®. Other seven types of fabrics with above three types of yarn
counts and varied weave densities were designed and manufactured in the weaving
lab of University of Manchester.

According to previous studies [51], the use of fine yarn improves the V50 of fabric
for ballistic protection. Therefore, to investigate the effect of weave density, the fine
yarn of 93tex was used to weave five different types of fabrics with varied weave
densities. The weave density was gradually increased by 1 end/cm start from
8.3ends/cm. During the manufacture, the actual weave density exhibited a little
inevitable difference from the designed value. The actual structure specification of
these five fabrics 8F, 9F, 10F, 12F and 13F were listed in Table 3.2. The middle yarn
110tex and the thick yarn 168tex were used to weave two fabric 10M and 7T. The
ballistic performance of these two types of fabric was also investigated for
comparison. Fabrics specifications were measured including the thickness, areal
density, cover factor, fabric tightness and crimp ratio.

1. Thickness and areal density

Woven fabric is formed by interlacing warp and weft yarns each other. Due to loose
filaments in yarns, the visual thickness of the fabric is much larger than the
compressed thickness under pressure (Appendix 1).

Taking the fabric 8F as an example, the visual thickness of fabric was measured
under an optical microscope (Leica DFC295). The fabric was made into the resin
slice and the cross-section was observed as shown in Fig. 3.2. For Twaron yarns,
filaments are assembled with little twists. The cross-section of yarns exhibited a
lenticular shape under tension between warp and weft yarns. The visual thickness of
8F was measured as 0.25mm.

Visual
thickness
0.25mm

Width of cross-
section 0.1134mm

Fig. 3.2 The cross-section of yarns in Twaron woven fabric 8F

66
Under a specific pressure, the fabric thickness decreases correspondingly. The nine
fabric samples were measured by callipers. The pressure was applied by hand to
compress the fabric tightly. The thickness values were related to the compressed
fabrics with less porosity in yarns, which were listed in Table 3.2. According to
measurement results, the thickness of 8F under pressure was measured as 0.20mm,
which is smaller than that of the visual thickness.

The areal density of all types of fabrics was measured on an electronic balance. The
average areal density of fabric was determined by five measurements that were
obtained on different regions of fabrics. The areal density of fabric samples were
listed in Table 3.2.

(b) Fabric tightness

In this study, five types of plain woven fabrics of 8F, 9F, 10F, 12F and 13F with yarn
count of 93tex were used to investigate the influence of weave density on ballistic
performance. With the increasing weave densities of fabrics, the weave structures
become tighter. To quantify the exact difference of weave structures in these five
fabrics, the fabric tightness were analysed.

For common fabric with twist yarns, the cover factor is usually used to describe the
fabric tightness. It is a calculated value relating to the geometry of the weave and
indicates the percentage of the gross surface area of a fabric that is covered by yarns
[56]. As twist yarns have round cross-sections, the cover factor can be easily
calculated according to the diameter of yarns and weave density of fabric. The
equations used to calculate the cover factor are showed as following [56].

(3-1)

(3-2)

(3-3)

( ) (3-4)

67
where is the width of warp or fill yarn in the fabric, is the weave density
of warp or fill yarns (ends per unit length), is the cover factor in warp or fill
yarns. is the cover factor of fabric. and is the area of fabric which is
covered by yarns and by gap between yarns. is the total area of
fabric.

However, due to the low twist of Twaron yarns used in soft body armour, filaments
in warp and weft yarns spread out to close gaps in the fabric. It is difficult to
calculate the diameter of a yarn. In this study, according to the proportion of the gap
area in fabric, the fabric cover factor can be approximately estimated as shown in
equations (3-3) and (3-4).

The size of gaps in five fabrics were measured by an optical microscope (Projectina
CCD-1300PQC) and listed in Table 3.3. In five fabrics, gaps between yarns decrease
obviously with the increasing weave densities as shown in Fig. 3.3. The cover factors
of yarns in five fabrics were all above 97%, which cannot exactly reflect the
difference between these five fabrics.

L=648.11µm

L=648.11µm

(a) 8F

(b) 9F (c) 10F

68
(d) 12F (e) 13F
Fig. 3.3 Gaps in five Twaron fabrics (a) 8F, (b) 9F, (c) 10F, (d) 12F, (e) 13F

During the Twaron fabric manufacture, the fabric tightness (K) is used to describe
weave structures. It is calculated accord ing to the yarn count and the weave density
of fabric as shown in equations (3-5) and (3-6) [145]. The fabric tightness K is
usually limited to less than 30 in the manufacture. If K exceeds 30, warps may be
squeezed together resulting in weave structure jams. Although there is no minimum
value for K, an extremely loose fabric cannot be manufactured. When the weave
density of fabrics decreases to a certain degree, warp yarns will produce distortion in
fabric formation. Therefore, for the fabric with yarn count of 93tex, when the K
reaches 30, the maximum weave density that can be engineered is 15.5ends/cm. For
the five fabrics used in this study, the fabric tightness was listed in Table 3.3. In this
way, fabric weave structures were quantified.

√ ×0.1 (3-5)

(3-6)

where K warp/weft is the fabric tightness along warp or weft direction; P is the weave
density (ends/cm); T is the yarn count (tex). K fabric is the total fabric tightness.
Table 3.3 Weave structures in different fabrics

Length of Width of Cover factor Fabric


Samples
gaps (µm) gaps (µm) (%) tightness
8F 134.54 76.31 97.56~100 16.01
9F 80.35 52.22 98.40~100 17.94
10F 52.22 13.05 99.31~100 20.06
12F 52.21 / ~100 22.37
13F 51.25 / ~100 24.30
69
(c ) Crimp ratio

Weave structures of fabric determine the crimp ratio of yarns. In this study, the effect
of crimp ratio on energy absorption of fabric was also investigated. According to
ASTM standard D3883 [139], crimp ratios of yarns in warp and weft directions were
calculated according to equation (3-7). The length of the crimp yarn in the fabric was
measured as L. And then the yarn was pulled out from the fabric and the straight
length was measured as P. The warp and weft crimp ratios were obtained by five
repeat measurements respectively. Due to different tensions of warp and weft yarns
during the weaving process, warp and weft crimp ratio exhibited a little difference.
The average value of warp and weft crimp ratios is the crimp ratio of fabric, which
was adopted in this study to simplify the analysis as shown in Table 3.2.

Crimp ratio= (3-7)

where P is the length of straight yarn pulled out of fabric (cm); L is the length of
crimp yarn in fabric (cm).

3.1.1.3 Dyneema UD

In this study, Dyneema UD laminate was used as a component for the hybrid panels.
Therefore, the material properties and ballistic response of Dyneema UD were
investigated in this chapter. Two types of Dyneema UD laminates with different
areal densities, Dyneema SB71 and Dyneema SB51, were analysed here, which were
provided by DSM®. A single layer laminate of Dyneema SB71 and Dyneema SB51
consists of four and six plies of uni-directional fibres oriented at 0o /90o respectively.
The specifications were listed in Table 3.4. The sample label contained a number,
short for product code of Dyneema UD, and a letter „U‟, short for UD.

Table 3.4 Specifications of Dyneema UD panels


The Areal
Sample Thickness
Material number of density
label (mm)
plies (g/m2 )
Dyneema
7U 6 0.24 186.94
UD SB71
Dyneema
5U 4 0.30 252.98
UD SB51

70
3.1.2 Mechanical properties

Mechanical properties of Twaron woven fabric and Dyneema UD laminate are


different from that of filaments due to the structure effect. It is crucial to determine
their mechanical properties for investigating the ballistic response and many other
impact behaviours of the multi- layer panel. In addition, these principle properties are
also required for FE numerical analysis of multi- layer panels under ballistic impact.

In this study, Twaron woven fabric 11F and Dyneema UD SB71 were conducted
tensile test and shear frame test to determine some tensile properties and shear
properties, which were conducted on a universal materials tester (Instron 4411) in the
laboratory at the University of Manchester.

3.1.2.1 Tensile test

Twaron yarns and Dyneema UD both have lower friction coefficient. During the
tensile test, inadequate gripping force will result in the severe slippage of the
specimen from the clamps. While excessive gripping force can lead to specimen
breakage at clamps due to stress concentration. In previous studies, non-standard
spciemen were adopted to avoid such problems in tensile test. In Shim‟s test of
Twaron CT 716 fabric [32], the specimen with 5 mm in width and a gauge length of
30 mm were used to conduct split Hopkinson tension. In Russell‟s [34] and Zok‟s
[124] tests of Dyneema laminate, a specimen with large grip region (100×60mm) and
small gauge section (50×4mm) was designed and used to avoid slippage.

In this study, to avoid slippage of specimen on the outer-sides with adequate gripping
force, the specimen was cut with the shape as illustrated in Fig. 3.4. According to
ASTM Standards D3039[125], the gauge length was 250mm. The effective width of
samples reduced to 12.5mm to ensure the effective grip.

A 0 =12.5mm

L0 =250mm
Fig. 3.4 The specimen of Dyneema UD SB71 for tension test

To avoid the severe slippage of the specimen from clamps, the capstan gripping
method was adopted in the tensile test. Two ends of the sample were wrapped around
two wooden rods with square cross-section outside two clamps as shown in Fig. 3.5.
71
The constant tensile speed of 200mm/min was used to provide the strain rate of 10-2 .
From the fracture configuration of the failed specimen as shown in Fig. 3.4, the
tension test was successful. Five repeat tests were conducted to determine the
mechanical properties.

Wooden
rods

F Clamp
s
Specimen

Fig. 3.5 Tensile test

According to the load and displacement of the specimen in the tensile test, the
responding stress and strain on the specimen can be converted by equations (3-8) and
(3-9). The tensile stress-strain curve of Dyneema UD was plotted as an example as
shown in Fig.3.6. The Young‟s Modulus was the slope of the linear part of the tensile
curve. Five repeat tests were conducted to obtain the average value of the stress,
strain and modulus, which was listed in Table 3.5.

(3-8)

(3-9)

where σ is the normal stress (GPa), F is the load on the specimen (KN), s is the
cross-section area of specimen (mm2 ), ε is the normal strain, ∆l is the displacement
(mm), L is the length of specimen (mm).
1

0.8
Stress (GPa)

0.6

0.4 Young‟s
Modulus
0.2

0
0 0.02 0.04 0.06 0.08 0.1
Strain (%)
Fig. 3.6 Tensile stress-strain curve of Dyneema UD
72
3.1.2.2 Shear frame test

Some shear properties of Twaron fabric and Dyneema UD laminates were obtained
by the shear frame test. Shear frame test, also named picture frame test, performs in-
plane shear deformation to specimen by applying a uniaxial tensile force at two
diagonally opposite corners of a frame. At small displacement, the deformation of
specimen in the plane was close to a pure shear case (shown in Fig.3.7).
F

Fig. 3.7 Pure shear deformation


The specimen with the effective size of 130×130mm was clamped within a square
shear frame as shown in Fig. 3.8 (a). The shear frame was installed on an Instron
4411. The constant tensile speed was 20mm/min [126]. Two opposite corners of the
frame were applied a displacement to induce shear deformation of fabric as shown in
Fig. 3.9. The shear angle α of the fabric is related to the displacement of the machine
∆l. When the fabric has a small deformation, the shear strain is approximately equal
to the tanα. According to the applied load and displacement of the specimen as
shown in Fig. 3.8 (b), the shear stress τ and shear strain γ can be obtained. The shear
stress-strain curve of Dyneema UD was plotted as an example as shown in Fig. 3.10.
Five repeat tests were conducted to obtain the average value as shown in Table 3.5.

F 1
l

α/2

L θ α/2

(a) The shear frame (b) Shear deformation


Fig. 3.8 Shear deformation of the specimen (a) The shear frame (b) Shear
deformation
73
Fig. 3.9 Shear frame test
0.7
0.6
Shear Stress (MPa)

0.5
0.4
0.3
0.2
0.1
0
0 0.1 0.2 0.3 0.4 0.5
Shear strain
Fig. 3.10 Shear tensile curve of Dyneema UD

2L  l
  acr cos( ) (3-10)
2L

   / 2  2 (3-11)

F
 12 
2 L  t  cos  (3-12)

  ta n 
(3-13)

where F is the load on the specimen, ∆l is the displacement of specimen, θ is the


fabric angle, α is the shear angle, t is the thickness of specimen, L is the length of
specimen side; τ is the shear stress, γ is the shear strain.

74
Table 3.5 Material properties of Twaron 11F and Dyneema UD SB71

Material σ (GPa) ε (%) E11 (GPa) E22 (GPa) G12 (MPa)

Twaron 11F 0.78  0.04 3.23  0.28 18.6  1.25 18.6  1.25 7.45  0.95

Dyneema 7U 0.91  0.08 6.33  0.25 38.8  1.34 38.8  1.34 69.5  0.58

Compared with previous studies test results [32-34], the failure stress and the
modulus of Twaron fabric and Dyneema UD that were tested in this study were
found to be lower and the failure strain larger. This was due to some inevitable
slippage of the specimen from clamps during tests. In particular for Twaron woven
fabric, it was difficult to apply the uniform gripping force on every yarn along the
direction of applying force, which results in yarns slippage out of clamps one by one
from the outer-sides and fractured at different times. However, the overall trend of
the stress-strain curve can reflect different properties characteristics of materials. The
failure stress and modulus of Twaron fabric were found lower than those of
Dyneema UD, which is consistent with the results in above previous studies [32-34].

3.2 Ballistic panels preparation

In this study, to optimise the construction of an armour panel, Twaron woven fabric
and Dyneema UD were used as components of the hybrid panel. A series of panel
samples were prepared for perforation and non-perforation ballistic tests to identify
ballistic performance, including woven panels, Dyneema UD panels and hybrid
panels.

3.2.1 Woven panels

3.2.1.1 Panels with a single layer fabric

To identify the effect of weave structures on ballistic performance of fabric, nine


types fabric of a single layer which were listed in Table 3.2 were conducted
perforation ballistic tests. The effect of the areal densityon energy absorption will be
analysed among nine types of fabrics. The effect of weave density and crimp ratio
will be investigated among five types of fabrics (8F, 9F, 10F, 12F, and 13F) with
yarn count of 93tex. Two types of fabric 10F and 10M were analysed for the effect of
yarn count.
75
3.2.1.2 Perforated multi-layer woven panels

To investigate ballistic performance of multi- layer woven panels, Twaron panels


composed of 8F and 11F with different layers were constructed to conduct
perforation ballistic tests. The specification of each sample was listed in Table 3.6. In
multi- layer panels, the subscript represents the number of layers.

Table 3.6 Perforated panels with varied number of layers


Areal density
Material Samples The number of layers
(g/m2 )
8F 1 155.35
8F3 3 466.05
8F6 6 932.10
8F9 9 1398.15
Twaron 11F 1 196.85
11F3 3 590.55
11F6 6 1181.10
11F9 9 1771.65
11F15 15 2952.75

To explore the effective construction of woven panels, two systems of multi- layer
panels with different constructions were designed at the close areal density as shown
in Table 3.7. The first system composes of two fabric panels with the same yarn
count and different weave densities namely 8F4 (93tex, 8.3 ends/cm) and 11F3 (93tex,
10.9 ends/cm). The second system includes two fabric panels 11F4 (93tex, 10.9
ends/cm) and 10M3 (110tex, 10.7 ends/cm) with different yarn counts and almost the
close weave density.

Table 3.7 Woven panels with different constructions

Material Sample label The number of layers Areal density (g/m2 )


8F4 4 621.40
Twaron 11F3 3 590.55
11F4 4 787.40
10M3 3 777.84

3.2.1.3 Non-perforated woven panels

To investigate ballistic performance of the non-perforated woven panel, several


fabric panels with more layers were constructed. Ballistic performance of BFS, the

76
number of perforated layers and the indentation volume can be analysed according to
ballistic results. The specification of each panel was listed in Table 3.8. For the non-
perforated panel, the woven panel with 11F was taken as the reference panel.

Table 3.8 Non-perforated woven panels with increasing layers


Materials Sample label The number of layers Areal density (g/m2 )

8F24 24 3728.40
11F19 19 3740.15
11F24 24 4724.40
Twaron
11F36 36 7086.60
11F48 48 9448.80
11F120 120 23622.00

3.2.2 UD panels

Dyneema UD is another one component of hybrid panels in this study. The ballistic
performance of Dyneema UD panels was also investigated in the perforation and
non-perforation case.

Table 3.9 Perforated Dyneema UD panels

Materials Sample label The number of layers Areal density (g/m2 )

7U 1 186.94
7U3 3 560.82
7U6 6 1121.64

Dyneema 7U9 9 1682.46


UD 5U 1 252.98
5U3 3 758.94
5U6 6 1517.88
5U9 9 2276.82

The multi- layer systems composed of few layers of 7U and 5 U were conducted
ballistic perforation tests to analyse performance of energy absorption. The non-
perforated Dyneema UD panels were constructed with even more layers to analyse
performance of BFS and indentation volume. The specifications of perforated and
non-perforated Dyneema UD panels were listed in Table 3.9 and Table 3.10.

77
Table 3.10 Non-perforated Dyneema UD panels
The number of
Materials Sample label Areal density (g/m2 )
layers
Dyneema 7U20 20 3738.80
UD 7U24 24 4486.56

3.2.3 Hybrid panels

To investigate the effect of hybridisation on ballistic performance, different


components, including Twaron woven fabrics with different weave structures (8F,
9Fm 13F and 7T) and Dyneema UD 7U, were combined in different manners. The
specification of each component was listed in Table 3.11.

Table 3.11 Specification of each component for hybridisation

Yarn count Weave density Areal density


Sample label Graphic
(tex) (ends/cm) (g/m2 )
8F 93 8.3 155.35
11F 93 10.9 196.85
13F 93 12.6 251.76
7T 168 6.7 260.41
7U / / 186.94

3.2.3.1 Different layering sequences of components

To investigate the effect of layering sequence on ballistic performance, different


perforated and non-perforated hybrid panels were constructed. The sample label of
hybrid panels contains sample labels of each component and the subscript number
which represents the number of layers. Each component was listed in the layering
sequence which is from the striking face to the exiting face and separated by the „/‟.
Hybrid structures panels and hybrid materials panels were prepared for perforation
ballistic tests. Woven fabric 8F and 11F were combined in reverse orders to
investigate the effect of hybrid structures. Dyneema 7U combining with 8F and 7T in
reverse orders were analysed the effect of hybrid material.

78
Table 3.12 Perforated hybrid panels with different layering sequences

Layering up The number Areal density


Sample
sequence of layers (g/m2 )

8F/11F 2 352.35

11F/8F 2 352.35

7U /8F 2 342.13

8F/7U 2 342.13

7U3 /8F3 6 1029.69

8F3 /7U3 6 1029.69

7U/7T 2 447.94

7T/7U 2 447.94

7U3 /7T3 6 1342.23

7T3 /7U3 6 1342.23

To identify the proper position of UD layers in hybrid panels, eight fabric layers of
8F and sixteen layers of Dyneema 7U were combined in four different layering
sequences. Dyneema UD layers were placed on the striking face, on the exiting face,
between fabric layers and outside fabric layers. The ballistic performance of BFS of
these hybrid panels were analysed in the non-perforation case.

Table 3.13 Non-perforated hybrid panels with different layering sequences

Layering up The number Areal density


Sample
sequence of layers (g/m2 )

8F8 /7U16 24 4248.80

79
7U16 /8F8 24 4248.80

7U8 /8F8 /7U8 24 4248.80

8F4 /7U16 /8F4 24 4248.80

3.2.3.2 Varied amount of each component

To identify the proper amount of Dyneema UD in the hybrid panel, Dyneema UD


layers and woven fabric layers were combined with varied mass fraction as shown in
Table 3.14.

Table 3.14 Hybrid panels with different amount of UD layers

Layering up The number Areal density


Sample
sequence of layers (g/m2 )

11F20 /7U4 24 4724.40

11F16 /7U8 24 4684.76

11F12 /7U12 24 4645.12

11F8 /7U16 24 4605.48

11F4 /7U20 24 4565.84

To identify the function and the range of the first group in the hybrid panel, hybrid
structures panels combining fabric layers of 8F, 13F with 11F were constructed. The
tight fabric 13F and loose fabric 8F were placed on the striking face in front of 11F
to identify the ballistic function of the first group. Varied amount of fabric layers 13F
were combined on the striking face to identify the exact range of the first group in
80
the hybrid panel. The specifications of three hybrid structures panels were listed in
Table 3.15.

Table 3.15 Hybrid structure panels

Layering up The number of Areal density


Sample
sequence layers (g/m2 )

13F3 /11F20 23 4695

8F3 /11F22 25 4793

13F9 /11F12 21 4623

3.2.3.3 Hybrid panels

According to the design principle of the hybrid panel, two hybrid panels composed
of woven fabrics and Dyneema UD laminates were constructed as shown in Table
3.16. The hybrid panel of 11F3 /8F8 /U6 were assessed using the perforation tests to
identify the degree of ballistic resistance. Another hybrid panel 13F3 /8F10 /U8 was
assessed using the non-perforation tests to analyse ballistic performance of BFS.

Table 3.16 The hybrid panels

Layering up The number of Areal density


Sample
sequence layers (g/m2 )

11F3 /8F8 /U6 17 2953

13F3 /8F10 /U8 21 3799

81
3.3 Ballistic test

3.3.1 Ballistic test equipment and procedure

Ballistic tests were conducted at the ballistic laboratory in the University of


Manchester. A firing range used for impact tests was shown in Fig. 3.11. A panel
was placed between two couples of IR sensors which were shown in Fig. 3.12(a). A
steel cylindrical projectile (5.5mm in diameter and height, 1.004 (+0.008) g) was
used as Fragment Simulation Projectiles (FSP) for impacting on panels in this study,
which was located in a plastic sabot as shown in Fig. 3.12(b). They were fired by a
machine simulating hand gun and propelled by gunpowder. When the projectile
passed through each pair of IR sensors, the responding time was recorded by the
front and back timer. A high speed camera was used to record the impact process.

Clay Panel

Fig. 3.11 Schematic of experimental ballistic range

(a) IR sensor (b) Cartridge, sabot and projectile

(c) Frame (d) Clay


Fig. 3.12 Ballistic test set-up
82
In the perforation test, a panel was clamped in a square steel frame of dimensions of
24cm×24cm with a circle of 15cm in diameter in the middle as shown in F ig. 3.12 (c).
The frame with the panel was placed in the firing range between two pairs of IR
sensors. Due to the variability of the exiting velocity of the projectile, the
performance of a perforated panel was determined by ten shots [7].

In the non-perforation test, a clay box with a size of 24×24×10cm was placed at back
of the panel as shown in Fig. 3.12 (d). The backing material is Roma Plastilina No.1
(RP#1) oil-based modelling clay [7]. It was used to memorize the configuration of
the indentation in its surface after impact. According to the NIJ standard, the clay
should be put into the oven at 38o C above 3 hours before the non-perforation ballistic
test. After the impact, an indentation was remained in the clay. As the non-
perforation test results were relative stable, the performance of a non-perforated
panel was determined by five repeat ballistic tests.

3.3.2 Data acquisition

3.3.2.1 Striking velocity and residual velocity

According to the distance between IR sensors and the time of the projectile passing
through, the striking velocity and residual velocity can be calculated based on the
equations (3-14) and (3-15). By taking account of the distance from the first IR
sensor to fabric, the decelerate velocity of projectile was calculated as 2m/s
according to equation (3-16) and (3-17). Therefore, the average striking velocity on
fabric was calculated as 483m/s, which is the velocity of the projectile when contacts
the panel. The residual velocity is the exiting velocity through the panel.

(3-14)

(3-15)

(3-16)

Y (3-17)

83
where s1 (47.0cm) and s2 (36.2cm) is the distance between first and second IR sensor
pair respectively. is the striking velocity and residual velocity respectively,
is the time of the projectile passing through two couple of IR sensors. a is the
accelerated velocity. y is the distance between two couples of sensors.

3.3.2.2 Energy absorption of the panel

According to the striking velocity and residual velocity of the projectile, the kinetic
energy dissipation of the projectile can be calculated. Considering the amount of the
striking velocity deceleration in the air even if without armour panels in the ballistic
range, the loss of kinetic energy 0.21J was measured by ballistic tests. Suppose the
energy dissipation due to the projectile deformation and acoustic losses are all
negligible. The kinetic energy dissipation of the projectile is assumed to be equal to
the energy absorption in the panel.

(3-18)

where is the kinetic energy dissipation of the projectile, m is the mass of the
projectile, is the striking velocity and residual velocity respectively.

To compare the energy absorption capacity of different panels, the normalised


energy absorption (SEA) was used to remove the effect of the areal density. This
reflects the energy absorption capacity of the panel at unit areal density.

(3-19)

where SEA is the specific energy absorption, EA is the energy absorption of panel,
Ar is the areal density of the panel.

3.3.2.3 Ballistic resistance tests

In this study, only one type of cartridge was provided to propel the projectile, the
ballistic limit velocity of panels cannot be obtained by adjusting the striking velocity.
Therefore, for a given striking velocity, V50 can be obtained by adjusting the number
of layers in the panel. Ballistic resistance performance of different panels can be
identified by comparing the limit number of layers. The limit number of layers can
be defined as when the panel is perforated at 50% chance at a given striking velocity

84
of 483m/s, the related number of layers is defined as the limit number of layers.
Fewer limit number of layers indicated better ballistic resistant capacity.

The limit number of layers was determined by a series of ballistic tests. The ballistic
test procedure referred to the test procedure of V50 tests. Six round ballistic tests
were conducted for determination of V50. If three panels were perforated and
another three panels were not perforated, the striking velocity of 483m/s was
identified as V50 and the number of layers was identified as the limit number of
layers of this panel at this striking velocity.

3.3.2.4 Quantification of the indentation in the clay

In the non-perforation case, a projectile was stopped in a panel. Some transmitted


kinetic energy of the projectile can produce an indentation in the clay behind the
panel. The indentations behind different panels exhibited different configurations. A
conical indentation was found behind Twaron woven panels with round edge in the
clay as shown in Fig. 3.13 (a). Behind Dyneema UD panels, the indentation exhibited
a conical bottom and square edge in the clay as shown in Fig. 3.13 (b).

(a) Behind Twaron woven panel (b) Behind Dyneema UD panel


Fig. 3.13 The indentations in the clay behind (a) Twaron woven fabric panel; (b)
Dyneema UD panel

To quantify different configurations of the indentations behind woven panel and


Dyneema UD panel, the exact mould of the indentation was taken by wax mould as
shown in Fig 3.14. The wax mould was cut into several sections. The profiles of
these sections were measured to identify a series of coordinates on the surface of the
wax mould. According to these coordinates, the three-dimension (3D) surface was
plotted in Matlab as shown in Fig. 3.15.

85
8F24 : 3782 g/m2
7U20 : 3740 g/m2

Fig. 3.14 Wax mould of the indentation behind Twaron woven panel 8F24 (left)
and Dyneema 7U20 (right)
25

16
20
14

12

10 15

6
10
4

0 5

25

20 0

15

10 -5

-10
0

-5

-15
-10
25
20
-15 15
10
5 -20
-20 0
-5
-10
-25 -15
-20
-25 -25
-25 -20 -15 -10 -5 0 5 10 15 20 25

(a) Indentations behind the Twaron woven panel 8F24

25

20

15

10
10

5
6

4
0

30 -5
0
25
20
20
-10
15 10
10
5 0 -15
0
-5 -10
-10 -20
-15 -20
-20
-25 -30 -25
-30 -20 -10 0 10 20 30

(b) Indentations behind the Dyneema UD panel 7U20


Fig. 3.15 3D surface of indentations behind the (a) Twaron woven panel 8F24 (b)
Dyneema UD panel 7U20

According to the observation, the indentation behind the Twaron woven panel was
deep and round, while the indentation behind the Dyneema UD panel was much
shallow and wider. The bottom of the indentation behind the Twaron woven panel
was larger than that of the Dyneema UD panel. With the close areal density of two
panels, the average depth of the indentation behind the Dyneema UD panel (11.81
mm) was less than a half as that of the Twaron woven panels (27.59 mm). The

86
average width of the indentation edge in the clay behind the Dyneema UD panel
(51.25 mm) was about one-third (36.19%) larger than that of the Twaron woven
panels (37.63 mm). Different characteristics of the indentation indicated the effect of
material properties on ballistic performance.

3.3.2.5 Backface Signature (BFS)

BFS is one of the most important criteria to evaluate ballistic performance of non-
perforated panels. It is the perpendicular distance from the reference plane of the clay
surface to the deepest point of the indentation [7]. After ballistic tests, an indentation
was produced in the clay behind armour panels. In this study, BFS can be directly
measured by a vertical calliper as shown in Fig. 3.16. The reference plane is the
original surface of the clay. Smaller BFS reflects better ballistic resistant capacities
of the non-perforated panel.

Reference
plane
BFS
Clay
Maximum
indentation

Fig. 3.16 BFS measurement

To remove the effect of areal density between different panels, the specific depth
(SD) was used to normalise the BFS behind panels with different areal density in this
study.

(3-20)

where the SD is the specified depth of the indentation in the clay (cm3 /g), the Ar is
the weight of the panel (g/cm2 ).

87
3.3.2.6 Indentation volume

The indentation in the clay behind different types of panels exhibited obviously
different configurations. Therefore, the smaller BFS does not directly indicate lower
transmitted energy and better energy absorption of the panel. The transmitted energy
is also determined by the configuration of the dent. For example, the indentation
behind Dyneema UD panel was found to be much shallower and wider than that of
Twaron woven panels. It is more reasonable to reflect the energy absorption capacity
of non-perforated panels by comparing the volume of the indentation in the clay. The
smaller volume of the indentation indicated little transmitted energy and better
energy absorption capacity of non-perforated panels. Such evaluation criterion is
more efficient and reliable for comparison between panels with different materials.

In this study, the volume of the indentation was used as a complementary criterion
for BFS to reflect the difference of energy absorption capacity between non-
perforated panels. It can be easily determined by measuring the volume of water in
the indentation as shown in Fig. 3.17. The top surface of water was kept the same
level as the reference plane. The edge of water was kept touching the end of marks in
the clay.

Fig. 3.17 Measurement of the indentation volume

3.4 Photographic observations

3.4.1 Impact process observations

Some ballistic responses of armour panels under impact can be directly observed by
a high speed camera. It recorded the impact process during ballistic impact. From the
photographic observations, many ballistic characteristics of panels were revealed,
such as the configuration of fabric transverse deformation at different times, the

88
strain evolution on yarns, and the impact angle and trajectory of the projectile. This
provides the physical evidences for the theoretical analysis and numerical simulation.

In this study, the impact process of some typical panels were recorded by the high
speed camera Photron FASTCAM SA1, including a single layer of Twaron fabric 8F,
Twaron woven panel 11F19 and 11F24 , Dyneema UD panel 7U20 . To quantify the
transverse deformation at different times during impact, the panels were marked with
a square grid on the outer layers with side length of 10mm as shown in Fig 3.12(c).

The high speed camera (Photron FASTCAM SA1) used in this study has frame rates
from up to 675,000 per second (fps). At the least time interval, each frame can be
taken every 5.5µs with the resolution 512 × 128 pixels. PFV (Photron FASTCAM
SA1) software was used to control the camera from the computer, including setting
camera options, shooting, and saving recorded data to the computer.

3.4.2 Post-impact panels observations

To investigate failure modes of each layer in the multi- layer system and analyse
failure mechanisms of soft armour panel, some typical post- impact panels were
observed after ballistic tests, including Twaron woven panel 11F 24 and Dyneema UD
panel 7U20 .

The global failure characteristics of post- impact panels were observed at first. The
fractured fabric layers were observed by naked eyes and optical microscope
(Projectina CCD-1300PQC) at a magnification from 5× times to 50× times. The
fractured yarns and fibres were observed by the optical microscope (Leica DFC295)
at a magnification of 50× times to 200× times.

In addition, the detailed inspection of the post- impact panels was carried out by
Scanning Electronic Microscope (SEM). Individual broken fibres and fractured yarns
were observed at higher magnifications to reflect different failure modes of each
layer in a panel. Before SEM observation, the samples from Twaron woven panels
and Dyneema UD panels were prepared and coated. Around the perforated hole of
each layer, the sample with the size of 20×20mm was cut from post- impact fabric
layer and mounted on carbon stubs. The samples were coated in a sputter coater with
a thin layer of gold. The Argon (Ar) pressure was 0.2torr, the electric current was set
10-15 mA and the sputtering time was 1-2 min. Samples were observed by Hitachi
89
S3000. The voltage of ion emission was 10-15KV. The magnifications were from 15
times to 1000 times.

3.5 Test data

Ballistic tests results were listed in Table 3.17-3.21. Perforation tests aims to reflect
the energy absorption (EA) performance of panels with a single layer or multi- layers
listed in Table 3.17-3.19. Ten specimens were used to obtain an average value. For
non-perforation tests, Backface Signature (BFS), the number of perforated layers and
the indentation volume (Vi) were obtained to reflect ballistic resistance performance
of panels as shown in Table 3.20-3.21. Five specimens were used to obtain an
average value.

The confidence of tests data was reflected by standard deviation (STD). To remove
the effect of areal density (Ar), the specific energy absorption (SEA) and the specific
Backface Signature (SD) were used to compare ballistic performance of panels with
different areal density. The number of broken layers in a non-perforated panel was
variable during ballistic tests. So the average perforated layer was used here.

Table 3.17 Ballistic performance of a single layer panel

Sample Ar EA EA SEA SEA


No. (g/m2 ) (J) STD (J.cm2 /g) STD

8F 155.35 6.07 0.86 390.73 15.76


9F 184.29 7.41 1.17 402.08 32.15
10F 198.67 6.38 1.13 321.14 10.98
11F 196.85 7.19 0.41 365.25 10.38
12F 227.95 6.33 1.23 277.69 46.52
13F 267.44 7.18 0.63 268.47 17.44
7M 153.25 6.06 0.69 395.43 18.26
10M 259.28 7.82 0.60 301.60 26.15
7T 260.41 6.15 0.48 236.17 30.29
7U 186.94 4.85 0.67 259.44 35.72
5U 252.98 4.57 1.25 180.65 49.31

90
Table 3.18 Ballistic performance of the perforated panels

Sample The number EA EA SEA SEA


No. of layers (J) STD (J.m2 /g) STD

3 18.43 1.46 395.45 19.35

Twaron 4 25.31 1.59 408.23 12.39


8F 6 35.69 2.98 382.90 20.13
9 40.48 6.51 289.53 22.46
3 20.16 1.70 339.65 14.26
Twaron 4 26.34 2.70 333.42 15.65
11F 6 38.33 2.37 322.89 13.71
9 45.84 5.99 257.43 15.66
Twaron
3 22.48 1.23 289.01 10.71
10M
3 15.41 1.99 274.78 35.58
Dyneema
UD 6 33.33 3.89 297.15 34.86
SB71
9 70.96 5.84 421.76 34.88
3 16.69 2.02 219.91 26.66
Dyneema
UD 6 36.07 5.69 237.63 37.48
SB51
9 76.63 11.08 336.57 48.64

Table 3.19 Ballistic performance of the non-perforated panels

Ave Ave
Ar Ave vs BFS Ave SD Vi
Panel BFS broken
(g/m2 ) (m/s) STD (cm3 /g) (cm3 )
(mm) layers

8F24 3728.40 485.71 27.59 1.35 7.40 12 10.0


11F19 3740.15 492.17 19.48 0.16 5.02 8 10.7
11F24 4724.40 480.07 15.25 1.12 3.23 8.5 9.2
11F36 7086.60 489.08 13.16 0.15 1.86 7 9.3
11F48 9448.80 484.71 9.52 0.23 1.01 6 8.0
11F120 23622.00 479.10 / / / 6 /
7U20 3738.80 485.37 11.81 0.19 3.16 10.5 8.0
7U24 4486.56 479.96 9.75 0.86 2.16 12 5.2

91
Table 3.20 Ballistic performance of the perforated hybrid panels

Sample Ar (g/m2 ) EA (J) STD


8F/7U 342.29 11.80 1.23
7U/8F 342.29 9.94 0.30
7T/7U 447.94 14.67 1.10
7U/7T 447.94 13.46 0.73
8F3 /7U3 1029.69 38.63 2.96
7U3 /8F3 1029.69 30.89 2.97
7T3 /7U3 1342.30 50.56 4.41
7U3 /7T3 1342.30 42.38 3.72

Table 3.21 Ballistic performance of the non-perforated hybrid panels

Ave BFS Ave Total


Ar Ave vs Vi
Panel BFS SD broken
(g/m2 ) (m/s) STD (cm3 )
(mm) (cm3 /g) layers
11F20 /7U4 4684.76 484.04 14.67 0.19 3.13 8 10.00
11F16 /7U8 4645.12 485.54 13.48 0.23 2.89 8.5 7.40
11F12 /7U12 4605.48 479.10 12.33 0.15 2.67 12 6.90
11F8 /7U16 4565.84 485.39 9.98 0.25 2.19 9 6.65
11F4 /7U20 4526.20 490.61 9.53 0.08 2.11 9 7.50
8F8 /7U16 4248.80 483.75 9.75 0.72 2.29 8.5 6.59
7U16 /8F8 4248.80 480.54 12.00 0.35 2.82 9 7.21
7U8 /8F8 /7U8 4248.80 482.49 10.25 1.26 2.41 10.5 6.97
8F4 /7U16 /8F4 4248.80 490.11 13.37 0.18 3.15 11.5 7.42
8F3 /11F22 4793.75 484.78 16.91 0.26 3.53 8.5 10.83
13F3 /11F20 4695.88 491.15 15.11 0.02 3.22 8 9.90
13F9 /11F12 4623.74 488.26 17.48 0.30 3.78 8.5 11.30

3.6 Summary

This chapter described experimental studies of ballistic armour panel. Two types of
ballistic materials, Twaron woven fabric and Dyneema UD laminate, were
investigated on the mechanical properties including yarns properties, fabric
properties and laminate properties. Different armour panels with a single layer and
multi- layer as well as hybrid panels were constructed with above two types of
ballistic materials to investigate ballistic performance in the perforation and non-
92
perforation case. Ballistic tests were introduced including ballistic tests equipment,
tests procedure and data acquisition. Impact processes of woven panels and UD
panels were observed by a high speed camera. The detailed failure characteristics of
each layer in post- impact panels were observed by naked eyes, optical microscope
and Scanning Electronic Microscope (SEM) to analyse the failure mechanisms of
soft armour panel. The observation results will be represented in Chapter 6.

The experimental studies directly reflected ballistic performance of different armour


panels. Based on these raw data, some influencing factors on performance of the
multi- layer system can be further analysed in the following chapters. In addition,
much physical evidences of ballistic responses were provided, which were vital for
the investigation on failure modes of armour panels and for the numerical simulation.

93
CHAPTER 4 Numerical modelling of armour panels under
ballistic impact

To gain insight into the different ballistic responses of each layer in the multi- layer
panel, Finite Element (FE) analysis is employed in this study. Some ballistic
characteristics of armour panels, such as stress wave propagation, transverse
deflection, and different forms of energy dissipation, can be identified and quantified
by this method. It is a vital complementary analysis technique to experimental
method to obtain a thorough and comprehensive understanding of mechanisms
responding to ballistic impact on soft armour panel. This chapter describes numerical
modelling of armour panels under ballistic impact. The primary aim of this chapter is
to create reliable and efficient FE models of armour panels under ballistic impact and
simulate the ballistic response.

In this chapter, some typical armour panels, such as Twaron woven panels and
Dyneema UD panels, under the perforation and non-perforation ballistic tests
described in Chapter 3 were simulated using Finite Element (FE) models.
Geometrical modelling of these panels under ballistic impact is introduced in section
4.1. In section 4.2, different material models and material properties of yarns, fabric,
UD laminate, projectile and clay, are defined and assigned to FE models. In section
4.2, the accuracy of these FE models is validated by experiment tests results. A
summary is presented in the last section.

4.1 Introduction of numerical modelling

In this study, Finite Element (FE) simulation was carried out by ABAQUS 6.11
commercial software. ABAQUS is engineering software that provides the facility to
create ABAQUS models, to run and monitor the progress of analysis, and to evaluate
the results from ABAQUS simulations, which was used by many researches [48, 82,
118, 127]. A complete Finite Element (FE) analysis is usually carried out in three
stages: pre-processing, simulation and post-processing. In the pre-processing stage,
an input file is created through ABAQUS/CAE environment to describe a physical
problem. In the simulation stage, a particular problem is defined and analysed. In the
post-processing stage, the analysis results are stored in an output file and evaluated
using the Visualisation module of ABAQUS/CAE.
94
Generally, FE model is usually built up in the following steps. First, the geometric
model of each part is created and assembled together. Then the loads setting, contact,
boundary conditions and meshing needs to be defined. In addition, material
properties, material model and damage criteria are identified and assigned to the
model. Finally, the analysis type and output request are defined [128].

In this chapter, Twaron woven panels and Dyneema UD panels under ballistic impact
were simulated in the perforation and non-perforation ballistic cases. Due to the short
dynamic response time of ballistic impact and comparatively large deformations of
armour panels, an explicit dynamic analysis was used to simulate the ballistic impact
using ABAQUS/Explicit. Three-dimensional (3D) FE models of armour panels
under ballistic impact were created using finite solid elements.

In the perforation case, FE models are composed of a projectile and a woven panel.
The fabric is interlaced by a specific number of individual warp and weft yarns into
the plain weave. The perforated panels are constructed with few fabric layers, which
were simulated at the yarn level. This modelling method can simulate some detailed
ballistic behaviour of yarns in fabric layers.

In the non-perforation case, a block of clay was added to the model at back of the
ballistic panel. For the sake of calculating efficiency, non-perforated panels with
more layers (above 20 layers) were simulated at the fabric level. The fabric was
defined to be a continuum and represented as a homogenous, transverse isotropic
plate. FE models at the fabric level focused on reflecting global ballistic
characteristics of each layer in the multi- layer system.

4.2 FE modelling of armour panels under ballistic impact

4.2.1 FE model of a single layer fabric

To simulate the armour panel under ballistic impact, FE modelling of a single layer
fabric as the basic unit of the multi- layer panel will be described as a starting point.
A single layer fabric was simulated at the yarn level. Yarns in the fabric were
assumed to be a 3D solid continuum cylinder with a defined crimp profile. The yarn
geometry in the FE model of the plain fabric has to be identified first.

95
One simple method is directly measuring the parameters of yarn geometry from a
slice of yarn cross-section by optical microscope (Chapter 3, Fig. 3.2). However, it
should be emphasised that the FE model of a yarn is different from the actual
construction of the yarn. The yarn model was simulated as a solid rod. The actual
yarns in fabric are composed of a bundle of fibres with voids. Therefore, using the
measured geometrical parameters of yarns will result in the increased amount of
materials in fabric model. In this study, the calculated geometrical parameters of a
yarn, including the wave length of a yarn, as well as the width and height of the yarn
cross-section, were used for FE modelling.

4.2.1.1 Geometry of yarns

For a given yarn count, the area of the yarn cross-section without voids should be a
constant value regardless of the shape of the yarn cross-section. It can be easily
calculated according to the yarn count and the volume density of fibres as shown in
Equation 4-1, which is an assumed extreme case. In fact, the area of the yarn cross-
section increases when the yarn porosity is considered. When fibres volume fraction
(Vf) in a yarn is taken into account, the area of the actual yarn cross-section
composed of fibres with voids should be revised as in Equation 4-2.

(4-1)

(4-2)

where s is the area of a yarn cross-section (cm2 ), s’ is the area of the actual yarn
cross-section. Ttex is the yarn count (tex), is the volume density of fibres (g/cm3 ).

s s

d d
(a) square packing (b) hexagonal packing
Fig. 4.1 Fibres packing in yarns (a) square packing, (b) hexagonal packing [128]
In a yarn with little twist, circular fibres are normally packed parallel in two manners
as shown in Fig. 4.1[129]. The square array of packing as shown in Fig. 4.1(a) gives
the maximum fibres volume fraction at 78.54%, while the hexagonal array packing

96
in Fig. 4.1(b) gives 90.69%. In most yarns, these two packing configurations of
fibres are mixed and exist at the same time. Therefore, the volume ratio of fibres in a
yarn is between 78.54% and 90.69% with compacted packing. In this study, a fibre
volume fraction of 85% was assumed and used to calculate the actual area of yarn
cross-section. Taking Twaron yarn of 93tex as an example, the area of yarn cross-
section is 7.58×10-2 mm2 .

To simplify the FE modelling of Twaron plain woven fabric, the weaves densities in
the warp and weft direction were assumed to be the same, regardless of a little
difference between warp and weft densities during the manufacture. Therefore, the
warp and weft yarns have the same geometrical parameters, such as the crimp wave
length and the width of cross-section as shown in Fig. 4.2.

Rm
h

Ri
θ
L

Fig. 4.2 Schematic of yarn cross-section

According to the photographic observations of the yarn cross-section (Chapter 3, Fig.


3.2), the yarn cross-section in Twaron woven fabric is similar as the lenticular shape.
The geometrical parameters of a yarn can be calculated by following equations.

(4-3)

(4-4)

(4-5)

(4-6)

√ ) (4-7)

97
(4-8)

where P is the weave density (ends/cm), L is the half crimp wave length (m), h is the
fabric thickness (m), Rm and Ri is the middle and inner radius of crimp arc of yarn (m)
respectively, W is the width of yarn cross-section (m). M is the areal density of fabric
(g/m2 ); ρ is the volume density of fabric (g/m3 ); S is the area of yarn cross-section
(m2 ); l is the length of yarn in fabric (g/m2 ); c is the crimp ratio; n is the number of
yarns in fabric.

For Twaron fabric 8F (93tex, 8.3ends/cm), the calculated geometrical parameters of


a yarn were shown in Table 4.1. The calculated values of the yarn cross-section were
close to the measured values of the actual yarn. In addition, the areal density of the
fabric was close to the actual value when the compressed thickness was used.
Therefore, these calculated parameters of a yarn were adopted to create the FE model
of the Twaron woven fabric. In the same way, the yarn geometry in the fabric 11F
was also determined. With the increasing weave densities of the fabric, the shape of
the yarn cross-section and the longitudinal configuration changed correspondingly.

Table 4.1 Yarns geometry in FE models of Twaron fabrics

8F (93tex, 8.3ends/cm) 11F (93tex, 10.9ends/cm)


Geometrical parameters
of Twaron fabric Measured Calculated Calculated
values values values
Crimp wave length (cm) 0.2556 0.2410 0.1835
Width of cross-section (cm) 0.1134 0.1161 0.0909
Thickness (cm) 0.25 0.1936 0.2723
Crimp ratio (%) 0.41 0.46 1.49
Areal density (g/m2 ) 155.35 163.71 220.78

4.2.1.2 FE modelling of fabric under impact

According to above calculated geometrical parameters of yarns, the yarn cross-


section with the lenticular shape was modelled as two symmetry arcs and remained
constant along the yarn axis. The yarn path was modelled as some connecting arcs
with crimp waves, which was shown in Fig. 6.3 (a). A fabric model was an assembly
of a certain number of warp and weft yarns according to the plain weave.

98
In the perforation ballistic tests, a fabric was clamped in a frame with four edges
fixed and impacted by a projectile perpendicular to the centre. In the FE mode l, the
fabric was created in the X-Z plane and the Y axis was the impact direction. The
projectile was assumed to impact the fabric at the centre of the fabric at 90 o to the
fabric plane. Due to the symmetry of the projectile- fabric system, only one-quarter of
the system was modelled. The actual sample was clamped in a square frame with a
circle of 150mm in the middle as shown in Fig. 3.13. To simplify the simulation, the
fabric was simulated as the square shape with the size of 75mm × 75mm.

(a) A single yarn (b) The fabric and projectile


Fig. 4.3 FE model of Twaron fabric (a) a single yarn, (b) the fabric and projectile

Assignment of the boundaries conditions ensure the fabric can only move along the
Y-axis. A fixed boundary condition was applied to the two outer edges away from
the impact area. The symmetry conditions were applied to another two edges
crossing at the impact area as illustrated in Fig. 4.3 (b). The projectile was assumed
to be a rigid body, due to no deformation produced by textile materials during impact.
An average velocity of 483m/s which was tested in the Chapter 3 was assigned to the
projectile along the Y-axis.

In the FE model, a general contact algorithm was used to define yarn-yarn interaction
and the projectile-fabric interaction. Simple coulomb friction was introduced for all
contact in the FE model.

Yarns in fabric and the projectile were meshed with eight node hexahedron elements
(C3D8R) with reduce integration and hour glassing in the model as shown in Fig. 4.4
(a). The yarn-cross section was meshed into ten elements. The yarn wavelength had
twelve elements as shown in Fig. 4.4 (b).

99
Symmetrical plane
(a) fabric and projectile (b) yarn
Fig. 4.4.The mesh of FE model (a) fabric and projectile, (b) yarn

4.2.2 FE model of perforated multi-layer panels

In the non-perforation case, FE models of the multi- layer panel can be developed
from the model of a single layer fabric. It was created by making copies of a single
layer fabric with a certain number of layers as shown in Fig. 4.5. All copies of fabric
keep the same parameters as the single layer fabric including fabric geometry,
material properties, boundary conditions, contact interaction conditions and mesh
properties.

Fig. 4.5 The FE model of a multi-layer panel

Correspondingly, the FE model of the multi- layer system contained high number of
elements (461,520 elements per layer). This could not be implemented due to the
higher calculation cost. Therefore, the number of elements of the multi- layer system
model has to be reduced. There are two measures that can be carried out including
model size reduction and the mesh density reduction.

1) Model size reduction

100
The model size of the multi- layer panel can be reduced. Three sizes of models (75cm,
50cm and 25cm) were selected to investigate the size effect on the energy absorption
of the model.

2) Mesh density reduction

The second alternative measure was reducing the mesh density of the model. In a
study by Roylance [45], the majority of ballistic impact energy was deposited in the
primary yarns. The secondary yarns were essentially ineffective. Therefore, the
hybrid mesh was adopted in the model to reduce elements. Primary yarns used the
same fine mesh that was used for the single layer fabric model. The yarn cross-
section was meshed into ten elements and the yarn path had six elements as shown in
Fig. 4.6 (a). The secondary yarns adopted the coarse mesh with four elements in the
yarn cross-section and four elements in the yarn path as shown in Fig. 4.6 (b).
Compared with the FE model of single layer fabric with 461,520 elements, the FE
model with the hybrid mesh only had 77,368 elements, a reduction of 83.3%.

(a) Primary yarn (b) Secondary yarn

(c) Hybrid mesh in fabric


Fig. 4.6 Hybrid mesh (a) the fine mesh of the primary yarn, (b) the coarse mesh
of the secondary yarn, (c) hybrid mesh in fabric

101
FE results of multi- layer panels with above two measures are validated by ballistic
test results in section 4.4. The better method was adopted in the FE model of the
multi- layer system.

4.2.3 FE model of non-perforated multi-layer panels

In the non-perforation case, the FE model of the multi- layer system was composed of
a projectile, an armour panel and clay. A non-perforated panel usually contains more
than 20 layers. If it is simulated at the yarn level, the FE model size will be too big to
run. In addition, the scope of the study on the non-perforated panel is to understand
the ballistic response of each layer in the multi- layer system, the detailed true-scale
representations of each yarn is unnecessary. Therefore, FE models of the non-
perforated panels were simulated at the fabric level. Twaron woven panel and
Dyneema UD panel were simulated at the case of the non-perforation ballistic tests.

(a) A single layer fabric plate (b) The armour panel

Coarse mesh
Amour panel
Fine mesh
30mm
50mm

Projectile
Clay

75mm
25mm
(c) The non-perforated panel under ballistic impact
Fig. 4.7 Quarter-symmetry FE model of the non-perforated panel under impact
102
A single layer fabric and UD laminate were modelled as a 3D solid continuum plate
as shown in Fig. 4.7. The non-perforated panel was layered up by copying the single
plate with the size of 75×75mm. The thickness of each layer is the same as the actual
thickness of woven fabric. The clay at back of the panel was modelled as a 3D solid
continuum square block with the size of 75×75×50mm.

In the non-perforation case, the panel was placed in front of the clay without
clamping. Therefore, in the FE model, the outer edges of the panel were set free. The
other two edges crossing at the impact area were applied symmetry conditions. For
the clay, it was placed in a backing material fixture. Therefore, the constraint
boundary conditions were assigned for the outer edges and the back of the clay. The
symmetry conditions were applied for the other two edges crossing at the impact area.

A general contact algorithm was used to define the interaction between layers and
projectile- fabric interaction. Simple coulomb friction was used for all contact in the
FE model.

To reduce the number of elements and maintain the accuracy of calculation, different
mesh sizes were used in different regions of the plate and the clay. Two elements
were applied through the thickness direction. In the plane of the plate, the fine mesh
0.446×0.446×0.12mm was used in the centre around the impact zone (25×25mm)
and the coarse mesh transited from 0.446×1.132×0.12mm to 0.446×4.582×0.12mm
was used in the regions remote from the impact point as shown in Fig. 4.7. For the
clay, in the central part (25×25×20mm), the fine mesh size of 0.446×0.446×0.455
mm was used. In the other region, the coarse mesh size from 0.446×0.455×1.505mm
to 1.329×3.000×6.020mm was adopted. The details of meshing and boundary
condition are shown in Fig. 4.7.

4.3 Material properties

4.3.1 Introduction of material assignment

For a FE model, material model selection and proper properties assignment are
essential to govern the simulation analysis. It must correctly reflect the precise
characteristics of materials. Assignment of material properties generally contains the

103
definition of the effective material response of the material, damage initiation
criterion and damage evolution law as shown in Fig. 4.8 [128].

Fig. 4.8 Typical stress-strain curve of elastic-plastic material [128]

Fig. 4.9 linear-elastic response of high performance fibres

For high performance fibres (Aramid or UHWMPE) used in the soft body armour,
the material behaviour exhibited typical linear-elastic response as shown in Fig.4.9.
The stress can be represented according to Equation 4-9.

(4-9)

104
As the ballistic fibres do not exhibit obvious yield behaviour, the onset of the damage
initiation was assumed to be the point where the stress reached the yield stress of
material. Abaqus/Explicit offers a variety of choices of damage initiation criteria.
These ballistic fibres are usually specified to be the ductile criterion. The ductile
criterion is a phenomenological model for predicting the onset of damage due to
nucleation, growth, and coalescence of voids.

The damage evolution law describes the rate of degradation of the material stiffness
once the corresponding initiation criterion has been reached. For damage in ductile
metals the stress tensor in the material is given by the Damage Equation 4-10 at any
given time during the analysis [128].

̅ (4-10)

Where D is the overall damage variable and ̅ is the effective (or undamaged) stress
tensor computed in the current increment. The material has lost its load-carrying
capacity when D=1.

After damage initiation, the material stiffness is degraded progressively according to


the specified damage evolution response. When material damage occurs, the stress-
strain relationship no longer accurately represents the material‟s behaviour.
According to Hillerborg‟s fracture energy theory, a stress-displacement response is
created after damage is initiated. G f, as a material parameter, is defined as the energy
required opening a unit area of crack. The softening response after damage initiation
is characterized by a stress-displacement response rather than a stress-strain response.
The damage evolution law can be specified in terms of equivalent displacement or in
terms of fracture energy dissipation Gf. For the high performance fibres, G f is usually
assigned 1000J [128].

̅ ̅
∫ ̅ ∫ ̅ (4-11)
̅

This expression introduces the definition of the equivalent plastic displacement, ̅̇ ,


as the fracture work conjugate of the yield stress after the onset of damage (work per
unit area of the crack). Before damage initiation ̅̇ ; after damage
initiation ̅̇ ̅̇ .

105
The damage evolution law describes the rate of degradation of the material stiffness
once the damage initiation criterion has been reached. The evolution of the damage
variable with the relative plastic displacement can be specified in tabular, linear, or
exponential form.

For the material assignment, it should be emphasised that there is always some
compromise between the complex dynamic phenomena during impact and materials
properties in term of the modelling accuracy and efficiency.

4.3.2 Projectile

In ballistic tests of this study, steel Right Circular Cylinder (RCC) projectiles were
used to impact panels. During impact, the projectile was not found any obvious
deformation. Therefore, in the FE model, the projectile was defined as a rigid body
without taking account of its deformation. Therefore, the projectile material was
assigned isotropic and homogeneous properties. The material properties of the steel
were listed in Table 4.2.

Table 4.2 Material properties of the projectile

Material properties Projectile

Young‟s Modulus E (GPa) 206.8


Poisson‟s ratio 0.3
Mass Density ρ (kg/m3 ) 7800

4.3.3 Twaron yarns and fabric

In this study, a single layer Twaron fabric and the multi- layer perforated Twaron
woven panels were modelled at the yarn level with Twaron yarn of 93tex. According
to previous researches, the isotropic and orthotropic material model of Twaron yarn
showed the same trend with very small differences in the stress distribution. The
longitudinal stress wave velocity was not sensitive to material models. The use of
isotropic material model is reliable in the calculation of energy absorption [130]. In
addition, the isotropic model was very economical in computing time.

To simplify the simulation, Twaron yarns were assumed to be isotropic material with
the same properties in all direction of a yarn. The linear elastic-plastic material

106
model was defined as the material response. The parameters of material properties
that were assigned for the FE models of yarns are listed in Table 4.3. The tensile
properties of Twaron yarns were referred to product properties of Teijin ® Company
(Chapter 3, section 3.1.1.1). The mass density of Twaron yarns was assumed as the
same as that of fibres (1440 kg/m3 ) to keep the same stress wave velocity in Twaron
material. The increased areal density of fabric has no influence on the specific energy
absorption (SEA). The Poisson‟s ratio was assumed the same as Twaron fibres.
According to tests results in previous studies [84, 131], the friction coefficient
between yarns and between projectile was assumed to be 0.2. Damage criterion of
the ductile damage is applied in the model. The damage evolution law was specified
in terms of fracture energy. The softening is defined as exponential form [128].

Table 4.3 Material properties of Twaron yarn and fabric

Material properties Twaron yarn Twaron fabric Dyneema UD


Yarn count (tex) 93 93 /
Yield stress σ (GPa) 3.6 1.2 1.5
Fracture strain ε (%) 4.3 3.5 4
Young‟s Modulus E (GPa) 80 20 40
Poisson‟s ratio 0.35 0.01-0.1 0.01-0.1
Mass density ρ (kg/m3 ) 1440 750 970
Friction coefficient 0.2 0.2 0.15
Fracture energy (J) 1000 1000 1000

In the non-perforation case, a fabric layer was the basic unit of Twaron woven panel,
which was modelled at the fabric level. According to the mass and volume of fabric,
the volume density of 750 kg/m3 was calculated for Twaron woven fabric.

The dynamic mechanical properties of Twaron fabric were highly strain- rate
dependent. In this study, the armour panel experienced very high stain rates (1000 s-1 )
under ballistic impact. The material properties of Twaron fabric at such high strain
rates cannot be measured. Therefore, the material properties of fabric that measured
at low impact speed were used for FE modelling here. According to Shim‟s tests [32],
when the strain rates increased from 10-2 to 495s-1 , the tensile strength increase from
0.507 to 1.23GPa, but the failure strain decreased from 5% to 3%. In this study, the
tested tensile strength of Twaron fabric is 0.78GPa and the failure strain is 3.23% as

107
shown in Table 3.5. For the FE model of Twaron fabric, the tensile strength 1.2GPa
was assigned as the failure stress and the failure strain was assumed to be 3.5%. High
yield stress and large failure strain will lead to higher energy absorption in the fabric.

The fabric was assumed to be homogeneous, transverse isotropic and elastic-plastic.


For the material direction, the 1- and 2-direction are assumed in the plane of fabric
along the warp and weft direction respectively. 3-direction is perpendicular to the
plane of the fabric. The behaviour of the material can be defined by specifying nine
engineering constants, three tensile moduli E11 , E22 and E33 ; Poisson‟s ratio v 12 , v13
and v 23 ; and the shear moduli G12 , G13 and G23 associated with the material‟s
principal directions.

(4-12)

As the same yarns were used in warp and weft direction, the material properties of
the Twaron fabric in the 1- and 2- direction were assumed to be the same.
Perpendicular to the plane of fabric, tensile properties are much lower. Therefore, in
this study, two longitudinal tensile moduli (E11 and E22 ) were assumed to be the same,
which are higher than the transverse modulus (E33 ). The two out-of-plane shear
moduli (G13 and G23 ) were assumed to be the same, which are higher than the in-plan
shear modulus (G12 ).

According to the tensile test results that listed in Table 3.5 (Chapter 3, Section 3.1.2),
the tensile modulus of 20GPa was assigned for the longitudinal tensile moduli of
Twaron fabric model. The transverse modulus (E33 ) was assumed to be one order of
magnitudes smaller than E11 [138]. The in-plan shear modulus (G12 ) was assumed to
be 0.07 GPa according to the tensile test results in Table 3.5. The shear moduli G13
and G23 can be approximately estimated to be two orders of magnitudes higher than
G12 [138]. Rao [131] proposed that the Poisson‟s ratio of the fabric can be set as zero
due to the fabric was the assembly of loose individual fibres. Therefore, the
Poisson‟s ratio of the fabric model was set as 0.01. All the parameters of the Twaron
fabric listed in Table 4.4 were modified within the same order of magnitude. Ductile

108
damage was defined as the material damage mode. The damage evo lution law was
specified in terms of fracture energy. The softening was defined as linear form.

(4-13)

Table 4.4 Material properties used in woven and UD models

Material E11 E22 E33 G12 G13 G23


v12 v13 v23
properties (GPa) (GPa) (GPa) (GPa) (GPa) (GPa)
Twaron
20 20 2 0.01 0.01 0.01 0.07 1.5 1.5
woven fabric
Dyneema UD
40 40 4 0.01 0.01 0.01 0.7 10 10
laminate

As the fabric exhibited different yield behaviours in different directions, Hill‟s


potential function was specified for the plastic definition in the material model. Hill
yield surfaces assume that yielding of the material is independent of the equivalent
pressure stress. It can be used to model anisotropic yield behaviour [128].

f   F    G    H  11   
2 2 2 2 2
22
 33 33
  11 22
2L 23
 2M 31
 2 N  12 (4-14)

where F, G, H, L, M, N are constants obtained from the results of tests carried out on
the material, in various directions. They are defined as:

 
2
0
 1 1 1  1 1 1 1 
F          2  (4-15)
 2   
  22  33  11
2 2 2 2
2  2  R 22 R 33 R 11 

 
2
0
 1 1 1  1 1 1 1 
G          2  (4-16)
 2   
  33  11  22
2 2 2 2
2  2  R 33 R 11 R 22 

 
2
0
 1 1 1  1  1 1 1 
H           (4-17)
 2   2 
  11  22  33
2 2 2 2
2  2  R 11 R 22 R 33 

2
3  
0
3
L     (4-18)
 
2 
2
23  2 R 23

2
3  
0
3
M     (4-19)
 
2   13
2
 2 R 13

2
3  
0
3
N     (4-20)
 
2   12
2
 2 R 12

109
Where each ̅ is the measured yield stress value when is applied as the only

nonzero stress component; is the user-defined reference yield stress specified for

the plasticity definition and  0



0
/ 3 . 11 , 22 , 33 , 12 , 13 , and 23 , are
anisotropic yield stress ratios in each direction. For engineer materials, they are
usually 1/20-1/50 of tensile moduli and shear moduli [128]. The six yield stress
ratios are defined as follows.

̅ ̅ ̅ ̅ ̅ ̅
, , , , ,
Table 4.5 Yield stress ratios of Twaron fabric and Dyneema UD

Material R 11 R22 R33 R12 R 13 R23

Twaron fabric 0.075 0.075 0.408 0.167 0.025 0.025


Dyneema UD 0.084 0.084 0.039 0.018 0.015 0.015

4.3.4 Dyneema UD laminate

In this study, the non-perforated Dyneema panel was modelled at the fabric level.
UD laminate contains several plies of uni-directional fibres oriented at 0°and 90°.
Therefore, the in-plane stiffness and strength in the 0° and 90° is approximately
equal. The stiffness and strength through thickness are much lower (<10%) than
those in-plane. Therefore, the FE model of Dyneema UD was simulated as an
transverse isotropic model [132].

As material properties of the Dyneema SB71 are unavailable in the published


literature, the material properties of Dyneema UD assigned for the FE model were
referred to the test results as shown in Table 3.5. Due to the inevitable slippage of the
sample during the tensile test, the tested tensile strength (0.91GPa) of Dyneema UD
decreased and the failure strain (6.33%) increased. According to Russell‟s tensile test
[34], the ultimate strength of the laminate was around 20% lower than that of yarn
and the yarn strength was 20% lower than that of fibre. For Dyneema fibres used for
ballistic impact materials, the tensile strength is around 3.3-3.9GPa (as shown in
Table 2.1). The tensile strength of Dyneema UD is estimated to be 2.1-2.5GPa.
Based on the test results in this study and previous study [34], the failure strength
was assumed to be 1.5GPa for this FE model. The failure strain of laminate was

110
assumed as the same as that of Dyneema fibres 4%. High tensile strength and large
failure strain will lead to higher energy absorption in UD laminate.

According to the test results as shown in Table 3.5, the initial tensile modulus of
40GPa was used for E11 and E22 in 0°and 90°directions in the plane of Dyneema
UD. The transverse modulus E33 was assumed to be 1/10 of the in-plane modulus E11
and E22 . The in-plan shear modulus G12 (0.7GPa) was referred to shear frame test
results as shown in Table 3.5. The out-of-plane shear modulus G13 and G23 was one
order of magnitude higher than the G12 . The material properties of Dyneema UD
model were listed in Table 4.4. Ductile damage was defined as the material damage
mode. The damage evolution law was specified in terms of fracture energy. The
softening was defined as linear form.

Mass density of Dyneema UD SB71 was calculated to be 970kg/m3 by the mass and
volume at the unit areal density. According to previous study, the friction coefficient
of Dyneema UD was found to be lower than that of Twaron woven fabric [24]. It was
assumed to be 0.15 in the FE model.

4.3.5 Clay

Clay was used at back of panels in non-perforation tests to memorize the


configuration of the indentation in its surface after impact. Roberts et.al assumed
material properties of the clay as the large-displacement, nonlinear behaviour [133].
According to material characterization tests in a study by Pamukcu, remoulded clay
behaves more like the von Mises material model [134]. It displayed no volume
change upon yield. The elastic and yield properties of the clay are strain- rate
dependent. Therefore, in this study, the clay (RP#1) was modelled as an isotropic,
homogeneous and also elastic-plastic material. The thermal responses of the clay
were considered. Material properties and the hardening response of clay were
referred to the tests in the research of Pamukcu‟s listed in Table 4.6 and 4.7.

Table 4.6 Material properties of clay [134]

Material Clay
Elastic modulus E (MPa) 6.58
Poisson‟s ratio 0.496
3
Density ρ (kg/m ) 1539
111
Yield stress σ (KPa) 60
Angle of internal friction 61o
Flow stress ratio 1.0
Dilation angle 0o

Table 4.7 Hardening properties [134]

Yield Stress (KPa) Plastic strain


60 0.0
72 0.000119
96 0.000128
120 0.000159
144 0.000188
240 0.000263
360 0.000343
479 0.000462

4.4 FE results and validation

4.4.1 Single layer fabric under impact

The FE model of a single layer fabric was validated according to performance of


fabric and ballistic response during ballistic impact process, including energy
absorption, fracture time, fractured yarns, stress distribution and transverse
deformation.

1) Energy absorption

The energy absorption in fabric is one of the most vital validation criteria for FE
model. The energy absorbed by the fabric was calculated from the striking velocity
and exiting velocity of the projectile. According to the FE results, the energy
absorption of a single layer fabric 8F and 11F were both close to the test results as
shown in Fig. 4.10.

112
9
FE simulation results
8
ballisitic test results
7

Energy absorption (J)


6
5
4
3
2
1
0
8F 11F
Fig. 4.10 FE results of a single layer Twaron fabric

The error of energy absorption in FE models was mainly due to two factors. First, the
friction between filaments was not taken account in the model due to the simulation
at the yarn level. This leads to the less friction energy absorption in the fabric in t he
FE simulation. In addition, yarn slipping from the frame was inevitable during
ballistic tests. This leads to the increasing energy absorption in fabric. As a result, the
energy absorption of the FE model was a little lower than that of tests results.

2) Fracture time and broken yarns

The fracture time of woven fabric can be approximately estimated from pictures
taken by high speed camera. For one example of fabric 8F, fractured yarns can be
observed on fabric at 11µs as shown in Fig. 4.11. Therefore, the fracture time of the
fabric must be less than 11µs.

(a) t=0 µs (b) t=5.5 µs (c) t=11 µs (d) t=16.5 µs


Fig. 4.11 Fracture time of a single layer Twaron fabric under impact

In the FE simulation, yarns in fabric failed one by one at different times. Therefore,
the fracture occurred over a period defined from the break of the first yarn to the
complete perforation of fabric. For Twaron woven fabric 8F, the fracture time was
from 2.4µs-7.2µs. For Twaron fabric 11F, the fracture time was from 1.5µs-4.4µs.
The complete perforation time of these two fabric models were both consistent with
the estimated fracture time (before 11µs) during ballistic test.
113
In the FE model of Twaron fabric, there were 2.5 primary yarns in a quarter-model as
shown in Fig. 4.12 (a). So in a whole FE model, there should be 5 fractured primary
yarns. This is consistent with the number of broken yarns in the post- impact fabric,
which usually had 5-6 fractured primary yarns as shown in Fig. 4.12 (b).

(a) FE model (b) post-impact fabric


Fig. 4.12 Fractured yarns in fabric 8F (a) FE model; (b) post-impact fabric

4) Stress wave and transverse deformation

Fig. 4.13 shows FE simulation of Twaron fabric 8F under ballistic impact. When the
projectile impacts the fabric, the fabric develops a transverse deflection. The yarns in
the impact area were pulled out of the original fabric plane. The pyramid shape of the
deformed area with a rhombus base was well represented as shown in Fig. 4.13 (b).
Stress is generated on the primary yarns and some secondary yarns. The coloured
area indicates the stress distribution. Different colours represented different stress
magnitudes.

(a) Fabric under ballistic impact

114
(b) Impact area
Fig. 4.13 FE simulation of Twaron fabric 8F under ballistic impact

The detailed stress distribution at different times can be visualized by stress contours
as shown in Fig.4.14. At the early contact time of 1.2µs, the stress waves are
generated from the impact point and propagated down the axis of the primary yarns.
At 2.4µs, the high stress was mainly concentrated in the impact area just around the
edge of the projectile. Due to the interaction between primary yarns and secondary
yarns at crossovers, the stress wave also developed in the secondary yarns, which can
be indicated from the varied colours on secondary yarns around the impact area as
shown in Fig.4.14 (b). When the stress magnitude of some elements exceeds the
yield stress value, the elements were deleted, which indicated that yarn fracture had
begun. In this model, the fabric began to fracture from 2.4µs to7.2µs. It can be seen
that at 4.8µs, the first primary yarn was fractured as shown in Fig.4.14 (c). At 7.2µs,
the fabric was perforated with the central part separation as shown in Fig.4.14 (d).

(a) t=1.2µs
115
(b) t=2.4µs

(c) t=4.8µs

(d) t=7.2µs
Fig. 4.14 Stress contours of Twaron fabric 8F at different impact time
116
Under ballistic impact, the stress and strain distribution of one primary yarn in the
fabric is plotted at different times as shown in Fig.4.15. Before the yarn fractures, the
impact area always had the highest stress and the largest strain. With the increasing
impact time, the magnitude of the stress and strain on the primary yarn away from
impact point gradually increased. In addition, the area of the stress wave propagation
enlarged. For example, at 1.2µs, the stress is concentrated around the edge of the
projectile. At 7.2µs, the stress wave has propagated almost to the edge of the fabric.
This indicates that more fabric material is under strain and producing strain energy.

3.0
t=1.2µs
2.5 t=4.8µs
2.0 t=7.2µs
Stress(GPa)

1.5

1.0

0.5

0.0
0 1 2 3 4 5 6 7 8
Distance from the impact point (cm)

(a)Stress distribution
4.0
t=1.2µs
3.5
3.0 t=4.8µs
t=7.2µs
Strain (%)

2.5
2.0
1.5
1.0
0.5
0.0
0 1 2 3 4 5 6 7 8
Distance fromt the impact point (cm)

(b) Strain distribution


Fig. 4.15 Stress and strain distribution of Twaron fabric 8F

During the impact process, the transvers deformation was produced in fabric before
fracture. According to observation of the impact process using a high speed camera,
117
the impact area of the fabric was shown in Fig. 4.16 (a). Every grid marked on the
fabric 8F was a square with the side length of 10mm. At 5.5μs, the maximum width
of the deformed area was around 10mm. In the FE model, the transverse deformation
of fabric at 5.6μs was displayed in Fig. 4.16 (b) and (c). The maximum width of the
deformed area was 13.8mm. The model simulated the characteristics of the fabric
transverse deformation in ballistic test very well.

10mm

13.8mm
(a) t=5.5 µs (b) t=5.6 µs
0.05
6.9mm
Transverse deflection (cm)

0
0 1 2 3 4 5 6 7 8
-0.05

-0.1

-0.15

-0.2
t=5.6 µs
-0.25
Distance from the impact point (cm)

Fig. 4.16 The transverse deformation of fabric 8F at 5.6µs

According to above analysis, FE simulation results showed a good agreement with


test results including energy absorption, stress and strain distribution, transverse
deflection and other failure characteristics of post-impact fabric. The results
indicated that the FE model of single layer Twaron woven fabric was valid. Thus this
model can be used to analyse the ballistic response of woven fabric under ballistic
impact. It can also be developed into the FE model of multi- layer panels in the
perforation case.

4.4.2 Perforated multi-layer panels

To ensure the implement of FE models of the multi- layer system, mesh elements
were reduced by two methods, namely model size reduction and mesh density

118
reduction. The better meshed method has the higher accuracy in predicting energy
absorption.

1) Model size reduction

According to the FE simulation results for single layer fabrics with different sizes,
energy absorption decreased significantly with the model size reduction as shown in
Fig. 4.17. It was speculated that this was because the small fabric model fractured
more quickly than the fabric models with the larger sizes, which resulted in a shorter
response time for energy absorption. In addition, the fabric with smaller size engaged
less material in energy absorption.

6.00 8.0
Fracture time (µs)
7.0 Energy absorption (J)
5.00
Energy absorption (J)

6.0

Fracture time (µs)


4.00
5.0
3.00 4.0
3.0
2.00
2.0
1.00
1.0
0.00 0.0
7.5cm 5.0cm 2.5cm

Fig. 4.17 Energy absorption in three sample sizes

To identify the effect of model size on energy absorption of the multi- layer system,
the small model size of 2.5cm was used for FE models of panels 8F 3 , 8F6 , and 8F9 .
FE simulation results indicated that the difference between energy absorption in FE
models and tests results had an increasing trend with more layers added in panels as
shown in Fig. 4.18.

50
Test results
FE simualtion
Energy absorption (J)

40

30

20

10

0
8F8F 8F3
8F3 8F6
8F6 8F9
8F9
Fig. 4.18 Energy absorption of Twaron panels with the small model size
119
Although the model size reduction can save computer costs, energy absorption in FE
models of multi- layer systems has been greatly affected. Therefore, this method is
not appropriate for FE modelling of the multi- layer system.

2) Mesh density reduction

Hybrid mesh was another measure used to reduce elements in the FE models of
multi- layer systems. A fine mesh density was used for primary yarns and a coarse
mesh density was used for secondary yarns in the fabric. For a single layer fabric, the
energy absorption (5.06J) of the FE model with hybrid mesh was almost the same as
that of the FE model with fine mesh (5.25J) as shown in Fig. 4.19. For the multi-
layer panel 8F3 , 8F6 and 8F9 , energy absorption of FE models with hybrid mesh was
more close to tests values as shown in Fig. 4.20.

6.00

5.00
Energy absorption (J)

4.00

3.00

2.00

1.00

0.00
FE model with uniform mesh FE model with hybrid mesh
Fig. 4.19 Energy absorption of the FE model with different mesh
50
Tests results
Energy absorption (J)

40
FE simulation

30

20

10

0
8F
1-layer 8F3
3-layer 8F6
6-layer 8F9
9-layer
Fig. 4.20 Energy absorption of Twaron panels with the hybrid mesh

With more layers added in the panel, the standard deviation of energy absorption in
FE models was inevitably enlarged. However, the FE model with hybrid mesh
showed higher accuracy of the energy absorption by comparing FE models with the

120
reduced model size. Therefore, the FE model with hybrid mesh was adopted to
analyse the ballistic response of the multi- layer panel in the perforation case.

4.4.3 Non-perforated multi-layer panels

In the non-perforation case, the number of perforated layers and BFS of the armour
panel were used as validation criteria of the FE model.

For the non-perforated Twaron woven panel 11F 24 , the number of perforated layers
was in the range of 7-10 layers in ballistic tests. While in the FE model the first
seven layers were perforated, which is consistent with test results.

When the projectile was stopped in the armour panel, a n indentation was left in the
clay as shown in Fig.4.21. According to the ballistic test results of Twaron panel
11F24 , the width of the indentation in the clay was 40-45mm and BFS was around
13.57-17.45mm.

40-45mm

13.57-17.45mm

Fig. 4.21 The clay behind the Twaron panel 11F24

Fig. 4.22 FE model of the clay behind the Twaron panel 11F24

The FE model of the clay showed almost the same configuration of the indentation as
that for ballistic tests as shown in Fig. 4.22. The half width of the indentation in the
clay was 22mm. This was consistent with the test results. BFS of the FE model was

121
measured as 8mm. The difference of BFS from the test results was due to the
additional impact force from the jacket outside the projectile in the ballistic test.

In most ballistic tests, the indentation was impacted by the projectile followed by the
jacket at the same position. In an accidental test, an individual indentation that
produced by the jacket was observed with the depth of 3.8mm as shown in Fig. 4.23.
When this value was considered, BFS of the FE model of Twaron panel 11F 24 was
close to the experiment results.

Fig. 4.23 The indentation produced by the jacket

4.5 Summary

This chapter described the numerical modelling of armour panels under ballistic
impact in this study. Several panels in perforation and non-perforation cases were
simulated by FE models.

Perforated panels with a single layer fabric and multiply layers were mode lled at the
yarn level. The yarn geometry in Twaron woven fabric was analysed first. Due to the
limit of the computer CPU, multi- layer panels adopted hybrid mesh densities with
fine mesh in primary yarns and coarse mesh in secondary yarns to reduce the
elements of models. Non-perforated panels were simulated at the fabric level. A
single fabric layer was simulated as a solid plate. A certain number of plates was
assembled and placed before the clay to form the non-perforated ballistic case.
122
Material properties was assigned and material models was selected for Twaron yarns,
fabric, UD laminate, projectile and clay.

The validation of the FE models was achieved by comparison with empirical test
results. The validation criteria included energy absorption, fracture time, transverse
deformation, stress and strain distribution, the number of perforated layers and BFS.
It was confirmed that these models were successful and can be used to analyse the
internal ballistic characteristics of soft armour panels.

123
CHAPTER 5 Exploration of the ballistic responses of the
multi-layer system

Following experimental studies and FE modelling, results analysis is developed in


chapter 5, chapter 6, and chapter 7. In this chapter, ballistic response of the multi-
layer system is investigated. According to previous studies, the ballistic resistant
efficiency of each layer has been shown not to be uniform, although this has not
clearly quantified. Therefore, combining the same fabric layers in a panel can‟t be
the most efficient in providing the better ballistic performance. In order to achieve
further improvement of ballistic performance and reductions in weight, optimising
the construction of armour panels by combining different components will be an
efficient and feasible option. To optimise the construction of panels for soft body
armour, some fundamental ballistic responses of each layer in the multi- layer system
should be understood firstly. This mainly includes different ballistic characteristics of
each layer, energy absorption distribution in the multi- layer system, related Backface
Signature (BFS), and failure modes of the armour panels under ballistic impact.
These fundamental understanding will provide a guide to develop the effective
hybrid design of armour panels.

In this chapter, some ballistic responses of Twaron woven panels are investigated.
Due to different boundary conditions, the perforated and non-perforated Twaron
woven panels are discussed in Section 5.1 and 5.2 respectively. According to
experiment tests results and FE results, different ballistic characteristics of each layer
are analysed and the energy absorption distribution in the multi- layer panels are
identified. A summary is presented in the Section 5.3.

5.1 The perforated woven panel

To identify ballistic response of each layer in perforated panels, woven panels 8F,
8F3 , 8F6 , and 8F9 were experimentally and numerically analysed , including the
fracture time, transverse deflection, stress distributions and energy distribution.

5.1.1 Fracture time

When a multi- layer panel was under ballistic impact, the projectile penetrated fabric
layers from the front layer on the striking face to the back layer on the exiting face.
124
According to FE results, with the increasing number of layers in armo ur panels 8F,
8F3 , 8F6 , and 8F9 , the projectile was decelerated more significantly. This can be seen
from the reduction rate of the residual velocity of the projectile through panels as
shown in Fig. 5.1.

490
Residual velocity (m/s) 480 8F
8F

470 Vr=472.02m/s 8F3


8F3

460 8F6
8F6
450 Vr=453.18m/s
8F9
8F9
440
Vr=436.19m/s
430
420
410
Vr=405.89m/s
400
0 5 10 15 20
Time (µs)
Fig. 5.1 The residual velocity history of Twaron panel

In FE simulation, for a single layer fabric, the fracture time of the fabric is a period
from the break of the first yarn to the complete fracture of the fabric. For a multi-
layer panel, the fracture time is a period from the fracture of the first layer to the last
layer. The fracture time of each layer in the multi- layer panel was determined by the
time corresponding to the maximum strain energy produced in the layer.

According to FE results, with more layers added in the panel, the fracture time of the
multi- layer panel increased as shown in Table 5.1. This indicated the interaction time
between the panel and the projectile increased with more additional layers.

In a panel, the fracture time of each layer was not the same. The back layers can
sustain longer than the front layers. For example, the first layer in the panel 8F 9 was
perforated at 0.3µs, while the last layer can sustain 11.4µs. In addition, with more
layers added in the panel, the front layer failed more quickly. In panels 8F and 8F 3 ,
the first front layer began to fracture at around 2µs. When the number of layers
increased to six and nine in panels of 8F 6 and 8F9 , both of the first front layers were
perforated at 0.3µs. Such results indicated that even if the same fabric material was
combined in the perforated panels, the sustain time of each layer was not the same.

125
Table 5.1 Fracture time of each layer in perforated Twaron panels
Fracture time
Sample The number Fracture time of
Layer of each layer
label of layers the panel (µs)
(µs)
8F 1 2.4-7.20 / 2.4-7.20
ply-1 1.98
8F3 3 1.98-7.02 ply-2 6.90
ply-3 7.02
ply-1 0.30
ply-2 5.33
ply-3 5.87
8F6 6 0.30-7.73
ply-4 7.20
ply-5 7.65
ply-6 7.73
ply-1 0.30
ply-2 1.20
ply-3 6.00
ply-4 7.27
8F9 9 0.30-11.40 ply-5 7.80
ply-6 8.30
ply-7 8.40
ply-8 9.50
ply-9 11.40

5.1.2 Transverse deformation

Under ballistic impact, fabric layers in the panel produced the transverse deflection
out of fabric and deformation in the plane of fabric. The evolution of the transverse
deformation of fabric layers was hard to be observed in the ballistic tests, but it can
be clearly identified and quantified by FE simulation. The transverse displacement of
each node on the centre of a primary yarn beneath the projectile was output and
plotted to represent the transverse deflection of each layer in a panel.

Fig. 5.2 showed the maximum transverse deformation before fracture of each layer in
four panels, which were related to different times during the impact process. Under
ballistic impact, each fabric layer produced different extents of transverse
deformation before fracture. With the position from front to back in the panel, the
maximum transverse deformation enlarged. The back layers can produce an even
wider transverse deformation area and much larger transverse deflection out of fabric

126
than that of the front layers. As a result, more fabric materials in back layers were
involved being in strain than that of front layers during the ballistic impact process.

0.0 0.5 1.0 1.5 2.0


0.0
ply1-7.4µs
Transverse deflection (cm)

-0.5

-1.0

-1.5

-2.0
Distance from the impact point (cm)

(a) 8F

0.0 0.5 1.0 1.5 2.0


0.0
ply1-1.0µs
ply2-6.0µs
Transverse deflection (cm)

ply3-7.0µs
-0.5

-1.0

-1.5

-2.0
Distance from the impact point (cm)

(b) 8F3

127
0.0 0.5 1.0 1.5 2.0
0.0
ply1-1.0µs

Transverse deflection (cm)


ply2-2.5µs
ply3-4.5µs
-0.5
ply4-5.0µs
ply5-7.0µs
ply6-8.0µs
-1.0

-1.5

-2.0
Distance from the impact point (cm)
(c) 8F6

0.0 0.5 1.0 1.5 2.0


0.0
ply1-1.0µs
ply2-2.5µs
Transverse deflection (cm)

ply3-5.0µs
-0.5
ply4-5.5µs
ply5-8.0µs
ply6-8.5µs
-1.0 ply7-9.5µs
ply8-10.5µs
ply9-11.5µs

-1.5

-2.0
Distance from the impact point (cm)
(d) 8F9
Fig. 5.2 The maximum transverse deflection of each layer in Twaron woven
fabric panels (a) 8F, (b) 8F3 , (c) 8F6 and (d) 8F9

In addition, compared with the single layer fabric 8F, the transverse d eformation of
some front layers in multi- layer panels of 8F3 , 8F6 and 8F9 significantly reduced. In
particular, for the first layer, the maximum transverse deformation was just localised
around the edge of the projectile, which was much smaller than that of the single
layer fabric. For example, in the panel 8F6 , the maximum transverse deflection of the
first layer was 60% less than that of the single layer fabric. The radius of the
128
transverse deformation area was only 27% as that of the single layer fabric. Such
results indicated that in the multi- layer panel, the transvers deformation of front
layers was constrained by back layers.

5.1.3 Stress distribution

In the FE simulation of the multi- layer panel, the stress in different regions of the
fabric is different. The stress of each node on the centre of a primary yarn beneath
the projectile was output to represent the stress distribution of each layer in a panel.
In the panel 8F3 , the stress of each layer was plotted before this layer fractured as
shown in Fig. 5.3, Fig. 5.4 and Fig. 5.5. The characteristics of stress distribution in
each fabric layer were shown different. In the first layer, the high stress concentrated
in an impact area around the edge of the projectile. The fabric failed very quickly
almost at 1.0 µs before the stress wave propagated. In the second layer and the third
layer, the fabric was perforated at around 6.0µs and 7.0µs respectively. The longer
interaction time with the projectile allowed the stress wave to propagate widely
almost to the edge of the fabric.

Ply1-0.5µs

Ply1-1.0µs

(a) Stress contours on the first layer of the panel 8F3


129
3.5
ply1(0.5µs)
3.0
ply1 (1.0µs)
2.5
Stress (GPa)
2.0

1.5

1.0

0.5

0.0
0 1 2 3 4 5 6 7 8
The distance from the center to the edge (cm)

(b) Stress magnitude on the first layer of the panel 8F3


Fig. 5.3 Stress distribution on the first layer of the panel 8F3

Ply2-1.0µs

Ply2-3.0µs

130
Ply2-3.0µs

(a) Stress contours on the second layer of the panel 8F3

3.5
ply2-1.0µs
3.0 ply2-3.0µs
2.5 ply2-6.0µs
Stress (GPa)

2.0

1.5

1.0

0.5

0.0
0 1 2 3 4 5 6 7 8
The distance from the center to the edge (cm)
(b) Stress magnitude on the second layer of the panel 8F3

Fig. 5.4 Stress distribution on the second layer of the panel 8F3

Ply3-1.0µs

131
Ply3-3.0µs

Ply3-7.0µs

(a) Stress contours on the third layer of the panel 8F3

4.0
ply3-1.0µs
3.5
ply3-3.0µs
3.0 ply3-7.0µs
Stress (GPa)

2.5
2.0
1.5
1.0
0.5
0.0
0 1 2 3 4 5 6 7 8
The distance from the center to the edge (cm)
(b) Stress magnitude on the third layer of fabric 8F3
Fig. 5.5 Stress distribution on the third layer of the panel 8F3

During the same impact time, for multi- layer panels with different number of layers,
the stress wave can propagate almost the same distance in the plane of each fabric
layer. As shown in Fig. 5.6, on the first layer in 8F 6 and 8F9 , the distance of the stress
wave propagation from the impact point both were around 0.75cm. However, in the

132
perpendicular direction to the panel, the stress gradient was exhibited on each layer.
At 0.3 µs (just before the first layer perforated), in the panel 8F6 , the stress wave was
already transmitted to the last layer. But for the panel 8F 9 , the last layer of fabric was
not applied any significant stress. Such results indicated that with the increasing
number of layers in a panel, the difference of the stress distribution on each layer will
become more prominent.

0.75cm

(a) 8F6

0.75cm

(b) 8F9
Fig. 5.6 Stress gradient on panels of (a) 8F6 , and (b) 8F9 at 0.3µs under impact

133
5.1.4 Energy absorption mechanisms

5.1.4.1 Energy absorption distribution

According to ballistic test results, energy absorption of armour panels increased with
the increasing number of layers as shown in Fig. 5.7. However, when the areal
density was taken into account, the specific energy absorption (SEA) decreased with
more layers added in panels as shown in Fig. 5.8. This indicated that the energy
absorption of the whole panel was not multiplied with the increasing number of
layers, which meant not every layer applied the same energy absorption capacity in a
panel as a single fabric layer.

60
8F 11F
50
Energy absorption (J)

40

30

20

10

0
Single layer 3-layer 6-layer 9-layer

Fig. 5.7 The energy absorption of woven fabric

500
Twaron 8F
SEA of panel (J.cm2/g)

400 Twaron 11F

300

200

100

0
0 1 2 3 4 5 6 7 8 9 10
The number of layers
Fig. 5.8 The specific energy absorption (SEA) of woven fabric

To identify different ballistic resistant efficiency of each layer in the multi- layer
panel, energy absorption of each layer was quantified by FE simulation. At a start
point, energy absorption mechanisms in FE model should be clarified at first. In FE
model, there is an energy balance for the entire model as following equation.
134
where is the internal energy, is the viscous energy dissipated, is the
frictional energy dissipated, is the kinetic energy, is the work done by the
externally applied loads, and , , and are the work done by contact
penalties, by constraint penalties, and by propelling added mass, respectively. The
sum of these energy components is . In the numerical model is only
approximately constant.

The internal energy is the sum of the recoverable elastic strain energy. It concludes
the energy dissipated through inelastic processes such as plasticity, the energy
dissipated through viscoelasticity or creep and the artificial strain energy. The
viscous energy is the energy dissipated by damping mechanisms, including bulk
viscosity damping and material damping. When a body is applied a force, the
external work is defined entirely by nodal forces and displacements. When a body is
applied the torque, the external work is defined by nodal moments and rotations.

During the ballistic impact, the kinetic energy of a projectile was mainly converted
into the fabric kinetic energy, yarn strain energy, frictional dissipation energy
between yarns and projectile-yarns, as well as other forms of dissipation energy. In
the FE model of a single layer fabric, the strain energy, kinetic energy and frictional
dissipation energy were the dominant forms of energy absorption of the fabric as
shown in Fig. 5.9, which was taken above 80% of total energy absorption in fabric.

100%
Other energy dissipation (%)
Energy absorption ratio (%)

90%
80% Frictional dissipation energy (%)
70% Yarn strain energy (%)
60% Yarn kinetic energy (%)
50%
40%
30%
20%
10%
0%
0.1 0.7 1.4 2.1 2.8 3.5 4.2 4.9 5.6 6.3 7.0 7.7
Time (μs)
Fig. 5.9 The proportion of the dominant forms of energy absorption in fabric 8F

In the FE simulation of the multi- layer system, the frictional energy dissipation is
correspond to the total amount of the integral system. The frictional energy
135
dissipation of each layer cannot be output. So only the kinetic energy and strain
energy of each layer were used to analyse the energy absorption distribution in a
panel.

3.0
Kinetic energy
Energy absorption (J)
2.5
Strain Energy
2.0

1.5

1.0

0.5

0.0
Single layer ply-1 ply-2 ply-3
The ith layer in the panel
(a) 8F3
3.0

2.5 Kinetic energy


Strain Energy
Energy absorption (J)

2.0

1.5

1.0

0.5

0.0
Single 6ply-1 6ply-2 6ply-3 6ply-4 6ply-5 6ply-6
layer
The ith layer in the panel
(b) 8F6
3.0

Kinetic energy
2.5
Strain Energy
Energy absorption (J)

2.0

1.5

1.0

0.5

0.0
Single 9ply-1 9ply-2 9ply-3 9ply-4 9ply-5 9ply-6 9ply-7 9ply-8 9ply-9
layer th
The i layer in the panel
(c) 8F9
Fig. 5.10 Energy absorption of each layer in panels of (a) 8F3 , (b) 8F6 and (c) 8F9

136
According to FE results of the multi- layer panels 8F3 , 8F6 , and 8F9 , energy
absorption of each layer was not uniform as shown in Fig. 5.8. It increased from
front to back layer in all three panels. The back layer always had the highest energy
absorption in the perforated panels.

Compared with the single layer fabric 8F, the energy absorption capacity of front
layers in the multi- layer panel decreased. For example, in panels 8F 3 , 8F6 , and 8F9 ,
energy absorption of the front layer was only about 30% as that of a single layer
fabric. While with more layers added in the panel, energy absorption of the last layer
in panels 8F6 and 8F9 can exceed than that of a single layer fabric. This can be
interpreted due to different characteristics of transverse deformation and stress
distribution of each layer in the multi- layer panel under impact. With the increasing
number of layers in the panel, the transverse deformation of front layers was
constrained by back layers. So less materials in front layers was involved into
deformation into energy absorption, while back layers can sustain longer time during
impact and produced larger transverse deformation area. As a result, more fabric
materials were in strain and produced more energy absorption.

In addition, with more layers added in a panel, more front la yers were constrained. In
the six- layer panel 8F6 , only the first two layers had lower energy absorption. While
in the nine- layer panel 8F9 , the front three layers were constrained with lower energy
absorption as shown in Fig. 5.10 (b) and (c). This indicated that the constraint effect
to front layer by back layers became more significant with the increasing number of
layers in the panel. However, the constraint effect to the front layer will reach a limit
value. For example, both the first layers in panels 8F6 and 8F9 had the same amount
of the energy absorption regardless of the total number of layers.

5.1.4.2 Energy absorption efficiency of each layer

According to FE results, compared with a single layer fabric, when the same fabric
material was combined in different positions in the multi- layer panel, energy
absorption efficiency of each fabric layer was varied. To quantify such variation of
energy absorption efficiency of each fabric layer in a panel, the energy absorption
ratio R was adopted in this study. The energy absorption of a single layer was the
reference baseline.

137
(7-1)

where R is the energy absorption efficiency ratio, is the total energy absorption
of the ith layer in a panel, is the energy absorption of a single layer fabric.

When R<100%, it indicates that the energy absorption capacity of the fabric layer
was constrained in the panel. When R=100%, it means that the fabric can make full
use of the energy absorption efficiency when it is placed in a panel. When R>100%,
it represents that the energy absorption efficiency of the fabric layers is improved.

160
8F
8F33
Energy absorption ratio (% )

140 8F
8F66
120 8F
8F99

100
80
60
40
20
0
0 1 2 3 4 5 6 7 8 9 10
The ith layer in the panel

Fig. 5.11 Energy absorption efficiency of each layer in panels 8F3 , 8F6 and 8F9

Fig. 5.11 showed the energy absorption efficiency of each layer in perforated panels
8F3 , 8F6 , and 8F9 . Compared with the single layer fabric 8F under impact, the energy
absorption efficiency of the first few layers in all panels decreased significantly. In
particularly, the energy absorption efficiency of the first layer only was 30% as that
of the single layer fabric. In addition, with more layers added in the panel, the
number of the front constrained layers increased. For example, in the panel 8F 3 , only
the first layer had R less than 100%. In panels 8F6 , and 8F9 , the energy absorption
efficiency of the front five and seven layers decreased respectively.

When the same fabric was placed at the back of the panel, the energy absorption
efficiency exceeds 100%. It indicated that energy absorption efficiency of back layer
was improved. This is due to the back layers can sustain longer and have more
freedom to produce transverse deformation. Therefore, in the perforation case, the
back position in an armour panel was beneficial to make best use of the energy
absorption efficiency of the fabric.
138
5.2 The non-perforated woven panel

To reflect energy absorption distributions in the non-perforated panel, ballistic


response of each layer in Twaron woven panel 11F 24 were experimentally and
numerically analysed, including transverse deformation, stress distribution, energy
absorption, the number of broken layers and Backface Signature (BFS).

5.2.1. Transverse deformation

When the non-perforated Twaron panel 11F 24 was under ballistic impact, the
perforated layers produced the maximum transverse deformation just before the
fracture of this layer. While the non-perforated layers can continue to produce the
transverse deformation until the moment of the projectile stopped.

-4 -3 -2 -1 0 1 2 3 4
1
ply1(6.0µs)
ply3(9.5µs)
Transverse deflection (cm)

0
ply6(16.5µs)
ply9(42.0µs)
-1
ply12(46.0µ s)
ply18(46.0µ s)
-2
ply24(46.0µ s)

-3

-4
Distance from the center (cm)
Fig. 5.12 The maximum transverse deflection of each layer in Twaron panel
11F24

According to FE results, each fabric layer in the non-perforated panel had different
extents of the maximum transverse deformation as shown in Fig. 5.12. For the front
few layers (ply-1, -2 and -3), the transverse deformation area in the fabric was just
localised around the edge of the projectile. While middle layers had an obvious wider
transverse deformation area. The transverse deformation area in the back layers
gradually reduced. Such difference of the transverse deformatio ns that produced in
each layer can be directly observed from the post- impact panel as shown in Fig. 5.13.
It can be inferred that with sufficient layers added in the panel, the back layers would
not have any transverse deformation due to the attenuation of the stress perpendicular
139
to the panel. Therefore, in the non-perforated panel, more fabric material in middle
layers was involved into the transverse deformation area and being in strain.

Ply-12 Ply-24

Fig. 5.13 The transverse deformation of fabric layers in Twaron panel 11F24

5.2.2 Stress distribution

According to FE results, stress distribution of some fabric layers in the non-


perforated panel was plotted as shown in Fig. 5.14. The stress distribution on front,
middle and back layers exhibited different characteristics.

On the front few layers, such as the front three layers, high stress concentrated in the
contact area around the edge of the projectile. It increased sharply and reached the
yield stress during less than 10 µs as shown in Fig 5.14 (a). When the stress exceeded
the yield stress of the material, the yarns were fractured. As the front layers were
perforated quickly, the stress cannot propagate out of the impact area.

1.4
ply1(6.0µs)
1.2
ply2(8.5µs)
1.0
Stress (GPa)

ply3(9.5µs)
0.8
0.6
0.4
0.2
0.0
0 2 4 6 8
Distance from the center to the edge (cm)
(a) The front layers

140
1.4
ply4(10.0µs)
1.2
ply6(16.0µs)
1.0

Stress (GPa)
ply8(21.5µs)
0.8
0.6
0.4
0.2
0.0
0 2 4 6 8
Distance from the center to the edge (cm)
(b) The middle layers
1.4
ply12(21.8µ s)
1.2
ply18(21.2µ s)
1.0
ply24(19.0µ s)
0.8
Stress (GPa)

0.6
0.4
0.2
0.0
0 2 4 6 8
Distance from the center to the edge (cm)
(c) The back layers
Fig. 5.14 The stress distribution on each layer in Twaron woven panel 11F24 (a)
front layers, (b) middle layers and (c) back layers

The middle perforated layers can sustain for longer time from 10 µs to 21.5µs as
shown in Fig. 5.14 (b). The fabric had longer responding time to the impact. So the
stress wave had enough time to propagate over a wider area from the impact point to
the edge of fabric. This resulted in more fabric material involved in the transverse
deformation area before fabric fracture and being in strain.

The back layers cannot be perforated and just produced transverse deformation until
the projectile stopped. These layers had even longer interaction time with the
projectile during the impact process. So the stress wave on back layers can propagate
to the edge of the fabric as shown in 5.14 (c). Due to the attenuation of the striking
velocity, the stress magnitude in back fabric layers was found much lower than that
of in the front layers.

The different characteristics of the stress distribution and the transverse deformation
on each layer in the multi- layer panel determined that the energy absorption capacity
of each layer cannot be same.
141
5.2.3 Energy absorption

5.2.3.1 Energy absorption distribution and efficiency

In the case of non-perforation, the projectile was stopped in the panel. Most of the
kinetic energy of the projectile was absorbed in fabric layers regardless of a little
amount was transmitted into the clay behind the panel. Total amount of the kinetic
energy and strain energy of each layer was analysed to identify the energy absorption
distribution in the non-perforated panel.

4.0

3.5
Energy absorption (J)

3.0

2.5

2.0

1.5

1.0

0.5

0.0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24
The ith layer in the panel
Fig. 5.15 Energy absorption of each layer in Twaron woven panel 11F24

Fig. 5.15 showed energy absorption of each layer in Twaron panel 11F 24 . For the
perforated layers, the maximum energy absorption was produced before the layer
fracture. For the non-perforation layers, it corresponded to the moment of the
projectile stopped. In the non-perforated panel, energy absorption in each layer
increased from front layer to the maximum value in the last perforated layer and then
decreased gradually in the following back layers. For Twaron panel 11F24 , the
seventh layer was the last perforated layer. It had the highest energy absorption.
While for the front few layers (front three layers) and some back layers (back ten
layers), energy absorption was rather lower. This is different from that of in the
perforated panel.

According to the equation (7-1), the energy absorption efficiency of each fabric layer
in Twaron panel 11F24 was analysed in detail as shown in Fig. 5.16, which was
compared with the reference fabric 11F.
142
90

Energy absorption efficiency


80
70
60
50

(% )
40
30
20
10
0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24
The ith layer in the panel

Fig. 5.16 Energy absorption efficiency of Twaron woven panel 11F24

According to the energy absorption efficiency of the multi- layer panel 11F24 at the
striking velocity of 483 m/s, the energy absorption efficiency of all layers in the non-
perforated panel was less than 100%. It indicated that the energy absorption
efficiency of all layer was limited. The seventh layer has the highest energy
absorption efficiency in the panel. It was only about 80% as that of the single layer
fabric 11F. While the front two layers and back 14 layers can only apply less than 50%
of the energy absorption efficiency as that of the single layer fabric.

According to above results, when the same fabric layer was combined in the panel,
the energy absorption efficiency of the fabric material was degraded. This was
mainly due to different boundary conditions under impact. In the non-perforation
case, the panel was placed before the clay. All fabric layers in the panel can‟t freely
produce the transverse deformation as that of the single layer fabric. So less fabric
material was involved into energy absorption.

In particular, for the front layers, the transverse deformation was just localised
around the impact area. Correspondingly, the high stress concentration resulted in
fabric layers failing quickly. While for back layers, due to the lower stress magnitude,
material properties of fabric were not fully applied to achieve higher energy
absorption. As a result, energy absorption of these layers was rather lower. For the
non-perforated panel under such ballistic impact conditions, the relatively better
position in the panel that was beneficial for fabric to apply the energy absorption
efficiency can be identified close to the last broken layers. Such fundamental
understanding of the energy absorption distribution in the multi- layer system
provided a vital guide to optimise the construction of an armour panel.
143
5.2.3.2 Influence of the number of layers

To improve the performance of energy absorption capacity of an armour panel, one


of the easiest ways is increasing the amount of material in the armour panel by
adding more layers. To identify the effective construction of the multi- layer system,
the effect of the total number of layers on energy absorption capacity of the non-
perforated panel was investigated. Another two panels 11F36 , and 11F48 with
different layers were used for FE analysis.

4.0

3.5
Energy absorption (J)

3.0

2.5

2.0

1.5

1.0

0.5

0.0
1 3 5 7 9 11 13 15 17 19 21 23 25 27 29 31 33 35
The ith layer in a panel
Fig. 5.17 Energy absorption of each layer in Twaron woven panel 11F36
4.0

3.5
Energy absorption (J)

3.0

2.5

2.0

1.5

1.0

0.5

0.0
1 3 5 7 9 11 13 15 17 19 21 23 25 27 29 31 33 35 37 39 41 43 45 47
The ith layer in a panel
Fig. 5.18 Energy absorption of each layer in Twaron woven panel 11F48

As shown in Fig. 5.17 and Fig. 5.18, when fabric layers in the panel increased to 36
and 48 layers, the pattern of energy absorption distribution in Twaron panels 11F 36
and 11F48 were found to remain the same as that of in the panel 11F 24 . It was not

144
affected by the number of layers in the panel. In another words, for a given fabric
material and structure of the armour panel, the increasing number of layers in the
panel can improve energy absorption, but have no influence on the pattern of energy
absorption distribution.

Table 5.2 Energy absorption of Twaron woven panels

Energy The number of


Multi- Total The bullet broken layers
Total energy absorption
layer number stopping
absorption (J) of front 24 Test FE
system of layers time (µs)
layers(J) results results
24 46.2 57.79 57.79 9 7
Twaron
woven 36 36.5 70.43 50.57 7 7
panel
48 35.9 73.64 47.26 6 7

According to FE analysis results as shown in Table 5.2, when the total number of
layers in the panels increased from 24 to 36, energy absorption of the whole panel
increased 17.95%. When the total number of layers continually increased from 36 to
48, energy absorption only increased 4.56%. Although the same amount of fabric
layers (12 layers) was added in the panel, the increment of energy absorption in the
panel reduced. With more layers added in the multi- layer system, the growth rate of
energy absorption of the whole panel will become less and less.

Due to different positions in the multi- layer system, the same amount of fabric
materials played different roles in energy absorption. For example, in the panel 11F 48 ,
energy absorption of front 24 layers was 64.18% of total energy absorption, which
was almost twice of the back 24 layers. Such results indicated that when the total
number of layers exceeds a certain value. The additional layers in the panel
contribute less to the total energy absorption.

In addition, energy absorption of front layers was constrained and reduced by the
additional layers. According to FE simulation, it was found that energy absorption of
front 24 layers in these three panels decreased with more layers added in the panel as
shown in Fig. 5.19. Comparing with the panel 11F 24 , energy absorption of front 24
layers in the panel 11F36 and 11F48 reduced 12.49% and 18.22% respectively. It
should be noted that with more layers added in the panel, energy absorption of front
layers became less and less. However, the reduction of energy absorption in front

145
layer would reach a lower limit with more additional layers added in the panel. Such
results were consistence with that in the perforated panel.

80
EA of back layers
70
Energy absorption (J) EA of front 24 layers
60
50
40
30
20
10
0
11F24
11F 24
11F36
11F 36
11F48
11F 48

Fig. 5.19 Energy absorption of Twaron panels

5.2.3.3 Effect of the striking velocity

According to previous studies, the energy absorption capacity of an armour panel


was obviously affected by the striking velocity. So the energy absorption distribution
in the panel 11F24 was investigated by FE simulation at different striking velocities,
including the low striking velocity of 300 m/s and the high striking velocity of 600
m/s. Fig. 5.20 showed energy absorption of each layer in the panel 11F 24 at these two
striking velocities.

2.5

2.0
Energy absorption (J)

1.5

1.0

0.5

0.0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24
The ith layer in a panel
(a) At the striking velocity of 300 m/s

146
5.5
5.0
4.5

Energy absorption (J)


4.0
3.5
3.0
2.5
2.0
1.5
1.0
0.5
0.0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24
The ith layer in a panel
(b) At the striking velocity of 600 m/s
Fig. 5.20 Energy absorption distribution at different striking velocities

According to FE results, with the increasing striking velocity, more fabric layers
were perforated due to the increasing impact energy. The energy absorption
distribution in the multi- layer panel still showed the same pattern. The last broken
layers in two panels always had the highest energy absorption at different striking
velocities. At the low striking velocity of 300 m/s, the projectile was stopped before
the panel with the first layer broken. The energy absorption of each layer decreased
from front to back in a panel. When the striking velocity increased to 600 m/s, front
twelve layers in the panel were perforated. The energy absorption increased from
front layer to the maximum value in twelfth layer and then decreased in back layers.

If at even higher striking velocity, the panel is completely penetrated. It becomes the
perforation case. The last layer has the highest energy absorption in a panel. So the
energy absorption distribution of each layer in the panel increases from front to back
layers. Therefore, the energy absorption distribution in the perforated panel can be
regarded as a specific case of the influence of the striking velocity on energy
absorption distribution in the multi- layer system. It followed the pattern of the
perforated layers in the non-perforated panel.

According to above analysis, the energy absorption distribution in the multi- layer
system showed the same pattern at different striking velocities. The last perforated
layer always has the highest energy absorption in a panel. The different thing is the
position of the last broken layer.

147
5.2.4 The number of broken layers

In the non-perforation case, only some layers in the multi- layer panel were
penetrated. The impact energy was mainly dissipated by yarn fracture and fabric
deformation. The number of broken layers can reflect the ballistic resistant capacity
of the panel. In addition, according to above analysis, the last broken layer
determines the position relating to the highest value of the energy absorption
distribution in the multi- layer system. Therefore, the number of broken layers in the
non-perforated panel was investigated.

For a given striking velocity of 483m/s, the non-perforated panels 11F19 , 11F24 , 11F36 ,
11F48 and 11F120 were conducted the ballistic tests. The average number of broken
layers was counted on the post-impact panel, which was shown in Fig. 5.21. With the
number of layers increased from 19 to 120 of non-perforated panels, the average
number of broken layers decreased from nine to six layers, which didn‟t exhibit
much difference. In addition, it seemed the number of broken layers will decrease to
a constant value when the total layers exceeded a certain value. For example, there
were always six perforated layers when the number of layers increased to 48 and 120.
According to FE simulation of non-perforated panels 11F24 , 11F36 , and 11F48 , FE
results also reflected that broken layers remained the same value when the panel was
added more layers from 24 to 48 layers as shown in Table 5.3.

10
9
Average perforated layers

8
7
6
5
4
3
2
1
0
0 20 40 60 80 100 120
Total number of layers in the non-perforated panel
Fig. 5.21 The number of broken layers in non-perforated panels

148
Table 5.3 Non-perforated Twaron woven panels

Experiment results FE results


Multi-layer Total number
system of layers Average number of The number of
broken layers broken layers
24 9 7
Twaron woven
36 7 7
panel
48 6 7

Such results indicated that at a certain striking velocity, the impact energy had
limited influence on fabric perforation. For example, at the striking velocity of
483m/s, the impact energy can only penetrate around 7-10 layers in the panel at the
non-perforation case, regardless of the total number of layers. So the position of the
last broken layer can be identified in a rather small region in the panel.

It should be noted that the impact energy is determined by the striking velocity and
the type of the projectile. Therefore, the influence of the impact energy on the
number of broken layers in the multi- layer system was numerically investigated at
different striking velocities. The non-perforated panel 11F24 was impacted at
different striking velocities of 300 m/s, 483 m/s, and 600 m/s. According to FE
results, the number of broken layers increased with the increasing striking velocity as
shown in Fig. 5.22. More fabric materials was required dissipating the increasing
impact energy. So with the increasing striking velocity, the position of last broken
layer in the multi- layer system moved back correspondingly until the panel was
completely penetrated.

14
The number of perforated

12

10

8
layers

0
0 100 200 300 400 500 600 700
The striking velocity (m/s)

Fig. 5.22 The number of broken layers in non-perforated panels at different


striking velocities
149
5.2.5 Backface Signature BFS and the indentation volume

Backface Signature (BFS) is one of the most important criteria of ballistic


performance of the non-perforated panel in many ballistic evaluation standards. In
addition, the indentation volume can also reflect the ballistic resistant capacity of the
armour panel. To investigate ballistic response of the multi- layer system in the non-
perforation case, different Twaron woven panels made of fabric 11F were carried out
non-perforation ballistic tests. The number of layers increased from 19 to 120. With
the increasing number of layers, the areal density of panels increased
correspondingly.

According to ballistic tests results, with more layers added in the non-perforated
panel, not only BFS behind panels decreased significantly as shown in Fig. 5.18, but
also the indentation volume in the clay reduced remarkably as shown in Fig. 5.19.
When the number of layers increased to 120 in the panel 11F 120 , no obvious
indentation was found in the clay behind the panel. Such tests results indicated that
the ballistic resistant capacity of an armour panel was obviously improved with the
increasing amount of fabric materials. When the number of layers exceeded to a
certain value, no impact energy was transmitted through the panel.

In addition, according to the regression of the test data, quadratic polynomial was
best fitted curve for BFS as shown in the Fig.5.23. For the indentation volume, the
linear fitted curve was better to express the relationship with the number of layers as
shown in the Fig.5.24. Such results indicated that the increasing amount of fabric
materials in the panel have more significant influence on BFS than on the indentation
volume.
25
y = 0.0018x 2 - 0.4315x + 26.019
20 R² = 0.985
Ave BFS (mm)

15

10

0
0 20 40 60 80 100 120
The number of layers in the panel

Fig. 5.23 BFS of non-perforated panels with different areal density


150
12
y = -0.1032x + 12.536
10 R² = 0.9838

Dent volume (mm3 )


8

0
0 20 40 60 80 100 120
The number of layers in the panel
Fig. 5.24 Indentation volume in the clay behind non-perforated panels

According to above analysis in the non-perforated panel, although the increasing


number of layers in the panel has little influence on improvement of the total energy
absorption, it has a great influence on minimizing BFS behind the panel.

5.3 Summary

In this study, different ballistic responses in the multi- layer system were
experimentally and numerically analysed in the perforation case and non-perforation
case, including the transverse deflection, stress distribution, energy absorption, the
number of perforated layers, BFS and indentation volume.

In the perforation case, the front layers had smaller transverse deformation with high
stress concentration. While back layers can produce large transverse deformation
with wider stress wave propagation. As a result, the energy absorption efficiency of
the front layers was rather lower due to the constrained transverse deformation by
back layers. As the back layers had more mobility and can sustain longer time before
fracture, these layers had higher energy absorption efficiency which even exceeded
that of a single layer fabric. So in the perforated panel, energy absorption of each
layer gradually increased from front to back.

In the non-perforation case, middle layers had an obvious wider transverse


deformation area than front and back layers. Due to the high stress concentration on
the front few layers and low stress magnitude on the back layers, the energy
absorption of these layers was rather lower. So in the non-perforated panel, energy
absorption of each layer increased from front layers to the maximum value in the last

151
perforated layer and then decreased gradually in the following back layers.
According to energy absorption efficiency, energy absorption capacity of all fabric
layers in the non-perforated panel was limited due to the constraint effect by the back
clay. For the given impact conditions in the non-perforated panel, the relatively
better positions that were beneficial to apply the energy absorption capacity of fabric
can be identified close to the last broken layers.

Some influencing factors on the energy absorption distribution in the multi- layer
system were also investigated, including the striking velocity and the number of
layers in the panel. It was found that the energy absorption distribution in the multi-
layer system remained the same pattern, but the position, in the thickness, of the peak
value in energy absorption was shifted more towards the back of the panel when
increasing the striking velocity. So the energy absorption distribution in the
perforated panel can be regarded as a specific case of the high striking velocity.

With the increasing number of layers in the non-perforated panel, energy absorption
of the panel increased but the growth rate reduced. The same amount of fabric
materials played different roles in energy absorption when they were placed in
different positions in a panel. The energy absorption of the front layers was
constrained and reduced by the additional layers. When the total number of layers
exceeds a certain value, the additional layers in the panel contribute less to the total
energy absorption.

The number of broken layers was also analysed to reflect the ballistic resistant
capacity of the multi- layer system. For a given impact conditions, with more layers
added in the non-perforated panel, the average number of perforated layers was in a
rather small range. In addition, it would reduce to a constant value when the total
layers exceeded a certain value. So the position of the last broken layer which had
the highest energy absorption can be identified in a small region in the panel. It
should be noted that the number of broken layers was affected by the impact energy,
which was determined by different striking velocities and types of projectile.

BFS of the non-perforated panel and the indentation volume both continually
reduced with the increasing number of layers. The regression of test data indicated
that the indentation volume was affected more significantly than that of BFS.
Although the increasing number of layers in the panel had little contribution to
152
improvement of the total energy absorption, it had a great influence on minimizing
BFS behind the panel.

According to above investigation, the energy absorption distribution in the armour


panel under ballistic impact was identified. Such fundamental understanding
provided an important guide to optimise the construction of the armour panel.

153
CHAPTER 6 Investigation on failure modes of armour
panels under impact

To optimise the construction of the multi- layer system, understanding different


ballistic characteristics of each layer is the key issue. Failure modes are one of
important aspects of ballistic response of an armour panel under impact. Examination
of the post- impact specimen can provide some information on how perforation
occurs and reflect failure modes of an armour panel. In addition, different ballistic
characteristics of each layer can be qualitatively identified by these physical
evidences.

In this chapter, failure characteristics of the post- impact Twaron woven panel and
Dyneema UD panel are thoroughly observed layer by layer at the fabric level, yarn
level and fibre level by naked eyes, optical microscope and Scanning Electronic
Microscope (SEM). Difference of failure characteristics between layers is qualitative
analysed. Failure modes in perforated and non-perforated Twaron woven panels are
discussed in Section 5.1. The failure modes of non-perforated Dyneema UD panels
are analysed in Section 5.2. Section 5.3 is the summary of this chapter.

6.1 Twaron woven panel

Twaron panels were made of Twaron woven fabric layer. Under ballistic impact,
fabric layers failed in forms of fibres and yarns fracture, yarn pull-out, yarn ravel,
and fabric deformation. To investigate different failure characteristics of each layer
in woven panel, three post-impact Twaron panels: a single layer fabric 11F, nine-
layer panel 11F9 , and 24-layer panel 11F24 , were observed at the fabric level, yarn
level and fibre level.

6.1.1 Failure characteristics of woven panel

6.1.1.1 A single layer woven fabric

As a basic unit of the multi- layer panel, a single layer fabric was investigated on
failure modes firstly. Fig. 6.1 showed the impact process of a single layer Twaron
fabric 11F, which was taken by the high speed camera. When the fabric was
impacted by the projectile at high striking velocity, the deformed area had a clear

154
rhombus base and the transverse deflection out of the plane. The shape resembled a
square pyramid with the impact point as the apex and four corners of the base lying
on the primary yarns. This is due to the woven fabric is the orthotropic material. The
stress wave can only propagate along primary yarns from the impact point.

For the type of projectile used in this study, there were 5-6 primary yarns that
directly contacted with the projectile in high tension and stretched to fracture under
impact. This was also observed from the post-impact fabric as shown in Fig. 6.2.

(a) t=0 µs (b) t=5.5 µs (c) t=11 µs (d) t=16.5 µs


Fig. 6.1 Impact process of a single layer Twaron fabric

Fig. 6.2 The perforated Twaron woven fabric

The fractured fibres in post-impact Twaron fabric were close observed by SEM as
shown in Fig. 6.3. The axial splitting was the dominant failure appearance of broken
fibres. This is the typical fracture morphology of fibres with highly oriented, highly
crystalline, linear polymer chain, such as para-aramid and UHMWPE fibres under
the high tensile stress [25].

Fig. 6.3 (a) (b) showed that some fibres exhibited the longitudinal striations on axial
surface. This is the initial appearance under the high tensile stress. Such longitudinal
striations can be developed more pronounced to form the axial crack with increasing
stress as shown in Fig. 6.3 (c) (d). And then the axial crack can be developed to
single or multiply axial splits as shown in Fig. 6.3 (e). In some situations, the axial
155
splitting proliferates into many fine strands as shown in Fig. 6.3 (f). Such fracture
morphologies was the dominant failure mode of Twaron fibres under impact, which
was commonly observed around the perforated hole on the post-impact fabric.

(a) (b)

(c) (d)

(e) (f)
Fig. 6.3 Failure morphology of Twaron fibres

To compare different failure modes, Twaron yarns were cut by scissors to induce the
shear force applying on fibres. The fractured yarns and fibres were observed by SEM
as shown in Fig. 6.4, which was very different from the axial splitting caused by the
stress induced by ballistic impact. The fractured yarns exhibited smooth axial surface
156
and the broken ends displayed the compressed flat surface. Such failure appearance
indicated the stress concentration in a small region. It can only cause the local
damage on fibres, while other region of fibres axis was not affected.

Fig. 6.4 Shear failure appearance of broken Twaron fibres

6.1.1.2 The perforated woven panel

During the impact process of the nine- layer Twaron panel 11F 9 , the transverse
deformation resembled as the pyramid shape was also found on fabric layers at the
initial contact with the projectile at 18µs. After the projectile penetrated the fabric,
the transverse deformation was gradually transformed into a conical shape, which
can be seen from pictures at 36µs and 54µs as shown in Fig. 6.5. With the increasing
number of layers, the fracture time increased correspondingly.

(a) t=0 µs (b) t=18 µs (c) t=36 µs (d) t=54 µs

Fig. 6.5 Impact process of Twaron woven panel 11F9

157
(a) ply-1 (b) ply-3

(c) ply-6 (d) ply-9


Fig. 6.6 Perforated layers in Twaron woven panel 11F9

In the perforation case, the woven panel produced the transverse deflection with four
sides clamped into the frame under impact. Under impact force, the primary yarns
were stretched and pulled out of the fabric. As a result, some creases were left along
primary yarns on the post- impact fabric layers. According to Shim, the creases on the
post-impact fabric give an indication of the extent of material deflection [53]. On the
last layer ply-9 of Twaron woven panel 11F 9 , more severity of crease can be
observed than that of on front layers as shown in Fig. 6.6. This suggested the last
back layer experienced more significant transverse deflection than that of front layers.
Such analysis confirmed previous FE simulation results that the transvers deflection
increased from front to back layers in the perforated panel (Chapter 5, Section 5.1.2).

In Twaron panel 11F9 , the appearance of the perforated hole on fabric layers was also
observed. The first layer on the impact site exhibited a perforated hole with clear and
orderly fractured yarns as shown in Fig. 6.7 (a). The number of broken yarns
decreased from front to back layer (Appendix 2). On the last layer, only one yarn was
fractured and a large gap between yarns was formed on fabric as shown in Fig. 6.7
(b). This indicated that on the front layer high stress concentration around the edge of
158
the projectile resulted in the broken primary yarns with orderly ends. While on back
layers, the primary yarns can sustain longer and were pulled out of fabric to form a
large gap. The large gap between yarns like a „trap door‟ allows the projectile to
easily slip through the fabric without breaking yarns.

(a) ply-1

(b) ply-9
Fig. 6.7 Perforated holes on fabric layers of Twaron panel 11F9

To identify different characteristics of each layer of Twaron panel 11F9 , one of


broken yarns were taken out from the first, the third, the sixth and the ninth fabric
layers to observe the global failure configuration by SEM at lower magnification.

The failure appearances of fractured yarns on each layer ca n be found some


difference as shown in Fig. 6.8. In front layers, ply-1 and ply-3, the broken yarns
have an orderly clear and even ends. This was caused by stress concentration on
yarns around the sharp edge of the projectile. With the position back in the panel, the
broken fibres on the fractured yarn had obvious different length. Such different
broken positions of fibres indicated the stress distributed over wider region on the
yarn.

159
(a) ply-1

(b) ply-3

(c) ply-6

(d) ply-9
Fig. 6.8 Fractured yarns in different layers in Twaron woven fabric panel 11F9

Different failure characteristics of broken fibres on each layer were observed by


SEM at higher magnification times. In the perforated Twaron woven panel, the axial
160
splitting is the dominant failure morphology of fibres on every layer. But the broken
fibres on each layer exhibited different extents of splitting.

On the front layers, such as ply-1 and ply-3, broken fibres exhibited even break ends
with some small tails as shown in Fig. 6.9 (a) (b). This was due to the high stress
concentration in a local region of fibres under impact. On the ply-3, some broken
fibres were observed twisted and cracked axially located below the end o f the broken
fibre as shown in Fig. 6.9 (b). Such twisted axial appearance of fractured yarns
indicated the interaction with the interlaced yarns under impact force.

(a) ply-1

(b) ply-3

(c) ply-6

161
(b) ply-9
Fig. 6.9 Broken fibres in fabric layers of Twaron panel 11F9

On layers of ply-6 and ply-9, broken fibres exhibited more pronounced axial splitting
as shown in Fig 6.9 (c) and (d). The severe multiple axial splitting extended over
long length of fibres. Some fibres exhibited multiple splits with fine splitting turing
around the fibre axis. Such pronounced failure appearance of fibres also suggested
the enlarged region that the stress applied on back layers.

6.1.1.3 The non-perforated woven panel

In the non-perforated Twaron woven panels 11F 24 , the front 10 layers were
perforated. Due to the free boundary, the fabric in the non-perforated panel produced
more significant movement in the fabric plane and the transverse deflection
compared with the perforated panel 11F 9 as shown in Fig. 6.10. Behind the fabric
panel, there was a deep and round indentation in the clay as shown in Fig. 6.11.

(a) t=0µs (b) t=18µs (c) t=36µs (d) t=54µs


Fig. 6.10 Impact process of the non-perforated Twaron woven panel 11F24

In the non-perforation case, when the projectile was stopped in the panel, the fabric
layers bounced back to form extensive creases as shown in Fig. 6.11. Due to the free
162
boundary conditions, more significant creases were observed in fabric layers of the
non-perforated panel as shown in Fig. 6.12. From the observation of each layer in
Twaron panel 11F24 , the last perforated layer ply-10 displayed the most significant
creases than that of the first layer ply-1 and back layer ply-18. The last back layers
ply-24 exhibited the least creases than that of front layers. This confirmed FE
simulation results that the middle layers had the largest transverse deformation than
that of front and back layers.

(a) Post-impact Twaron panel (b) The indentation in the clay


Fig. 6.11 The non-perforated Twaron panel 11F24 (a) post-impact Twaron panel
and (b) the indentation in the clay

(a) The 1st layer (right) and the 24th last layer (left)

(b) The 10th layer (right) and the 24th last layer (left)
163
(c) The 18th layer (right) and the 24th last layer (left)
Fig. 6.12 Different extent of crease on fabric layers in Twaron panel 11F24

The impact area of each layer in the non-perforated panel 11F24 was close observed.
On the first front perforated layers ply-1 the damage area was just localised around
the perforated hole as shown in Fig. 6.13 (a). Due to the high stress concentration
around the edge of the projectile, these layers were perforated too quickly to develop
any obvious transverse deflection before fracture.

On the last perforated layer ply-10, there was a clear bulge in the impact area with
only one broken yarn as shown in Fig. 6.13 (b). Around the bulge, the fabric
exhibited some obvious transverse deformation. Such failure appearance indicated
the stress distribution propagated more widely on fabric, which involved more
material into transverse deformation.

On back layers, such as ply-18 and ply-24, fabric cannot be perforated and only
produced transverse deformation as shown in Fig. 6.13 (c) and (d). This reflected that
back layers were applied lower stress which can only engage yarns in strain but can‟t
break them.

These observations got the consistent conclusions with the previous FE simulation
results. When the non-perforated panel was under impact, the high stress was
concentrated on the front few layers. On the middle perforated layers, the stress can
reach the yield stress value of the material and propagate widely on fabric. While on
back layers, stress magnitude was lower than the yield stress of the material, but it
can propagate most widely.

164
(a) ply-1 (b) ply-10

(c) ply-18 (d) ply-24


Fig. 6.13 The impact site of fabric layers in Twaron panel 11F24

The debris from the penetrated hole was also observed. Macroscopically, the debris
was loose and fluffy fibre hair without clear structure. Close inspection by SEM
showed some fractured fibres became the flat belt with some ribbons and strips as
shown in Fig. 6.14. Such appearance suggested the severe compressive stress
combining with the tensile stress applied on fibres during the interaction with the
projectile.

Fig. 6.14 Debris of broken yarns in the Twaron panel 11F24

165
The broken fibres in the non-perforated panel 11F24 were close observed by SEM.
On the perforated layers, the different extent of the axial splitting between the first
perforated layer and last perforated layer can also be identified as shown in Fig. 6.15.
But the difference was not as significant as that of in the perforated panel 11F 9 . The
damage region of broken fibres was rather localised around broken ends on every
perforated layer. The extent of axial splitting was not severe in all perforated layers.
Many fractured fibres on perforated layers had flat ends with some fine strands
extended to the same direction.

(a)

(b)
Fig. 6.15 Broken fibres in the non-perforated panel 11F24 (a) ply-1 (b) ply-10

Such failure appearance of broken fibres indicated the stress applied in a local region
in the perforate layers under impact. This is due to the constrained transverse
deformation of perforated layers by back layers and clay behind the panel. The flat
broken end of fractured fibres indicated the significant compressive stress and the
applied direction under impact. Such failure appearance was similar as that observed
in the front layer in the perforation case. In the non-perforation case, the compress
force on the panel increased due to the back clay.

166
6.1.2 Failure modes of the woven panel

6.1.2.1 Failure modes

When a projectile impacted on a fabric layer, the primary yarns were stretched and
produced a transverse deflection. The longitudinal stress waves were generated and
propagated along fibre axis from the impact point. The secondary yarns were pulled
out of the original fabric plane by primary yarns. This resulted in the transverse
deformation in the plane of the fabric and the transverse deflection out of fabric
under impact.

At the initial contact with the projectile, the primary yarns under the projectile were
compressed and sheared transversely. This was confirmed from the observation of
the flat broken fibres ends and the twisted broken ends as shown in Fig. 6.9 and Fig.
6.15. At the same time, the stress propagated quickly from the impact point along the
primary yarns. The fibres on the primary yarns were stretched and experienced
increasing tensile stress at high strain rate, which was indicated in Fig. 6.16.

Different failure morphologies of fractured yarns in the panel indicated d ifferent


stress forms applied on fibres under impact. During ballistic impact, the yarn fracture
was usually caused by the combined tensile stress, shear stress and compressive
stress which worked together. When the combined stress on the yarn exceeded the
limit value of the yield stress of the material, yarns were fractured and the fabric was
perforated.

Fig. 6.16 Schematic of the stress propagation

At the microscopic level, Twaron fibre has tough molecular chain. For broken fibres,
the breakage of primary bonds needs even more fracture energy compared to the
intermolecular slip. According to failure modes of polymer, material always fails
preferentially via the mechanism that requires the least amount of the energy. Due to
the discontinuity of defect on the surface of the fibre or internally, this will give rise
to a shear stress in the fibre. Even a small shear stress around the flaw or defect can
overcome the weak intermolecular bonds before the large tensile stress breaks the

167
covalent bonds within the chain molecules as shown in Fig. 6.17 [25]. As a result, at
the macroscopic level, the axial splitting is the dominant morphology of broken
Twaron fibres. This can be found on any perforated Twaron layers in the perforated
and non-perforated panels as shown in Fig. 6.9 and Fig. 6.15.

Intermolecular slip

Polymer chain

Fig. 6.17 Schematic of the fibre splitting development [25]

6.1.2.2 Transverse deformation and stress distribution

After the panel was penetrated or partial penetrated, some creases were left along
primary yarns on the post- impact fabric layers. According to the extent of crease on
the fabric, the transverse deflection of each layer can be indicated. In the perforated
panel, more severity of crease was observed on the last layer than that of on front
layers as shown in Fig. 6.6. This suggested the last back layer experienced more
significant transverse deflection than that of front layers. In the non-perforated panel,
more significant creases were observed on the last perforated layer than that of the
first layer and back layer. The last back layers ply-24 exhibited the least creases than
that of front layers. Such analysis confirmed previous FE simulation results that the
transvers deflection increased from front to back layers in the perforated panel.
While in the non-perforated panel, the transvers deflection increased from front to
the last perforated layer and then decreased to back layers.

Under ballistic impact, the stress wave propagated along fibres of primary yarns from
the impact point. If the stress magnitude exceeds the yield stress value of the material,
fibres will break at the position relating to the highest stress value. It means the
broken position of fractured fibres can only be in the region that the stress applied.
Therefore, on post-impact panels, different broken positions of fibres and fracture
appearance of yarns on each layer suggested varied stress distribution in the panel.

168
In the perforation case, the orderly and even ends of broken yarns on the front layer
reflected the same broken positions of fibres. Such failure appearance indicated that
the high stress concentration just around the perforated hole as shown in Fig. 6.7 and
Fig. 6.8. While on back layers, the obvious length difference of broken fibres
reflected that the fracture positions of broken yarns distributed over a larger region.
This indicated that the wider stress propagation along yarns on back layers. In
addition, the large gap between yarns in the perforated hole with only one broken
yarn indicated the reduced impact force on the back layer.

In the non-perforated panel, the impact area on each layer exhibited different
characteristics as shown in Fig. 6.13. The first layer only exhibited a perforated hole
in the impact area. On the last perforated layer, some obvious transverse deformation
of fabric was observed around the bulge of the perforated hole at the impact site.
While on the back non-perforated layers, only transverse deformation was displayed
with yarns pulled out of the plane of the fabric at the impact site. The stress on fabric
propagated more widely from front to back layers during impact. Such failure
appearance indicated the extent of fabric damage decreased.

Such analysis of failure modes provided the consistent conclusions with previous FE
simulation results. In the perforated panel, the transverse deformation and the stress
distribution increased from front to back layers. While in the non-perforated panel,
the transverse deformation and the stress distribution increased from the front layer
to the last perforated layer and decreased to back layers.

6.1.2.3 Energy absorption

Under impact, the kinetic energy of the projectile is mainly dissipated through yarn
fracture and fabric deformation. The stress distribution and the extent of the
transverse deflection of each layer also gave some indication of energy absorption of
each layer in the multi- layer system. The significant transverse deflection suggested
the fabric material was under higher stress. While the wide stress distribution on
fabric indicated that more fabric material was in strain. As a result, the fabric layer
that formed prominent transverse deflection with wide stress distribution before
fracture can produce more strain energy. According to above qualitative analysis of
the stress distribution and transverse deflection of each layer in perforated and non-
perforated panels, energy absorption of each layer can be identified different. In the
169
perforated panel, back layers exhibited more evidence of higher energy absorption
than that of front layers. In the non-perforated panel, the last perforated layer seemed
to have higher energy absorption than that of front and back layers.

In addition, the yarn fracture is an important form of energy dissipating. It was


reported that three-quarters of the strain energy was mainly converted to create
fracture surfaces and non-recoverable plastic deformation despite of the little
acoustic emissions [135]. According to Griffith fracture energy theory, the fracture
surface energy to create two new fracture surfaces per crack tip is equal to the change
of the potential elastic strain energy in the body. The surface energy represents
energy absorbed in crack growth, while some stored strain energy is released as the
crack extends due to unloading of regions adjacent to the new fracture surfaces. So
the larger fracture surface indicates more strain energy dissipation.

For Twaron fibres, the axial splitting was the dominant fractured morphology. In the
perforated panel, it was observed that the back perforated layers exhibited obviously
more prominent axial splitting than the front layers as shown in Fig. 6.9. In the non-
perforated panel, the different extent of the axial splitting between the first perforated
layer and last perforated layer can also be identified, although it was not as
significant as that of in the perforated panel as shown in Fig. 6.14. The severer
splitting of fibres suggested much more fracture surface energy was required during
fracture. Therefore, more strain energy was produced in back perforated layers than
that of in front perforated layers both in the perforated and non-perforated panels.

According to above investigations on failure modes of each layer in the post- impact
Twaron woven panels, some different ballistic characteristics of each layer in the
multi- layer system were reviewed by the observation and fractography analysis. Such
qualitative analysis provides a lot of physical evidence for the theoretical analysis to
quantify the difference of ballistic resistant efficiency of each layer in the multi- layer
system.

6.2 Dyneema UD panel

Dyneema UD as a kind of compliant composite has a different structure from the


woven fabric. Due to the superior ballistic resistance capacity and low mass desity, it
has been increasingly applied in the armour panel in recent years. In this study, it was
170
used as a component of the hybrid panel. Therefore, failure modes of Dyneema UD
were also investigated to identify ballistic characteristics under impact. The post-
impact perforated Dyneema UD panel 7U9 and the non-perforated Dyneema UD
panel 7U24 were observed layer by layer.

6.2.1 Failure modes of Dyneema UD

During the impact process of Dyneema UD, a flash was observed on a single layer
UD laminate as shown in Fig. 6.18. This has also been found in the 24- layer panel
7U24 in the non-perforation case as shown in Fig. 6.19. Such phenomenon has also
been found in other studies including on thick laminates. It was analysed due to the
generated heat during impact [74].

(a) t=0 µs (b) t=5.5 µs (c) t=11 µs (d) t=16.5 µs


Fig. 6.18 Impact process of a single Dyneema UD layer

(a) t=0 µs (b) t=38 µs (c) t=76 µs (d) t=114 µs


Fig. 6.19 Impact process of non-perforated Dyneema UD panel 11F20

On the post- impact Dyneema UD panel 7U9 laminate, the perforated hole was close
observed. The fractured fibres on Dyneema UD were contracted and formed a clear
perforated hole with a square shape and even peripheral as shown in Fig. 6.20.
Around the perforated hole, the fractured fibres exhibited some shrivelled fibres end
with some solidified particles. These particles were so strong that can‟t be separated
easily by tweezers. Such appearance reflected the contraction of fibres.

171
The contraction is the typical characteristics of the thermoplastic fibres in the melt. It
is due to extended or stretched molecules contracted into a random coil conformation.
According to Prevorsek [104], the HPPE fibres will undergo a 100-fold contraction
when exposed to temperatures exceeding the melting point. Such fibres contraction
around the perforated hole on Dyneema UD gave an indication of the thermal
damage under ballistic impact.

Generator strip

Fig. 6.20 Perforated Dyneema UD

On the post- impact UD layer, the severe generator strip and ply splitting was found
on the striking and exit face as shown in Fig. 6.20. The generator strip was the
typical ply delamination in the thin laminate when the fracture occurred [19].
Critescu stated the generator strip was the principal mechanism of spreading the
deformation from the impact point to other regions in the composite laminates. [110]
According to Shim [53], ply splitting in the thin laminate was induced by the shear
through the thickness and combined modes of tensile failure.

From SEM observation on the exit face of Dyneema UD panel 7U9 , in different
regions the ply splitting exhibited different failure characteristics. In the region 2
away from the impact site as shown in Fig. 6.21 (b), the resin exhibited the peel
morphology with apparent fibres tracks. In the region 1 as shown in Fig. 6.21 (c) at
the impact site, the resin exhibited the melting and gross drawing. These differe nt
appearances indicated that the thermal damage occurred in a local region around the
impact area.

172
2 1

(a) Ply splitting

(b) Region 1

(c) Region 2
Fig. 6.21 Ply splitting on the exit face of the post-impact Dyneema UD 7U9 :
(a) Ply splitting, (b) Region 1, and (c) Region 2

Around the perforated hole of the Dyneema UD panel 7U9 , the fractured Dyneema
fibres were observed by SEM at higher magnification. Most broken fibre ends had a
smooth, partly globular ends as shown in Fig. 6.22 (a) and (b). Some fibres ends
displayed a mushroom appearance as shown in Fig. 6.22 (c) and (d). Such fracture
morphology suggested the material underwent a phase of softening and plasticity
flow. It is the typical thermal damage appearance of the material. Such failure

173
characteristics confirmed that all fractured Dyneema fibres experienced the
significant thermal damage under ballistic impact.

(a) ×750 times (b) ×850 times

(c) ×1800 times (d) ×3200 times


Fig. 6.22 Fractured fibres on Dyneema UD panel 7U9 at different magnification:
(a)×750 times, (b) ×850 times, (c) ×1800 times, and (d) ×3200 times

In the non-perforated Dyneema UD panel 7U24 , the globular fractured fibres ends can
be observed in all perforated layers around the edge of the perforated hole. It
indicated that the thermal damage was accompanied with the perforation regardless
of the position of the perforated layer in a panel.

Fig. 6.23 showed fractured fibres on the last perforated layer ply-9. The projectile
was stopped before this layer. It only produced a small gap on this layer but didn‟t
penetrate through this layer. However, the fractured fibres ends also exhibited the
thermal damage. This indicated that the interaction between the edge of the projectile
and the UD layer can produce sufficient heat to cause the thermal damage.

174
(a) ×25 times (b) ×500 times

(c) ×800 times (d) ×1000 times


Fig. 6.23 Fractured fibres of ply-9 in the post-impact Dyneema UD panel 7U24 at
different magnification: (a)×25 times, (b)×500 times, (c)×800 times, and
(d)×1000 times

In addition to the thermal damage, the axial splitting was another failure mode of the
Dyneema UD under the high tensile stress. On the axis of fractured fibres, sever axial
splitting was observed with some ribbon bounded fibres as shown in Fig. 6.24. This
indicated Dyneema fibres have experienced the high stress during impact.

Fig. 6.24 Broken fibres with axial splitting in Dyneema UD panel 7U24
Around the impact site, the kink bands or rings has been found around the individual
fibre circumferences as shown in Fig. 6.25. Such appearance indicated that fibres
175
experienced the compressive stress. This failure mode was also observed by
Marissen. The kink banding is attributed to molecular bucking. It is a characteristic
of the material such as Dyneema fibre with low compressive strength (about 1% of
tensile strength) [136] .

Fig. 6.25 Broken fibres with kink bands in Dyneema UD panel 7U24

6.2.2 Mechanisms of thermal damage

When a projectile impacted on Dyneema UD panel, the first layer was compressed
and stretched to form transverse deformation. The material that was contacted with
the flat surface of the projectile was pushed aside and quickly moved out of the
surface of the projectile at high velocity as shown in Fig. 6.26. This induced the
friction between the edge of the projectile and Dyneema UD. As a result, the heat
was generated at the interface. Due to the short interaction timescale and the low
thermal conductivity of the polyethylene, the localised generated heat on a fibre due
to the friction increased very quickly to form a „hot spot‟ on a fibre. In addition to the
lower melting point of UHMWPE fibres, the generated heat on the „hot spot‟ was
sufficient to cause softening or melting of fibres. According to previous researches,
although the exact temperature of the generated heat on the fabric under ballistic
impact was hard to be measured, the heat can be definitely identified to cause the
thermal damage of the UHMWPE fibres in Dyneema UD laminate.

Fig. 6.26 The interaction between the projectile and UD layer under impact

According to the SEM observation of each perforated layers in Dyneema UD panels,


all fractured fibres around the perforated hole on Dyneema UD exhibited contracted

176
ends with globular and melting surface as shown in Fig. 6.20 and Fig. 6.22. Such
failure morphology indicated that fractured fibres suffered prominent thermal
damage under impact.

In addition, the generated heat on the Dyneema UD layer was attenuated from the
impact site instead of evenly distributed. In the plane of Dyneema UD layer, the
extent of thermal damage in different regions was varied. This can be indicated from
different appearance of the ply splitting between in the impact site and in the region
away from the impact site as shown in Fig. 6.21. It can be inferred that the generated
heat was attenuated from the front to back layer in Dyneema UD panel due to the
attenuated striking velocity. The frictional interaction between back layers and
projectile became less significant. This was also analysed in Prevorsek‟s study [103]

There was another possibility that was proposed in previous researches to induce the
thermal damage of fibres on Dyneema UD. The heat can be generated after the fibre
fracture. Due to the friction between the side face of the projectile and broken fibres
ends around the perforated hole, the heat increased during the projectile sliding
through the layer. However, according to the observation of the non-perforated
Dyneema UD panel in this study, the projectile was stopped before the last
perforated layer ply-9 and didn‟t pass through it. In this case, there was no
interaction between the side face of the projectile and fractured fibres ends. The
fractured Dyneema fibres in this layer still exhibited obvious characteristics of
melting damage as shown in Fig. 6.29. Such observation suggested that the generated
heat can‟t only occur after the fibre fracture on the Dyneema UD layer.

According to the investigation on the failure modes of Dyneema UD panels, the


thermal damage was found as the dominant failure characteristics of Dyneema UD
laminate under ballistic impact. It was most likely to be caused by the frictional
interaction between the edge of the projectile and Dyneema UD layers before the UD
layer fracture.

6.3 Summary

To investigate failure modes of soft armour panels under ballistic impact, failure
characteristics of post-impact Twaron panels were observed layer by layer at the
fabric level, at the yarn level and at the fibre level. According to SEM observation,
177
the axial splitting is the dominant failure mode of Twaron fibres under ballistic
impact. It was due to the intermolecular slip between molecular chains under the
coupled work of different stress forms.

According to the investigation on failure modes of each layer in the post- impact
Twaron woven panels, some different ballistic characteristics of each layer in the
multi- layer system were identified by the observation and fractography analysis. The
different failure characteristic of each layer, including the broken ends of fibres, the
failure appearance of yarns and the impact region of fabric, indicated that the area of
the stress distribution and the extent of the transverse deflection of each layer were
different. This determined different energy absorption in each layer of the panel. In
addition, based on the theory of fracture energy, different severity of fibre splitting
on fabric layers suggested different strain energy was dissipated to create new
fracture surfaces. Such qualitative analysis of failure modes got the consistent
conclusions with previous FE simulation results.

For Dyneema UD, due to the lower melt point, the thermal damage was the dominant
failure mode of Dyneema fibres under ballistic impact. This can be clearly identified
from the thermal damage appearance of globular and melt broken fibres ends on
every perforated layer. The heat was mainly generated by the friction between the
edge of the projectile and Dyneema UD. Due to the decreased impact force, the
generated heat was attenuated from the impact area to other region of fabric and from
the front to back layer.

Such observation of failure modes of panels under ballistic impact indicated some
different ballistic characteristics of each layer in an armour panel, which provided a
lot of physical evidence for the theoretical analysis.

178
CHAPTER 7 Design of hybrid armour panels

According to above analysis in chapter 5 and chapter 6, the ballistic response of each
layer in an armour panel has been demonstrated not the same. Combining uniform
layers in a panel is not an efficient construction to achieve the improved ballistic
protection. In addition, no one kind of component can provide all the expected
properties to meet requirements for ballistic resistance. Therefore, the construction of
the multi- layer system can be optimised by combining different components of
ballistic materials in the most effective positions, which is the hybridisation design.

This chapter aims to establish a procedure for constructing a hybrid armour panel. At
first, the mechanisms of hybridisation are investigated and identified including the
materials selection and analysis of layering sequence effect, which are discussed in
Section 7.1 and Section 7.2 respectively. Based on such fundamental understanding,
a new design principle of hybrid panels for soft body armour is explored in Section
7.3. The ballistic performance of two hybrid panels is evaluated by ballistic
resistance tests and non-perforation tests in Section 7.4. The summary for this
chapter is presented in Section7.5.

7.1 Selection of components for hybridisation

In this research, Twaron woven fabric and Dyneema UD laminate, two of the most
popular ballistic materials, were used as components for hybridisation. To design the
hybrid panel, ballistic performance of these two proposed components were
investigated to identify their advantages and disadvantages for ballistic resistance.
Ballistic performance including energy absorption in the panel and Backface
Signature (BFS) was analysed experimentally and numerically.

7.1.1 Ballistic performance of fabric with different weave structures

Woven fabric can provide better energy absorption capacity for ballistic protection
with retaining flexibility, lightweight and comfort. So it is commonly used in soft
body armour. In this study, nine types of Twaron woven fabric referred to table 3.2
were used to investigate the effect of weave structure on ballistic performance. Some
influencing factors of weave structures on ballistic performance were investigated,
such as areal density, yarn count, weave density, and crimp ratio.
179
7.1.1.1 Effect of areal density

The areal density of fabric is the amount of material in unit area. According to
ballistic test results of woven fabrics with different weave structures, energy
absorption increased with the increasing areal density of fabrics as shown in Fig. 7.1.
This means that more fabric material in unit area of panel can provide better ballistic
performance of energy absorption.

500 20
Specific energy absorption

18 SEA
400 16 EA

Energy absorption (J)


14
(J.cm2 /g)

300 12
10
200 8
6
100 4
2
0 0
100 150 200 250 300
2
Areal density (g/m )
Fig. 7.1 Energy absorption of woven fabric with varied areal density

According to the theory of strain energy, for a linear elastic material, the work done
by external forces in causing deformation is stored within the body in the form of
strain energy [137]. The strain energy of material is not only related to tensile
properties of material but also the volume of the undeformed material as shown in
equation (7-1) and (7-2). Obviously, the heavyweight fabric can engage more
materials in strain energy absorption than that of the lightweight fabric during impact.

Strain energy density: (7-1)

Strain energy: ∫ (7-2)

where U is the strain energy stored within an entire body, V is the original or
undeformed volume of the elastic body, U0 is the strain energy per volume. σ x , σx ,
and σx are the stresses applied on the body in the x, y and z direction, ε x , εy , and εz
are the strain in the x, y and z direction.

180
However, for the unit areal density, energy absorption of heavyweight fabric was
lower than that of lightweight fabric as shown in Fig. 7.1. With the increasing areal
density of fabric, the energy absorption at unit areal density decreased. This indicated
that the energy absorption efficiency of lightweight fabric was better than that of
heavyweight fabric for a given weight.

Given the same areal density, the effect of the weave structure on energy absorption
of fabric was investigated. Perforation test results for three woven fabric, 13F, 10M
and 7T, with varied yarn counts and weave densities were shown in Fig. 7.2.
Statistical analysis of the test data was carried out. T-test results indicated that the
mean of the specific energy absorption between 13F, 10M and 7T both differs
significantly with a confidence level of 95%.

Such results indicated that weave structures have a great influence on energy
absorption capacity of fabric regardless of the areal density was the same. To further
investigate the effect of weave structures on energy absorption of fabric, some
dominant parameters were discussed individually in the following section.

350
13F: 93tex, 12.6ends/cm,
Specific energy absorption

300 251.76 g/m2


10M: 110tex, 10.7ends/cm,
250 259.28 g/m2
7T: 168tex, 7.0ends/cm,
(J.cm2/g)

200 260.41 g/m2


150

100

50

0
13F 10M 7T

Fig. 7.2 Effect of areal density on energy absorption of woven fabric

7.1.1.2 Weave structures in a single layer fabric

According to ballistic test results, the yarn count used in woven fabrics has great
influence on energy absorption capacity. Given the weave density held almost
constant (around 10.5ends/cm), the specific energy absorption (SEA) of the fabric
with different yarn counts was shown in Fig. 7.3. Statistical analysis indicated that
the mean of the specific energy absorption between 10F and 10M differs
significantly with a confidence level of 95%. This indicated that the fabric with fine

181
yarns possesses better energy absorption capacity at unit areal density. Such results
confirm the conclusions of the previous study [51].

350 10 SEA
Specific energy absorption 9 EA
300
8

Energy absorption (J)


10F: 93tex, 10.4ends/cm,
250 7 202.08 g/m2
(J.cm2 /g)

200 6 10M: 110tex, 10.7ends/cm,


5 259.28 g/m2
150 4
100 3
2
50
1
0 0
10F 10M
Fig. 7.3 Energy absorption of 10F (10.4ends/cm) and 10M (10.7ends/cm)

The energy absorption of fabric was also influenced by the weave density. For a
given yarn count of 93tex, the performance of energy absorption in five woven
fabrics 8F, 9F, 10F, 12F and 13F, were analysed by ballistic tests. Results in Fig.7.4
indicated that the specific energy absorption of fabrics decreased gradually with the
increasing weave densities. Correspondingly, the crimp ratio increased when the
fabric become tight (Chapter 3, Section 3.1.1.2.). Fig. 7.5 showed the correlation
between the specific energy absorption and the increasing crimp ratio. Such results
reflected that the energy absorption capacity of the tight fabric was lower than that of
the loose fabric.

500

400
absorption(J.cm2 /g)
Specific energy

300

200

100

0
8 9 10 11 12 13
Weave density (ends/cm)
Fig. 7.4 Energy absorption of woven fabrics for a given yarn count of 93tex

182
500

Specific energy absorption


400

300

(J.cm2 /g)
200

100

0
0.0 0.5 1.0 1.5 2.0 2.5
Crimp ratio (%)
Fig. 7.5 Energy absorption of woven fabrics with varied crimp ratios

To take insight into the mechanism responding to ballistic impact on woven fabrics
with different weave structures, FE simulation of single layer fabrics 8F and 11F
under ballistic impact were employed to analyse the fracture time, transverse
deformation and stress distribution.

According to FE results, the loose fabric 8F and the tight fabric 11F were completely
perforated at 7.2µs and 4.4 µs. The same Twaron yarns in the loose fabric can sustain
longer than the yarns in the tight fabric. In addition, the maximum transverse
deformation area on loose fabric 8F exhibited obviously larger than that of 11F as
shown in Fig. 7.6. Under ballistic impact, it was easier for loose fabric to produce
deformation due to the better mobility of yarns in fabric. Fabric deformation greatly
reduced the stress concentration in the local impact area. Correspondingly, the stress
wave propagates widely on fabric.

0.1

0.0
0.00 1.00 2.00 3.00 4.00 5.00
Transverse deflection (cm)

-0.1

-0.2 8F: 93tex, 8.3ends/cm


11F: 93tex, 10.9ends/cm
-0.3

-0.4

-0.5
8F-7.4µs
-0.6
11F-4.4µs
-0.7

-0.8
The distance from the center to edge (cm)
Fig. 7.6 The maximum transverse deformation of 8F and 11F
183
Fig. 7.7 showed the stress distribution along a primary yarn in fabric 8F and 11F
before the first yarn fractured. During the same time interval, the stress on fabric 8F
can propagate over more widely than that of on fabric 11F. This indicated that the
stress wave velocities on fabrics with different weave structures were also found
different.

3.5
8F-0.6µs
3.0
8F-1.8µs
2.5 8F-2.4µs
Stress (GPa)

2.0

1.5

1.0

0.5

0.0
0 1 2 3 4
The distance from the center (cm)
(a) 8F
3.5
11F-0.6µs
3.0
11F-1.8µs
2.5
11F-2.4µs
Stress (GPa)

2.0

1.5

1.0

0.5

0.0
0 1 2 3 4
The distance from the center (cm)
(b) 11F
Fig. 7.7 Stress distribution on a primary yarn of fabric (a) 8F and (b) 11F

To obtain the exact stress wave velocity on two fabrics, the stress distribution history
was analysed. Before the first yarn fractured during impact, the stress wave
propagated from the impact point to the edge of fabric. When the stress wave along
the specific path on fabric reached a certain value, such as 0.5GPa marked on Fig.
7.7, the corresponding time and the corresponding position were plotted as the time-
displacement curve as shown in Fig. 7.8. The slope of the time-displacement curve
184
was the stress wave velocity. On the loose fabric 8F, the stress wave velocity was
4798m/s. It was a little higher than the stress wave velocity of 4238m/s in the tight
fabric 11F. As a result, during the same time interval, the stress wave can propagate
more widely on the loose fabric than that of on the tight fabric. The material with
high stress wave velocity can engage more materials in strain and produce more
strain energy during the same impact time.

2.5
y = 0.4798x + 0.9657 8F
2.0 R² = 0.9934
Displacement (cm)

11F
1.5

1.0

0.5 y = 0.4238x + 0.1386


R² = 0.9924
0.0
0 0.5 1 1.5 2 2.5 3
Time (µs)
Fig. 7.8 Stress wave velocity on two fabrics 8F and 11F

According to above investigation of weave structure, it can be concluded that the


lightweight fabric has higher energy absorption capacity than that of the heavyweight
fabric at unit areal density. The lightweight fabric can be obtained by using fine
yarns or reducing the weave density. It should be noted that to obtain the lightweight
fabric, adjusting the weave structure can only be in a certain range. The range was
determined by the fabric tightness during the fabric manufacture, which was
discussed in detail in Chapter 3, Section 3.1.1.1.

7.1.1.3 Construction of the multi-layer panel

To explore the better construction of the multi- layer panel, the performance of woven
panels composed of different types of fabrics were analysed, which includes energy
absorption and BFS. The compared panels were designed with the same areal density.

For a given areal density of a woven panel, the specific energy absorption of the
woven panel 8F4 and 11F3 with the yarn count of 93tex was shown in Fig.7.9.
Statistical analysis of T-test indicated that the mean of the specific energy absorption
between 8F4 and 11F3 differs significantly with a confidence level of 95%. The

185
woven panel 8F4 composed of four layers of loose fabric was 18.67% higher than
that of the woven panel 11F3 with three layers of tight fabric.

With almost same areal density of a woven panel, the specific energy absorption of
the woven panel 11F4 and 10M3 with the constant weave density (around 11ends/cm)
was shown in Fig.7.10. Statistical analysis indicated that the mean of the specific
energy absorption between 11F4 and 10M3 differs significantly with a confidence
level of 95%. With the constant areal density, energy absorption of the four- layer
panel 11F4 with fine yarns was 15.37% higher than that of the three-layer panel 10M3
with thick yarns as shown in Fig. 7.10.

450
Specific energy absorption

400 8F4 : 93tex, 8.3ends/cm,


621.40 g/m2
350
11F3 : 93tex, 10.9ends/cm,
300 590.55g/m2
(J.cm2/g)

250
200
150
100
50
0
8F4
8F4 11F3
11F3

Fig. 7.9 Effect of weave density on energy absorption capacity of woven panels
8F4 and 11F3
400
Specific energy absorption

350 11F4 : 93tex, 10.9ends/cm,


787.4g/m2
300 10M 3 : 110tex, 10.7ends/cm,
777.84g/m2
(J.cm2/g)

250
200
150
100
50
0
11F4
11F4 10M3
10M3

Fig. 7.10 Effect of yarn count on energy absorption capacity of woven panels
11F4 and 10M 3

These ballistic test results indicated two points. First, the construction of the multi-
layer system has a great influence on ballistic performance. Even with the same
amount of fabric material, energy absorption capacity of multi- layer panels can be
186
different due to different constructions. The better construction of the multi- layer
panel can provide improved performance of energy absorption without increasing the
weight of an armour panel.

Second, the armour panel with more layers of lightweight fabric was better for
energy absorption than the panel combining heavyweight fabric with fewer layers.
Either using the lower weave density or fine yarns in fabric, the lightweight fabric
always displayed the higher energy absorption capacity in the multi- layer panel,
which wasn‟t affected by the interaction between layers. It can be inferred that such
construction will exhibit more significant improvement of energy absorption with
more layers added in the panel.

According to non-perforation tests results, for a given areal density of panels, the
woven panel that composed of the lightweight fabric 8F produced larger BFS than
the panel with heavyweight fabric 11F as shown in Fig 7.11. Statistical analysis
indicated that the mean of the BFS between 8F24 and 11F19 was significantly
different with a confidence level of 95%. According to the FE analysis (Section
7.1.1.2), this was due to the loose fabric 8F has larger transverse deflection than that
of 11F as shown in Fig. 7.7. Such results indicated that the lightweight fabric had
poor performance of BFS. This makes the lightweight fabric impossible to be
directly put into use for ballistic protection.

35

30 8F24 : 3728.40 g/m2


25 11F19 : 3740.15g/ m2
7U20 : 3738.80g/m2
BFS (mm)

20

15

10

0
8F24
8F24 11F19
11F19 7U20
7U20
Fig. 7.11 BFS of woven panels and Dyneema UD panel

According to above analysis, woven fabric panels can achieve improvement of


energy absorption by combining lightweight fabric with more number of layers
replacing of heavyweight fabric with few layers. However, the related BFS of the

187
woven panel composed of lightweight fabric increased correspondingly due to the
large transverse deflection during the ballistic impact.

7.1.2 Ballistic performance of Dyneema UD

Dyneema UD is another popular material for ballistic protection used in soft body
armour. Due to superior material properties and laminate structures, Dyneema UD
panel exhibited different ballistic characteristics from that of Twaron woven panel.

7.1.2.1 Energy absorption capacity

According to ballistic test results of perforated panels, SEA of Dyneema UD panels


had an increasing trend with more layers added in the panel as shown in Fig. 7.12.
Such trend was very different from that of woven panel. In particular, when the
number of layers was above 9 layers, SEA seemed to have a sharp increase.

One of the possible reasons can be analysed as the performance degradation of


Dyneema UD caused by the generated heat on the striking face during the impact
process, which was also reflected in ballistic tests (section 7.2.2.1). According to the
analysis in Chapter 5, the generated heat was not evenly distributed in a panel. It
attenuated from the front to the back layer. When the number of UD layers exceeds a
certain value, the generated heat has little influence on back layers than that of on
front layers on the striking face.

500
Specific energy absorption

400 Twaron 8F
Twaron 11F
300 Dyneema UD SB71
(J.cm2/g)

Dyneema UD SB51
200

100

0
0 1 2 3 4 5 6 7 8 9 10
Number of layers

Fig. 7.12 The specific energy absorption of Twaron and Dyneema UD panel

188
Due to the bullet has much bigger size than the projectile used in this study, it should
be noted that the region which was affected by the generated heat in UD panels will
be enlarged when impacted by the bullet. As a result, the energy absorption
performance of UD panel may not be the same.

7.1.2.2 Advantage of small BFS

Compared with the woven fabric, Dyneema UD panel has an obvious advantage of
lower BFS as shown in Fig. 7.11. Statistical analysis (T-test) indicated that the mean
of the BFS behind 7U20 was significantly different from that of 8F24 and 11F19 with a
confidence level of 95%. According to non-perforation test results, BFS behind
Dyneema UD panel 7U20 was 57.19% smaller than that of Twaron woven panel 8F 24
and 39.37% smaller than that of the panel 11F 19 with the same areal density of the
panels. The indentation behind the Dyneema UD panel was observed much
shallower and wider than that of Twaron woven panel (Chapter 3, section 3.3.2.6).
2.5cm
0.494cm

(a) Twaron panel 11F24 (Ar=4724.4 g/m2 )

2.5cm
0.480cm

(b) Dyneema UD panel 7U25 (Ar=4673.5 g/m2 )


Fig. 7.13 Stress distribution contours at 3µs

To reflect different ballistic responses of Dyneema UD panel and Twaron woven


panel, FE analysis of two non-perforated panels Twaron 11F 24 and Dyneema 7U25
with the same areal density was carried out. With the high modulus and lo wer
density, the stress wave speed of Dyneema UD was higher than that of Twaron.
According to FE results as shown in Fig 7.13, at 3µs after impact, the stress wave on
UD panel propagated more widely in the plane of the laminate layer and

189
perpendicular to the panel than that of Twaron panel. Fig 7.14 showed the detailed
stress distribution in the plane of two panels on the first and the sixth layer at 0.5µs
and 3µs before these layers fracture. Fig 7.15 showed the stress magnitude
perpendicular to the panels at 0.5 µs and 3 µs. It was obvious that Dyneema UD had
advantages of the quick stress propagation than that of Twaron woven panel.

The higher stress wave speed on the UD laminate results in the wider transverse
deformation area and small transverse deflection. FE simulation results reflected
different extent of the maximum transverse deformation of the last layer in Dyneema
UD and Twaron woven panel as shown in Fig. 7.16. The last layer in Dyneema UD
panel produced obviously wider and shallower deformation area than that of Twaron
woven fabric panel. This resulted in the smaller BFS behind the Dyneema UD panel.

1.6
UD-ply1-0.5µs
1.4 UD-ply6-3µs
W-ply1-0.5µs
1.2
W-ply6-3µs
Stress (GPa)

1.0
0.8
0.6
0.4
0.2
0.0
0 0.5 1 1.5 2 2.5 3
The distance from the center (cm)
Fig. 7.14 Stress distribution in the plane of Twaron panel and Dyneema UD
panel
1.6
1.4 UD-0.5µs
UD-3µs
1.2 W-0.5µs
1.0 W-3µs
Stress (GPa)

0.8
0.6
0.4
0.2
0.0
0 0.2 0.4 0.6
The distance from the center(m)
Fig. 7.15 Stress distribution perpendicular to the plane of Twaron panel and
Dyneema UD panel

190
The distance from the center (cm)
-4 -3 -2 -1 0 1 2 3 4
1

Maxmium transverse
deflection (cm)
0
ply25 of UD panel

-1 ply24 of Twaron
woven panel
-2

-3

-4

Fig. 7.16 The maximum transverse deformation area of the last layer in Twaron
woven panel and Dyneema UD panel

According to above analysis, Dyneema UD panel showed obvious advantage of


lower BFS. The quick stress wave speed and wide propagation results in the lower
transverse deflection. Therefore, Dyneema UD was used as one component of hybrid
panel to minimize BFS behind the woven panel.

Based on above investigation on ballistic response of different Twaron fabric weave


structures and Dyneema UD, the components for hybridisation were selected. Due to
the better energy absorption capacity, the lightweight Twaron fabric can be adopted
in the panel with more layers replacing of the heavyweight fabric panel with few
layers to achieve improvement of energy absorption. To minimize the related
increasing of BFS behind the lightweight fabric panel, Dyneema UD is considered to
be combined behind to constrain the large transverse deflection of the lightweight
fabric, due to its obvious advantage of lower BFS.

7.2 Layering sequence of hybridisation

According to previous studies, the layering sequence of the hybrid panel also has a
great influence on ballistic performance of hybrid panel. In this research,
mechanisms of hybridisation were investigated by analysing the hybrid structures
and hybrid materials.

7.2.1 Hybrid structures

To investigate the hybrid structures, two layers of woven fabric, lightweight fabric
8F and heavyweight fabric 11F were combined in reverse sequences to construct two
191
hybrid panels, 8F/11F and 11F/8F. The hybridisation effect on energy absorption and
stress distribution was experimentally and numerically analysed.

7.2.1.1 Energy absorption

According to ballistic test results, Fig. 7.17 showed energy absorption of above two
hybrid panels with two layers of woven fabrics. When the lightweight fabric 8F was
placed behind the heavyweight fabric, the energy absorption has a little improvement
of 5.47%. T-test indicated that the mean of the energy absorption between 8F/11F
and 11F/8F was not shown significantly different. This was due to there were only
two layers in hybrid panels. The difference of energy absorption cannot be
significant.

18
16 FE results
Energy absorption (J)

14 Test results

12 8F: 93tex, 8.3ends/cm


11F: 93tex, 10.9ends/cm
10
8
6
4
2
0
8F/11F 11F/8F

Fig. 7.17 Experiment results of energy absorption of two hybrid panels

To confirm above effect of the layering sequence on energy absorption, FE


simulation of these two hybrid panels under ballistic impact were employed. FE
results also exhibited the difference of energy absorption between two hybrid panels
as shown in Fig. 7.17.

To further investigate the effect of layering sequence, energy absorption of each


layer in two hybrid panels of 8F/11F and 11F/8F was compared according to FE
simulation results as shown in Fig. 7.18. On the striking face of both hybrid panels,
energy absorption of the first layer was rather lower. The difference between the first
layers in two panels was not obvious. However, on the exiting face of both hybrid
panels, the back layers displayed significant different energy absorption. The energy
absorption of the back layer 8F in the panel 11F/8F was 28.5% higher than the back

192
layer 11F in the panel 8F/11F. Such results indicated that combining the lightweight
fabric at the back of the panel was beneficial to make full use of its energy
absorption capacity.

Energy absorption (J)


7
6
5
4
3
2
1
0
8F/11F-ply1 8F/11F-ply2 11F/8F-ply1 11F/8F-ply2

Fig. 7.18 FE results of energy absorption of each layer in two hybrid panels

7.2.1.2 Stress distribution

To interpret the effect of layering up sequence in the hybrid panel, the stress
distribution of each layer in two hybrid panels was analysed by FE simulation.

In two hybrid panels of 8F/11F and 11F/8F, both of the first layers had high stress
concentration in a rather local impact area around the edge of the projectile at the
early contact. Therefore, the fabric on the first layer failed very quickly (1.5µs-2.0µs)
as shown in Fig. 7.19 (a) and Fig. 7.20 (a). This resulted in lower energy absorption
on the first layer in panels.

3.5
ply1-0.5µs
3.0
ply1-1.0µs
2.5
Stress (GPa)

ply1-2.0µs
2.0
1.5
1.0
0.5
0.0
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08
Distance from the center (m)
(a) The first layer 8F

193
3.5
ply2-1.0µs
3.0
ply2-2.0µs

Stress (GPa)
2.5

2.0 ply2-3.0µs

1.5

1.0

0.5

0.0
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08
Distance from the center (m)
(b) The second layer 11F
Fig. 7.19 The stress distribution on each layer of 8F/11F
3.0
ply1-0.5µs
2.5
ply1-1.0µs
Stress (GPa)

2.0 ply1-1.5µs

1.5

1.0

0.5

0.0
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08
Distance from the center (m)
(a) The first layer 11F

3.0
ply2-1.0µs
2.5
ply2-3.0µs
ply2-6.0µs
Stress (GPa)

2.0
ply2-9.0µs
1.5

1.0

0.5

0.0
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08
Distance from the center (m)
(b) The second layer 8F
Fig. 7.20 The stress distribution on each layer of 11F/8F

194
But the second layers of two hybrid panels had more freedom to produce the
transverse deformation during impact. So the fabric 8F can sustain much longer (up
to 9.0µs) before fracture than the fabric 11F (only 3.0µs). This resulted in the stress
wave on the lightweight woven fabric 8F propagated even more widely before
fracture than that of on the heavyweight fabric 11F as shown in Fig. 7.19 (b) and Fig.
7.20 (b). Correspondingly, more materials of the fabric 8F were involved into energy
absorption than that of the fabric 11F when they were placed in the exiting face of
the hybrid panel.

According to above experimental and numerical analysis, for the hybrid weave
structure, it is beneficial to place the lightweight fabric behind the heavyweight
fabric to achieve improvement of energy absorption capacity.

7.2.2 Hybrid materials

The layering sequence was also analysed in hybrid materials. Twaron woven fabric
and Dyneema UD were combined in reverse sequences to construct hybr id panels.
Twaron woven fabric includes two types of weave structures: fabric 8F with higher
energy absorption capacity (390.73 J.cm2 /g) than that of Dyneema UD (259.44
J.cm2 /g) and fabric 7T with lower energy absorption capacity (236.17 J.cm2 /g).
Ballistic performance including energy absorption and BFS of hybrid panels was
experimentally investigated to reflect the effect of layering sequence.

7.2.2.1 Energy absorption

According to a series of ballistic tests results, the layering sequence of Dyneema UD


and woven fabric has a great influence on energy absorption of hybrid panels.
Statistical analysis of the test data presented in Fig. 7.21 was carried out. T-test
indicated that the mean of the energy absorption between 8F/7U and 7U/8F, as well
as 7T/7U and 7U/7T differs significantly with a confidence level of 95%. Compared
with the panel 7U/8F, energy absorption of 8F/7U increased 18.71%. When the
fabric 7T with the lower performance of energy absorption capacity was combined
with Dyneema UD, energy absorption of the hybrid panel 7T/7U was still improved
8.99% than the panel 7U/7T.

195
400
350

Specific energy absorption


300
250

(J.cm2 /g)
200
150
100
50
0
8F/7U 7U/8F 7T/7U 7U/7T
Fig. 7.21 Specific energy absorption of hybrid material panels

To confirm the effect of the layering sequence, the number of layers increased to six
layers in hybrid panels with three layers of woven fabric and three layers of
Dyneema UD. Ballistic tests results were shown in Fig.7.22. Statistical analysis (T-
test) indicated that the mean of the energy absorption between 8F 3 /7U3 and 7U3 /8F3 ,
as well as 7T3 /7U3 and 7U3 /7T3 also differs significantly with a confidence level of
95%. Energy absorption of the hybrid panel 8F 3 /7U3 increased 25.06% than that of
the hybrid panel 7U3 /8F3 . While the energy absorption of the hybrid panel 7T3 /7U3
increased 19.30% than that of 7U3 /7T3 .

450
400
Specific energy absorption

350
300
(J.cm2 /g)

250
200
150
100
50
0
8F3 /7U3
8F3/7U3 7U3 /8F3
7U3/8F3 7T3 /7U3
7T3/7U3 7U3 /7T3
7U3/7T3

Fig. 7.22 Specific energy absorption of six-layer hybrid material panels

According to the analysis in Section 7.2.1, the fabric with higher energy absorption
capacity should be placed at back to achieve improvement of energy absorption.
However, such experiment tests results indicated that the back position of the hybrid
panel was always better for Dyneema UD regardless of combining with fabric of
higher or lower energy absorption capacity. The reason can be explained due to the

196
performance degradation of Dyneema UD, which was caused by the thermal damage
on the striking face during impact. Therefore, Dyneema UD was not appropriate to
be combined on the striking face in a hybrid panel.

7.2.2.2 Backface Signature (BFS)

For the non-perforated panel, the layering sequence of hybridisation also has
influence on BFS of the hybrid panel. When Twaron fabric layers and Dyneema UD
layers were combined in hybrid panels, there are four types of layering sequences.
Fig. 7.23 showed BFS behind four types of different hybrid panels. Statistical
analysis (T-test) indicated that the mean of the BFS between 8F8 /7U16 and 7U16 /8F8 ,
as well as 8F8 /7U16 and 8F4 /7U16 /8F8 differs significantly with a confidence level of
95%. The mean of BFS between 8F8 /7U16 and 7U8 /8F8 /7U8 was not significantly
different.

These results indicated that when Dyneema UD layers were placed behind woven
fabric layers, the hybrid panel has lower BFS than that of other layering sequences.
Such results indicated that the layering sequence of combining woven fabric layers
on the striking face and UD layers on the exiting face was beneficial to minimize
BFS of the hybrid panels.

16
14
12
BFS (mm)

10
8
6
4
2
0
8F8/U16
8F8 /7U16 U16/8F8
7U16 /8F8 U8/8F8/U8
7U8 /8F8 /7U8 8F4/U16/8F4
8F4 /7U16 /8F4

Fig. 7.23 Effect of the layering up sequence on the BFS of panels

7.3 Design of hybrid armour panels

Based on above ballistic performance of different components and the effect of


layering sequences, the hybrid armour panel was designed to achieve improvement
of ballistic performance of an armour panel and reductions in weight. The
197
construction of the multi- layer system was optimised by combining different
components in the most effective positions to make full use of different material
properties and to avoid weak points.

7.3.1 Design principle of the hybrid armour panels

According to the experimental and numerical investigation on ballistic responses of


the multi- layer system in Chapter 5 and Chapter 6, ballistic characteristics of each
layer were demonstrated not to be uniform. Combining the same fabric layers in the
armour panel is not efficient to make best use of energy absorption capacity of fabric
and material properties. Therefore, the construction of the armour panel can be
optimised by combining different components in the most effective positions. The
key issue is to identify the most effective positions for each component. In this study,
the fabric layers in an armour panel were divided into several groups based on
different characteristics of each layer, which was identified from the transverse
deformation, stress distribution, and energy absorption.

In a non-perforated armour panel, the front, middle and back layers were found to
have different extents of transverse deformation and stress distribution. The middle
layers had an obvious wider transverse deformation area than front and back layers.
On the front few layers, the transverse deformation was just localised around the
edge of the projectile. This results in high stress concentration in the impact area.
While for back layers, transverse deformation become less obvious due to the low
stress magnitude. Such different characteristics of each layer were not only
quantified by FE simulation, but also confirmed by photographic observations of
post-impact panels, which provide a great deal of physical evidence that was
identified from different failure appearance of fibres, yarns and fabric on each layer.

In addition, energy absorption capacity of each layer in a panel was not the same. It
increased from front to the maximum value in the last perforated layer and then
decreased gradually to the following back layers. Those layers close to the last
perforated layer have higher energy absorption. Due to the high stress concentration
on the front few layers and low stress magnitude on back layers, energy absorption of
these layers were rather lower. This can also be indicated from different extents of
the fibre splitting on each layer on post- impact panel, which was related to different
fracture energy absorption.
198
Such pattern of energy absorption distribution was not affected by the striking
velocity and the total number of layers in the panel. But the position of the peak
value in energy absorption was shifted more towards the back of the panel when
increasing the striking velocity. The perforation case can be regarded as a specific
case at the high striking velocity. Energy absorption of each layer in the perforated
panel increased from front to the maximum value in the last back layers, which
followed the pattern of the perforated layers in the non-perforated panel.

According to energy absorption efficiency, energy absorption capacity of all fabric


layers in the non-perforated panel was limited, which was due to the constraint effect
by the back clay. For given impact conditions, the re latively better positions in the
non-perforated panel that was beneficial to apply the energy absorption capacity of
fabric can be identified close to the last broken layers.

Group 1
Group 2 a woven panel
Group 3

Fig. 7.24 Schematic of a panel division

Based on different ballistic characteristics and energy absorption efficiency of each


layer, a panel were discretely divided into three groups including front, middle and
back layers as shown in Fig. 7.24.

In the multi- layer system, the first group composed of front few layers, which had
high stress concentration and fractured very quickly under ballistic impact. Taking
the panel 11F24 as example, the energy absorption efficiency of the first two or three
layers was rather lower, which was no more than -50%. Therefore, these layers were
classified as the first group.

Some of middle layers close to the last broken layer can sustain for longer time
before fracture and the stress wave can propagate widely. These layers have higher
energy absorption efficiency. According to the energy absorption efficiency of the
panel 11F24 , the last broken fabric layer can applied 80% energy absorption capacity
as that of a single layer fabric. Those layers close to the last broken layer have a

199
relatively higher energy absorption capacity. So these middle layers were roughly
classified into the second group.

All back layers are classified into the third group. These back layers are non-
perforated layers. On these fabric layers, the stress wave can propagate across the
fabric and stress magnitude is lower. The energy absorption efficiency of the panel
11F24 was below -50%. So these layers were roughly classified into the third group.

7.3.2 The function and the range of each group in a panel

Following the group division of the multi- layer system, the proper components for
each group were selected and the specific range of each group were analysed. In this
study, the hybrid panel were designed by combining Twaron woven fabric and
Dyneema UD laminate.

7.3.2.1Heat resistant layers

The first group was on the striking face of a panel. During impact, heat was
generated in a local impact area of fabric on the striking face. According to failure
modes in the soft armour panel, the generated heat can cause the thermal damage of
Dyneema UD during impact process due to the lower melt point (Chapter 6, section
6.2). In addition, hybrid panels combining UD layers on the striking face showed
obvious degradation of ballistic performance (Chapter 7, section 7.2.2.1). As a result,
it was better to place woven fabric on the striking face of the armour panel.

According to FE simulation results, the first few layers (two or three layers) in the
multi- layer panel fractured very quickly due to the high stress concentration. For
example, the fracture time of the first layer in the panel 8F 9 was only 0.3 µs. During
such short responding time, fabric material that was involved in strain was just
confined in small local impact area regardless of any weave structures of fabric. In
this case, for a given amount of transverse deformation area, the heavyweight fabric
with higher absolute amount of energy absorption was beneficial to improve energy
absorption. Therefore, the heavyweight fabric was placed on the striking face of a
panel.

The function and the range of the first group in a panel were experimentally
investigated at the non-perforation case. The woven panel 11F24 was taken as the
200
reference panel. Two hybrid panels, 13F3 /11F20 and 8F3 /11F22 , were constructed by
combining three layers of heavyweight fabric 13F and the lightweight fabric 8F on
the striking face before some layers of fabric 11F. These two hybrid panels have the
close areal density as the reference panel. To remove the little difference of areal
density between panels, specific depth (SD) was used here to normalise the Backface
Signature (BFS).

According to test results, the material on the striking face has an influence on the
BFS and the specific depth (SD). Statistical analysis of the test data presented in Fig.
7.25 and Fig.7.26 was carried out. T-test indicated that the mean of the BFS and the
specific depth (SD) between 13F3 /11F20 and 8F3 /11F22 differs significantly with a
confidence level of 95%.

The hybrid panel 13F3 /11F20 with combining the heavyweight fabric on the striking
face achieved better performance of BFS, which was 36.07% lower than that of the
reference panel 11F24 as shown in Fig. 7.25. When the lightweight fabric was
combined on the striking face, BFS of the hybrid panel 8F 3 /11F22 became larger.
Taking account into the little difference of areal density between these panels, the
hybrid panel 13F3 /11F20 still had lower specific depth in the clay than other two
panels as shown in Fig.7.26. Such results demonstrated that combining the
heavyweight fabric on the striking face of the panel can achieve better ballistic
resistance.

20.0
11F24 : Ar=4724.40 g/m2
18.0 13F3 /11F20 : Ar=4695.88 g/m2
16.0 8F3 /11F22 : Ar=4793.75 g/m2
14.0
BFS (mm)

12.0
10.0
8.0
6.0
4.0
2.0
0.0
11F
11F24
24 13F3 /11F20
13F3/11F20 8F3 /11F22
8F3/11F22

Fig. 7.25 BFS of panels with different components in the first group

201
4.0
3.5

Specific depth (cm 3 /g)


3.0
2.5
2.0
1.5
1.0
0.5
0.0
11F24
11F24 13F3 /11F20
13F3/11F20 8F3 /11F22
8F3/11F22

Fig. 7.26 Effect of different weave structure in the first group

To identify the range of the first group, different layers of heavyweight fabric 13F
were placed in the first group of the hybrid panel. The test results were shown in Fig.
7.27. T-test indicated that the mean of the BFS and the specific depth (SD) between
13F3 /11F20 and 13F9 /11F12 differs significantly with a confidence level of 95%.
When the heavyweight fabric 13F increased to nine layers, the positive hybridisation
effect of minimizing BFS of the panel cannot be observed. Instead, the specific depth
(SD) of the hybrid panel 13F 9 /11F12 increased 17.03% by comparing with the
reference panel 11F24 .

4.5
11F: 93tex, 10.9ends/cm
4.0 13F: 93tex, 12.6ends/cm
3.5
Specific depth (cm 3 /g)

3.0
2.5
2.0
1.5
1.0
0.5
0.0
11F24
11F24 13F3 /11F20
13F3/11F20 13F9 /11F12
13F9/11F12
Fig. 7.27 Effect of the range of the first group on specific depth

Such test results demonstrated the ballistic function of the first group and indicated
that the first group in a hybrid panel only composed of two or three layers. This
confirmed the panel division according to the energy absorption efficiency of each
layer. When the number of layers in the first group exceeded a certain range,
combining the heavyweight fabric in the first group was not beneficial to minimize
202
BFS. This can be analysed as a result of the lower energy absorption capacity of the
heavyweight fabric.

7.3.2.2 Energy absorption layers

The second group composed of some middle layers in a panel. The stress can
propagate widely and fabric can sustain longer to produce more strain energy. These
layers can make best use of energy absorption capacity. Therefore, it is better to
place the component with high energy absorption capacity in this group.

According to the investigation of fabric weave structures (Chapter 7, section 7.1.1),


for a given unit areal density, the lightweight fabric has higher energy absorption
capacity than that of the heavyweight fabric. For example, with the same areal
density in the multi- layer system, specific energy absorption (SEA) of the
lightweight fabric panel 8F4 was around 20% higher than that of heavyweight fabric
panel 11F3 . So the lightweight fabric was selected for the second group. The
lightweight fabric can be obtained by using the fine yarn and /or reducing the weave
density.

Perforated panels Non-perforated panels


16
The number of perforated layers

14

12

10

0
0 20 40 60 80 100 120 140
Total number of layers in a panel

Fig. 7.28 The number of perforated layers in the multi-layer system

Based on energy absorption efficiency of each layer in the multi- layer system, the
layers close to the last broken layer were roughly classified into the second group. To
investigate the specific range of the second group, the position of the last perforated
layer in a non-perforated panel needs to be identified.

203
According to ballistic tests results, Fig. 7.28 showed the correlation between the
number of perforated layers and the total number of layers in a panel. For a given
striking velocity, there is a limit number of layers in a panel between the perforation
and non-perforation case. When the number of layers was less the limit value, the
panel was completely perforated. The number of perforated layers increased linearly
with the increasing striking velocity. When the number of layers in the panel
exceeded the limit value, the panel was partially perforated. Only front several layers
were perforated. In addition, with more layers added in the panel, the number of
perforated layers decreased correspondingly and then reached a constant value.

In the non-perforation case, for a given striking velocity, the type of projectile and
material properties of the panel, the amount of fabric material required to dissipate
the kinetic energy of the projectile by fracture is constant. The number of perforated
layers can‟t be affected by additional layers when the number of layers exceeded the
limit value. Therefore, the position of the last perforated layer was ide ntified in a
certain region of a non-perforated panel.

According to the number of layers in first group N F, the position of the last
perforated layer N p and the total number of layers N Total in the multi- layer system, the
number of layers in each group can be identified according to following equations.

Group 1: NF (7-3)

Group 2: N M= 2×( N p - N F-1)+1 (7-4)

Group 3: N B= NT otal - N F - N M (7-5)

where N F is the number of layers in Group 1, N M is the number of layers in Group 2,


N p is the number of perforated layers in a panel, N B is the number of layers in Group
3, N T otal is the total number of layers in a panel.

For a non-perforated panel, the total number of layers N Total is determined by the
requirement of BFS. It can be obtained by non-perforation ballistic tests or FE
simulation. The number of layers in the first group NF is usually three or four. It has
already been identified by ballistic tests and FE simulation. The number of perforated
layers in a panel N p can be obtained by ballistic tests or FE simulation. According to
N F, Np and NT otal, the number of layers in each group can be identified. It should be

204
noted that for different armour panels, NF, Np and NTotal may be changed, due to
different materials and weave structures in a panel.

The division of the reference panel 11F 24 was listed in Table 7.1. According to the
analysis of the first group (Section 7.2.2.1), N F was identified to be three layers. In
ballistic tests of 20 shots on the panel 11F 24 , front seven to ten layers were perforated.
The probability of the number of perforated layers was shown in Fig. 7.29.
According to the highest probability, N p was assumed to be eight. So the number of
layers in each group was identified according to equations. The division of the panel
was shown in Fig. 7.30.
Table 7.1 Division of the reference panel 11F24
The non-perforated Twaron woven panel
panel 11F24
Region in a panel Group 1 Group 2 Group 3

Np=7 7 14
Np=8 9 12
The number of layers 3
Np=9 11 10
Np=10 13 8

50%
45%
40%
35%
Probability

30%
25%
20%
15%
10%
5%
0%
7 8 9 10
The number of perforated layers
Fig. 7.29 The probability of the number of perforated layers in the panel 11F24

205
0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24
-10

Energy absorption efficiency (%)


-20

-30

-40

-50

-60

-70

-80
Group 1 Group 2 Group 3
-90
The number of layers
Fig. 7.30 Division of three groups in the non-perforated panel 11F24

7.3.2.3 Backface Signature (BFS) constraining layers

The third group composed of all back layers behind the second group. Dur ing
ballistic impact, these fabric layers can‟t be perforated, but just produce a certain
transverse deformation. According to the observation of post- impact panels (Chapter
3, section 3.3.2.4) and FE simulation (Chapter 7, section 7.1.2.2), Dyneema UD has
obvious advantage of smaller transverse deflection, which was beneficial to constrain
BFS of woven fabric layers.

In addition, according to experimental results of hybrid materials discussed before,


ballistic performance of hybrid panels was degraded when UD layers were placed on
the striking face (Chapter 7, section 7.2.2.1). Combining Dyneema UD in back layers
can avoid the thermal damage caused by the generated heat during the impact.
Therefore, Dyneema UD was used in the third group in a hybrid panel. In the multi-
layer system, all layers at back of the second group were classified into the third
group. The number of layers in the back group can be determined based on equation
(7-5).

To utilize the advantage of Dyneema UD in hybrid panels, the proper amount of


Dyneema UD in the panel was also investigated. Several hybrid panels combining
different numbers of UD layers behind Twaron fabric layers were carried out non-
perforation tests. The specific depth (SD) of the indentation was used here to
normalise BFS of different panels. Experimental test results showed that with the
increasing amount of UD layers, specific depth (SD) of the indentation behind non-
206
perforated panels decreased as shown in Fig. 7.31. In addition, BFS of the hybrid
panel must be between those panels made of single components. Such results
confirm the positive effect of Dyneema UD on constraining the transverse deflection
of woven fabric.

(7-6)
100% 4.0 UD ratio
Woven ratio
90% 3.5 Test SD
80%
3.0

Specific depth (cm3 /g)


Areal density ratio

70%
60% 2.5

50% 2.0
40% 1.5
30%
1.0
20%
10% 0.5

0% 0.0
11F24 11F20/U4 11F16/U8 11F12/U12 11F8/U16 11F4/U20 U24

Fig. 7.31 The specific depth of different hybrid panels

According to the regression of test data, it exhibited a linear relationship between the
specific depth of the indentation and mass fraction of Dyneema UD as shown in Fig.
7.32. Such results indicated that with more layers of Dyneema UD added, BFS of the
hybrid panel can decrease linearly. Therefore, Dyneem UD layers were used for all
layers in the third group.

3.50

3.00
Specific depth (cm3/ g)

2.50

2.00
y = -1.6542x + 3.4067
1.50 R²= 0.9644
1.00

0.50

0.00
0.0 0.2 0.4 0.6 0.8 1.0
Mass fraction of Dyneema UD
Fig. 7.32 The specific BFS of hybrid panel composed of 11F and Dyneema UD

207
7.4 Ballistic performance evaluation of the hybrid panels

7.4.1 Hybrid panels manufacture

According to the design principle of the hybrid panel, two better hybrid panels were
constructed to conduct the perforation and non-perforation ballistic tests. The striking
velocity of the projectile was around 460-500m/s.

The first designed hybrid panel were constructed for perforation ballistic tests to
evaluate the ballistic resistant capacity. Based on the division of fabric layers in the
reference panel 11F24 , the first group composed of front three layers. The second
group contained nine layers and the last group had twelve layers. The hybrid panel
was designed as the following manner. Three layers of heavyweight fabrics 11F as
heat resistant layers were placed in the first group. Due to the lightweight fabric 8F
has higher energy absorption capacity (SEA), eight layers of lightweight fabric 8F as
energy absorption layers were combined in the second group. To keep almost the
same weight as reference panel 11F24 , 15 layers Dyneema UD layers were placed in
in the third group. Such hybrid panel can‟t be perforated with the current impact
conditions due to the better components were combined in the hybrid panel. The
performance of the hybrid panel will be definitely improved with comparing with the
reference panel 11F24 .

Ballistic resistance tests that were similar as V50 tests were conducted referring the
procedure introduced in Chapter 3, Section 3.3.2.3. The number of layers of the
hybrid panel was reduced to the limit number of layer to achieve the 50% chance of
perforation at the striking velocity of 483m/s. When there were three perforated
panels and other three non-perforated panels in the six shot, the related number of
layers in this hybrid panel was identified as the limit number of layers. Finally, the
total limit number of layers of the hybrid panel was identified to be 17 layers.
Correspondingly, six layers of Dyneema UD were identified to be combined in the
third group to constrain BFS. As a result, the hybrid panel 11F3 /8F8 /U6 was
constructed. The ballistic resistant capacity of the hybrid panel was compared with
two reference woven panels 11F 15 and 8F19 with the same areal density.

The second hybrid panel was constructed for non-perforation tests to reflect the
positive effect on BFS. Correspondingly, the total number of layers in the hybrid
208
panel increased. To reinforce the constraining effect of BFS, heavyweight fabrics
13F were combined in the first group. Ten layers of lightweight fabrics 8F were
placed in the second group to achieve the improvement of energy absorption. Eight
Dyneema UD layers were placed in the third group to minimize BFS of whole panel.
Two reference woven panels 13F15 and 8F24 keep the same areal density as the hybrid
panels.

7.4.2 Ballistic performance evaluation of the hybrid panels

7.4.2.1 Ballistic resistance tests

The perforation tests results were listed in Table 7.2. With the same weight of panels,
the hybrid panel (11F3 /8F8 /U6 ) can stop 50% projectile at the average striking
velocity of 490.70m/s. Both woven panels 11F 15 and 8F19 were perforated even at
lower average striking velocities of 483.02m/s and 486.24m/s. It can be concluded
that the hybrid panel were not likely to stop the projectile compared with the uniform
woven panel with the same areal density.

Table 7.2 Ballistic test results of the hybrid panel and woven panel
Number
Hybrid Areal Striking Average of BFS
STD Status
panel density velocity vs (m/s) perforated (mm)
(g/m2 ) (m/s) layers
493.70 Non-Perforated 8 16.28
487.05 Non-Perforated 7 15.38
489.07 Non-Perforated 10 17.03
11F3 /8F8 /U6 2954.99 490.70 2.66
489.58 Perforated / /
493.70 Perforated / /
491.12 Perforated / /
485.04 Perforated / /
481.56 Perforated / /
481.06 Perforated / /
11F15 2952.75 483.02 2.89
483.75 Perforated / /
487.29 Perforated / /
479.41 Perforated / /
487.55 Perforated / /
484.54 Perforated / /
485.04 Perforated / /
8F19 2951.65 486.24 2.81
482.55 Perforated / /
490.61 Perforated / /
487.16 Perforated / /

209
7.4.2.2 Backface Signature (BFS) tests

Table 7.3 listed non-perforation test results of reference panels and hybrid panel. The
hybrid panel exhibited the significant positive hybridisation effect on BFS. With
almost the same areal density, the woven panel 13F15 was perforated. The hybrid
panel 13F3 /8F10 /U8 and woven panels 8F24 and 11F24 successfully stopped the
projectile of every shot. Compared with the woven panel 8F24 , the average BFS of
the hybrid panel 13F3 /8F10 /U8 showed 46.32% reduction with the same weight.
Compared with the reference panel 11F24 at the same range of striking velocity, BFS
of the hybrid panel 13F3 /8F10 /U8 decreased 16.20% in addition to the areal density
reduced 19.48%. Obviously, the hybrid panel can achieve the improvement ballistic
performance and reduction of weight.

Table 7.3 Ballistic test results of the hybrid panel and woven panel
Number Ave
Areal
Hybrid Ave vs of BFS BFS
density Status STD
panel (m/s) broken (mm) (mm)
(g/m2 )
layers
Non-Perforated 10 13.04
13F3 /8F10 /U8 3804.30 488.23 Non-Perforated 9 12.44 12.78 0.31
Non-Perforated 8 12.85
Perforated / /
13F15 3776.40 482.51 Perforated / / / /
Perforated / /
Non-Perforated 12 26.35
8F24 3728.40 485.71 Non-Perforated 12 27.38 27.59 1.35
Non-Perforated 12 29.03
Non-Perforated 8 15.75
11F24 4724.40 480.07 Non-Perforated 10 13.57 15.25 1.12
Non-Perforated 8 14.22

7.5 Summary

To optimise the construction of ballistic panel for soft armour panel, a procedure for
hybridisation design was established in this chapter. The multi- layer panels
composed of different components of ballistic materials in proper positions was
designed and engineered to achieve improvement of ballistic performance and
lightweight.

210
The components for the hybrid panel were selected based on their advantages and
disadvantages of material properties for ballistic protection. Due to the better energy
absorption capacity, the lightweight Twaron fabric was adopted in the woven panel.
To minimize the related increasing of BFS, Dyneema UD was combined behind to
constrain the large transverse deflection of the lightweight fabric, due to its obvious
advantage of lower BFS. According to the investigation of materials hybridisation
and weave hybrid structures, it is beneficial to place the lightweight fabric at back of
the heavyweight fabric to achieve better energy absorption capacity of the hybrid
panel, and combine Dyneema UD layer at the back of the panel to minimize BFS.

According to different ballistic responses of each layer in the front, middle and back
of the multi- layer system, a panel were discretely divided into three groups. The first
group composed of first few layers on the striking face. The heavyweight fabrics as
heat resistant layers were used to resist the generated heat. The second group
composed some middle layers close to the last perforated layers. The lightweight
fabric was combined to make full use of the energy absorption capacity. All back
layers were classified into the third group. Dyneema UD laminates were placed to
minimize BFS of the panel. The specific range of each group can be identified
according to the number of layers in the first group, the number of perforated layers
in a panel and total number of layers in a panel.

Based on above design principle, two hybrid panels were designed. The ballistic
performance was evaluated in the perforation and non-perforation ballistic tests. In
the case of perforation, the hybrid panel exhibited higher probability of stopping the
projectile compared with the uniform woven panel with the same areal density. In the
case of non-perforation, the hybrid panel achieved significant lower BFS and lower
areal density compared with the reference panel. Such ballistic tests results
confirmed the positive effect of the hybrid design. The ballistic performance of
ballistic resistant capacity and BFS of hybrid panels were found significantly
improved in addition to the reduction of areal density.

211
CHAPTER 8 Conclusions and future work

8.1 Conclusions

This research systematically revealed the different ballistic characteristics of each


layer in different positions of an armour panel and the way of energy absorption in
the panel. Such findings contribute to the understanding of different ballistic
responses in different positions of an armour panel under ballistic impact. According
to such theoretical understandings, a new hybrid design concept was put forward.
The positive hybridisation effect has been proved by a series of ballistic tests. Such
hybrid design makes best use of different available materials to achieve an
improvement of ballistic performance and reductions in weight. It has a practical
significance for the armour panel design.

Objectives set out for this PhD research include: (1) a comprehensive investigation
of different ballistic characteristics of each layer in armour panels, including the
transverse deflection, stress distribution, energy absorp tion efficiency and failure
modes; (2) optimising the construction of armour panels by hybridisation. This
included the investigation on ballistic performance of different components (Twaron
woven fabric and Dyneema UD laminate), analysis of layering seque nce,
development of a new hybrid design concept and constructing the hybrid armour
panels with the improved performance. Two complementary methodologies, namely
the empirical method and Finite Element (FE) simulation, were used to reflect the
ballistic response of armour panel under impact.

This research has led to the following conclusions:

(a) During ballistic impact, the ballistic resistant efficiency of each layer was not
the same at the front, middle and back positions of an armour panel.

(1) When the armour panel was under ballistic impact, the ballistic characteristics of
each layer were numerically quantified. It was found that in the armour panel the
fabric layers at the front, middle and back positions exhibited different characteristics.
In the front few layer, high stress concentrated in the impact area with the smallest
transverse deformation. While the middle layers close to the last perforated layer had
wider transverse deformation area than the front and the back layers. The stress in
212
middle layers can propagate widely. The non-perforated back layers have low stress
magnitude although the stress can propagate to the edge of fabric. These different
characteristics were also qualified from different failure morphologies of fibres,
yarns and fabric on each layer by photographic observations of the post- impact
panels.

(2) In a non-perforated panel, energy absorption of each layer increased from front to
the maximum value in the last perforated layer and then decreased gradually in the
following back layers. Such energy absorption distribution in the multi- layer system
remained the same pattern at different striking velocities. Only the position of the
peak value in energy absorption was shifted towards the back of the panel w ith the
increasing striking velocities. Therefore, the energy absorption distribution in the
perforated panel can be regarded as a specific case of the high striking velocity.

(3) According to energy absorption efficiency, energy absorption of the front layers
was constrained and reduced by additional layers. The same amount of fabric
materials played different roles in energy absorption due to different positions in the
panel. With the increasing number of layers in a panel, the energy absorption of the
panel increased but the growth rate decreased. When the total number of layers
exceeded a certain value, the additional layers in the panel contributed less to the
total energy absorption.

(4) In a non-perforated panel, the average number of perforated layers was found to
be in a rather small range. In addition, it would reduce to a constant value when the
total number of layers exceeded a certain value. Therefore, the position of the last
perforated layer which has the highest energy absorption can be identified in a small
region in the panel. It should be noted that the number of broken layers was affected
by the impact energy, which was determined by different striking velocities and the
types of projectile.

(5) The Backface Signature (BFS) of the non-perforated panels and the indentation
volume both continually decreased with the increasing number of layers. The
increasing number of layers in the panel has a great influence on minimizing BFS
behind the panel, although it has little contribution to the improvement of the total
energy absorption.

213
Such findings contribute to the theoretical understandings of different ballistic
requirements in different positions of an armour panel under ballistic impact. It
provides a guide for the hybrid soft armour panel design. The different materials can
be selected for different positions of an armour panel according to different ballistic
requirements.

(b) The hybrid design of an armour panel can provide better constructions to
achieve higher ballistic resistant capacity and lower Backface Signature (BFS)

(1) The components for hybridisation were selected based on their advantages and
disadvantages of material properties for ballistic protection. According to the ballistic
tests results, the lightweight Twaron fabric exhibited the better energy absorption
capacity than the heavyweight fabric. The woven panel can achieve the improvement
of energy absorption performance by combining more lightweight fabric layers
instead of utilising fewer heavyweight fabric layers for a given weight. However, the
related Backface Signature (BFS) of lightweight fabric panels increased significantly.

Dyneema UD laminate displayed an obvious advantage of lower BFS than that of the
woven fabric. However, due to the low melt point, the generated heat on the striking
face of an armour panel can cause the thermal damage of Dyneema fib res during
impact. When the UD laminates were placed on the striking face, the ballistic
performance was shown obvious degradation.

(2) According to the investigation on the layering seque nce of hybridisation, it is


beneficial to place the lightweight fabric behind the heavyweight fabric to achieve
the better energy absorption capacity of the hybrid panel. To avoid the ballistic
performance degradation due to the thermal damage, Dyneema UD laminates were
considered to be combined behind the woven fabric to constrain the transverse
deflection and minimize BFS.

(3) Based on the energy absorption efficiency of each layer in an armour panel under
ballistic impact and the ballistic characteristics of different component, a new
systematic hybrid design concept was developed and a procedure for hybridisation
design was established.

An armour panel was divided into three groups. The first group was composed of the
first few layers on the striking face. The heavyweight fabrics as the heat resistant
214
layers were used to resist the generated heat on the striking face. The second group
included some middle layers close to the last perforated layers. In this group, the
lightweight fabric was placed to make full use of the energy absorption capacity. All
back layers were classified into the third group, which combined Dyneema UD
laminates to minimize the BFS of the panel. The number of layers in each group can
be identified according to the number of layers in the first group, the number of the
perforated layers and the total number of layers in a reference panel.

(4) Two hybrid panels were constructed and evaluated. In the perforation ballistic
tests, compared with several Twaron woven panels with the same areal density, the
hybrid panel were more likely to stop the projectile at an impact velocity of 483m/s.
In the case of non-perforation ballistic tests, the hybrid panel achieved significant
lower BFS with the reduced weight.

This part of work is original and has a practical significance for the armour panel
design. Such hybrid design makes best use of different available materials to achieve
a balance ballistic performance. This is very meaningful for material selection for
ballistic protection. As a result, the hybrid panel can provide a balance and stable
ballistic performance.

8.2. Recommendations for future research work

To further investigate the ballistic response of soft body armour and perfect the
hybrid design of the armour panel, some future recommendations could be set as
continuations to the current work.

The material properties of Twaron fabric and Dyneema UD laminate need to be


accurately measured, in particular for the transverse modulus perpendicular to the
fabric layer and in-plan shear modulus. A scientific test procedure for material
properties should be further explored to provide more reliable parameters for the FE
model of the multi-layer panel at the fabric level.

In the future work, FE model of the hybrid panel combined with Twaron woven
fabric and Dyneema UD laminate should be explored. Different ballistic
characteristics of each component layer in a hybrid panel should be further analysed.
To improve the computation efficiency, the shell model can be considered for

215
Twaron woven fabric and composite model can be used for Dyneema UD laminate
as another alternative approach for modelling the multi- layer panel.

The hybrid design concept will be developed at different threat levels. The best
group division of a panel for different threat levels need to be further investigated to
provide the theoretical guide for hybrid design. The different impact conditions such
as different types of projectile or bullet, and impact trajectory will be taken into
account.

In order to perfect the hybrid design concept, the material selection will be extended
to a large range of ballistic materials for hybridisation, such as Flex (Kevlar UD),
Spectra, Dyneema woven panel and so on. In this way, different current available
materials can be made full use to provide the balance ballistic performance.

Evaluation of ballistic performance of the hybrid panel needs to be more


systematically. Different impact conditions should be taken into account to ensure
the necessary adaptability during usage.

216
REFERENCE

1. Starley, D., Determining the Technological Origins of Iron and Steel. Journal.
of Archeology. Science., 1999. 26: p. 1127-1133.

2. David, N.V., Gao, X., and Zheng, J., Ballistic Resistant Body
Armor :Contemporary and Prospective Materials and Related Protection
Mechanisms. Applied Mechanics Reviews, 2009. 62: p. 1-20.

3. Cavallaro, P.V., Soft Body Armor: An Overview of Materials, Manufacturing,


Testing, and Ballistic Impact Dynamics, in NUWC-NPT Technical Report
12,057. 2011, Naval Undersea Warfare Center Division: Newport, Rhode
Island.

4. Azrin Hani, A.R., Roslan, A., Mariatti, J., and Maziah, M., Body Armor
Technology: A Review of Materials, Construction Techniques and
Enhancement of Ballistic Energy Absorption. Advanced Materials Research,
2012. 488-489: p. 806-812.

5. Bajaj, P., and Sriram, Ballistic protective clothing: An overview. Indian


Journal of fibre &textile research, 1997. 22: p. 274-291.

6. Lane, R.A., Hign performance fibers for personnel and vehicle armor s ystems.
The AMPTIAC Quarterly, 2005. 9: p. 3-9.

7. Ballistic Resistance of Body Armor NIJ Standard-0101.06.,U.S. Department


of Justice, National Institute of Justice, 2008, National Institute of Justice:
Washington.

8. McConnell, V.P., Ballistic protection materials a moving target. Reinforced


plastics, 2006(12): p. 20-25.

9. Utracki, L.A., Rigid ballistic composites, Archive, National research council


Canada Publications Archive, 2010.

10. Lamothe, D., Corps to field two new body armour vests, in Marine Corps
Times (Gannett). 2009. p. 19.

11. Karahan, M., Kus, A., and Eren, R., An Investigation into Ballistic
Performance and Energy Absorbtion Capabilities of Woven Aramid Fabrics.
J. Impact Eng, 2008. 35(6): p. 499–510.

12. Park, J.L., Chi, Y.S., Hahn, M.H., and Kang, T.J., Kinetic Dissipation in
Ballistic Tests of Soft Body Armors. Experimental Mechanics, 2012. 52(8): p.
1239-1250.

13. Poh, C.W., Investigation of new materials and methods of construction of


personnel armour. Master thesis. Naval postgraduated school, Monterey,
California, 2004,p. 95.

14. Jacobs, M.J.N., and Van Dingenen, J.L.J., Ballistic Protection Mechanisms in
Personal Armour. Journal of Materials Science, 2001. 36: p. 3137 – 3142.
217
15. Dusablon, L.V., The Casualty Reduction Analysis Model for Personnel
Armor Systems. Army Natick Research, Development and Engineering
Center. 1972.

16. Carrol, A., and Soderstrom, C., A new nonpenetrating ballistic injury. Ann
Surg, 1978. 6: p. 753-757.

17. Johnson, W., Collins, C., and Kindred, F., A Mathema tical Model for
Predicting Residual Velocities of Fragments After Perforating Helmets,
Ballistic Research Laboratories Technical, 1968, p.1705.

18. Ivanov, I., and Tabiei, A., Loosely Woven Fabric Model With Viscoelastic
Crimped Fibres for Ballistic Impact Simulations. Journal of Numerical.
Methods Engineer., 2004. 61: p. 1565-1583.

19. High-performance fibres, ed. Hearle, J.W.S. Cambridge, England: Woodhead


Publishing Limited. 2001, p.337.

20. Languerand, D.L., Zhang, H.,Murthy, N.S.,Ramesh, K.T.,Sansoz, F., Inelastic


behavior and fracture of high modulus polymeric fiber bundles at high strain-
rates. Materials Science and Engineering A. 2009. 500: p. 216-224.

21. Prevorsek, D.C., Ballistic Armor Material from Spectra ® fibre, in the 33rd
International SAMPE Symposium 1988.

22. Chocron, S., From Fibers to Composites, in Conference at Airbus Military.


Southwest Research Institute: San Antonio, Texas. 2009.

23. Bhatnagar, A., Lightweight ballistic composites. Military and law-


enforcement applications. Cambridge (England): Woodhead Publishing in
Materials. 2006.

24. Prevorsek, D.C., and Lewin, M., High Technology Fibers Part D, ed. Preston,
Marcel Dekker, Inc. USA. 1996.

25. Hearle, J.W.S., ed. Atlas of fibre fracture and damage to textiles,Woodhead
publishing limited, Cambridge,1998, p. 329.

26. Prevorsek, D.C., Chin, H.B., Kwon,Y.D., and Field, J.E., Strain rate effects in
ultrastrong polyethylene fibers and composites. Journal of Applied Polymer
Science, 1991. 47: p. 45-66.

27. Wang, Y., and Xia, Y., Experimental and Theoretical Study on the Strain
Rate and Temperature dependence of Mechanical Behavior of Kevlar Fiber.
Composites , Part A, 1999. 30: p. 1251-1257.

28. Gu, B., Analytical Modeling for the Ballistic Perforation of Planar Plain-
woven Fabric Target by Projectile. Composites Part B, 2003. 34B(4): p. 361-
371.

29. Zeng, X., Numerical analysis of fabric armour under ballistic impact. PhD
thesis, National University of Singapore. 2006.

218
30. Creasy, T.S.,Modeling Analysis of Tensile Tests of Bundled Filaments with a
Bimodal Weibull Survival Function. Journal of Composite. Material. 2002.
36: p. 183–194.

31. Cheng, M., Chen, W., and Weerasooriya, T., Mechanical Properties of
Kevlar® KM2 Single Fiber. Journal of Engineering Materials and Technology.
2005. 127: p. 197-203.

32. Shim, V.P.W., Lim, C.T., and Foo, K.J., Dynamic mechanical properties of
fabric armour. Impact Engineering, 2001. 25(1): p. 1-15.

33. Chocron, S., Rodríguez, J., Martí


nez, M.A., Sánchez-Gálvez, V., High strain
rate properties of aramid and polyethylene woven fabrics composites.
Composites-Part B: engineering, 1996. 27B: p. 147-154.

34. Russell, B.P., Karthikeyan, K., Deshpande, V.S., Fleck, N.A., The high strain
rate response of Ultra High Molecular-weight Polyethylene: From fibre to
laminate. International Journal of Impact Engineering, 2013. 60: p. 1-9.

35. Koh, C.P., Shim, V.P.W., Tan,V.B.C., Tan, B.L., Response of a high-strength
flexible laminate to dynamic tension. International Journal of Impact
Engineering 2008. 35: p. 559-568.

36. Govaert, L.E., and Peijs, T., Tensile strength and work of fracture of oriented
polyethylene fibre. Polymer, 1995. 36: p. 4425-4431.

37. Marzena Fejdyś, Marcin Łandwijt, and Struszczyk, M.H., Effect of


Accelerated Ageing Conditions on the Degradation Process ocomf
Dyneema® Polyethylene Composites. Fibres & Textiles in Eastern Europe
2011. 19(1 (84) ): p. 60-65.

38. Huang, W., Wang, Y., and Xia, Y., Statistical dynamic tensile strength of
UHMWPE-fibers. Polymer, 2004. 45: p. 3729–3734.

39. Cheeseman, B.A., and Bogetti, T.A., Ballistic impact into fabric and
compliant laminates. Composite Structures, 2003. 61: p. 161-173.

40. Roylance, D., and Wang, S., Penetration mechanics of textile structures.In:
Laible RC, editor. Ballistic Materials and Penetration Mechanics. Elsevier
Publishing New York. 1980.

41. Roylance, D., Stress wave-propagation in fibers-effects of crossovers. Fiber


Science Technology. 1980. 13(5): p.385-95.

42. Cunniff, P., Dimensionless parameters for optimization of textile-based body


armour systems, in 18th International symposium on ballistics conference.
San Antonio, USA.1999: p.1303-1310.

43. Smith, J.C., McCrackin, F.L., and Schiefer, H.F., Stress–strain relationships
in yarns subjected to rapid impact loading. Part V: wave propagation in lo ng
textile yarns impacted transversely. Journal of Research of the National
Bureau of Standards, 1958. 28: p. 288-302.

219
44. Cunniff, P.M., An analysis of the system effects in woven fabrics under
ballistic impact. Textile Research Journal, 1992. 62(9): p. 495-509.

45. Roylance, D., Wilde, A., and Tocci, G., Ballistic impact of textile structures.
Textile.Research.Journal, 1973: p. 34-41.

46. Duan, Y., Keefe, M., Bogetti, T.A., Cheeseman, B.A., Modeling friction
effects on the ballistic impact behavior of a single-ply high- strength fabric.
International Journal of Impact Engineering 2005. 31: p. 996-1012.

47. Lee, B.L., Walsh, T. F., Won, S. T., Patts, H. M., Song, J. W.,and Mayer, A.
H. Penetration failure mechanisms of armor- grade fiber composites under
impact. Journal of Composite Materials, 2001. 35(18): p.1605–33.

48. Tan, V.B.C., Lim, C.T., and Cheong, C.H., Perforation of high-strength fabric
by projectiles of different geometry. International Journal of Impact
Engineering, 2003. 28(2): p. 207-222.

49. Rao, Y. and Farris, R.J., A Modeling and Experimental Study of the
Influence of Twist on the Mechanical Properties of High-Performance Fiber
Yarns. J. Applied. Polymer. Science., 2000. 77: p. 1938–1949.

50. Yang, H.H., Kevlar aramid fibre. 1st ed. Wiley: New York. 1993, p.210.

51. Stein, W., Construction and action of bullet resistant vests. Melliand Textile,
1981. 62(6): p. 595-602.

52. Hearle, J.W.S., Leech, C.M., Adeyefa, A., and Cork, C.R., Ballistic impact
resistance of multi- later textile fabrics. Department of Textile Technology,
UMIST. 1981, p. 244.

53. Shim, V.P.W., Tan, V.B.C., and Tay, T.E., Modelling deformation and
damage characteristics of woven fabric under small projectile impact.
International Journal of Impact Engineering, 1995. 16(4): p. 585-605.

54. Prosser, R.A., Cohen, S.H., and Segars, R.A., Heat as a Factor in the
Penetration of Cloth Ballistic Panels by 0.22 Caliber Projectiles. Text ile.
Research. Journal, 2000. 70: p.709-722.

55. Lim, C.T., Tan, V.B.C., and Cheong, C.H., Perforation o f high-strength
double-ply fabric system by varying shaped projectiles. International Journal
of Impact Engineering. 2002. 27: p. 577-591.

56. Chitrangad, H., Midlothian, V., Hybrid ballistic fabric. Patent. DuPont de
Nemours and company, Wilmington, Del. 1993

57. Cork, C.R., and Foster, P.W., The ballistic performance of narrow fabrics.
International Journal of Impact Engineering. 2007. 34: p.495–508.

58. Scott, R.A., ed. Textiles for protection, Woodhead Publising Limited, 2005,
p.532-533.

220
59. Zhou, Y., Development of Lightweight Soft Body Armour for Ballistic
Protection. University of Manchester: Manchester. 2013, p. 218.

60. Mcdaniels, K., Downs, R.J., Meldner, H., Beach, C., Adams, C., High
strength-to-weight ratio non-woven technical fabrics for acrospace
applications. Cubic Tech Corp, 2009: p. 1-9.

61. Tan, V.B.C., Shim, V.P.W., and Zeng, X., Modelling crimp in woven fabrics
subjected to ballistic impact. International Journal of Impact Engineering
2005. 32: p. 561–574.

62. Karahan, M., Comparison of Ballistic Performance and Energy Absorption


Capabilities of Woven and Unidirectional Aramid Fabrics. Textile Research
Journal 2008. 78: p. 718.

63. Wetzel, E.D., Kirkwood, K.M., and Kirkwood, J.E., Yarn Pull-Out as a
Mechanism for Dissipating Ballistic Impact Energy in Kevlar KM-2 Fabric,
Part I: Quasi-Static Characterization of Yarn Pull-Out. Text. Res. J., 2004: p.
920-928.

64. Tabiei, A., and Nilakantan, G., Ballistic Impact of Dry Woven Fabric
Composites: A Review. Applied Mechnical Reviews, 2008. 61: p. 1-13.

65. Zeng, X., Tan, V.B.C., and Shim, V.P.W., Modelling inter-yarn friction in
woven fabric armour. Impact Engineering. 2006. 66: p. 1309-1330.

66. Bazhenov, S., Dissipation of energy by bulletproof aramid fabric. Mater ial
Science. 1997. 32(15): p. 4167-4173.

67. Ahmad, M.R., Ballistic Response of Natural Rubber Latex Fabric Systems.
Journal of Rubber Research, 2007. 10(4): p. 207-221.

68. Briscoe, B.J., and Motamedi, F., The Ballistic Impact Characters of Aramid
Fabrics:The Influence of Interface Friction. Text Res J, 1992. 158: p. 229-247.

69. Barnes, H.A., Shear-Thickening in Suspensions of Nonaggregating Solid


Particles Dispersed in Newtonian Liquids. Journal of Rheology, 1989. 33: p.
329-366.

70. Struszczyk, M.H., Puszkarz, A.K., Wilbik-Halgas, B., Cichecka, M., Litwa,
P., Urbaniak-Domagala, W., Krucinska, I., The surface modification of
ballistic textiles using plasma-assisted chemical vapor deposition (PACVD).
Textile Research Journal, 2014. 0(00): p. 1-9.

71. Sun, D., Chen, X., Plasma modification of Kevlar fabrics for ballistic
applications. Textile Research Journal, 2012. 82: p. 1928-1934.

72. Lee, Y.S., Wetzel, E.D., and Wagner, N.J., The ballistic impact
characteristics of Kevlar woven fabrics impregnated with a colloidal shear
thickening fluid. Journal. Material. Science., 2003. 38: p. 2825-2833.

73. Hogenboom, E.H.M., and Bruinink, P., Combinations of Polymer Filaments


or Yarns.patent. Netherland. 1991: p. 3.
221
74. Chocron, S., King, N., Bigger, R., Walker, J.D., Heisserer, U., and van der
Werff, H., Impacts and Waves in Dyneema HB80 Strips and Laminates
Journal of Applied Mechanics, 2013. 80: p. 031806-1-10.

75. Gray, W. and Zabal, P.H., Quality Assurance and Quality Control to
Thermophysical Properties, in Interantional encyclopaedie, VCH Publishers
Ins: New York, USA. 1991, p. 125-140.

76. Montgomery, T.G., Grady, P.L., and Tomasino, C., The effects of projectile
geometry on the performance of ballistics fabrics. Text Res J, 1982. 52(7): p.
442-450.

77. Zhang, G.M., Batra, R.C., and Zheng, J., Effect of frame size, frame type, and
clamping pressure on the ballistic performance of soft body armor.
Composites: Part B, 2008. 39: p. 476-489.

78. Chocron, S., Figueroa, E., King, N., Kirchdoerfer, T., Nicholls, A.E.,
Sagebiel, E., Weiss, C., and Freitas, C.J., Modeling and validation of full
fabric targets under ballistic impact. Composites Science and Technology,
2010. 70: p. 2012-2022.

79. Mamivand, M., and Liaghat, G.H., A model for ballistic impact on multi-
layer fabric targets. International Journal of Impact Engineering, 2010. 37:
p.806-812.

80. Zeng, X., Shim, V.P.W., and Tan, V.B.C., Influence of boundary conditions
on the ballistic performance of high-strength fabric targets. International
Journal of Impact Engineering, 2005. 32: p. 631-642.

81. Porwal, P.K., and Phoenix, S.L., Modeling system effects in ballistic impact
into multi- layered fibrous materials for soft body armor International Journal
of Fracture, 2005. 135: p217-249.

82. Shockey, D.A., Erlich, D.C., and Simons, J.W., Improved Barriers to Turbine
Engine Fragments: Interim Report III. 2004.

83. Novotny, W.R., Cepus, E., Shahkarami, A., Vaziri, R., Poursartip, A.,
Numerical investigation of the ballistic efficiency of multi-ply fabric armours
during the early stages of impact. International Journal of Impact Engineering,
2007. 34: p. 71-88.

84. Gogineni, S., Gao,X., David, N.V., Zheng, J., Ballistic Impact of Twaron
CT709 Plain Weave Fabrics. Mechanics of Advanced Materials and
Structures, 2012. 19: p. 441-452.

85. Cunniff, P.M., Decoupled response of textile body, in Proceedings of the 18th
International Symposium on Ballistics. 1999: San Antonio, Texas, US.

86. Joo, K., and Kang, T.J., Numerical Analysis of Energy Absorption
Mechanism in Multi-ply Fabric Impacts. Textile Research Journal, 2008. 78:
p. 561.

222
87. Joo, K., and Kang, T.J., Numerical Analysis of Multi-Ply Fabric Impacts.
Textile Research Journal. 2007. 77: p. 359.

88. Prosser, R.A., Penetration of Nylon Ballistic Panels by Fragment-Simulating


Projectiles Part I: A Linear Approximation to tlic Relationship between the
Square of the V50 or Vc Striking Velocity and the Number of Layers of
Cloth in the Ballistic Panel. Textile Research Journal.1988. 2: p. 61-85.

89. Zohdi, T.I., Modeling and simulation of progressive penetration of


multilayered ballistic fabric shielding. Computational Mechanics. 2002. 29: p.
61-67.

90. Shen, W., Niu, Y., Laurence,P., Link,N., Characterizing the Interaction
Among Bullet, Body Armor, and Human and Surrogate Targets. Journal of
Biomechanical Engineering, 2010. 132: p. 1-11.

91. Chabba, S., Vanek, D., and Cordova, D., Multilayerd material sheet for use in
soft ballistics. 2010: US.

92. Heinecke, J., From fibre to armor, in Law Enforcement Technology magazine.
2007.

93. Zufle, T.T., Soft body armor. United States patent. 1993.

94. Porwal, P.K., and Phoenix, S.L., Effects of layer stacking order on the V50
velocity of a two-layered hybrid armor system. Journal of mechanics of
Material and structures, 2008. 3: p. 627-639.

95. Park, J.L., Chi, Y.-S., and Kang, T.J., Ballistic performance of hybrid panels
composed of unidirectional/woven fabrics. Textile Tesearch Journal, 2012.
0(00): p. 1-16.

96. Rahman, M., Morozov, E.V., Shankar, K., Fien, A., Computational Analysis
of Low Velocity Impact Response of Composite Panels. Applied Mechnism.
Material, 2012. 157: p. 1135-1138.

97. Bicerano, J., Prediction of polymer properties. 2nd ed. 1996, New York:
Marcel Dekker Inc.

98. Alger, M.S.M., Polymer science dictionary. 2nd ed. 1997, London SE1 8HN,
UK: Chapman&Hall. 631.

99. Kausch, H.-H., Polymer Fracture. 2nd ed. 1987, Berlin Heidelberg: Springer-
Verlag 456.

100. Lee, B.L., Song, J.W., and Ward, J.E., Failure of spectra polyethylene fiber
reinforced composites under ballistic impact loading. J Compos Mater, 1994.
28(13).

101. Termonia, Y., Meakin, P., and Smith, P., Theoretical Study of the Influence
of Strain Rate and Temperature on the Maximum Strength of Perfectly
Ordered and Oriented Polyethylene. Macromolecules, 1986. 19(1): p. 154-
159.
223
102. Greenhalgh, E.S., Bloodworth, V.M., Iannucci, L., Pope, D., Fractographic
observations on Dyneema composites under ballistic impact. Composites:
Part A, 2013. 44: p. 51-62.

103. Iremonger, M.J., Polyethylene composites for protection against high velocity
small arms bullets, in Proceedings of the 18th International Symposium on
Ballistics. San Antonio, Texas. 1999:p. 946-953.

104. Prevorsek, D.C., Kwon, Y.D., and Chin, H.B., Analysis of the temperature
rise in the projectile and extended chain polyethylene fiber composite during
ballistic impact and penetration. Polymer Engineer Science. 1994. 34(2). p.
141-152.

105. Chitrangăd, H., Ballistic Structure. patent. DuPont De Nemours And


Company. 1994,

106. Prosser, R.A., Penetration of Nylon Ballistic Panels by Fragment-Simulating


Projectiles Part II: Mechanisms of Penetration. Textile. Research. Journal,
1988. 3: p. 161-165.

107. Chen, X., Zhu, F., and Wells, G., An analytical model for ballistic impact on
textile based body armour. Composites: Part B 2013. 45: p. 1508-1514.

108. Iremonger, M., Ballistic impact of fibre composite armours by fragment-


simulating projectiles. Composites, Part A 1996. 27A: p. 575–581.

109. Scott, B., The penetration of compliant laminates by compact projectiles. in


Proceedings of the 18th International Symposium on Ballistics. San Antonio,
Texas,USA. 1999.

110. Critescu, N., Malvern, L.E., and Sierakowski, R.L., Failure Mechanisms in
Composite Plates Impacted by Blunt-Ended Penetrators. Foreign Object
Impact Damage to Composites.American Society for Testing and Materials
STP568, 1975: p. 159-172.

111. Zhang, Z., Shen, S., Song, H., Zhang, D., Ballistic penetration of Dyneema
fiber laminate. Journal of Material Science Technology, 1998(14): p. 265-268.

112. Coffey, C.S., and Armstrong, R.W., Description of „Hot Spots‟ Associated
With Localized Shear Zones in Impact Tests. Naval Surface Weapons Center:
Dahlgren,Virginia. 1980,

113. Powell, D.A., and Zohdi, T.I., Attachment mode performance of network-
modeled ballistic fabric shielding. Composites: Part B, 2009. 40: p. 451-460.

114. Xing, Z., Wang, M., liu, W., Hua, J., Wu, G., Crystal structure and
mechanical properties of UHMWPE-g-PMA fiber prepared by radiation
grafting. Radiation Physics and Chemistry, 2013. 86: p. 84-89.

115. Grujicic, M., Arakere, G., He, T., Gogulapati, M., Cheeseman, B. A., A
numerical investigation of the influence of yarn-level finite-element model on

224
energy absorption by a flexible- fabric armour during ballistic impact, in Proc.
IMechE. Part L: J. Materials: Design and Applications. 2008. p. 259-276.

116. Grujicic, M., Grujicic, M., Arakerea, G., Hea, T., Bell, W.C., Cheeseman,
B.A., Yen, C.F., Scott, B., A ballistic material model for cross-plied
unidirectional ultra-high molecular-weight polyethylene fiber-reinforced
armor- grade composites. Materials Science and Engineering A, 2008. 498: p.
231-241.

117. Shockey, D.A., Erlich, D.C., Simons, J.W., Improved barriers to turbine
engine fragments,, in US department of transportation. SRI International,
Menlo Park. 2000:

118. Gu, B.H., Ballistic Penetration of Conically Cylindrical Steel Projectile Into
Plain-Woven Fabric Target: A Finite Element Simulation. Compos.Mater.,
2004. 38(22).

119. Ting, C., Ting, J., Cunniff, P., Roylance, D.K., Numerical Characterization of
the Effects of Transverse Yarn Interaction on Textile Ballistic Response. in
Proceedings of the 30th International SAMPE Technical Conference. 1998.

120. Scott, B.R., and Yen, C.F., Analytic design trends in fabric armor.
Proceedings of the 22nd international ballistics symposium, 2005.

121. Phoenix, S.L., and Porwal, P.K., A new membrane model for the ballistic
impact response and V50 performance of multi-ply fibrous systems.
International Journal of Solids and Structure, 2003(40): p. 6723-6765.

122. Parga- Landa, B., and Hernandez-Olivers, F., An analytical model to predict
impact behaviour of soft body armors. International Journal of Impact
Engineering 1995. 16(3): p. 455-466.

123. Ben-Dor, G., Dubinsky, A., and Elperin, T., Ballistic Impact: Recent
Advances in Analytical Modeling of Plate Penetration Dynamics–A Review.
Applied Mechanics Reviews, 2005. 58: p. 355-371.

124. Zok, F., Fleck, N., and Deshpande, V., A Design Tool for Robust Composite
Structures. The University of California, Santa Barbara: Santa Barbara, CA
93106. 2010, p. 22.

125. ASTM D3039, Standard Test Method for tensile test standard for composite,
ASTM International, West Conshohocken, PA, 2004.

126. Mohammed, U., Lekakou,C., Dong, L., Bader, M.G., Shear deformation and
micromechanics of woven fabrics. Composites: Part A 2000(31): p. 299–308.

127. Lim, C.T., Shim, V.P.W., and Ng, Y.H., Finite-element modeling of the
ballistic impact of fabric armor. Impact Engineering, 2003. 28(1).

128. Abaqus Analysis User's Manual, ed. 6.7, V. 2007, Availalbe at:
http://abaqus.ethz.ch:2080/v6.11/books/gsa/default.htm

225
129. Kaw, K.A., Mechanics of composite materials. 2nd ed. Taylor&Francis
Group. 2005, p.490.

130. WANG,Y., High Strain Rate Damage Mechanism in 2D/3D Ballistic Fabrics.
Final report. EPSRC/DSTL project EP/G042829/1, 2013.

131. Rao, M.P., Duan,Y., Keefe, M., Powers, B.M., Bogetti, T.A., Modeling the
effects of yarn material properties and friction on the ballistic impact of a
plain-weave fabric Composite Structures. 2009(89): p. 556-566.

132. Hayhurst, C., Leahy, J., van der Jagt-Deutekom, M., Jacobs,M., Kelly,P.,
Development of material models for numerical simulation of ballistic impact
onto polyethylene fibrous armour. Personal armour systems symposium,
2000. Sep: p.5-8

133. Roberts, J.C., Ward, E. E., Merkle, A. C., O‟Connor, J.V., Assessing Behind
Armor Blunt Trauma in Accordance With the National Institute of Justice
Standard for Personal Body Armor Protection Using Finite Ele ment
Modeling. TRAUMA, 2007. 62(5): p.1127-1133.

134. Pamukcu, S., Naito, C., Pervizpour, M., Cui, Q., Trasborg, P. A., Medina, C.,
Mentzer, M., Distributed Fiber-Optic Sensing and Numerical Simulation of
Shock Wave Response of Manufactured Clay. Aberdeen Proving Ground,
MD:U.S. Army Research Laboratory.2013.

135. Bunsell, A.R., The tensile and fatigue behavour of Kevlar-49(PRD-49) fibre.
Journal of Material science, 1975. 10: p. 1300-1308.

136. Marissen, R., Design with Ultra Strong Polyethylene Fibers. Ma terials
Sciences and Applications, 2011. 2: p. 319-330.

137. Ugural, A.C., and Fenster S. K., Advanced strength and applied elasticity.
Third Edition ed. PTR Prentice-Hall: New Jersey. 1995, p.89.

138. Bullido, A. P., Numerical analysis of impact behaviour on aeronautical


composite protections. University Carlos III De Madrid: Madrid.2011, p53

139. ASTM D3883-04(2012), Standard Test Method for Yarn Crimp and Yarn
Take-up in Woven Fabrics, ASTM International, West Conshohocken, PA,
2012,

140. Sabit Adanur, Handbook of weaving, CRC Press LLC, N.W. Corporate Blvd.,
Boca Raton, Florida, 2000, p 245

226
Appendix 1 The thickness of specimen

Table 1 The thickness of specimen

Sample Compressed Thickness Visual thickness


Material
label (mm) (mm)
8F 0.20 0.25
9F 0.22 0.27
10F 0.25 0.28
11F 0.26 0.31
Twaron
woven 12F 0.30 0.34
fabric
13F 0.32 0.36
7M 0.20 0.25
10M 0.34 0.36
7T 0.30 0.34

Dyneema 7U 0.24 0.26


UD 5U 0.30 0.35

227
Appendix 2 Fractured yarns in Twaron woven panel 8F 9

Table 2 Fractured yarns in the Twaron woven panel 8F9

Layer number No. of fracture yarns

ply1 5
ply2 5
ply3 5
ply4 5
ply5 5
ply6 5
ply7 3.5
ply8 2.5
ply9 1

228

You might also like