You are on page 1of 324

Impact and Energy Absorption of

Straight and Tapered Rectangular Tubes

By

Gregory Nagel, B.Eng. (Hons.)

A THESIS SUBMITTED TO THE SCHOOL OF CIVIL ENGINEERING


QUEENSLAND UNIVERSITY OF TECHNOLOGY IN PARTIAL
FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF
DOCTOR OF PHILOSOPHY

FEBRUARY 2005
Keywords

Tapered tubes; Energy absorption; Axial, oblique impact; Finite element analysis;
Computer simulation; Crashworthiness.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes i
Abstract

Over the past several decades increasing focus has been paid to the impact of
structures where energy, during the impact event, needs to be absorbed in a
controlled manner. This has led to considerable research being carried out on energy
absorbers, devices designed to dissipate energy during an impact event and hence
protect the structure under consideration. Energy absorbers have found common
usage in applications such as vehicles, aircraft, highway barriers and at the base of
lift shafts. A type of energy absorber which has received relatively limited attention
in the open literature is the tapered rectangular tube. Such a structure is essentially a
tube with a rectangular cross-section in which one or more of the sides are inclined to
the tube’s longitudinal axis.

The aim of this thesis was to analyse the impact and energy absorption response of
tapered and non-tapered (straight) rectangular tubes. The energy absorption response
was quantified for both axial and oblique loading, representative of the loading
conditions typically encountered in impact applications. Since energy absorbers are
commonly used as components in energy absorbing systems, the response of such a
system was analysed which contained either straight or tapered rectangular tubes as
the energy absorbing components. This system could typically be used as the front
bumper system of a vehicle.

Detailed finite element models, validated using experiments and existing theoretical
and numerical models, were used to assess the energy absorption response and
deformation modes of straight and tapered tubes under the various loading
conditions. The manner in which a thin-walled tube deforms is important since it
governs its energy absorption response.

The results show that the energy absorption response of straight and tapered
rectangular tubes can be controlled using their various geometry parameters. In
particular, the wall thickness, taper angle and the number of tapered sides can be
effectively used as parameters to control the amount of absorbed energy. Tapered
rectangular tubes display less sensitivity to inertia effects compared with straight

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes ii
rectangular tubes under impact loading. This is beneficial when the higher crush
loads associated with inertia effects need to be reduced. Furthermore, though the
energy absorption capacity of thin-walled rectangular tubes diminishes under oblique
impact loading, the capacity is more maintained for tapered rectangular tubes
compared with non-tapered rectangular tubes.

Overall, the results highlight the advantages of using tapered rectangular tubes for
absorbing impact energy under axial and oblique loading conditions. Understanding
is gained as to how the geometry parameters of such structures can be used to control
the absorbed energy. The thesis uses this knowledge to develop design guidelines for
the use of straight and tapered rectangular tubes in energy absorbing systems such as
for crashworthiness applications. Furthermore, the results highlight the importance of
analysing thin-walled energy absorbers as part of an energy absorbing system, since
the response of the absorbers may be different to when they are treated on their own.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes iii
Table of Contents

Keywords …………………………………………………………….………...….... i

Abstract …...………………………………………………………...…………....... ii

Table of Contents ……………………………………………………….…...……. iv

List of Figures …………………………………………………………..…….......... x

List of Tables ………………………………………………………….………..... xix

Publications …………...………………………………………………….……..... xx

Statement of original authorship ...………...……………………………...….... xxii

Acknowledgements ……………...………………………………………...…… xxiii

1 Introduction ........................................................................................................ 1
1.1 Impact energy absorption of vehicle structures............................................ 1
1.2 Use of thin-walled tubes for energy absorption ........................................... 2
1.2.1 Comparison of straight and tapered rectangular tubes ......................... 3
1.2.2 Oblique loading response of tapered rectangular tubes ....................... 4
1.2.3 Finite element modelling of tapered rectangular tubes ........................ 5
1.3 Thesis aims and methodology ...................................................................... 6
1.3.1 Thesis aims........................................................................................... 6
1.3.2 Thesis methodology ............................................................................. 7
1.4 Contributions of the thesis............................................................................ 8
1.5 Thesis outline ............................................................................................. 10

2 Literature Review............................................................................................. 13
2.1 Impact mechanics and structural crashworthiness ..................................... 13
2.1.1 Impact mechanics ............................................................................... 14

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes iv
2.1.1.1 Inertia effects.................................................................................. 14
2.1.1.2 Material strain rate sensitivity........................................................ 18
2.1.2 Structural crashworthiness ................................................................. 21
2.2 Energy absorbers........................................................................................ 24
2.2.1 Principles of operation of energy absorbers....................................... 24
2.2.1.1 Energy absorbed per unit mass ...................................................... 25
2.2.1.2 Mean crushing stress ...................................................................... 26
2.2.1.3 Energy absorbed per unit crush length........................................... 30
2.2.1.4 Specific crushing stress .................................................................. 30
2.2.1.5 Crush force efficiency.................................................................... 30
2.2.1.6 Stroke efficiency ............................................................................ 31
2.2.1.7 Dynamic amplification factor ........................................................ 31
2.2.1.8 Stroke length per unit mass ............................................................ 32
2.2.2 Energy absorber types and their applications .................................... 32
2.2.2.1 Thin-walled tubes and columns ..................................................... 33
2.2.2.2 Tapered tubes ................................................................................. 45
2.2.2.3 Honeycombs................................................................................... 47
2.2.2.4 Foams ............................................................................................. 49
2.2.2.5 Isolators .......................................................................................... 52
2.3 Analysis of energy absorbers ..................................................................... 53
2.3.1 Analytical modelling.......................................................................... 53
2.3.2 Empirical modelling........................................................................... 55
2.3.3 Factorial studies ................................................................................. 56
2.4 Finite element modelling of energy absorbers ........................................... 57
2.4.1 History................................................................................................ 57
2.4.2 Principles of finite element modelling ............................................... 62
2.4.2.1 Analysis.......................................................................................... 62
2.4.2.1.1 Linearity and nonlinearity ........................................................ 62
2.4.2.1.2 Static and dynamic analysis ..................................................... 63
2.4.2.2 Finite elements ............................................................................... 65
2.4.2.3 Material models.............................................................................. 66
2.5 Summary and need for further research ..................................................... 66

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes v
3 Development and Validation of the Finite Element Model for Axial
Loading...................................................................................................................... 70
3.1 Introduction ................................................................................................ 70
3.2 Development of the finite element model for axial loading....................... 71
3.2.1 Geometry and finite element mesh..................................................... 71
3.2.2 Loading, interaction and boundary conditions ................................... 73
3.2.3 Material models.................................................................................. 76
3.3 Existing models for the axial crushing of straight and tapered tubes......... 78
3.3.1 Theoretical model for straight and double-tapered tubes................... 78
3.3.2 Numerical model for frusta ................................................................ 80
3.4 Validation of the finite element model for axial loading ........................... 80
3.4.1 Validation results for the straight tube ............................................... 82
3.4.2 Validation results for the double-tapered tube ................................... 84
3.4.3 Validation results for the frusta.......................................................... 87
3.4.4 Deformation modes for the straight and tapered tubes ...................... 87
3.5 Summary .................................................................................................... 88

4 Experimental Testing and Finite Element Model Validation for Oblique


Loading...................................................................................................................... 89
4.1 Introduction and background ..................................................................... 89
4.2 Experimental testing of straight and tapered rectangular tubes under
oblique loading....................................................................................................... 91
4.2.1 Specimen details and experimental program ..................................... 91
4.2.2 Observed deformation modes............................................................. 95
4.2.3 Energy absorption response ............................................................... 97
4.3 Development of the finite element model for oblique loading ................ 104
4.3.1 Model geometry and material properties ......................................... 104
4.3.2 Element mesh and initial imperfections ........................................... 105
4.3.3 Interactions, loading and boundary conditions ................................ 112
4.4 Validation of the finite element model..................................................... 116
4.4.1 Load-deflection and mean load-deflection response........................ 117
4.4.2 Deformation modes .......................................................................... 121
4.5 Summary .................................................................................................. 125

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes vi
5 Quasi-Static Response of Straight and Tapered Rectangular Tubes under
Axial Loading ......................................................................................................... 126
5.1 Introduction .............................................................................................. 126
5.2 Response variables and tube dimensions ................................................. 127
5.3 Load-deflection response for straight and tapered rectangular tubes ...... 128
5.4 Parametric study of straight and tapered tubes under quasi-static loading130
5.4.1 Effect of wall thickness.................................................................... 130
5.4.2 Effect of taper angle ......................................................................... 134
5.4.3 Effect of the number of tapers.......................................................... 140
5.4.4 Effect of wall thickness, taper angle and number of tapers on other
design parameters............................................................................................. 143
5.4.4.1 Energy absorbed per unit crush length, Ecl .................................. 144
5.4.4.2 Energy absorbed per unit mass, Em .............................................. 145
5.5 Developing an empirical relation between the geometry and energy
absorption response of tapered tubes ................................................................... 148
5.5.1 Developing the empirical relation.................................................... 149
5.5.2 Testing the empirical relation for various aspect ratios, slenderness
ratios and relative crush distances.................................................................... 152
5.6 Deformation modes of straight and tapered rectangular tubes under quasi-
static axial loading................................................................................................ 154
5.6.1 Effect of the number of tapers on the deformation mode ................ 154
5.6.2 Effect of the taper angle on the deformation mode.......................... 155
5.6.3 Effect of wall thickness on the deformation mode........................... 158
5.7 Conclusions and design guidelines .......................................................... 159

6 Dynamic Response of Straight and Tapered Rectangular Tubes under


Axial Impact Loading ............................................................................................ 162
6.1 Introduction .............................................................................................. 162
6.2 Tube dimensions and loading variables ................................................... 163
6.3 Effect of impact velocity and impact mass .............................................. 165
6.3.1 Effect of impact velocity.................................................................. 165
6.3.2 Effect of impact mass....................................................................... 171
6.3.3 Determining the presence of axial inertia effects............................. 173

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes vii
6.4 Effect of geometry parameters on the dynamic response of straight and
tapered rectangular tubes...................................................................................... 176
6.4.1 Effect of wall thickness .................................................................... 176
6.4.2 Effect of taper angle ......................................................................... 180
6.4.3 Combined effects of wall thickness, taper angle and number of tapers
on the absorbed energy..................................................................................... 181
6.4.4 Effect of wall thickness, taper angle and number of tapers on other
design parameters ............................................................................................. 183
6.4.4.1 Dynamic initial peak load, Fbd ..................................................... 183
6.4.4.2 Ratio of dynamic mean load to initial peak load, Fmd / Fbd .......... 185
6.4.4.3 Energy absorbed per unit crush length, Ecl................................... 185
6.4.4.4 Energy absorbed per unit mass, Em .............................................. 188
6.5 Comparison between static and dynamic response .................................. 190
6.5.1 Effect of impact velocity on the DAF .............................................. 190
6.5.2 Effect of wall thickness on the DAF ................................................ 192
6.5.3 Effect of taper angle on the DAF ..................................................... 194
6.5.4 Effect of the number of tapers on the DAF ...................................... 194
6.6 Use of a factorial study for energy absorber design................................. 197
6.7 Deformation modes of straight and tapered rectangular tubes under axial
dynamic impact loading ....................................................................................... 200
6.7.1 Effect of impact energy on the deformation mode........................... 201
6.7.2 Effect of velocity on the deformation mode- constant impact energy
203
6.8 Impact pulse loading ................................................................................ 207
6.8.1 Introduction ...................................................................................... 207
6.8.2 Effect of impact surface properties .................................................. 208
6.9 Conclusions and design guidelines .......................................................... 211

7 Dynamic Response of Straight and Tapered Rectangular Tubes under


Oblique Impact Loading........................................................................................ 215
7.1 Introduction .............................................................................................. 215
7.2 Loading arrangement and description of the finite element model.......... 216
7.3 Tube geometries ....................................................................................... 218
7.4 Effect of load angle on tube response ...................................................... 219

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes viii
7.5 Effect of wall thickness on the tube response under oblique loading ...... 229
7.6 Effect of taper angle on the tube response under oblique loading ........... 232
7.7 Effect of breadth on the tube response under oblique loading................. 233
7.8 Effect of length on the tube response under oblique loading................... 238
7.9 Effect of impact velocity on the tube response under oblique loading .... 242
7.10 Conclusions and design guidelines .......................................................... 245

8 Application: Dynamic Impact Simulation and Energy Absorption of a


Vehicle Bumper System......................................................................................... 248
8.1 Introduction .............................................................................................. 248
8.2 Description of the energy absorbing system ............................................ 250
8.3 Experimental testing................................................................................. 252
8.3.1 Quasi-static test setup....................................................................... 252
8.3.2 Experimental results......................................................................... 255
8.4 Development and validation of the finite element model ........................ 257
8.4.1 Description of the finite element model........................................... 257
8.4.2 Comparison between the finite element and experimental results... 259
8.5 Parametric study....................................................................................... 260
8.5.1 Symmetric loading ........................................................................... 262
8.5.2 Oblique loading................................................................................ 270
8.6 Conclusions and design guidelines .......................................................... 280

9 Conclusions ..................................................................................................... 284


9.1 Contributions of the thesis ....................................................................... 284
9.2 Recommendations for future work........................................................... 286

References………………………………………………………...……..……….. 288

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes ix
List of Figures

Figure 1.1-1 The economic loss caused by road crashes (Lu & Yu 2003). ................. 1
Figure 1.1-2 Typical scene of a vehicle incident (Royal Automobile Club of
Queensland 2004)................................................................................................. 2
Figure 1.2-1 A vehicle structure with front chassis rails which deform to absorb
impact energy (image from www.alcanautomotive.com [Accessed June 2004]).
.............................................................................................................................. 3
Figure 2.1-1 Permanent deformed profile of a tube subjected to an (a) quasi-static
load and (b) impact load (Jones 1989). .............................................................. 15
Figure 2.1-2 Two classes of structures: (a) Type I and (b) Type II ........................... 16
Figure 2.1-3 Energy absorbing cellular structure from a passenger vehicle: (a) quasi-
static axial testing and (b) final deformed profile. ............................................. 22
Figure 2.1-4 Floor structures in a helicopter which are designed to absorb energy
during a crash landing (Kindervater & Georgi 1993). ....................................... 23
Figure 2.2-1 Load-deflection curve for the axial crushing of a thin-walled rectangular
tube made of mild steel. ..................................................................................... 27
Figure 2.2-2 A thin-walled mild steel tube with a square cross-section before and
after quasi-static axial crushing (Jones 1989). ................................................... 29
Figure 2.2-3 Axial progressive deformation modes of thin-walled square tubes: (a)
symmetric mode, (b) extensional mode and (c) asymmetric mixed B-type mode
(Jones 1989). ...................................................................................................... 35
Figure 2.2-4 Basic collapse elements: (a) Type I- inextensional and (b) Type II-
extensional (Jones 1989). ................................................................................... 36
Figure 2.2-5 (a) Non-compact and (b) compact deformation of a tube subject to an
axial load. ........................................................................................................... 37
Figure 2.2-6 Crush triggers or initiators in the side walls of a thin-walled tube: (a)
undeformed profile and (b) final deformed profile. ........................................... 38
Figure 2.2-7 Load-deflection profile and examples of inverted tubes ....................... 39
Figure 2.2-8 Load-deflection profile and examples of split tubes. The top specimens
have been subsequently curled (Reid 1993)....................................................... 40
Figure 2.2-9 Typical plastic hinge formed during bending of a rectangular tube...... 41

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes x
Figure 2.2-10 Schematic showing the bending collapse of aluminium hat profiles
(Chen 2001)........................................................................................................ 42
Figure 2.2-11 Lateral indentation of tubes (Alghamdi 2001). ................................... 44
Figure 2.2-12 Schematic of straight (non-tapered) and tapered rectangular tubes. ... 45
Figure 2.2-13 (a) Schematic showing the three major axes of a honeycomb structure,
and (b) Generalised stress-strain curve for laterally loaded honeycombs.......... 48
Figure 2.2-14 Hydraulic isolator mounted to the front of a vehicle........................... 53
Figure 2.3-1 Theoretical model for a quadrant of a square tube under axial symmetric
folding: (a) with concentrated hinges and (b) with plastic zones....................... 54
Figure 2.4-1 Deformed profiles of single- and multi-cell columns under a quasi-static
axial load, as predicted using the FE code PAM-CRASH................................. 60
Figure 2.4-2 FE model of the oblique loading collapse of mild steel square columns.
Symbol Φ is the load angle (Han & Park 1999). ............................................... 60
Figure 2.4-3 Experimental and simulated progressive collapse of a square frusta
(Mamalis et al. 2001). ........................................................................................ 61
Figure 2.4-4 Major finite element families available in ABAQUS/Explicit (Hibbit,
Karlsson & Sorensen, Inc. 2002a; b). ................................................................ 65
Figure 3.2-1 Geometries of the straight and tapered rectangular tubes. .................... 71
Figure 3.2-2 Finite element mesh and loading arrangement for straight and tapered
tubes. .................................................................................................................. 72
Figure 3.2-3 Displacement curve for the supporting rigid body................................ 74
Figure 3.2-4 Tensile test apparatus. ........................................................................... 76
Figure 3.2-5 True static stress-strain curve for mild steel.......................................... 77
Figure 3.4-1 Model geometry for the (l) straight and (r) tapered tubes. .................... 81
Figure 3.4-2 Comparison of results for the straight tube under quasi-static axial
loading (here h = 1.5 mm).................................................................................. 82
Figure 3.4-3 Comparison of results for the straight tube under dynamic axial loading
(here h = 1.5 mm, V = 15 m/s and M = 90 kg)................................................... 83
Figure 3.4-4 Comparison of results for the double-tapered tube under quasi-static
axial loading (here h = 1.5 mm and θ = 100)...................................................... 85
Figure 3.4-5 Comparison of results for the double-tapered tube under dynamic axial
loading (here h = 1.5 mm, V = 15 m/s and M = 90 kg)...................................... 86
Figure 4.2-1 Tinius Olsen universal testing machine................................................. 92
Figure 4.2-2 Test rig used for oblique loading........................................................... 93

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes xi
Figure 4.2-3 (a) Straight and tapered tube specimens, (b) and (c) oblique loading test
apparatus............................................................................................................. 94
Figure 4.2-4 Typical final deformed profiles for the (a) straight and (b) tapered tube,
for a 100 load angle............................................................................................. 96
Figure 4.2-5 Typical final deformed profiles for the (a) straight and (b) tapered tube,
for a 300 load angle............................................................................................. 96
Figure 4.2-6 Lobe formed near fixed base of each specimen for a 300 load angle. ... 97
Figure 4.2-7 Load-deflection curve for the (a) straight and (b) tapered tubes, for a 100
load angle (results for all three test specimens are shown). ............................... 98
Figure 4.2-8 Load-deflection curves for the (a) straight and (b) tapered tubes, for a
300 load angle. .................................................................................................... 99
Figure 4.2-9 Mean load-deflection curves for the (a) straight and (b) tapered tubes,
for a 100 load angle........................................................................................... 100
Figure 4.2-10 Mean load-deflection curves for the (a) straight and (b) tapered tubes,
for a 300 load angle........................................................................................... 101
Figure 4.2-11 Initial peak load, Fbs and mean load, Fms for each tube geometry as
functions of load angle. .................................................................................... 102
Figure 4.3-1 Element mesh for the (a) straight tube and (b) tapered tube. .............. 106
Figure 4.3-2 Corner detail in the finite element mesh.............................................. 106
Figure 4.3-3 Fabrication warp used in finite element mesh for the tapered tube..... 107
Figure 4.3-4 Imperfection eigenmodes for 100 loading: (a) first eigenmode for
straight tube and (b) second eigenmode for tapered tube................................. 108
Figure 4.3-5 Imperfection trigger geometry for 300 loading: (a) straight tube, (b)
tapered tube and (c) trigger detail. ................................................................... 109
Figure 4.3-6 Effect of trigger depth on (a) peak load and (b) mean load................. 111
Figure 4.3-7 Load-deflection curves for various trigger depths: (a) straight tube, (b)
tapered tube. ..................................................................................................... 112
Figure 4.3-8 Oblique loading model. ....................................................................... 113
Figure 4.3-9 Effect of step time on the (a) load-deflection response and (b) mean
load-deflection response................................................................................... 114
Figure 4.4-1 Experimental vs. numerical load-deflection response for the straight
tube: (a) 100 load angle and (b) 300 load angle. ............................................... 117
Figure 4.4-2 Experimental vs. numerical load-deflection response for the tapered
tube: (a) 100 load angle and (b) 300 load angle. ............................................... 118

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes xii
Figure 4.4-3 Experimental vs. numerical mean load-deflection response for the
straight tube: (a) 100 load angle and (b) 300 load angle. .................................. 119
Figure 4.4-4 Experimental vs. numerical mean load-deflection response for the
tapered tube: (a) 100 load angle and (b) 300 load angle. .................................. 120
Figure 4.4-5 Comparison of the predicted and experimental initial peak loads. ..... 120
Figure 4.4-6 Experimental and numerical deformed profiles under 100 loading..... 122
Figure 4.4-7 Experimental and numerical deformed profiles under 300 loading..... 123
Figure 4.4-8 Progressive deformation at the loaded end for each tube.................... 124
Figure 4.4-9 Negative curvature of the straight tube under 300 loading: comparison of
experimental and predicted result. ................................................................... 124
Figure 5.2-1 Model geometry for the straight and tapered tubes............................. 128
Figure 5.3-1 Load-deflection response for straight and tapered rectangular tubes.. 129
Figure 5.4-1 Effect of wall thickness on the quasi-static mean load (here θ = 100). 131
Figure 5.4-2 Effect of wall thickness on the absorbed energy (here θ = 100).......... 132
Figure 5.4-3 Effect of wall thickness on the initial peak load (here θ = 100). ......... 133
Figure 5.4-4 Ratio of mean load to initial peak load vs. wall thickness .................. 133
Figure 5.4-5 Effect of taper angle on the quasi-static mean load............................. 134
Figure 5.4-6 Effect of taper angle on the absorbed energy (here h = 1.5 mm). ....... 136
Figure 5.4-7 Effect of taper angle on the initial peak load....................................... 136
Figure 5.4-8 Ratio of mean load to initial peak load vs. taper angle. ...................... 138
Figure 5.4-9 Relative effect of wall thickness and taper angle on the initial peak load.
.......................................................................................................................... 139
Figure 5.4-10 Relative effect of wall thickness and taper angle on the ratio of mean
load to initial peak load. ................................................................................... 140
Figure 5.4-11 Effect of the number of tapered sides on the initial peak load .......... 141
Figure 5.4-12 Effect of the number of tapered sides on the mean load ................... 141
Figure 5.4-13 Effect of the number of tapered sides on the mean load ................... 142
Figure 5.4-14 Effect of the number of tapers on the absorbed energy..................... 143
Figure 5.4-15 Effect of the number of tapers on the absorbed energy..................... 143
Figure 5.4-16 Effect of wall thickness and number of tapers on Ecl (θ = 50)........... 144
Figure 5.4-17 Effect of wall thickness and number of tapers on Em (θ = 50)........... 146
Figure 5.4-18 Effect of wall thickness and taper angle on Em for the double-tapered
tube................................................................................................................... 147
Figure 5.4-19 Relative effect of wall thickness and taper angle on Em.................... 147

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes xiii
Figure 5.5-1 Double-tapered rectangular tube geometry. ........................................ 149
Figure 5.5-2 Variation of crush force efficiency with relative crush distance and
slenderness ratio for the double-tapered tube (here θ = 150 and the aspect ratio,
Wt / Bt = 1). ....................................................................................................... 150
Figure 5.5-3 Prediction of the initial peak load using Equation 5.5-2 and the finite
element model. ................................................................................................. 153
Figure 5.6-1 Collapse sequence of straight and tapered rectangular tubes under a
quasi-static axial load (here h = 1.5 mm and θ = 100)...................................... 155
Figure 5.6-2 Dependence of fold length on taper angle for the double-tapered tube
and frusta (h = 1.5 mm).................................................................................... 156
Figure 5.6-3 Dependence of deformation mode on the taper angle for the triple-
tapered tube (h = 1.5 mm). ............................................................................... 157
Figure 5.6-4 Resultant force created in the triple-tapered tube due to its unsymmetric
cross-section. .................................................................................................... 158
Figure 6.2-1 Model geometry for the (l) straight and (r) tapered tubes. .................. 164
Figure 6.2-2 Arrangement for axial impact loading (straight tube quarter model
shown). ............................................................................................................. 165
Figure 6.3-1 Effect of impact velocity on the load-deflection response .................. 166
Figure 6.3-2 Effect of impact velocity on the dynamic mean load .......................... 168
Figure 6.3-3 Effect of impact velocity on the dynamic absorbed energy ................ 169
Figure 6.3-4 Relative effect of strain rate and inertia effects on the dynamic response
(here h = 1.5 mm, θ = 100 and M = 80 kg). ...................................................... 171
Figure 6.3-5 Effect of impact mass on the dynamic load-deflection response ........ 172
Figure 6.3-6 Load-time curves for the straight tube (here h = 1.5 mm)................... 174
Figure 6.3-7 Load-time curves for the double-tapered tube..................................... 175
Figure 6.4-1 Effect of wall thickness on the dynamic mean load (here θ = 100)..... 177
Figure 6.4-2 Effect of wall thickness on the dynamic absorbed energy .................. 178
Figure 6.4-3 Effect of wall thickness on the maximum crush distance. .................. 179
Figure 6.4-4 Effect of taper angle on the dynamic mean load for the double-tapered
tube (here h = 1.5 mm). .................................................................................... 180
Figure 6.4-5 Dynamic absorbed energy vs. taper angle and number of tapers
(calculated at d = 100 mm)............................................................................... 182
Figure 6.4-6 Effect of each geometry parameter on the dynamic initial peak load
(here V = 20 m/s and M = 80 kg). .................................................................... 184

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes xiv
Figure 6.4-7 Effect of each geometry parameter on the dynamic crush force
efficiency (here V = 20 m/s and M = 80 kg). ................................................... 186
Figure 6.4-8 Energy absorbed per unit crush length vs. taper angle, number of tapers
and wall thickness. ........................................................................................... 187
Figure 6.4-9 Energy absorbed per unit mass vs. wall thickness, number of tapers and
taper angle. ....................................................................................................... 189
Figure 6.5-1 Effect of impact velocity on the DAF (here h = 1.5 mm and θ = 100).191
Figure 6.5-2 Effect of wall thickness on the DAF (here θ = 100 and V = 20 m/s). .. 193
Figure 6.5-3 Effect of taper angle on the DAF (here h = 1.5 mm and V = 20 m/s). 195
Figure 6.5-4 Effect of the number of tapers on the DAF ......................................... 196
Figure 6.5-5 Effect of the number of tapers on the DAF ......................................... 197
Figure 6.6-1 Relative effect of wall thickness, taper angle and impact velocity on the
(a) initial peak load and (b) mean load for each tube....................................... 199
Figure 6.7-1 Deformation profiles for the straight tube, double-tapered tube and
frusta at high and low impact velocities (h = 1.5 mm and θ = 100). ................ 202
Figure 6.7-2 Deformation profiles for the triple-tapered tube at high and low impact
velocities (h = 1.5 mm and θ = 100)................................................................. 203
Figure 6.7-3 Final deformed profiles for the straight tube, double-tapered tube and
frusta for a constant impact energy (h = 1.5 mm and θ = 100)......................... 204
Figure 6.7-4 Final deformed profiles for the triple-tapered tube for a constant impact
energy (h = 1.5 mm and θ = 100). .................................................................... 205
Figure 6.7-5 Influence of impact velocity on the energy absorbed up to a given
deflection (here h = 1.5 mm, θ = 100 and impact energy = 9 kJ)..................... 206
Figure 6.8-1 Input acceleration-time curves for pulse durations of 20, 30 .............. 209
Figure 6.8-2 Energy absorbed under pulse loading(here h = 1.5 mm and θ = 100). 210
Figure 6.8-3 Effect of surface properties (shown as reduction of absorbed energy
when the pulse duration increases from 20 to 40 ms)...................................... 211
Figure 7.2-1 Model arrangement for oblique loading.............................................. 216
Figure 7.3-1 Tube geometries compared for oblique loading.................................. 218
Figure 7.4-1 Effect of load angle on the load-deflection response. ......................... 220
Figure 7.4-2 Typical deformation modes for straight and tapered rectangular tubes
under oblique loading (deformed profiles shown at d = 200 mm). Straight tube
(top), single-tapered tube (middle), and double-tapered tube (bottom). .......... 222
Figure 7.4-3 Effect of load angle on the mean load-deflection response................. 223

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes xv
Figure 7.4-4 Effect of load angle on the energy absorption-deflection response. ... 224
Figure 7.4-5 Effect of load angle on the mean load under dynamic impact loading
(calculated up to d = 200 mm). ........................................................................ 225
Figure 7.4-6 Reduction of absorbed energy with increasing load angle. ................. 227
Figure 7.4-7 Deformation modes of the plastic hinge formed during global collapse.
.......................................................................................................................... 228
Figure 7.5-1 Effect of wall thickness on the mean load with increasing load angle.
.......................................................................................................................... 229
Figure 7.5-2 Energy absorbed per unit mass vs. load angle and wall thickness. ..... 231
Figure 7.5-3 Energy absorbed per unit mass vs. load angle and number of tapers
(here h = 2 mm)................................................................................................ 232
Figure 7.6-1 Effect of taper angle and load angle on the percentage of energy
absorbed under axial loading............................................................................ 233
Figure 7.7-1 Effect of breadth on the mean load with increasing load angle. ......... 234
Figure 7.7-2 Reduction of absorbed energy with increasing load angle. ................. 236
Figure 7.7-3 Energy absorbed per unit mass vs. load angle and breadth................. 237
Figure 7.8-1 Effect of length on the mean load with increasing load angle. ........... 239
Figure 7.8-2 Reduction of absorbed energy with increasing load angle. ................. 240
Figure 7.8-3 Energy absorbed per unit mass vs. load angle and length. .................. 241
Figure 7.9-1 Effect of load angle on the DAF.......................................................... 243
Figure 7.9-2 Effect of load angle on the quasi-static mean load.............................. 244
Figure 7.9-3 Reduction of absorbed energy with increasing load angle under quasi-
static loading. ................................................................................................... 245
Figure 8.2-1 Energy absorbing system consisting of tapered rectangular tubes (green
component labelled “energy absorber”). .......................................................... 250
Figure 8.2-2 Trigger/ energy absorber deformation mode under an axial load. ...... 251
Figure 8.2-3 Comparison of absorber response when incorporating the trigger...... 251
Figure 8.3-1 Test specimen of the bumper system welded to its supporting ........... 252
Figure 8.3-2 Strain gauge location. .......................................................................... 253
Figure 8.3-3 LVDT (circled) to measure deflection of the energy absorbers. ......... 254
Figure 8.3-4 Quasi-static test setup for the bumper system. .................................... 254
Figure 8.3-5 Load-deflection and mean load-deflection response of the bumper
system............................................................................................................... 255
Figure 8.3-6 Deformation mode of the energy absorber. ......................................... 256

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes xvi
Figure 8.3-7 Principal strains of the right-hand-side components. .......................... 256
Figure 8.4-1 Finite element model of the bumper system. ...................................... 257
Figure 8.4-2 Displacement curve for the loading rigid body. .................................. 258
Figure 8.4-3 Experimental and predicted deformed profiles of the bumper system.
.......................................................................................................................... 259
Figure 8.4-4 Comparison of experimental and predicted quasi-static response for the
bumper system. ................................................................................................ 260
Figure 8.5-1 Dynamic loading conditions for the bumper system (top view). ........ 262
Figure 8.5-2 Velocity-deflection response of the bumper system with a straight tube
as the energy absorber...................................................................................... 263
Figure 8.5-3 Change in velocity of the bumper system up to a deflection .............. 264
Figure 8.5-4 Load-deflection response of the bumper system with a straight tube as
the energy absorber. ......................................................................................... 265
Figure 8.5-5 Load-deflection response of the bumper system with a tapered tube as
the energy absorber. ......................................................................................... 266
Figure 8.5-6 DAF at various deflections with either a straight or tapered tube as the
energy absorber. ............................................................................................... 268
Figure 8.5-7 Crush sequence of the bumper system under an axial impact load..... 269
Figure 8.5-8 Effect of load angle on the mean load-deflection response of the bumper
system............................................................................................................... 271
Figure 8.5-9 Effect of load angle on the energy absorption-deflection response of the
bumper system. ................................................................................................ 272
Figure 8.5-10 Effect of load angle on the mean load of the bumper system under
dynamic impact loading (calculated up to d = 120 mm).................................. 273
Figure 8.5-11 Reduction of absorbed energy with increasing load angle................ 274
Figure 8.5-12 Effect of channel wall thickness on the response of the bumper system
under oblique loading....................................................................................... 275
Figure 8.5-13 Deformed profile of the bumper system with a straight tube energy
absorber (top view shown). .............................................................................. 276
Figure 8.5-14 Deformed profile of the bumper system with a tapered tube energy
absorber (top view shown). .............................................................................. 277
Figure 8.5-15 oblique loading response of the bumper system when the channel’s
wall thickness is reduced to 3 mm (with a straight tube as the energy absorber).
.......................................................................................................................... 278

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes xvii
Figure 8.5-16 oblique loading response of the bumper system when the channel’s
wall thickness is reduced to 3 mm (with a tapered tube as the energy absorber).
.......................................................................................................................... 278
Figure 8.5-17 Dependence of the bumper system’s deformation mode on the
channel’s wall thickness under oblique loading (here α = 300). ...................... 279

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes xviii
List of Tables

Table 2.1-1 Values of D and q for various metals (Jones 1989)................................ 20


Table 2.2-1 Energy absorbed per unit mass and tube length needed to absorb 31.7 kJ,
for various materials (Lin & Mase 1990)........................................................... 26
Table 3.2-1 True stress-plastic strain data points used for steel in the FE model...... 77
Table 3.4-1 Tube dimensions used for validating the finite element model.............. 81
Table 3.4-2 Comparison of results for the frusta finite element model with those of
Mamalis et al. (2001). ........................................................................................ 87
Table 4.2-1 Average section dimensions of all specimens used in the oblique loading
experiments. ....................................................................................................... 92
Table 4.2-2 Summary of experimental results. ........................................................ 103
Table 4.3-1 Dimensions used to model the straight and tapered tubes under oblique
loading.............................................................................................................. 104
Table 4.3-2 Wall thickness used in the finite element model for oblique loading. . 105
Table 4.3-3 Trigger dimensions. .............................................................................. 110
Table 5.2-1 Model dimensions used in the quasi-static parametric analysis. .......... 128
Table 5.4-1 Levels used in the factorial study. ........................................................ 139
Table 5.4-2 Increase in tube mass with wall thickness and taper angle................... 145
Table 6.2-1 Model dimensions used in the parametric analysis. ............................. 164
Table 6.6-1 Levels used in the factorial study. ........................................................ 198
Table 7.3-1 Range of each geometry parameter used in the oblique loading
parametric analysis........................................................................................... 219
Table 8.5-1 Absorber dimensions used in the parametric study. ............................. 261

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes xix
Publications

Refereed International Journal Papers

Nagel, G. M. and Thambiratnam, D. P. 2004. A numerical study on the impact


response and energy absorption of tapered thin-walled tubes. International Journal of
Mechanical Sciences. 46 (2): 201-216.

Nagel, G. M. and Thambiratnam, D. P. 2004. Dynamic simulation and energy


absorption of tapered tubes under impact loading. International Journal of
Crashworthiness. 9 (4): 389-399.

Nagel, G. M. and Thambiratnam, D. P. 2005. Dynamic simulation and energy


absorption of tapered thin-walled tubes under oblique impact loading. International
Journal of Impact Engineering. (Accepted and in press).

Nagel, G. M. and Thambiratnam, D. P. 2005. Computer simulation and energy


absorption of tapered thin-walled rectangular tubes. Thin-Walled Structures.
(Submitted).

International Conference Papers

Nagel, G. M. and Thambiratnam, D. P. 2002. Energy absorption and performance of


a Vehicle Impact Protection system. In Jones, N., Brebbia, C. and Rajendran, A.
(eds), Proc. Seventh International Conference on Structures Under Shock and
Impact, Montreal, Canada, 27-29 May 2002. Southampton: WIT Press.

Nagel, G. M. and Thambiratnam, D. P. 2002. Development of energy absorbing


mechanisms for a Vehicle Impact Protection system. In Loo, Y., Chowdhury, S. and
Fragomeni, S. (eds), Advances in Mechanics of Structures and Materials- Proc. 17th
Australasian Conference on the Mechanics of Structures and Materials, Gold Coast,
Australia, 12-14 June 2002. Tokyo: A. A. Balkema Publishers.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes xx
Nagel, G. M. and Thambiratnam, D. P. 2003. Use of thin-walled frusta energy
absorbers in protection of structures under impact loading. In Ohno, T.,
Krauthammer, T. and Pan, T.-C. (eds), Proc. DAPSIL 2003- The First International
Conference on Design and Analysis of Protective Structures against
Impact/Impulse/Shock Loads, Tokyo, Japan, 16-18 December 2003. Tokyo: Daioh
Co. Ltd.

Nagel, G. M. and Thambiratnam, D. P. 2004. Performance of thin-walled frusta


energy absorbers in structures under impact loads. In Zingoni, A. (ed), Proc. of the
2nd International Conference on Structural Engineering, Mechanics and
Computation, Cape Town, South Africa, 5-7 July 2004. Singapore: A. A. Balkema
Publishers.

Local Conference Papers

Nagel, G. M. and Thambiratnam, D. P. 2001. Development of a Vehicle Impact


Protection System for passenger vehicles. In Clark, B. (ed), Proc. 8th PIC
Postgraduate Conference- Innovative Research for the 21st Century, QUT, 3-4
December 2001. Brisbane, Australia: Physical Infrastructure Centre, School of Civil
Engineering, Queensland University of Technology.

Nagel, G. M. and Thambiratnam, D. P. 2002. Energy absorption performance of thin-


walled frusta under dynamic impact loading. In Wake-Dyster, K. (ed), Proc. 10th PIC
Postgraduate Conference- Research for Changing Environments- A Civil
Engineering Perspective, QUT, 12-13 December 2002. Brisbane, Australia: Physical
Infrastructure Centre, School of Civil Engineering, Queensland University of
Technology.

Non-refereed Research Publications

Nagel, G. and Thambiratnam, D. 2002. Development of a Vehicle Impact Protection


System for passenger vehicles. In Clark, B. (ed), Digest: Innovative Research in
Civil Engineering from the Physical Infrastructure Centre. 12 (1): January 2002.
Brisbane, Australia: Physical Infrastructure Centre, School of Civil Engineering,
Queensland University of Technology.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes xxi
Statement of original authorship

The work contained in this thesis has not been previously submitted for a degree or
diploma at any other higher education institution. To the best of my knowledge and
belief, the thesis contains no material previously published or written by another
person except where due reference is made.

Gregory Nagel

14 February 2005

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes xxii
Acknowledgements

I would like, indeed am indebted, to thank the following people, without whose
support, the undertaking and completion of this thesis would not have been possible.

Firstly, I would like to thank my Principal Supervisor, Professor David


Thambiratnam, and my Industry Supervisor, Mr Simon Orton. Their practical
guidance, patience and wisdom were invaluable for helping me through the many
stages and milestones of the thesis.

I would also like to give a big thank-you to my industry partner, TJM Products Pty
Ltd (www.tjmproducts.com.au), who financed the work related to this thesis, and
provided the vision and incentive for the thesis. Included in this big thank-you are the
many staff who supported me throughout the project, notably those in R & D
(Research and Development department), who manufactured the specimens for the
experiments. In particular I owe a big thanks to Mr Arthur Vlahogenis, General
Manager of TJM Products Pty Ltd, for his invaluable support and understanding. I
would also like to thank my original Industry Supervisor, Mr Mark McGuiness.

Another big thank-you goes to the staff at the School of Civil Engineering, QUT,
who have supported me in many and varied ways during the course of my PhD. In
particular I would like to thank the members of the review panel for my thesis;
Associate Professor Andy Tan, Professor Rod Troutbeck and Associate Professor
Doug Hargreaves.

I am also most thankful for the company and support of my friends, fellow
colleagues and post-graduate students.

Lastly, but definitely not least, I would like to thank my family, particularly my
mother, for their vital support, patience and encouragement during the course of my
PhD.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes xxiii
1 Introduction

1.1 Impact energy absorption of vehicle structures

Modern society is placing increased reliance on transportation systems. This is


perhaps most clearly reflected by the continuously increasing number of vehicles
over the last century. Furthermore, the range of transportation means is wide, while
new and novel means of transport are appearing more frequently on the market.

With the increasing number of vehicles have come higher speeds and larger vehicles,
such as large trucks and aircraft. Consequently, these vehicles themselves are costly
and the damage they can potentially cause to people and the environment during an
accident will be more serious. Motor vehicle related accidents are a major worldwide
health problem. For example, Americans between the ages of 1 and 34 are killed by
vehicle crashes more than by any other source of injury or type of disease (Lu & Yu
2003). Road crashes also constitute a great economic loss to society. As an
illustration, Figure 1.1-1 shows the annual costs of road crashes in Australia by crash
type. The annual cost of road crashes in Australia for 1996 was about AUD $15
billion (Lu & Yu 2003). Thus, the detrimental health and economic impact on
modern society caused by vehicle accidents is a significant problem.

Annual cost of road crashes in Australia in 1996


(in AUD $ billion)

This figure is not available online.


Please consult the hardcopy thesis
available from the QUT Library

Figure 1.1-1 The economic loss caused by road crashes (Lu & Yu 2003).

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 1
In the light of these detrimental effects, a greater demand has been placed on
personal and public protection associated with the use of transportation systems. To
this end, over the past several decades increasing focus has been paid to the use of
impact energy absorbers, devices designed to dissipate energy during an impact
event and hence protect the structure and occupants under consideration. For
instance, energy absorbers are widely used in motor vehicle body structures to absorb
energy during a collision (Figure 1.1-2).

This figure is not available online.


Please consult the hardcopy thesis
available from the QUT Library

Figure 1.1-2 Typical scene of a vehicle incident (Royal Automobile Club of


Queensland 2004).

1.2 Use of thin-walled tubes for energy absorption

Thin-walled tubes have been extensively used as energy absorbers, and most
commonly exist as either square or circular in cross-section. For instance, thin-walled
circular tubes are desirable as energy absorbers since they are inexpensive, efficient
and versatile (Jones 1989). Such desirable energy absorption characteristics have led
to them being used as energy absorbers in a diverse range of applications, notably
behind car bumpers and train buffers, and at the base of lift shafts.

A widespread application of thin-walled tubes for absorbing impact energy is as the


front chassis rails of a vehicle (Figure 1.2-1). Such tubes permanently (or plastically)
deform to mitigate the impact energy and loads transmitted to the vehicle, thus

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 2
reducing the decelerations experienced by the occupants. Furthermore, these tubes
act “sacrificially” by deforming in preference to other components of the vehicle
structure, thus reducing repair costs and maintaining the integrity of the passenger
compartment.

This figure is not available online.


Please consult the hardcopy thesis
available from the QUT Library

Figure 1.2-1 A vehicle structure with front chassis rails which deform to absorb
impact energy (image from www.alcanautomotive.com [Accessed June 2004]).

Despite the amount of published research available on the energy absorption of thin-
walled tubes, there are various areas requiring further investigation, several of which
are outlined below.

1.2.1 Comparison of straight and tapered rectangular tubes

Much of the research on thin-walled tube energy absorbers has focused on those of
circular or square cross-section, particularly those in which all the side-walls are
parallel to the longitudinal axis of the column (herein termed “straight” tubes).
Recently, increased focus has been given to tapered tubes in which one or more sides
of the tube are oblique to the longitudinal axis. Such structures have been considered
preferable to straight tubes since they are more likely to provide a desirable constant
mean load-deflection response under dynamic impact loading, and are capable of
withstanding oblique impact loads as effectively as axial loads (Reid & Reddy 1986).
Circular tapered tubes (termed frusta or frustra), have been found to possess a load-
deflection response which is in general more stable than that of straight circular

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 3
tubes (Mamalis & Johnson 1983; Mamalis et al. 1984; 86a; b). In the context of
vehicle crashworthiness the energy absorber is commonly subjected to both axial and
oblique loads, thus tapered tubes may prove advantageous in such applications.
However, compared with straight tubes, relatively few studies have been reported on
the energy absorption performance of tapered tubes, particularly those of rectangular
cross-section. Further investigation is required on the comparison of straight and
tapered rectangular tubes, the effects of parameters such as the number of tapered
sides, the taper angle (i.e. the inclination of the tapered side), wall thickness, and
whether such parameters can be used to effectively control the absorbed energy.
Furthermore, while the inertia effects on the impact energy absorption response of
straight tubes has been quantified (see for example Karagiozova & Jones 2004; 2001;
Karagiozova et al. 2000; Langseth & Hopperstad 1996; Langseth et al. 1999), such
effects on the response of tapered rectangular tubes remain to be assessed.

1.2.2 Oblique loading response of tapered rectangular tubes

In the context of vehicle crashworthiness the energy absorber is commonly subjected


to both axial and oblique (off-axis) loads. Compared with axial loading conditions,
relatively few studies have been conducted on the energy absorption response of
thin-walled tubes under oblique loads. Nevertheless, several recent studies have been
conducted on aluminium square and circular tubes under combined loading (Borvik
et al. 2003; Reyes et al. 2004; 2002). To the author’s knowledge, the only study on
the oblique loading response of tapered rectangular tubes is that carried out by Reid
and Reddy (1986). In the study, tubes with one and two tapered sides were
experimentally impacted at a load angle of 100 to the tube’s longitudinal axis. The
tapered tubes were found to be almost as effective under oblique impact as they were
under axial impact. However, further investigation is required on the effect of load
angle and number of tapers on the energy absorption response of the tapered tubes. In
lieu of this, an associated question would be: how does the energy absorption
response of tapered rectangular tubes compare with that of straight tubes as the load
angle increases? Furthermore, what is the relative influence of the various geometry
and loading parameters (such as load angle and impact velocity), on the response of
straight and tapered rectangular tubes under oblique loading? These are important

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 4
questions when the relative merits of the two structures for use as impact energy
absorbers need to be compared.

1.2.3 Finite element modelling of tapered rectangular tubes

Thin-walled tubes deform via complex modes under either axial, transverse or
combined loading. These modes and the ensuing energy absorption response have
been extensively studied using experimental, theoretical and empirical techniques.
Recently, the use of detailed finite element (FE) analysis has become common for
investigating the energy absorption response of thin-walled tubes. This has primarily
been the result of increased accuracy of FE codes, significantly increased
computation power and reduced computer hardware cost. High-end FE codes can
account for factors such as inertia and material strain rate effects under impact
loading, and the complex contact conditions which occur within the tube as it
deforms. Compared with experimental, theoretical and empirical techniques, there
are still relatively few studies in the published literature which use FE analysis to
study the energy absorption response of thin-walled tubes. However, a number of
studies which use FE analysis have been conducted on the axial impact loading of
square tubes (Karagiozova & Jones 2004; Langseth & Hopperstad 1996; Langseth et
al. 1999), and circular tubes (Karagiozova & Jones 2001; Karagiozova et al. 2000),
and on the oblique loading of square tubes (Han & Park 1999; Reyes et al. 2004;
2002) and circular tubes (Borvik et al. 2003). FE analysis has also been used to
examine the axial energy absorption of circular tapered tubes (Alghamdi et al. 2002),
and square tapered tubes (Mamalis et al. 2001). However, overall there are not many
studies which use FE analysis to study the axial loading of tapered tubes. To the
author’s knowledge, there are no studies in the open literature which use detailed FE
analysis to assess the impact response of tapered rectangular tubes under oblique
loading conditions. The development and use of a FE model to analyse the energy
absorption response of tapered tubes under both axial and oblique loading would be a
powerful tool. Such a model would provide a cost-effective means of assessing the
effect of various geometry and loading parameters on tube response, while including
inertia and material strain rate effects under impact loading. This knowledge could be

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 5
used as a basis of design recommendations for using tapered rectangular tubes as
energy absorbers such as in crashworthiness applications.

1.3 Thesis aims and methodology

1.3.1 Thesis aims

Overall, there are several areas which require further study pertaining to the use of
tapered thin-walled rectangular tubes for impact energy absorption such as in
crashworthiness applications. To address this need, the primary aim of this thesis was
as follows:

To generate research information on the impact and energy absorption of


tapered thin-walled rectangular tubes, to facilitate their application in energy
absorbing systems.

The subsequent aims of this thesis were as follows:

(1) Develop FE models of tapered rectangular tubes and calibrate these models
with experiments and existing theoretical and numerical models.
(2) Develop relationships to quantify the energy absorption of these tubes in
terms of their geometry and loading parameters. The geometry parameters
include the tube wall thickness, taper angle, and number of tapers, while the
loading parameters include the impact mass, velocity and duration, and the
angle of applied load.
(3) Compare the energy absorption response of tapered rectangular tubes with
straight rectangular tubes, to determine their relative performance for use as
impact energy absorbers.

The primary outcome of the thesis was design information for using tapered thin-
walled rectangular tubes as energy absorbers in impact applications.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 6
1.3.2 Thesis methodology

To achieve these aims, this thesis adopted extensive finite element modelling,
coupled with theoretical, empirical and experimental techniques. Three separate FE
models were developed. The first model simulated the deformation and energy
absorption of straight and tapered rectangular tubes under axial loading, for both
quasi-static (slowly applied) and dynamic impact loading regimes. The second model
provided the same assessment under oblique loading conditions. The third FE model
involved the application of the tapered tubes in an energy absorbing system.

The FE models were validated to ensure they simulated the response of thin-walled
tubes with sufficient accuracy. The FE model of the quasi-static and impact response
of straight and tapered tubes under axial loading was validated using existing
theoretical and numerical models. Once validated, the FE model could be used to
predict the deformation and energy absorption response of the straight and tapered
rectangular tubes, for variations in their geometry parameters such as wall thickness,
taper angle, cross-section dimensions, length and number of tapered sides. For axial
impact loading, the impact mass and velocity could be varied to assess their
influence on tube response. Additionally, controlling the impact duration allowed the
effect of the properties of the impact surface to be studied. Parametric studies
conducted on the quasi-static energy absorption response of the straight and tapered
rectangular tubes allowed empirical relations to be developed. Using these relations,
the quasi-static energy absorption response of tapered rectangular tubes can be easily
calculated based on the geometry, dimensions and material properties of the tubes.

Common to all the FE models was the definition of a material model. The material
used for all simulations was hot rolled mild steel. This is commonly used in thin-
walled tubes for energy absorption due to its relative ductility. A proper description
of the strain hardening properties of the steel is necessary for good validation of the
FE models. To this end, standard tensile tests were conducted to determine the post-
yield behaviour of the material, for inclusion in the FE models. Hot rolled mild steel
is a strain rate sensitive material, and it is recognised that strain rate effects can
significantly affect the dynamic response of thin-walled tubes. To account for such

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 7
effects, the FE models for the dynamic impact loading of straight and tapered
rectangular tubes incorporated strain rate dependency in the material definition using
the Cowper-Symonds constitutive equation. The values used for the material
parameters in the equation were the same as those used in previous studies of the
dynamic axial crushing of mild steel tubes (Abramowicz & Jones 1984; 86; Reid &
Reddy 1986; Reid et al. 1986).

The FE model used to simulate the oblique loading response of straight and tapered
rectangular tubes was validated using quasi-static experimental tests. Owing to the
difficulty of performing precision oblique impact experiments on thin-walled tubes,
the FE model was modified to simulate oblique impact loading by including strain
rate effects in the material definition via the Cowper-Symonds equation. The same
material parameter values were used as in the FE model for axial impact loading, and
inertia effects were accounted for in the oblique impact model.

Finally, the FE model of the energy absorbing system was validated using axial
quasi-static tests. The model was adjusted to simulate impact and oblique loading
conditions based on the validated FE models of the individual tubes.

All modelling in this thesis was conducted using the high-end FE codes
ABAQUS/Standard and ABAQUS/Explicit version 6.3 (Hibbitt, Karlsson &
Sorensen, Inc. 2002a; b). The extensive simulations conducted in this thesis
demonstrated the ability of advanced FE codes to predict the non-linear deformation
of the thin-walled tubes under quasi-static and dynamic impact loading. Finally, all
experimental testing conducted as part of this thesis was performed using a screw-
type Tinius Olsen Universal Testing machine.

1.4 Contributions of the thesis

The thesis makes various contributions towards further understanding of the response
of tapered rectangular tubes under quasi-static and dynamic impact loading. These
contributions are briefly summarised below, while the specific results are detailed at
the end of each chapter, and their practical implications are expressed in Chapter 9.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 8
(1) Validated finite element models have been developed which can be used to
predict the quasi-static and dynamic impact response of straight and tapered
rectangular tubes under axial and oblique loading conditions. The models
include a variety of tube geometry and loading parameters, thus the models
are a cost-effective means for performing parametric studies on the tubes as
part of a “virtual” design cycle, for example.

(2) Using the finite element models, the energy absorption response of tapered
rectangular tubes is quantified under both axial and oblique loading. This
leads to a better understanding of such structures for energy absorption, in the
following specific ways:

a. The response is compared with that of straight rectangular tubes, thus


highlighting the advantages of tapering the sides of a rectangular tube
to achieve a desired energy absorption response. Furthermore,
understanding is gained as to how the geometry parameters of such
structures can be used to control the absorbed energy. Such
knowledge is useful for design purposes.

b. It is well understood that the response of straight (non-tapered) thin-


walled tubes can be significantly influenced by inertia effects under
dynamic axial impact loading. This thesis contributes to such
understanding by analysing the effect which tapering the sides of a
thin-walled tube has on its dynamic response.

c. The thesis sheds new light on the response of tapered rectangular


tubes under oblique loading. The knowledge generated shows how the
energy absorption of such structures is influenced by the various
loading and geometry parameters under oblique loading. This
knowledge can then be used to control the energy absorption of
straight and tapered rectangular tubes under oblique loading
conditions.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 9
(3) Based on the parametric studies, design guidelines have been developed
which can be used for developing straight and tapered rectangular tubes for
use in energy absorbing systems. Furthermore, as part of these guidelines an
empirical relation has been developed which allows prediction of the quasi-
static energy absorption response of tapered rectangular tubes under axial
loading, based on their geometry parameters. This relation can be used as a
simple tool to compare the response of different tube geometries for a
particular application.

(4) Energy absorbers are most commonly used as components in energy


absorbing systems. This thesis uses advanced finite element modelling to
analyse the energy absorption response of such a system which incorporates
either straight or tapered rectangular tubes as the energy absorbing
components. Important findings arise from this analysis, and these findings
are used as the basis of design guidelines for the use of straight and tapered
thin-walled tubes in systems designed to absorb impact energy.

1.5 Thesis outline

Chapter 2 provides a review of the literature related to the aims and scope of this
thesis. Areas where further research is required are identified, thereby positioning the
aims of this thesis. Topics reviewed include impact mechanics and crashworthiness,
energy absorbers (with particular focus on thin-walled tubes), analysis and testing of
energy absorbers, and finally finite element modelling of thin-walled tubes.

Chapter 3 describes the development and validation of the finite element model used
for simulation of the straight and tapered rectangular tubes under axial loading. The
steps used to develop the model are covered, as well as the techniques used to
simulate quasi-static and dynamic impact loading conditions. Development of an
appropriate material model for the tubes is also covered in this chapter. The existing
theoretical and numerical models used to validate the axial loading finite element
model are explained. Finally, the validation of this model is carried out by comparing
its predictions of the energy absorption response of the straight and tapered tubes

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 10
with the predictions of the existing models. The validated FE model developed in
this chapter serves as a basis for parametric studies conducted in subsequent chapters
on the axial loading response of the straight and tapered rectangular tubes.

Chapter 4 describes the development and validation of the finite element model used
for simulation of the straight and tapered rectangular tubes under oblique loading.
Detailed description is provided of the quasi-static oblique loading tests used to
validate the FE model. The development and subsequent validation of the FE model
using the experimental results is subsequently covered.

Chapter 5 performs a parametric study of the energy absorption response of straight


and tapered tubes under quasi-static axial loading. Using the validated FE model for
quasi-static loading detailed in Chapter 3, the parametric study examines the relative
effect of various geometry parameters on the quantified energy absorption and
deformation response of the tubes. This provides practical guidance as to how the
various geometry parameters can be used to control the tube response.

Chapter 6 follows the same procedure as in Chapter 5, though applied to dynamic


impact axial loading. The validated FE model for dynamic impact loading detailed in
Chapter 3 is used to study the effect of geometry and impact velocity, mass and
duration on the tube response. Design guidelines are developed for the use of straight
and tapered rectangular tubes as energy absorbers under axial impact loading.

Chapter 7 extends the analysis of straight and tapered tubes to take account of
oblique loading effects. Using the validated FE model developed in Chapter 4, the
influence of load angle on the response of the tubes is examined under dynamic
impact loading conditions. The findings are used to develop design guidelines for
maximising the energy absorption of straight and tapered rectangular tubes under
oblique impact loading.

Chapter 8 applies the knowledge generated in previous chapters to the assessment of


an energy absorbing system which comprises either straight or tapered rectangular
tubes. Development and validation of the FE model of the system using quasi-static
tests are described. A parametric study is then conducted to quantify the energy

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 11
absorption response of the system for different absorber geometries and loading
conditions. Based on the findings of this chapter, design recommendations are drawn
as to the usage of tapered rectangular tubes in energy absorbing systems such as for
crashworthiness applications.

Finally, Chapter 9 summarises the main conclusions of the thesis and their practical
implications.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 12
2 Literature Review

This chapter reviews the background literature which is pertinent to the research
conducted in this thesis. Areas where relevant research is lacking are identified,
thereby establishing the need for this thesis.

The topics addressed in the literature review are broadly contained as follows:

(1) Impact mechanics and structural crashworthiness,


(2) Energy absorbers,
(3) Analysis of energy absorbers, and
(4) Finite element modelling of energy absorbers

2.1 Impact mechanics and structural crashworthiness

Fundamental to the development and analysis of energy absorbing structures is an


understanding of the topic of impact mechanics. Once in its working environment, an
energy absorber would be subject to impact loading. The frontal collision of a
vehicle is a typical and common example. Analysis (such as numerical modelling) of
impact energy absorbers must therefore take into account the nature of such loading
and the ensuing response of the absorber. This section provides an overview of
impact mechanics, and also focuses on a relatively new field of science known as
structural crashworthiness. This is concerned with how a structure withstands an
impact, and as such is the more direct application of impact mechanics to the study
and development of energy absorbing systems.

In terms of availability of information on the field of impact mechanics, little


theoretical background was found in relation to energy absorbers in general, however
an invaluable source was Jones (1989), from which much of the information in this
chapter was sourced. The reader is also referred to the books by Lu and Yu (2003),
Kinslow (1970), Johnson (1972), Blazynski (1987), Gupta (1993) and Stronge (2000)

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 13
for information on impact mechanics, and to Jones and Wierzbicki (1993) for
information on structural crashworthiness. In journal articles there exists an
unfathomable wealth of experimental, theoretical and, more recently, computer finite
element work on the impact loading of specific energy absorbers. However, such
analytical approaches vary from article to article, even when the same mechanism is
reviewed under similar scenarios. The in-depth review of such particular cases is not
warranted unless it is intended to research a certain energy absorber. As such, for the
scope of this thesis, it is only worthwhile presenting a concise background to the
theory of impact mechanics. The next section of this chapter provides a
comprehensive coverage of the types of energy absorbers encountered in the
literature.

2.1.1 Impact mechanics

The collapse response of a structure subject to a dynamic impact load differs from
the response under a quasi-static (slowly applied) load due to two physical
phenomena known as inertia effects and strain rate effects. The energy absorption
response of thin-walled energy absorbers is particularly influenced by these
phenomena, and they will be addressed in turn.

2.1.1.1 Inertia effects

When a structure is loaded quasi-statically, the load is applied sufficiently slowly


such that inertia effects have no influence on the structure’s response. In the case of a
thin-walled tube subjected to axial quasi-static loads, the loads are applied
sufficiently slowly such that neither the axial nor the lateral inertia effects of the tube
wall play a significant role during the response (Jones 1989). However, if the same
tube is subjected to a sufficiently severe dynamic axial load, then structural inertia
effects will influence the tube’s response. The final deformed shape of the loaded
structure differs depending on whether it is subject to a quasi-static or a dynamic
load. This can be seen in Figure 2.1-1 which shows the permanently deformed profile
of a thin-walled tube subjected to both a quasi-static load and an impact load, applied
along the axis of the tube.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 14
This figure is not available online.
Please consult the hardcopy thesis
available from the QUT Library

Figure 2.1-1 Permanent deformed profile of a tube subjected to an (a) quasi-


static load and (b) impact load (Jones 1989).

In the case of the specimen under a quasi-static load, the deformation is localised at
one end in the form of progressive folds. This is known as progressive buckling. On
the other hand, the dynamically loaded specimen is wrinkled over its entire length.
Generally speaking, it has been observed that the influence of inertia forces on rods,
rings, plates and shells loaded dynamically results in the development of lateral
displacement fields with high mode numbers.

Inertia can influence impact response in different ways, depending on the nature of
the structure. There are two generic types of structures which plastically deform to
absorb energy, as identified by Calladine and English (1984). Termed Type I and
Type II structures, they are distinguished by the shape of their overall static load-
deflection curves. The load-deflection response of these structures is shown in Figure
2.1-2, along with particular specimens tested by Calladine and English (1984). Type
I has a relatively “flat-topped” curve, while Type II has an initial peak load followed
by a “steeply falling” curve. The Type I structure is modelled as a laterally
compressed circular tube, while the Type II structure is modelled as a column with
two separate bars fixed together at clamped supports, in which deformation is
assumed to take place at plastic hinges. Experiments conducted on these structures
showed that the deformation of Type II structures is much more sensitive to impact
velocity than that of Type I. Furthermore, keeping the kinetic energy of the striking
mass constant, higher impact velocities cause smaller final deflections, and this
phenomenon is much more significant for Type II than for Type I structures.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 15
F F

d d

(a) (b)

Figure 2.1-2 Two classes of structures: (a) Type I and (b) Type II
(from Lu & Yu 2003).

Further studies showed that the deformation of the Type II structure specimen
examined by Calladine and English has an initial phase of collapse which is initiated
by the effect of lateral inertia forces on the plates (Tam 1990). A considerable
portion of the striker’s kinetic energy was absorbed during this phase via axial
compression of the plates. As a consequence, the deformation under impact loading
was significantly less than that corresponding to the same amount of absorbed energy
under quasi-static loading. This phenomenon was evidenced by impact-to-static
energy ratios being greater than unity, even after accounting for the effect of strain
rate. In a subsequent study it was demonstrated that the deformation of Type II
structures has two phases: the first phase involves only plastic compression of the
specimen and the second phase involves sole rotation of the plastic hinges. As a
consequence, lateral inertia is the dominant effect in the first phase and the behaviour
of the second phase is more sensitive to strain rate (Tam & Calladine 1991). Su et al.
(1995a) extended the theoretical model of the Type II structure and confirmed the
findings of other studies that the dynamic response of the structure and the final
deflection are dominated by the effective mass ratio (ratio of impact mass to
specimen mass), rather than the impact velocity. Such a structure may therefore be

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 16
referred to as “velocity sensitive” rather than “mass sensitive” or “inertia sensitive”.
Furthermore, the initial peak load under dynamic loading was greater than that under
quasi-static loading. A companion study also found that this dynamic effect is
enhanced by strain rate effects if the material is strain rate sensitive (Su et al. 1995b).
Additionally, the initial peak load was found to decrease as the initial kink angle of
the kinked plates increases. A similar effect is achieved in the automotive industry by
adding mechanically imposed triggers to the vehicle structural members to reduce
inertia effects and promote a stable crushing mode during collapse (Langseth &
Hopperstad 1996; Langseth et al. 1999).

It has been shown that thin-walled square tubes deforming under an axial impact load
are characteristic of a Type II structure. For example, the dynamic axial crushing of
square aluminium tubes was found to be influenced by inertia effects (Langseth et al.
1999). Increasing the impact velocity led to an increase of the initial peak (buckling)
load. Furthermore, lateral inertia forces tended to “support” the side-walls of the tube
in the early phase of deformation, thus increasing the axial deformation before
buckling. Such behaviour is known as dynamic plastic buckling, and occurs under
certain dynamic loading conditions when a considerable portion of the initial kinetic
energy is absorbed by axial compression of the shell due to the influence of inertia
effects during the impact event. The dynamically loaded specimen shown in Figure
2.1-1(b) is an example of dynamic plastic buckling. Various recent studies on this
phenomenon have been carried out for both circular and square tubes. Such studies
have revealed that square and circular tubes are sensitive to both velocity and mass
when subject to axial impact loading, such that larger energies can be absorbed by a
shell for high-velocity impacts when decreasing the striking mass. This is similar to
the response exhibited by the Type II structure discussed earlier and shown in Figure
2.1-2(b). Furthermore, the initial buckling mode of the tubes is governed by the
propagation of axial stress waves along the tube’s length (Karagiozova & Jones
2004). In turn, this stress wave propagation is determined by the inertia
characteristics and material properties of the shell. Other studies on the dynamic
plastic buckling of square and circular tubes have been made by Karagiozova et al.
(2000) and Karagiozova and Jones (2001). Karagiozova and Jones (2001) examined
the effect of impacting a tube into a rigid wall compared with impacting a stationary
tube with a finite mass having the same energy and velocity. It was found that only

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 17
the buckling shape of the tube, and not the energy absorbed per unit crush length
(defined in Section 2.2.1.3) depended on the inertia of the striker.

For certain cases, the inertia forces associated with impact loading tend to reduce the
maximum force required to crush an energy absorber. For example, in the case of the
dynamic inversion of a metal tube onto a die, the inertia of the inverted metal tends
to aid in the inversion process, such that less load is required to compress the tube
compared to when it is loaded quasi-statically (Harrigan et al. 1999). On the other
hand, Singace et al. (2001) analysed the energy absorption performance of right
circular frusta (cones) under dynamic loading and compared it to that under quasi-
static loading. They discovered that, due to inertia effects, the amount of energy
absorbed under dynamic loads was greater than that under quasi-static loads.
Furthermore, Reid and Reddy (1986), on investigating the static and dynamic
crushing of tapered metal tubes, discovered that the mean crushing force increased
under dynamic loading due to material strain rate sensitivity. It can therefore be
observed that the energy absorption response caused by inertia effects depends
heavily upon the type of energy absorber.

Overall, this section shows that the response of thin-walled tubes under dynamic
impact loading conditions is sensitive to impact mass and velocity due to inertia
effects.

2.1.1.2 Material strain rate sensitivity

It has been observed that the properties of many materials under dynamic loading
conditions are different to the corresponding quasi-static property values (Jones
1989). The stress-strain relations in particular are sensitive to the speed of a test. This
phenomenon, known as strain rate sensitivity or viscoplasticity, can significantly
influence the dynamic response of energy absorbing structures, and as such is
considered in this section.

The yield criteria or plastic flow of various materials is sensitive to strain rate, such
that the yield and ultimate stress of the material tends to increase as the strain rate
increases. The fact that the structure essentially strengthens as the material strain rate

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 18
increases can be unacceptable behaviour for energy absorbers used to enhance
vehicle crashworthiness, since the system could transmit detrimental decelerations
and forces to the human body (Jones 1989). Many constitutive equations have been
developed to theoretically explain the strain rate sensitive behavior of materials.
Determining the various coefficients in these equations requires careful experimental
work. Much work has been carried out to establish the characteristics of these
equations, however there is still uncertainty and lack of reliable data even for
common materials.

A constitutive relation which shows reasonable agreement with the available


experimental data for various metal is the Cowper-Symonds constitutive equation,
defined as:

ε& p = D{(σ d / σ s ) − 1}q for σ d ≥ σ s (2.1-1)

where σd is the dynamic flow stress at a uniaxial plastic strain rate, ε& p and σs is the

associated static flow (or yield) stress. The constants D and q are material
parameters. According to Reid and Reddy (1986), the Cowper-Symonds relationship
is widely used to account for strain rate effects in dynamic structural plasticity
problems. The dynamic flow stress agrees reasonably well with the dynamic uniaxial
tension and compression test results on several materials. Equation 2.1-1 can also be
written as:

σ d / σ s = 1 + (ε& p / D )1 / q (2.1-2)

Collectively, Equation 2.1-1, its logarithmic form and Equation 2.1-2 provide a
reasonable estimate of the strain rate sensitive uniaxial behavior of mild steel. Table
2.1-1 gives typical and commonly used values of D and q for certain metals. It
should be noted that, for the values of D and q for mild steel in Table 2.1-1, the
Cowper-Symonds equation is only suitable for small strains (e.g. 2-4 %), whereas the
strains in many structural crashworthiness applications are much larger.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 19
Table 2.1-1 Values of D and q for various metals (Jones 1989).

This figure is not available online.


Please consult the hardcopy thesis
available from the QUT Library

In this case, the coefficient D in Equation 2.1-2 would be larger and the associated
strain rate and load enhancement would be less. To account for this, an alternative
has been to modify Equation 2.1-2 as follows (Jones 2000):

σ d / σ s = 1 + {ε& p /( B + Cε )}1 / q (2.1-3)

where ε is the strain and B and C are constants. The literature indicates that the
Cowper-Symonds equation has been modified in other ways to suit the strain and
strain rates typical of crashworthiness scenarios, however the strain rate
characteristics of many materials for larger strains are still not well understood.

Another alternative which has been adopted to take account of the larger strains in
crashworthiness applications is to base Equation 2.1-2 on the static ultimate stress
rather than the static yield stress. The values D = 6844 s-1 and q = 3.91 were found to
fit experimental data for the ultimate stress of steel specimens examined by
Campbell and Cooper (1966). In other words, these values for D and q were found
suitable for mild steel tubes which deform with strains close to those associated with
the static ultimate tensile stress. A number of studies on the crushing of mild steel
tubes have used these values (Abramowicz & Jones 1984; 86; Reid & Reddy 1986;
Reid et al. 1986).

Further background on the strain rate sensitive behaviour of materials is provided in


Jones (2000; 1999; 89).

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 20
2.1.2 Structural crashworthiness

Structural crashworthiness refers to the impact performance of a structure, and in


recent times has become an integral part in the design of vehicular structures where
passengers are involved, such as automobiles. It is necessary to ensure the structure
is capable of withstanding impact forces and is able to absorb impact energy without
excessive damage to itself or injury to the occupants it contains. According to
Mamalis et al. (2001):

Crashworthiness studies are in high demand for designing transportation systems.


Energy absorption capability in vehicle collisions is an important parameter for the
development of passive safety systems dealing with the protection of passengers and
cargo departments in the event of an accident.

Lu and Yu (2003, 11) define crashworthiness as:

…the quality of response of a vehicle when it is involved in or undergoes an impact.


The less damaged the vehicle and/or its occupants and contents after the given event,
the higher the crashworthiness of the vehicle or the better its crashworthy
performance.

Johnson and Mamalis (1978) provide a concise and practical overview of


crashworthiness as applied to automobiles, trains, aircraft and ships.

As mentioned in Chapter 1 of this thesis, energy absorbers are commonly used as


structural members in vehicles to maintain crashworthiness. Perhaps the most
obvious example is the front chassis rails of a vehicle, which deform to absorb
energy during a frontal impact. Different energy absorbing components are used to
absorb energy for different types of impacts. For high speed frontal impacts the front
chassis rails and engine bay structure are designed to deform and absorb energy. It is
desirous to maintain the structural integrity of the passenger compartment during an
impact event, particularly during side impacts. The latter usually involves reinforcing
the doors and side-pillars to withstand a high speed impact, thus reducing intrusion
of the passenger compartment. Many countries use comprehensive testing procedures
to assess the crashworthiness of new vehicles on the market. For high speed impacts
both frontal offset and side impact crash tests are frequently used. The former

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 21
involves crashing the vehicle into a barrier with a crushable aluminium face which
extends across approximately 40 % of the vehicle’s front. In a side impact test the
vehicle is stationary and a trolley with a crushable aluminium face impacts the side
of the vehicle. Recently provisions have been made in European test programs for
separate tests to assess pedestrian protection.

For relatively low speed frontal impacts, the front “crush boxes” deform to absorb
energy. These crush boxes are usually separate components which are attached to the
vehicle structure. In a low speed frontal impact the crush boxes may permanently
deform, thus avoiding or reducing damage to the rest of the vehicle. The crush boxes
can then be removed and replaced. Different energy absorbers are also used to absorb
impact energy during low speed frontal or rear collisions. Figure 2.1-3 shows a
cellular structure made of Ethylene Vinyl Acetate (EVA), used as the front energy
absorber in a popular commercial passenger vehicle. The structure is shown in Figure
2.1-3(a) being tested under a quasi-static axial load to determine its load-deflection
response. Figure 2.1-3(b) shows the final deformed profile of the structure. The
buckling of the individual cell walls as load was applied was clearly evident, though
is difficult to see in Figure 2.1-3(b) due to rebound of the walls. The use of cellular
and honeycomb structures as energy absorbers is further considered in Section
2.2.2.3.

(a) (b)
Figure 2.1-3 Energy absorbing cellular structure from a passenger vehicle: (a)
quasi-static axial testing and (b) final deformed profile.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 22
Isolators have also been used in early and current vehicle models for energy
absorption during low speed frontal and rear impacts, and these types of absorbers
are treated in Section 2.2.2.5.

Other common examples where energy absorbers are used to maintain vehicle
crashworthiness are as follows:

(1) Aircraft floor structures which are designed to crumple during crash landings
and reduce impact decelerations transmitted to the occupants (Figure 2.1-4).
(2) Energy absorbing structures at the front of passenger trains. These structures
crush to reduce impact forces transmitted down the length of the train during
a frontal collision, and thus localise deformation towards the front of the
train.
(3) Thin-walled tubing at the base of lift shafts to arrest the lift in a controlled
manner in the event of a cable-breakage or over-wind. Tube inversion, which
will be discussed in Section 2.2.2.1, has been used as an energy absorption
mode in this application.

Figure 2.1-4 Floor structures in a helicopter which are designed to absorb


energy during a crash landing (Figure 9 on page 203 of: Kindervater, C. and
Georgi, H. 1993. Composite strength and energy absorption as an aspect of
structural crash resistance. In Jones, N. and Wierzbicki, T. (eds), Structural
Crashworthiness and Failure. Essex: Elsevier Applied Science Ltd.)

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 23
2.2 Energy absorbers

An energy absorber (in the context of crashworthiness) is a device that converts


impact kinetic energy into another form of energy, thus reducing the peak damaging
forces transmitted to the structure being protected. This energy conversion can be
reversible, as in the case of a compressible fluid, or irreversible, such as involving
plastic dissipation energy associated with the permanent deformation of a solid.
Another common name used in the literature for an energy absorber is an Energy
Absorbing (EA) mechanism, however the term energy absorber is adopted in this
thesis.

There is a wealth of information in research journals, technical books and


commercial realms as to the types, development and uses of energy absorbers for
various applications. To develop an energy absorbing system it is necessary to
review existing energy absorbers to determine which will be suitable as a means of
absorbing impact energy, and whether such absorbers will need to be modified.
Often, at the absorber selection stage the primary aspect of interest is the energy
absorbing capacity. Other aspects which may be later considered are mounting of the
chosen energy absorbers, their complexity and availability of materials. This section
provides a comprehensive review of the energy absorbers covered in the literature,
with a view to identifying areas for future research.

2.2.1 Principles of operation of energy absorbers

In the first place it is worth establishing the basic principles of operation of energy
absorbers in general. The open literature reports several quantitative criteria used to
evaluate the energy absorbing capacity of energy absorbers. Such criteria are
typically used in the initial design stages of energy absorbing systems. The following
is a list of such criteria, some of which are used in the subsequent analyses chapters
of this thesis:

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 24
(1) Energy absorbed per unit mass,
(2) Mean crushing stress,
(3) Energy absorbed per unit crush length,
(4) Specific crushing stress,
(5) Crush force efficiency,
(6) Stroke efficiency,
(7) Dynamic amplification factor, and
(8) Stroke length per unit mass

These will be addressed in turn.

2.2.1.1 Energy absorbed per unit mass

The energy absorbed per unit mass, Em (kJ/kg), also called the specific energy, is
given as:

E m = E abs / m (2.2-1)

where Eabs is the absorbed energy, and m is the original mass of the undeformed
absorber. The energy absorbed per unit mass of a particular energy absorber depends
upon the mode in which it fails. For instance a hollow tube crushed axially has a
much higher Em than a tube of the same mass indented laterally, since more material
is deformed for the axially crushed tube and hence more energy is absorbed through
deformation. The specific energy allows a comparison between different energy
absorbers to determine which is more efficient in terms of the amount of energy they
absorb for a given mass, and is a useful parameter when weight reduction is
important.

A word of caution is worth making. Even though the energy absorbed per unit mass
for a particular energy absorber or material may be very high, this does not
necessarily mean that the absorber can be smaller or shorter than an energy absorber
with a lower Em to absorb the same amount of energy. This is because the energy
absorbed per unit mass also depends upon the density of the absorber material. For

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 25
instance, aluminium tube has a higher energy absorbed per unit mass than steel tube
of the same dimensions. However, since aluminium tube has a lower mass per unit
length, a longer length of tube is required for it to absorb the same amount of energy
as the steel tube. Table 2.2-1 from Lin and Mase (1990) compares the energy
absorbed per unit mass for different materials, and the length of tube needed to
absorb 31.7 kJ.

Table 2.2-1 Energy absorbed per unit mass and tube length needed to absorb
31.7 kJ, for various materials (Lin & Mase 1990).

This table is not available online.


Please consult the hardcopy thesis
available from the QUT Library

2.2.1.2 Mean crushing stress

The mean crushing stress, σ cr (MPa) is the mean crushing force, Fmean divided by
the original cross-sectional area, A0 of the energy absorber:

σ cr = Fmean / A0 (2.2-2)

The mean crushing force is the average force over which the energy absorber
deforms in a stable manner, and is obtained for a given deflection by dividing the
absorbed energy by the crush distance.

The load-deflection curve is used extensively to analyse the crushing behavior of


energy absorbers. This shows the load required to deform the energy absorber versus

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 26
the deflection as the absorber deforms, and is characteristic for each energy absorber.
Figure 2.2-1 shows the load-deflection curve for a thin-walled steel tube of square
cross-section as it is crushed under an axially applied quasi-static load. The curve
will be analysed to explain the associated terminology. The curve is the result of a
computer simulation conducted by the author, while the mean load represented by
the dashed line is obtained from a theoretical model (Reid & Reddy 1986), which
will be described in Chapter 3.

160
140 Fpeak
120
100
Load (kN)

Fmean
80
60

40
20
0
0 25 50 75 100 125 150 175 200 225 250
Deflection (mm)

Figure 2.2-1 Load-deflection curve for the axial crushing of a thin-walled


rectangular tube made of mild steel.

Following the graph from left to right, the deformation process begins with the initial
application of load to the energy absorber, shown by the sharp rise in force over a
relatively small distance. This corresponds to the elastic region of the column
material, where the yield point has not yet been reached. The peak crushing force,
Fpeak is reached when either the material yields or the tube buckles elastically. Which
of these mechanisms occurs depends on the slenderness of the tube (ratio of wall
thickness to tube length). If the slenderness ratio is small then elastic buckling
occurs, whereas if the slenderness ratio is relatively large then the tube will yield
(plastically collapse). After this point the load drops somewhat dramatically to enter
the post-buckling region where the column undergoes progressive deformation, in
which folds or lobes are formed down the length of the tube. How the load fluctuates
in this region determines the energy absorbing characteristics of the energy absorber.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 27
The ideal case is for the load to remain constant, thus offering constant resistance and
hence deceleration throughout the entire stroke. In reality, however, the load always
fluctuates about a steady state crushing force as the tube is crushed. These
fluctuations depend predominantly on the mode of deformation, which in turn
depends upon the cross-sectional dimensions of the energy absorber, the boundary
conditions and the material properties. If the cross-sectional dimensions of the tube
are such that it buckles globally by Euler-type buckling then the load will
dramatically decrease, leading to a reduction in the amount of energy absorbed. As
such, local and axial progressive deformation is favored for maximum crushing and
hence energy absorption, and the dimensions of the energy absorber are tailored for
such.

For convenience, designers often disregard the load fluctuations as the tube crushes
and use a mean value for the load, Fmean, as indicated in Figure 2.2-1. Both
theoretical and empirical methods have been developed to calculate the mean load
for thin-walled tubes of a variety of cross-sections and profiles, as will be discussed
later in this chapter. For experimental and numerical investigations, the mean load
can be calculated using the energy absorption and crush deflection of the tube. The
absorbed energy is obtained from the integral of the crushing force with respect to
the displacement.

As described above, the most efficient use of the tube material (in terms of
maximising energy absorption) is obtained when as much of the tube as possible is
crushed, thus maximising the volume of material which reaches plasticity. This is
illustrated in Figure 2.2-2 for a thin-walled tube with a square cross-section (Jones
1989). According to Chen and Wierzbicki (2001) the axial progressive folding
deformation of thin-walled metal tubes, as shown in Figure 2.2-2, is recognised as an
efficient energy absorbing mode. Referring back to Figure 2.2-1, the load starts to
rise towards the maximum crush distance, dmax. This is due to the folds stacking up
on each other at the base of the tube, and is known as compaction or fold
consolidation. Such a phenomenon prevents 100 % energy absorption efficiency
from being possible in actual thin-walled energy absorbers undergoing progressive
deformation (Yu-Hallada et al. 1998).

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 28
This figure is not available online.
Please consult the hardcopy thesis
available from the QUT Library

Figure 2.2-2 A thin-walled mild steel tube with a square cross-section before and
after quasi-static axial crushing (Jones 1989).

This has led to a feature commonly know as the effective crushing distance, which
states that the deformation in any given fold in a tube is restricted by the interference
of adjacent folds and strain hardening (Abramowicz & Jones 1984). This causes an
increase in the mean load. Theoretical models for the axial crushing of thin-walled
tubes can take account of the effective crushing distance when calculating the mean
load, and this will figure in the next chapter.

If the tube in this example were to be impacted by a mass with a sufficiently high
velocity such that inertia effects influenced the response of the absorber, a larger
peak load would be observed due to the influence of inertia effects and instantaneous
plastic strains at the beginning of the response.

The last feature to point out on the load-deflection curve is the energy absorbed
through deformation of the energy absorber. Obviously, through basic mechanics,
this is simply the area under the curve given by the expression:

d max

Es = ∫ F .dx
0
(2.2-3)

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 29
where the energy absorbed (kJ) has been calculated up to the maximum crush
distance, dmax (mm), and F is the crushing force (kN).

2.2.1.3 Energy absorbed per unit crush length

The energy absorbed per unit crush length, Ecl (kJ/m) is important when the available
crush zone is limited. For impact loading it can be calculated using the impact energy
divided by either the undeformed length of the absorber or the maximum crush
distance reached for a combination of loading and geometry parameters.

2.2.1.4 Specific crushing stress

The specific crushing stress (MPa.m3/kg) is simply the mean crushing stress divided
by the material density.

2.2.1.5 Crush force efficiency

The crush force efficiency, FE is the mean crushing force divided by the peak
crushing force, or:

FE = Fmean / F peak (2.2-4)

It is desirable to maximise the crush force efficiency for energy absorbers used in
crashworthiness design when the protection of occupants is a priority. Peak impact
forces and decelerations are transmitted through the front energy absorbing crash
boxes to the passenger compartment of a vehicle during a frontal impact, and can be
sufficiently high so as to exceed the tolerable limits of the passengers. As such, crush
triggers, or indentations in the side walls of the energy absorber have been
traditionally introduced to reduce the peak forces and hence increase the crush force
efficiency. These triggers act as stress raisers as the initial load is applied, thus
assisting in the deformation of the energy absorber.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 30
2.2.1.6 Stroke efficiency

The stroke efficiency, SE is a measure of how much of the total length of the energy
absorber is actually used up in deformation and subsequent absorption of energy. The
stroke efficiency is given as:

S E = d max / l (2.2-5)

where dmax is the maximum crush distance of the absorber at which compaction
occurs and the energy absorbed is at a maximum, and l is the absorber’s original
length. In many cases deformation occurs over a majority of the length of the energy
absorber, yet further deformation is hindered due to the build-up of compacted
material towards the end of the crush process, as discussed in the case of the thin-
walled steel tube. On the other hand, cellular materials such as honeycombs and
foams can attain a much higher stroke efficiency. This is because the large amount of
space within the cells allows greater compressibility of the material. Clearly, the aim
is to develop an energy absorber which deforms over as much of its length as
possible, thus maximising energy absorption. The stroke efficiency is a useful
criterion in applications when there are restrictions on the available space over which
energy is to be absorbed. Furthermore, maximising the length available for energy
absorption reduces the required retarding force for a given input energy, according to
the principle that work done by a force is equal to its magnitude times the resulting
displacement along the line of the force.

2.2.1.7 Dynamic amplification factor

The dynamic amplification factor, DAF is the ratio of the energy absorbed under
dynamic loading to the energy absorbed under quasi-static loading up to a given
crush distance, and is given as:

Ed
DAF = (2.2-6)
Es

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 31
This relatively uncommon parameter provides a simple way of assessing the dynamic
effects due to inertia and strain rate effects on energy absorption of the energy
absorber.

2.2.1.8 Stroke length per unit mass

The stroke length per unit mass (mm/kg) is simply the maximum crush distance
divided by the total mass of the energy absorber. This is similar to the dependence of
the energy absorbed per unit mass upon the mass of an energy absorber, and provides
another way of comparing the energy absorbed for different materials of different
densities.

Another important criterion for an energy absorber is that it must possess a stable and
repeatable deformation mode. When in service, an energy absorber may be subject to
working loads which are uncertain in their magnitude, pulse shape, direction and
distribution (Lu & Yu 2003). The absorber should have a stable and repeatable
deformation mode which is insensitive to the above loading uncertainties, while
ensuring the required energy-absorption capacity is met. For instance, the crush cans
at the front of a vehicle must be designed to effectively absorb energy under axial as
well as oblique impact loads, over a range of vehicle velocities.

Finally, energy absorbers must be easy to manufacture, install and maintain in order
to be cost-effective. This is particularly the case for energy absorbers which are used
on a “once only” basis, i.e. once having been deformed they are discarded and
replaced.

2.2.2 Energy absorber types and their applications

The previous section established the various criteria that would allow evaluation and
comparison of the reviewed energy absorbers based on their energy absorbing
capacity. In this section the main types of energy absorbers identified in the open
literature will be analysed.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 32
Energy absorbers encountered in the literature were classified into the following
major groups:

(1) Thin-walled tubes and columns,


(2) Tapered tubes,
(3) Honeycombs,
(4) Foams, and
(5) Isolators

The energy absorption of other structures and materials has been studied, such as
rings, ring systems, elastomers, wood, and the bending of channel and angle sections.
However, to simplify the present study these areas are omitted. Nevertheless, to
complement the information provided in this section, the reader is referred to the
papers by Alghamdi (2001) and Abramowicz (2003) for recent reviews of collapsible
thin-walled impact energy absorbers, and the book by Lu and Yu (2003) for a
broader discussion on energy absorption of structures and materials.

2.2.2.1 Thin-walled tubes and columns

Thin-walled tubes are considered the most common type of energy absorber
(Alghamdi 2001). This is primarily due to their relative simplicity compared with
other energy absorbers. In terms of use as energy absorbers they have a high energy
absorption capacity and therefore have been extensively used in modern vehicle
frames, designed to absorb impact energy via material deformation through crushing.
Jones (1989) observed that thin-walled circular tubes are desirable as energy
absorbers since they are inexpensive, efficient and versatile. Such desirable energy
absorption characteristics have led to them being used as energy absorbers in a
diverse range of applications, notably behind car bumpers and train buffers, and at
the base of lift shafts.

The most common materials used for thin-walled tubes are steel, aluminium alloy,
and Fiber Reinforced Composites (FRCs). The deformation of tubes made from
metals has been studied for several decades, whereas tubing made from composite
materials is still a relatively new area. As mentioned before, the deformation mode of

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 33
the energy absorber is important since it governs how much energy is dissipated,
how quickly, and in what manner. Metallic and composite tubing each deform and
dissipate energy in markedly different ways. Metals are ductile and absorb energy
through gross plastic deformation, whereas most composite materials are brittle and
absorb energy through a complex combination of fracture mechanisms including
matrix cracking, delamination, and fiber breakage (Beard & Chang 2000). Due to the
common nature of metallic thin-walled tube energy absorbers, they will be
considered in this section. However, it is worth noting that FRCs have displayed
increasing popularity for use in thin-walled energy absorbers due to their good
energy absorption behaviour. For instance, polymer composite materials have much
higher energy absorption per unit mass compared with metals such as mild steel and
aluminium. For further information on the use of FRCs as energy absorbers, refer to
Lu and Yu (2003), Haug and De Rouvray (1993), Kindervater and Georgi (1993),
Rechnitzer et al. (1996), Arnaud and Hamelin (1998) and Mamalis et al. (2004).

Though there a variety of deformation modes through which metallic thin-walled


tubes can dissipate plastic energy, the five most common, which will be examined
here, are:

(1) Axial crushing,


(2) Axial inversion,
(3) Axial splitting,
(4) Lateral bending,
(5) Lateral indentation, and
(6) Lateral flattening

Axial crushing

Axial crushing of tubing is one of the most commonly used tube deformation modes
since it provides a reasonably constant crushing force. Also, it has comparatively
high energy absorbing capacity since under axial loading almost all of the material
participates in the absorption of energy by plastic work. Perhaps the most commonly
used thin-walled tubes for energy absorption by axial crushing are those of circular,
square or rectangular cross-section. An axially crushed circular tube can undergo one

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 34
or more of three distinct deformation modes: global buckling, concertina mode of
deformation, and diamond mode of deformation. The latter two are local deformation
modes and are favorable over the global deformation mode since they promote the
deformation of more material, and hence dissipate more energy. Which of the three
modes occurs when a circular tube is axially crushed depends on its ratio of diameter
to thickness, and its ratio of length to wall thickness. Other factors which influence
the axial deformation mode of thin-walled tubes are the material and end constraints.
Reference to further works on the axial crushing of circular tubes is provided in Lu
and Yu (2003), and Jones (1989).

When a thin-walled square tube is axially crushed, it collapses in either a symmetric


or non-symmetric mode, depending on the ratio of mean side width, C to mean wall
thickness, H. For example, the symmetric (progressive) crushing mode (Figure
2.2-3(a)), occurs for square thin-walled tubes with approximately C / H > 40.8 (Jones
1989). For thick square tubes with approximately C / H < 7.5, an extensional
buckling mode can occur (Figure 2.2-3(b)). For 7.5 ≤ C / H ≤ 40.8, an asymmetric
mixed B-type progressive buckling mode occurs (Figure 2.2-3(c)). However, the
difference between the theoretical crushing forces associated with a symmetric mode
and an asymmetric B mode is small such that either may occur in a square tube
specimen which has slight imperfections (Jones 1989).

This figure is not available online.


Please consult the hardcopy thesis
available from the QUT Library

Figure 2.2-3 Axial progressive deformation modes of thin-walled square tubes:


(a) symmetric mode, (b) extensional mode and (c) asymmetric mixed B-type
mode (Jones 1989).

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 35
This figure is not available online.
Please consult the hardcopy thesis
available from the QUT Library

Figure 2.2-4 Basic collapse elements: (a) Type I- inextensional and (b) Type II-
extensional (Jones 1989).

Each of the mechanisms in Figure 2.2-3 can be theoretically represented using the
basic collapse elements illustrated in Figure 2.2-4. The Type I element consists of
adjacent inward and outward forming faces, such that the mid-line running across
each face does not theoretically extend. The Type II element, on the other hand,
consists of both the adjacent faces moving outwards, such that the mid-line extends.
This requires greater plastic deformation than the Type I element, and hence a
greater crush force. Each layer of lobes in a symmetric crushing mode is idealised
with four of the Type I basic collapse elements, while the extensional mode of
crushing is idealised with four of the Type II basic collapse elements. Finally, the
asymmetric mixed B-type progressive buckling mode is idealised as two adjacent
layers of lobes having seven Type I basic collapse elements and one Type II basic
collapse element.

The elastic and plastic half-wavelengths also dictate the deformation mode of the
tube. For square and rectangular tubes, the elastic and plastic half-wavelengths are
equal to or dependent on the side lengths (width) of the tube. For rectangular tubes
the values of these wavelengths are intermediate between the side widths. In these
cases the elastic half-wavelength depends on the aspect ratio (tube width to tube
length), while the plastic fold length also depends on the strain hardening capacity of
the material (Reid & Reddy 1986). For square and rectangular tubes, when the elastic
half-wavelength is greater than the plastic half-wavelength, non-compact behavior
occurs, as shown in Figure 2.2-5(a).

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 36
This figure is not available online.
Please consult the hardcopy thesis
available from the QUT Library

(a) (Reid et al. 1986) (b) (Jones 1989)


Figure 2.2-5 (a) Non-compact and (b) compact deformation of a tube subject to
an axial load. (Figure (a) reprinted from Internatoinal Journal of Mechanical
Sciences, Vol. 28, No. 5, Reid, S. R., Reddy, T. Y. and Gray, M. D., Static and
dynamic axial crushing of foam-filled sheet metal tubes, p. 295-322, 1986, with
permission from Elsevier.)
This non-compact deformation mode consists of folds which tend to form at the
sections of peak deformation of the elastic buckling pattern, leaving the folds
separated by curved, relatively undeformed panels. This leads to irregular behavior
and the frequent triggering of an Euler-type instability in the tubes (Reid et al. 1986).
Non-compact response is thus undesirable from an energy absorption viewpoint, and
can be avoided by filling the tube with foam (Reid et al. 1986), or by tailoring the
ratio of wall thickness to width, to produce the more regular and stable compact
deformation mode shown in Figure 2.2-5(b).

Crush load uniformity has been the focus of much research on the axial crushing of
tubes. As mentioned in Section 2.2.1.2, it is desirable for an axially loaded tube to
have a stable load-deflection response, such that energy absorption is maximised and
a constant retarding force is obtained. Adding grooves to the side-walls of axially
loaded tubes is an effective means of stabilising the load-deflection response, since
the grooves force the plastic deformation to occur at predetermined intervals along
the tube. Such has been studied for corrugated metal thin-walled tubes (Singace &
El-Sobky 1997). The pre-formed corrugations in the wall of the tube helped promote
bending moments as the tube crushed, thus reducing the fluctuations in crush force.

The load-deflection and energy absorption response of thin-walled steel circular


tubes can be controlled by changing the distances of circumferential grooves cut

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 37
alternately inside and outside the tubes (Hosseinipour & Daneshi 2003). El-Sobky
and Singace (1999) investigated the axial compression of profiled polymer pipes and
found that the profiles improved the uniformity of the load. Often, thin-walled tubes
which are used as energy absorbing elements at the front of automobiles have local
geometry imperfections in their side-walls, such as that shown in Figure 2.2-6.
Known as “crush initiators”, these imperfections promote localised deformation as
the tube starts to deform under an axial load, hence reducing the initial peak load.

(a) (b)
Figure 2.2-6 Crush triggers or initiators in the side walls of a thin-walled tube:
(a) undeformed profile and (b) final deformed profile.

In more recent times, tubes of less conventional cross-sections have become the
focus of attention for use as impact energy absorbers. This particularly applies to
multi-corner columns which have more than four sides. Multi-corner columns have
found much use in recent times as energy absorbers in automobiles. Aluminium
extrusion technology allows internal stiffening ribs to be integrally extruded with the
tube, such that cross-sections consisting of multiple cells are commonly used in the
automotive industry, such as the tube shown in Figure 2.2-6. Steel thin-walled tubes
of octagonal cross-section subjected to quasi-static and low-speed axial impact
loading have been studied using experiments and finite element simulation (Mamalis
et al. 2003). It was hypothesised that using an octagonal shape in an energy absorber
such as for crashworthiness would provide better energy absorption than a
rectangular tube. However, the cited work made no direct comparison between the
energy absorption characteristics of the two structures. Finally, Chen and Wierzbicki
(2001) studied the axial crushing of single-cell, double-cell and triple-cell aluminium
extrusions, both analytically and numerically. They found that for a given energy

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 38
absorption the energy absorbed per unit mass increased from the single- to the triple-
cell columns. Furthermore, increasing the number of cells led to an increase in the
average crushing force.

Axial inversion

Introduced by General Motors in 1969, tube inversion, as the name suggests,


involves the inverting of a tube over a die as it is pressed onto the die (Figure 2.2-7).
The thin metal tube is made of a ductile metal (usually aluminium alloy, however
steel tubes have also successfully been inverted), and is either turned inside out or
outside in, as the tube edge is forced over the smooth corner on the die. This
deformation mode results in a generally constant inversion load for uniform
thickness tubes. Thin-walled tubes with a tapering wall thickness have also been
inverted, and tests have shown an increase in absorbed energy per unit mass of up to
50% compared with uniform thickness tubes (Chirwa 1993).

Figure 2.2-7 Load-deflection profile and examples of inverted tubes


Reprinted from Internatoinal Journal of Mechanical Sciences, Vol. 35, No. 12,
Reid, S. R., Plastic deformation mechanisms in axially compressed metal tubes
used as impact energy absorbers, p. 1035-1052, 1993, with permission from Elsevier.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 39
Axial splitting

Tube splitting involves loading the tube onto a die with a radius larger than that used
for inversion. The plastic energy is dissipated via plastic bending, stretching and
tearing of the metal into strips which can be subsequently curled to absorb further
energy. Axially splitting tubes is efficient in terms of energy absorption, and the
tubes can sustain long stroke (up to 90 % of the tube length), with an almost constant
crush load. Both circular and square tubes can be successfully split, and Figure 2.2-8
shows an example of the former.

Miscow F. and Al-Qureshi (1997) investigated the quasi-static and dynamic


inversion of copper and brass tubes. They demonstrated using experiments that the
mode of deformation, be it crushing, inversion, splitting or global (Euler) buckling,
depends primarily on the radius of the die, and the ratio of die radius to tube outer
diameter.

Figure 2.2-8 Load-deflection profile and examples of split tubes. The top
specimens have been subsequently curled (Reid 1993).
Reprinted from Internatoinal Journal of Mechanical Sciences, Vol. 35, No. 12,
Reid, S. R., Plastic deformation mechanisms in axially compressed metal tubes
used as impact energy absorbers, p. 1035-1052, 1993, with permission from Elsevier.
G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 40
Lateral bending

Realising that axial progressive folding of thin-walled energy absorbers rarely acts
alone in vehicle crash events, researchers have recently turned their attention to the
bending collapse mode of thin-walled members. Their motivation has been the
observation that this is the most predominant failure mode of structural members in
vehicle crashes. Bus rollover is an example where the thin-walled tubes comprising
the bus frame are subject to bending collapse behavior. The associated energy
absorption of the structure is localised at plastic hinges. Figure 2.2-9 shows a typical
plastic hinge formed during bending collapse of a rectangular tube which has been
fixed at its base. The bending mechanism firstly involves bulging of the side webs
followed by development of the collapse mechanism with traveling hinge lines.
These hinge lines define the growth of the inward and outward forming buckles. As
the rotation angle of the tube increases, the traveling hinge lines eventually stop and
additional hinge lines develop. The bending mechanism stops when jamming occurs
between the two buckled halves of the compression flange.

The first theoretical treatment of this problem was provided by Kecman (1983).
Validated using extensive experimental tests, a theoretical model was developed
which related the hinge moment to the angle of rotation for the bending collapse of
rectangular and square section tubes. Chen (2001) performed experimental and
numerical analyses on the bending collapse of aluminium hat profiles (Figure
2.2-10). Introducing foam filler was found to improve the load carrying capacity and
hence the energy absorption of the column by 30 to 40 % over non-filled members,
when bent over realistic distances as encountered in vehicle crashes.

Side web Compression flange

Fixed base of tube

Figure 2.2-9 Typical plastic hinge formed during bending of a rectangular tube.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 41
Figure 2.2-10 Schematic showing the bending collapse of aluminium hat profiles
Reprinted from Internatoinal Journal of Solids and Structures, Vol. 38,
Chen, W., Experimental and numerical study on bending collapse of aluminium
foam-filled hat profiles, p. 7919-7944, 2001, with permission from Elsevier.

Tubes of hat section have found common usage as energy absorbing elements in
vehicle body structures (White & Jones 1999). Under axial compression, such tubes
behave similarly to square and rectangular tubes, while the load-deflection curves
largely resemble those for other sections.

Real-world crashworthiness applications seldom involve purely axial or bending


loading. Rather, oblique loading occurs which involves a combination of axial and
transverse bending loads. Such loading acts at an angle to the thin-walled tube’s
longitudinal axis, causing the tube to deform via a combination of both axial and
global bending (Euler) collapse modes. As such it is important to understand how
thin-walled tubes respond and absorb energy under oblique loading. Relatively few
studies on the oblique loading response of thin-walled tubes have been published in
the open literature. Nevertheless, several recent studies have been conducted on
square and circular tubes under combined loading. Han and Park (1999) performed
numerical investigations on the oblique loading of mild steel square columns. They
achieved oblique loading conditions by axially impacting the column at an inclined
rigid wall. Contact between the column and wall was assumed frictionless. Results
showed that there is a critical angle at which a transition occurs from the axial
collapse mode to the bending collapse mode. Furthermore, there was a significant
reduction in the mean load associated with this transition.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 42
The crushing response of thin-walled square columns subjected to combined
compression-bending loading has been studied using both numerical and theoretical
methods (Kim & Wierzbicki 2001). Both methods compared well, while the results
of the study could be applied to the development of simplified crash-oriented design
tools. Reyes et al. (2002) examined the oblique loading response of square
aluminium extrusions using experimental and numerical techniques. Tests were
carried out by clamping the columns at one end and oblique loading conditions were
achieved by applying a quasi-static load at different angles to the centreline of the
column. A validated numerical model was used to perform a factorial study on
oblique impact with variations in parameters such as load angle, column length,
thickness, alloy heat treatment and impact velocity. The response parameter was the
mean load, while the velocity range was 5-15 m/s. The studies showed that the quasi-
static peak crush load dropped drastically over the load angle range considered, i.e.
when bending was introduced into the column. The mean load also decreased as load
angle increased. The factorial analysis showed that the thickness was the most
dominant input parameter, and its effect increased as the load angle was increased. In
like manner, empty and foam-filled circular (Borvik et al. 2003) and square (Reyes et
al. 2004) aluminium tubes have been studied under axial and oblique quasi-static
loading. The results showed that the relative reduction in peak force and energy
absorption as the load angle increases is greater for foam-filled columns than for
empty ones. Other studies involving combined loading of thin-walled tubes are the
combined torsion and bending of circular tubes (Reddy et al. 1996), and the
combined bending and compression of square tubes (Kim & Wierzbicki 2001).

Lateral indentation

Lateral indentation of tubes (Figure 2.2-11) involves applying a point load


perpendicular to the longitudinal axis of the tube. The tube is supported at two
points, and energy absorption is achieved firstly by local indentation at the point
where the load is applied, and subsequently by global bending collapse of the tube.
Indentation can also be achieved using a blunt wedge running across the face of the
tube. Although lateral indentation has been applied to such areas as automobile
bumpers, little interest has been shown in this mechanism from an energy absorption
point of view due to the limited amount of material that participates in plastic

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 43
Figure 2.2-11 Lateral indentation of tubes.
Reprinted from Thin-Walled Structures, Vol. 39, Alghamdi, A., Collapsible impact
energy absorbers : an overview, p. 189-213, 2001, with permission from Elsevier.

deformation, and hence restricted energy absorption capacity. Further background


and reference to specific research on lateral indentation can be found in Lu and Yu
(2003).

Lateral flattening

The final deformation mode of tubes used as energy absorbers which will be
considered is lateral flattening. Intuitively, this involves crushing the entire length of
the tube parallel to its longitudinal axis. This mechanism allows more material
deformation than for lateral indentation however not as much as for axial crushing.
The energy absorbed by an axially crushed tube is one order of magnitude more than
for a laterally flattened tube (Johnson & Reid 1978). The energy absorbed by
laterally flattened tubing has been found to increase when internal bracing is
mounted inside the tube, and when the sides of the compressed tube are constrained.
Constraining the tubes laterally causes more plastic hinges to form during collapse,
thus increasing the energy absorption compared with unconstrained tubes.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 44
2.2.2.2 Tapered tubes

The axial crushing of thin-walled metal tubes has received considerable attention
over the past three to four decades. Much of the literature pertaining to such
investigations have related to axial loading alone. However, in the context of vehicle
crashworthiness, off-axis or oblique loads are unavoidable. Few experimental results
exist to prove the effectiveness of conventional tubing and columns under even
slightly oblique loads (Reid & Reddy 1986).

Recently, increased focus has been given to tapered tubes in which one or more sides
of the tube are inclined to the longitudinal axis. Such structures have been considered
preferable to straight (non-tapered) tubes since they are more likely to provide a
desirable constant mean load-deflection response under dynamic loading (Reid &
Reddy 1986). Also, tapered tubes are capable of withstanding oblique impact loads
as effectively as axial loads, and are less likely to fail by global buckling than
straight tubes. In the context of vehicle crashworthiness the energy absorber is
commonly subjected to both axial and oblique loads, thus tapered tubes may prove
advantageous in such applications. Tapered thin-walled tubes can be circular, square
or rectangular in cross-section. Figure 2.2-12 shows a schematic of straight and
tapered rectangular tubes.

θ
θ

Straight Double taper Triple taper Frusta


(four tapered sides)

Figure 2.2-12 Schematic of straight (non-tapered) and tapered rectangular


tubes.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 45
Compared with straight tubes, relatively few studies have been reported on the
energy absorption performance of tapered tubes. However, several experimental
investigations have been carried out to compare the load-deflection and collapse
mode response of both straight and tapered circular tubes (essentially cones and are
termed frusta or frustra) under axial quasi-static crushing. See for example Mamalis
and Johnson (1983), and Mamalis et al. (1984; 86a; b). Such studies found that the
initial peak and post-buckling (mean) load increase in a broadly parabolic manner
with increasing slenderness ratio (ratio of wall thickness to large end outer diameter).
However, both the peak and mean load decreased with increasing taper angle.
Perhaps the most interesting finding from these studies was that the load-deflection
curves of the frusta were, in general, more stable than those of the straight tubes.
More recent studies on circular frusta have focused on different axial deformation
modes (Alghamdi et al. 2002; Alghamdi 2002a; b), the effect of end constraints on
both collapse mode and energy absorption under axial impact loading (Singace et al.
2001), and the use of composite materials (Karbhari & Chaoling 2003; Mamalis et al.
1997).

Studies on rectangular or square tapered tubes are further limited. However, Reid and
Reddy (1986) investigated the quasi-static and dynamic crushing of straight, single-
tapered and double-tapered rectangular mild steel tubes using experimental and
analytical methods. Quasi-static axial compression tests revealed that the load–
deflection characteristics of the tapered tubes were more stable than for straight
tubes, and the former was less likely to fail via Euler-type buckling. Dynamic tests
performed both normal and oblique to the longitudinal axis of the tapered tubes
revealed that they were almost as effective under oblique impact as they were under
axial impact. More recent studies have focused on the axial crushing of square tubes
with four tapered sides (Mamalis et al. 1989; 2001), wherein such structures were
found to be potential energy absorbers.

Studies have also been carried out on the axial inversion of a cylindrical tube which
has a constant internal diameter and a tapering wall thickness which varies along the
tube’s length (Chirwa 1993). Known as an inverbucktube, it has been universally
used in rail buffers, aircraft landing gears and in energy absorbers that require high
energy absorption capability in a short length. Since the focus of this thesis is on

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 46
tapered tubes which have varying side dimensions, the inverbucktube will not be
further studied here.

2.2.2.3 Honeycombs

The honeycomb (also termed “cellular structure”) would arguably be one of the most
extensively used energy absorbers in the real world. Honeycomb panels are used in a
variety of applications for absorbing impact energy, their use in automobile bumper
bars perhaps being the most well known. Honeycombs are placed at the front of the
French TGV locomotive to reduce forces transmitted to the conductor’s survival cell
during an impact. Other applications include impact energy absorption in vehicle
accidents and aircraft supply drops.

Honeycomb energy absorbers are made from a variety of materials, predominant


ones being metals such as sheet steel or aluminium alloy, thermoplastics, injection
molded polyolefin, and elastomeric material. Most honeycomb cells are hexagonal in
section, however other shapes are possible such as triangular, square, rhombic or
circular. Wierzbicki (1983) has emphasised the advantages of metal honeycombs for
use as energy absorbers:

Metal honeycombs have long been recognized as an excellent lightweight structural


material due to their strength and energy absorption properties. Hexagonal cell
structures are characterized by a considerable rigidity in shear, high crushing stress,
almost constant crushing force, long stroke, low weight and relative insensitivity to
the overall loss of stability.

Figure 2.2-13(a) shows the three major axes of a typical honeycomb structure. The
literature reports considerable testing for crushing in both the out-of-plane or axial
(x3) direction and in-plane or lateral (x1 and x2) directions. The structural response of
the honeycomb cells governs the global stress-strain response of the honeycomb
structure as a whole. The generalised stress-strain curve for in-plane loading is
shown in Figure 2.2-13(b). The response consists of three main regions. The first
corresponds to the linear- elastic response of the honeycomb as the load is initially
applied and the cell walls bend elastically with small deflections.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 47
x3
x2

x1

Stress
Strain

(a) (b)
Figure 2.2-13 (a) Schematic showing the three major axes of a honeycomb
structure, and (b) Generalised stress-strain curve for laterally loaded
honeycombs.

Once a critical stress is reached the stress levels and remains at a relatively constant
value over a large deflection range. This is referred to as the “plateau” region, and
corresponds to the buckling of the cell walls. This region may be governed by one of
three failure mechanisms of the cell walls: elastic buckling, plastic collapse or brittle
fracture. The first two depend on the slenderness of the cell walls (ratio of cell wall
thickness to length), such that cell walls with small slenderness ratios buckle
elastically, while those with large slenderness ratios collapse plastically,
corresponding to material yield. Brittle fracture occurs for honeycombs having brittle
base material with small critical strains, such that excessive strains build-up in the
cell walls, causing them to fracture.

As the walls buckle they eventually build up upon themselves such that the stress
rises sharply due to the densification of the cell wall material. This corresponds to the
final region of the profile in Figure 2.2-13(b).

The energy absorbing capacity of honeycombs is tuned through varying geometry


parameters such as the cell wall thickness, length, width and height. Lorenzo et al.
(1997) investigated the design of a collapsible honeycomb energy absorber
integrated with the inner door panel of a passenger car. Such a device would absorb
the impact load of a side-on collision, thus increasing protection to occupants. They

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 48
found that increasing the mesh density, or the number of cells per unit area, resulted
in the load-deflection profile becoming more uniform for dynamic loading scenarios.

The dynamic impact crushing behavior of multi-layer honeycomb sandwich panels


has been investigated (Yasui 2000), as both uniform and pyramid types. The uniform
type consists of multiple layers of the same overall length and width stacked upon
each other, whereas in the pyramid type the above-lying layers reduce in length and
width. The pyramid type built from two or three basic panels was observed to have
the better energy absorbing capacity and specific energy (in terms of deformed
mass), while both types proved promising as energy absorbers.

Zhao and Gary (1998) investigated the experimental crushing behavior of aluminium
honeycombs under impact loading. Their work involved the use of the Split
Hopkinson Pressure Bar (SHPB), a device commonly used to investigate the
dynamic response of energy absorbers under impact loading. The experimental
results showed that only the out-of-plane crushing behavior is affected by the loading
rate.

A comprehensive treatment of the properties of honeycomb structures and other


cellular solids is provided in the book by Gibson and Ashby (1997).

2.2.2.4 Foams

The energy absorption mode of foams is similar to that of honeycombs in that it


involves the crushing of a cellular structure. The cells in foams, however, are several
orders of magnitude smaller than the cells in honeycombs. Furthermore, the cells in
honeycombs are two-dimensional, whereas in foams the cells are three-dimensional.
Foam materials have been used extensively in the packaging industry. They possess
the ability to absorb impact energy and maintain the peak force on the packaged
object below the limit which will cause damage or injury (Zhang and Ashby 1994).
To absorb energy at a near-constant load, the foam must have the correct cell-wall
material and relative density. The relative density is the ratio of the foam density to
the density of the solid material from which it was made. If the relative density is too
low the cells will be crushed before sufficient energy has been absorbed. If the

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 49
relative density is too high the stress will exceed a critical value before sufficient
energy can be absorbed.

There are two predominant types of foams used as energy absorbers: polymer and
metal foams. Polymer foams are commonly made by blowing gas bubbles into a
liquid monomer or hot polymer. These bubbles are allowed to grow and stabilise
before the polymer solidifies. Metal foams are commonly made by mixing organic
beads into the metal melt in an inert atmosphere, followed by cooling and
solidification of the metal during which the carbon burns off to leave a cellular
matrix.

Metal foams are a relatively new class of material and have great potential for use as
energy-absorbing structures. The most commonly used metal foams at present are
those made from aluminium or nickel. A recent summary on metal foams is provided
by Ashby et al. (2000). Aluminium foam has been identified as one of the highest
efficiency energy absorbers since its load-deflection curve closely approximates the
square profile which represents maximum energy absorption efficiency (Yu-Hallada
et al. 1998). However, its efficiency and total energy absorption decrease with off-
axis loads. Aluminium foam is also a relatively new material and is therefore
expensive. For further information on aluminium foam as an energy absorber refer to
Lorenzi et al. (1997).

A material that comes close to aluminium foam in terms of energy absorption


capacity is rigid friable polyurethane (PU) foam. This material has relatively low
density, high efficiency, and good moldability for greater design flexibility. Rigid
friable PU foam is however unrecoverable, such that it can only withstand a single
impact. Recoverable (semi-rigid) PU foam has been developed for energy absorbers
requiring light or minimal covering (Yu-Hallada et al. 1998). This has shown to be
an excellent energy absorber, and the fact that it is recoverable has led to it being
used for covering vehicle interior surfaces, and as bumperettes attached to the front
of Vehicle Frontal Protection Systems (Bignell 2004). For further information on
polyurethane as an energy absorber refer to Kelly and Rucker (1997), while for
further information on foams in general as energy absorbers refer to the book by
Gibson and Ashby (1997).

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 50
Significant research has been carried out on the crushing of thin-walled tubes filled
with foams under quasi-static and dynamic loading regimes. The philosophy behind
introducing an ultralight foam metal core is that the rigidity and thus the energy
absorption of the thin-walled member are increased without significantly increasing
its weight. The interaction between the foam and tube walls helps enhance the energy
absorption. This interaction is such that the foam provides constraint when the tube
wall buckles inwardly. The reader is referred to Reid et al. (1986) and Reddy and
Wall (1988) for examples of tubes filled with polyurethane foam, and to Hanssen et
al. (2000a; b) for examples of square and circular tubes filled with aluminium foam.

Metal fillers such as aluminium foam are also introduced into thin-walled tubes in
order to increase their energy absorbed per unit mass, Em. The crushing behavior of
axially compressed columns filled with aluminium foam has been examined (Santosa
& Wierzbicki 1998; Santosa et al. 2000; Hanssen et al. 2000a; b). In some cases the
energy absorbed per unit mass improved with the addition of foam filler, whereas in
other cases the peak crushing force and total energy absorbed increased, however due
to the added weight of the filler, Em did not change significantly. Furthermore,
introducing foam filler tends to reduce the maximum crush distance since the foam
becomes locked between the folds formed in the walls of the deforming column, and
becomes incompressible. Filling corrugated tubes with polyurethane foam and
subjecting them to axial compression was found to produce a more uniform crushing
force (Singace & El-Sobky 1997). Hanssen et al. (2001) found that optimised foam-
filled square aluminium columns displayed smaller cross-section dimensions and
reduced weight compared to traditional non-filled columns, thus allowing reductions
in mass, length and volume using aluminium foam. Furthermore, as discussed in
Section 2.2.2.1, filling an empty tube with foam can beneficially alter the
deformation mode from non-compact to compact (Reid et al. 1986).

Chen and Wierzbicki (2001) investigated the axial crushing of aluminium foam-
filled single and multi-cell columns. Using aluminium foam of varying density, they
found that the average crushing force and specific energy increased with the addition
of aluminium foam. As the foam density increased for a given column section, the
average crushing force also increased due to the interaction between the tube wall

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 51
and foam core. The foam-filled sections generally displayed higher specific energy,
and therefore were more weight-efficient in terms of energy absorption.

Another recent innovative study has been made by Chen et al. (2001). They
investigated the torsional crushing behavior of foam-filled thin-walled square
columns. The motivation of their work was the recognition that, in crashworthiness
studies, less attention has been given to the load bearing capacity of thin-walled
prismatic beams under torsion or combined loading than has been given under axial
loading. They found that introducing aluminium foam filler changed the torsional
collapse mechanism, leading to higher torsional resistance.

2.2.2.5 Isolators

Isolators act similar in principle to shock absorbers in vehicle suspension systems in


that they absorb impact by forcing fluid through an orifice located inside a piston.
Some isolators use hydraulic fluid while others use silicone gels as the energy
absorbing media. Isolators are commonly used as minor impact attenuators in
automobiles, mounted between the bumper and the chassis (Figure 2.2-14).
Hydraulic isolators are relatively efficient however a major disadvantage is that they
are impact velocity sensitive: as the impact velocity increases, the peak load also
increases. At high speeds the force can rise considerably before sufficient
displacement occurs, resulting in forces potentially injurious to occupants being
transmitted through to the vehicle.

Alternative isolator systems have been conceived, such as a type that incorporates a
spring mechanism (Watson 1974). For minor impacts the spring elastically deforms,
while the isolator only begins to absorb energy through hydraulic work once a certain
speed threshold is reached. This allows the isolator to be effective over a larger speed
range. Another isolator principle involves a cylinder containing a gel which is forced
out of the cylinder due to an advancing ram. Such a system is advantageous since it
is velocity insensitive. However, an inherent drawback of isolators designed to
absorb impact energy is that their effectiveness is primarily restricted to external
loads applied down the isolator’s axis.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 52
Figure 2.2-14 Hydraulic isolator mounted to the front of a vehicle
Murray, N. 1994. When it comes to the crunch - the mechanics of car collisions
Singapore: World Scientific Publishing Co. Ptd. Ltd.
Figure 8.6, p. 116.

2.3 Analysis of energy absorbers

The energy absorption response of energy absorbing structures has traditionally been
studied using analytical, empirical and experimental techniques. These techniques
have in recent times been complemented with finite element analysis (FEA). FEA is
a powerful tool, particularly for conducting parametric design studies. A convenient
method of conducting a parametric study is to use a factorial study. This section
briefly considers analytical, empirical and factorial techniques, while the
experimentation used in this thesis will be described in subsequent chapters. FEA
was a primary component of this thesis, and its background is treated separately in
the next section.

2.3.1 Analytical modelling

The development of theoretical models has been commonly used to examine the
energy absorption response of a variety of thin-walled energy absorbers under
different loading conditions. These models typically start off with a simplified
deformation mechanism whose deformation is defined using the “yield line” or
“hinge line” approach. In this approach, a kinematically admissible model of the
local buckle geometry is developed based on observations of the physical collapse
mode. An example of such a model is the inextensional (Type I) basic collapse

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 53
element of an axially crushed square tube, shown in Figure 2.3-1(a). The direction of
applied load is shown by the thick arrows, while the original height of the
undeformed element is 2H, representing the axial length of a single fold in the tube.
The model assumes all plastic deformation to occur at plastic hinges separated by
panels representing the undeformed portions of the tube wall. These panels are
allowed to translate and/ or rotate, provided these motions are kinematically
admissible. The hinge lines are denoted by lines AB, BC, BD and BE in Figure
2.3-1(a). Commonly, the plastic hinges are assumed to have rigid, perfectly plastic
material behaviour, however material strain hardening is sometimes included to
improve correlation of the model with experimental results. The discrete plastic
hinge lines can either be fixed or travel to allow deformation of the buckle. If the
rigid, perfectly plastic idealisation of the material is used, the magnitude of the
bending moment per unit length must equal the fully-plastic bending moment per
unit length for the sheet metal, given as M0 = σ0h2/4, σ0 being the yield stress of the
material, and h the wall thickness of the tube. Once the collapse mode of the local
buckle geometry has been developed, the mean load-deflection response is obtained
by equating the external work to internal work.

The theoretical model for the crushing of square tubes has been improved by
extending the plastic deformation into plastic zones instead of concentrated hinges,
thus allowing the plastic deformation to propagate, as shown in Figure 2.3-1(b).

D
C

A
B

Plastic deformation
occurs in these regions

(a) (b)
Figure 2.3-1 Theoretical model for a quadrant of a square tube under axial
symmetric folding: (a) with concentrated hinges and (b) with plastic zones.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 54
During deformation, the four plane trapezoidal plates move as rigid bodies. A
treatment of the energy dissipation associated with this modified model is provided
in Lu and Yu (2003).

The above approach has been used to develop theoretical expressions for the axial
crushing of both circular and square tubes. The FE model developed in this thesis for
axial loading of straight rectangular tubes was based on a theoretical model for the
axial crushing of square tubes (Abramowicz & Jones 1986). Reid and Reddy (1986)
modified this model and applied it to the axial crushing of tapered rectangular tubes.
This modified model was used to validate the FE model of the tapered rectangular
tubes examined in this thesis. Both the theoretical models for straight and tapered
tubes have also been modified to take account of the effective crushing distance and
strain rate effects for dynamic loading (Abramowicz & Jones 1984; 86; Reid &
Reddy 1986). Such models have been further modified to include the effect of filling
the tubes with foam (Reid et al. 1986). Using the same procedures, theoretical
models have been developed for other thin-walled structures such as the axial
crushing of multi-cell thin-walled columns (Chen & Wierzbicki 2001), and the
bending collapse of box columns (Kecman 1983).

2.3.2 Empirical modelling

An alternative approach to theoretical analysis of the energy absorption response of


thin-walled tubes is to use empirical modelling. This involves applying linear or non-
linear multivariable regression analysis to the results of extensive FE analysis or
experimental testing. Expressions can then be developed relating the energy
absorption response of the tubes to key geometry parameters. These expressions can
then be used to develop guidelines for energy absorption response. Empirical
modelling has been used in many studies for a variety of thin-walled sections and
loading conditions. Of relevance are studies conducted on the axial quasi-static and
dynamic crushing of circular frusta (Mamalis & Johnson 1983; Mamalis et al. 1984;
86a; b), which will be referenced throughout this thesis.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 55
2.3.3 Factorial studies

The energy absorption response of thin-walled tubes can be analysed based on a one-
factor-at-a-time methodology, where the effect of one parameter on the tube
response is investigated by varying this single parameter and keeping all remaining
parameters fixed. Hence, the method provides an estimate of the effect of a single
variable at selected fixed values of the other variables. However, for such an estimate
to be generally meaningful it must be assumed that the effect is the same at other
values of the other variables. If, in reality, the effect is not the same, then the
variables are said to “interact”, and such interaction cannot be detected and estimated
using the one-factor-at-a-time approach to analysing energy absorber response.

An alternative to this more traditional approach to analysing data is factorial design.


Such an approach allows estimation of so-called main and interaction effects. The
“effect” of a factor refers to the change in response of the system due to the factor as
that factor moves from a low to a high value. For instance, the approach allows the
average effect of wall thickness on the mean load to be determined as the wall
thickness is increased from say 1 mm to 2 mm, over all conditions of the other
variables. Such is known as a main effect. An interaction effect is where, for
instance, the effect of wall thickness on the mean load may be greater for one taper
angle than for another taper angle; the wall thickness and taper angle are then said to
“interact”. This is known as a two-factor interaction where two factors interact.
Intuitively, a three-factor interaction arises in which the average of two two-factor
interactions is sought over two values of a third variable. The interaction between
three variables can thus be quantified. For a detailed explanation of factorial design
the reader is referred to Box et al. (1978). Instances where factorial design has been
used for analysing energy absorbers is provided in Nagel and Thambiratnam (2004a)
and Reyes et al. (2002).

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 56
2.4 Finite element modelling of energy absorbers

2.4.1 History

High-end Finite Element (FE) software packages have become an invaluable tool in
the design and development of energy absorbers. Most energy absorbers deform and
fail via highly complex modes which, to say the least, are difficult to accurately
model using traditional analytical techniques. With the use of FE software, designers
can model to a sufficient degree of accuracy the behavior of an energy absorber as it
deforms to absorb energy. This section briefly considers instances where finite
element analysis (FEA) has been used to model particular energy absorbers, and
which provided a guide for the modelling procedures adopted in this thesis.

A host of research projects are documented in the open literature which use FEA to
model energy absorber behavior. A substantial amount of these projects consist of
experimental testing followed by FEA of the energy absorber. Results obtained by
experiment are used to calibrate or “validate” the FE model, to ensure it modelled the
experiment with sufficient accuracy. The numerical model can then be used to
perform parametric studies by changing the loading and geometry of the energy
absorber to determine how various parameters influenced its energy absorption, and
to optimise the absorber’s energy absorption capacity.

During the initial design of energy absorbing structures such as for vehicles, it is
usually cost-effective to use specialised computer based design tools rather than the
larger dynamic FE codes. Such specialised software codes are based on some of the
analytical collapse models discussed in Section 2.3.1, and are used specifically to
predict the collapse mode and energy absorption response for thin-walled tubes of
arbitrary cross-sections. An example of such a code is CRASH CAD (Abramowicz
2003), which is used for the bending and axial collapse of thin-walled energy
absorbing columns in vehicles. Once a desired cross-section has been selected for the
columns, a dynamic FE code can then be used to perform a more detailed analysis of
the crash response of the larger vehicle structure.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 57
FE codes operate using either an explicit or implicit solution method, and the theory
behind the two will be covered in the next section. Explicit FE codes allow
simulation of the highly complex response of thin-walled tubes under dynamic
impact loads. Langseth et al. (1999) used the explicit code LS-DYNA to model the
axial impact of square aluminium tubes, and studied the effect of impact mass and
velocity on the tube response. Taking account of tube symmetry, only one quarter of
the tube specimens needed to be modelled, thus reducing the number of degrees of
freedom and hence the computation time. The FE model, which was validated using
quasi-static and dynamic impact experimental tests, could successfully simulate
inertia effects on the tube’s energy absorption response and deformation mode. The
dynamic elastic-plastic buckling of thin-walled circular and square tubes under high
velocity axial impact has been successfully studied using the explicit code
ABAQUS/Explicit in several studies (see for example Karagiozova et al. 2000;
Karagiozova & Jones 2004). The propagation of elastic-plastic stress waves along
the length of the tube could be simulated, as well as the influence of the impact
velocity and striking mass on the development of the buckling shape. The predictions
for the initiation of buckling showed good agreement with experimental results.

Implicit codes can also be used to simulate the energy absorption response of thin-
walled tubes, however the models must be simplified as much as possible to avoid
excessive computation time. Karagiozova and Jones (2001) used the implicit FE
code ABAQUS/Standard version 5.8 to model the crushing behavior of aluminium
and steel cylindrical tubes under axial impact loading. FEA allowed investigation of
the influence of the material properties, shell geometry, boundary conditions and
loading techniques on the energy absorbed and the deformation mode. Kinematic
strain hardening and self-contact of the tubes during deformation were included in
the simulation. Furthermore, the Cowper-Symonds equation was used to account for
strain rate effects on the response of steel tubes. Since an axisymmetric buckling
mode was assumed, a 2D axisymmetric solid model was used, which consists of a
two-dimensional “slice” of the tube, such that a portion of the tube can be modelled
in one plane, rather than modelling the whole tube. This technique simplifies the
model to be built, and reduces computation time since less degrees of freedom are
used than if the whole tube was modelled. This technique has also been used by

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 58
Aljawi (2000) to compare the experimental and FE analysis of the quasi- static and
dynamic inversion of plastic tubes.

In recent times, high-end FE packages have allowed the material models for
crushable foams to be added to simulations. Numerous studies are published which
use FE modelling to simulate the axial, bending and oblique loading collapse of
metallic thin-walled tubes filled with aluminium foam. Options are also available to
model fracture of the foam using a simple criterion. Such FE models can be used to
develop closed-form solutions of phenomena involving tube crushing. For instance,
Santosa and Wierzbicki (1998) and Santosa et al. (2000) used numerical analysis to
investigate the effect of foam filling of thin-walled metal columns under axial
crushing. Based on their numerical simulations, simple closed-form solutions were
developed to calculate the average crushing force of foam-filled square sections.
Further examples of studies which use FE modelling of aluminium foam-filled tubes
are Chen and Wierzbicki (2001) for axial loading, Chen (2001) for bending collapse,
Chen et al. (2001) for torsional crushing, and Reyes et al. (2004) for oblique loading.

The true capabilities of high-end FE software are demonstrated in the modelling of


Fibre Reinforced Composites (FRCs). Such materials undergo complex deformation
modes when permanently deformed, making them extremely difficult to model
analytically. High-end FE software packages contain parameters which allow FRCs
to be modelled depending on their construction (number of lamina, fibre orientation
etc), and the matrix and fibre properties. The deformation, be it elastic or plastic, can
be subsequently modelled. Beard and Chang (2000) performed experimental and FE
modelling of the crushing of braided composite tubes. A model was proposed for the
crushing behavior and energy absorption characteristics of the tubes and simulated
using ABAQUS, in order to compare the model with experimental data. Predictions
from the model correlated well with the experimental results.

The buckling of multicorner columns has also been investigated using FE code
software packages. Chen and Wierzbicki (2001) used the explicit, non-linear code
PAM-CRASH to simulate the axial crushing of single- and multi-cell columns filled
with aluminium foam. Based on the FE results, closed-form solutions were derived
to calculate the mean crushing strength of the foam-filled sections.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 59
Figure 2.4-1 Deformed profiles of single- and multi-cell columns under a quasi-
static axial load, as predicted using the FE code PAM-CRASH
Reprinted from Thin-Walled Structures, Vol. 39, Chen, W. and Wierzbicki, T.,
Relative merits of single-cell, multi-cell and foam-filled thin-walled structures
in energy absorption, p. 287-306, 2001, with permission from Elsevier.

Figure 2.4-1 shows examples of deformed finite element meshes used to model the
tubes. Compare these with the experimental deformed profiles of the tube shown in
Figure 2.2-6.

Han and Park (1999) performed numerical investigations on the oblique loading of
mild steel square columns using PAM-CRASH. They achieved oblique loading
conditions by axially impacting the column with an initial velocity at an inclined
rigid wall (Figure 2.4-2). The end of the column was constrained to translate along
the column’s axis, while an impact mass was concentrated via a rigid body at the end
of the column. Contact between the column and wall was assumed frictionless.

Figure 2.4-2 FE model of the oblique loading collapse of mild steel square
columns. Symbol Φ is the load angle.
Reprinted from Thin-Walled Structures, Vol. 35, Han, D. C. and Park, S. H.,
Collapse behavior of square thin-walled columns subjected to oblique loads,
p. 167-184, 1999, with permission from Elsevier.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 60
The FE model allowed the effect of various geometry parameters on the mean crush
load to be estimated and compared over a range of load angles. Based on these
results, empirical relations were developed relating the mean crush load under
oblique loading to the load angle.

As a final case, Mamalis et al. (2001) carried out a FE simulation of the axial
compression of metallic thin-walled square frusta. The FE code LS-DYNA was used,
which allowed the deformation mode and energy absorption characteristics such as
the load-deflection curve to be estimated for various geometries. Modelling the frusta
material using true stress-strain data allowed close correlation between simulation
and experimental results. Figure 2.4-3 shows the experimental and simulated
deformation mode progression for one of the frusta geometries.

Figure 2.4-3 Experimental and simulated progressive collapse of a square frusta


(Mamalis et al. 2001).

FE analysis can also be applied to energy absorbing systems, in which one or more
energy absorbing elements are combined with other structural members. An example
is the front crush cans and bumper beam of an automobile. Hanssen et al. (2000c)
studied the dynamic axial impact loading of a car bumper system consisting of crash
boxes filled with aluminium foam. Experiments were used to validate a finite
element model which simulated symmetric and asymmetric loading of the system.
The explicit, non-linear finite element program LS-DYNA was used to model the

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 61
bumper system with existing constitutive models, and it was able to represent the
complete crushing process with sufficient accuracy.

2.4.2 Principles of finite element modelling

In this section the principles involved in finite element modelling of energy


absorbers will be detailed. Application of these principles to the FE modelling
conducted in this thesis will be demonstrated in Chapters 3 and 4.

2.4.2.1 Analysis

2.4.2.1.1 Linearity and nonlinearity

All structural problems in FE modelling are either linear or nonlinear (Hibbit,


Karlsson & Sorensen, Inc. 2002b). A linear analysis is where there is a linear
relationship between the load applied to, and the response of, the structure. A
nonlinear analysis involves a structure whose stiffness changes as it deforms. All
structures in the real world are nonlinear, such that a linear analysis is inadequate for
many structural simulations and is used only as an approximation of the structure’s
true behavior. The progressive plastic deformation of energy absorbers is evidently a
nonlinear process, as such nonlinear behavior was included in the FE modelling of
energy absorbers in this thesis.

Structural mechanics simulations contain three sources of nonlinearity: material


nonlinearity, boundary nonlinearity and geometry nonlinearity. Material nonlinearity
is where the stress-strain relationship of the material becomes nonlinear, such as at
stresses above the yield stress in metals. Strain rate dependent material data, as exists
in impact loading scenarios, is a form of material nonlinearity. Boundary nonlinearity
occurs if the boundary conditions change during the analysis, and geometry
nonlinearity occurs whenever the magnitude of the displacements affects the
response of the structure.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 62
Most FE packages use the Newton-Raphson method to obtain solutions for nonlinear
problems. The solution for a nonlinear analysis cannot be calculated by solving a
single system of equations as for a linear analysis. Rather, the solution is found by
gradually applying the specified loads incrementally over time, thus working towards
the final solution. The simulation is broken into a number of load increments and the
approximate equilibrium solution is found at the end of each load increment. The
sum of all the incremental responses is the approximate solution for the nonlinear
analysis.

2.4.2.1.2 Static and dynamic analysis

In this thesis, the FE codes ABAQUS/Explicit version 6.3 and ABQUS/Standard


version 6.3 were used to model the response of energy absorbers under both quasi-
static and dynamic impact loading. ABAQUS is a suite of highly versatile yet
powerful engineering simulation programs that can simulate problems in such
diverse areas as structural analysis, heat transfer, mass diffusion, coupled thermal-
electrical analyses, acoustics, soil mechanics and piezoelectric analysis. FE codes
operate using dynamic integration operators which are broadly characterised as either
implicit or explicit. Implicit methods use a time increment with no inherent limitation
on it size, this size being determined from accuracy and convergence considerations
(Hibbit, Karlsson & Sorensen, Inc. 2002a). As such, implicit simulations typically
take several orders of magnitude fewer increments than explicit simulations.
However the cost per increment for an implicit simulation is far greater since a
global set of equations must be solved for each increment. Explicit methods, which
ABAQUS/Explicit uses, require a small time increment whose size depends only on
the highest natural frequencies of the model, as will be discussed later. Simulations
generally take on the order of 10,000 to 1,000,000 increments however the
computational cost per increment is relatively small.

Numerical codes such as ABAQUS/Explicit which operate using explicit dynamic


methods are suited for modelling high-speed dynamic events where the solution is
dominated by inertia effects, and which are extremely costly to analyse using implicit
programs. Explicit codes are also suited for unstable post-buckling problems, where
the stiffness of the structure changes drastically as the loads are applied.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 63
ABAQUS/Explicit is well suited for analysing the transient dynamic response of
structures that are subject to impact loads and which undergo complex contact
interactions within the structure. Most of the energy absorbers encountered in the
literature involve these phenomena as they deform to absorb energy either under
quasi-static or impact loading. As such, ABAQUS/Explicit was used to model
nonlinear loading of the energy absorbers analysed in this thesis.

ABAQUS/Explicit uses a central difference integration scheme to integrate the


equations of motion explicitly through time, using the kinematic conditions at one
increment to calculate the kinematic conditions at the next increment. Out-of-balance
forces are propagated as stress waves between neighbouring elements while a state of
dynamic equilibrium is solved at the beginning of each increment. This allows the
nodal accelerations at the beginning of each increment to be determined. The explicit
procedure always uses a diagonal or lumped mass matrix, such that there are no
simultaneous equations to solve. The acceleration of a node is determined only by its
mass and the net force acting on it, such that nodal calculations are very inexpensive.
Knowing the accelerations, the velocity and displacement of each node can be
determined “explicitly” through time. “Explicit” means that the state at the end of the
increment is based solely on the displacement, velocities and accelerations at the
beginning of the increment. Whereas a program using the implicit procedure must
iterate to determine the solution to a nonlinear problem, the explicit procedure allows
the solution to be determined without iterating, by explicitly advancing the kinematic
state from the previous increment.

For the explicit procedure to produce accurate results, the time increments must be
sufficiently small so that the accelerations are nearly constant through an increment.
Analyses typically require many thousands of increments due to their relatively small
size, however each increment is inexpensive since there are no simultaneous
equations to solve, as discussed earlier. The maximum size of a time increment is
determined by the stability limit, ∆tstable which is defined in terms of the highest
frequency in the (undamped) system:

2
∆t stable =
ω max

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 64
The time increment must remain below the stability limit for the state to remain an
accurate representation of the problem, otherwise a numerical instability and
unbounded solution may result. Such instabilities can be avoided by increasing the
stable time increment using mass scaling, or by making the mesh as uniform as
possible. ABAQUS/Explicit uses automatic time incrementation throughout an
analysis such that the stability limit is not exceeded. For further information on time
incrementation, the stability limit and mass scaling in ABAQUS/Explicit, the reader
is referred to Hibbit, Karlsson & Sorensen, Inc. (2002a).

2.4.2.2 Finite elements

The ABAQUS/Explicit code contains a large and diverse library of finite elements
which are used to build a particular model, the major families of which are shown in
Figure 2.4-4.

Shell elements have been used in this thesis to model thin-walled tubes. Shell
elements are suited when the thickness of the modelled part is less than 1/10 of a
typical “characteristic length” of the part, such as the height. Also, shell elements can
be used as opposed to solid elements when the material behaviour across the
thickness of the element is not important.

Figure 2.4-4 Major finite element families available in ABAQUS/Explicit


(Hibbit, Karlsson & Sorensen, Inc. 2002a; b).
Reproduced here with the permission of ABAQUS, Inc. (previously known as
Hibbitt, Karlsson & Sorensen, Inc.)

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 65
Shell elements are designated by the code SiRj where i corresponds to the number of
nodes, and j is an optional parameter corresponding to the number of degrees of
freedom. For example S4R5 is a linear thin-shell, general-purpose element suitable
for geometrically linear behaviour. If the geometry behaviour is nonlinear, then S4R
elements are used which is interpreted as a quadrilateral shell element with four
nodes. Such elements allow transverse shear deformation, making them suitable for
buckling and post-buckling analyses involving large strains.

2.4.2.3 Material models

Though high-end FE packages provide a comprehensive suit of material models,


only one principal material model has been used in this thesis; the metal plasticity
model. The metal plasticity model requires the plastic material properties to be
included as part of the material definition of the physical part to be modelled. Metal
plasticity was used in this thesis together with the linear elasticity model which
defines the Young’s modulus, in-plane Poisson ratio and density, while the principal
plastic property required is the yield strength. Such information allows a material
with perfect plasticity to be defined, where the post-yield stress remains constant at
the yield stress value. Inclusion of true stress-strain data such as from a standard
tensile test allows an isotropic material to be defined, which allows the FE code to
more accurately model the material’s post-yield behaviour. ABAQUS/Explicit can
take account of strain rate-dependent material behaviour using various constitutive
equations. In this thesis the rate-dependent behaviour of mild steel under impact
loading was modelled using the Cowper-Symonds equation as a power law relation
together with the isotropic hardening metal plasticity model. Chapter 3 details the
material model used in the FE analysis of this thesis.

2.5 Summary and need for further research

This literature review has provided a background of the topics relevant to this thesis.
Firstly the topic of impact mechanics was covered. Since the response of thin-walled
energy absorbers under impact loading can differ significantly to that under quasi-
static loading, impact mechanics is essential for an understanding of the physical

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 66
response of energy absorbers under dynamic impact loading. Practical applications of
energy absorbers in the field of crashworthiness were then discussed, followed by
classification of the major groups of energy absorbers. Finally, the use of theoretical,
empirical and numerical methods for simulating energy absorber response was
discussed.

The findings of the literature review show that there is a substantial body of
published information on the energy absorption response of thin-walled tubing, in
particular those of square and circular cross-section. Furthermore, much of this
information relates to the dynamic impact loading of thin-walled tubes, while FEA
has in recent times become a common tool for studying the response of such
structures. However, much of the research on thin-walled tubes has focused on tubes
in which all the side-walls are parallel to the longitudinal axis of the tube (herein
termed “straight” tubes). In comparison, relatively little information has been
published on the energy absorption response of tapered thin-walled tubes, in which
one or more of the side-walls are oriented obliquely to the longitudinal axis.

Studies on rectangular tapered tubes are further limited. As identified in Section


2.2.2.2, rectangular tapered tubes have been considered preferable to straight
rectangular tubes since they are more likely to provide a desirable constant mean
load-deflection response under dynamic loading. Also, tapered rectangular tubes are
capable of withstanding oblique impact loads as effectively as axial loads, and are
less likely to fail by global buckling than straight tubes. In the context of vehicle
crashworthiness the energy absorber is commonly subjected to both axial and oblique
loads, thus tapered rectangular tubes may prove advantageous in such applications.
Nevertheless, to the author’s knowledge, other than the study by Reid and Reddy
(1986) there is no published research on the energy absorption response of tapered
rectangular tubes under oblique loading. Indeed, Lu and Yu (2003, 143) identified
that …there seems to be little study on the dynamic bending of thin-walled tubes.
Furthermore, the study by Reid and Reddy (1986) only considered oblique loading at
100 to the tube’s axis, and the only geometry parameter varied was the number of
tapers (one and two tapers). The relative effect of the various geometry and loading
parameters (load angle, impact velocity) on the oblique loading response of tapered
rectangular tubes remains to be determined.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 67
Furthermore, it remains to be found how tapered rectangular tubes compare with
straight rectangular tubes under axial and oblique loading, for both quasi-static and
dynamic impact conditions. Such is an important question when the benefits of either
structure must be “weighed in the balance” for a particular application. It has been
shown that the dynamic response of straight tubes of square and circular cross-
section is sensitive to inertia effects. However, it appears from the literature review
that there is little or no understanding of the influence of inertia effects on the impact
response of tapered rectangular tubes.

The literature review also shows that, while experimental, empirical and theoretical
techniques have been used to study the response of rectangular and square tapered
tubes, relatively few studies have been reported which use finite element codes to
simulate the energy absorption response of such structures. As has been shown in
this chapter, FEA is a powerful tool for analysing the response of thin-walled energy
absorbers due to its accuracy (via appropriate model validation) and ability to
perform parametric studies. Thus, it would be useful to study the energy absorption
response of tapered thin-walled rectangular tubes using FEA.

Finally, while thin-walled tubes have been used in a variety of energy absorbing
systems, the use of tapered rectangular tubes in such systems has not been widely
published. Therefore, it would be beneficial to study the energy absorption response
of a protective structure which uses tapered rectangular tubes for absorbing impact
energy. Such a study would provide design recommendations for the use of tapered
rectangular tubes in energy absorbing systems such as in crashworthiness
applications.

In the light of these areas requiring further research, the following series of questions
may be asked:

How do tapered thin-walled rectangular tubes respond under impact loading? In


particular, what is the effect of impact velocity and mass on the response of these
tubes? How do these tubes compare under impact scenarios in which the
orientation of applied load is variable? How can the load-deflection, deformation
mode and energy absorption response of such tubes be controlled using their

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 68
geometry parameters for such loading conditions, and which of these parameters
have the most influence on the absorbed energy? Furthermore, how can tapered
rectangular tubes be used effectively in an energy absorbing system? Finally,
how does the energy absorption response of tapered rectangular tubes compare
with that of straight rectangular tubes?

The aim of this thesis was to answer these questions.

This thesis studies the impact behaviour of tapered thin-walled rectangular tubes
under axial and oblique loading, for both quasi-static and dynamic impact
conditions. In particular, the effect of various loading and geometry parameters
on the energy absorption and deformation response of these tubes is examined.
The loading variables include the impact mass, velocity and duration, while the
relative effect of the various geometry parameters on the energy absorption
response is quantified. The response of the tapered tubes is also compared with
that of straight rectangular tubes. Such knowledge will lead to design guidelines
which allow control of the energy absorption response. The energy absorption
response of a protective structure which uses straight and tapered rectangular
tubes for absorbing impact energy is studied. Such a study provides design
recommendations for the use of straight and tapered rectangular tubes in energy
absorbing systems such as for crashworthiness applications. Finite element
analysis is used as the primary method of studying energy absorber response,
coupled with experimental, theoretical and empirical techniques.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 69
3 Development and Validation of the Finite
Element Model for Axial Loading

3.1 Introduction

For a finite element model to be sufficiently accurate it must first be validated, and
this is commonly accomplished by comparing the predictions of the model with
those of a previously established solution. In this thesis the finite element models for
the axial and oblique loading of straight and tapered rectangular tubes were validated
using existing theoretical and numerical models available in the open literature, as
well as a series of experimental tests conducted as part of this thesis. The outcome of
the validation process were finite element models which could predict with sufficient
accuracy the response of straight and tapered rectangular tubes under axial and
oblique loading, for both quasi-static and dynamic loading regimes. Such validation
of the tubes was achieved for variations in their wall thickness, taper angle, number
of tapered sides, and impact mass and velocity for dynamic loading. These finite
element models could then be used to perform parametric analyses on the tubes.

This chapter describes the development and validation of the finite element model
for axial loading, while Chapter 4 presents the model used for oblique (off-axis)
loading. As described in Chapter 2, the codes ABAQUS/Explicit version 6.3 and
ABAQUS/Standard version 6.3 were used for all finite element modelling in this
thesis. These codes were used in conjunction with the pre- and post-processor
ABAQUS/CAE version 6.3, which provides the facilities to build the finite element
model and view the output results in graphical format. The modelling techniques
adopted in this thesis followed those recommended for analysing the crushing of
thin-walled tubes with the ABAQUS codes (Hibbitt, Karlsson & Sorensen, Inc.
2002a). The result is a validated finite element model which can be used to perform a
parametric analysis on the axial loading of straight and tapered tubes having a variety
of geometries, for both quasi-static and dynamic loading conditions.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 70
3.2 Development of the finite element model for axial
loading

In this section the development of the finite element model used for axial loading is
described by addressing key aspects of the model, namely geometry, finite element
mesh, loading, interaction and boundary conditions, and finally material properties.
The existing theoretical and numerical models used for validation will be described
in the next section.

3.2.1 Geometry and finite element mesh

In this thesis the axial quasi-static and dynamic impact energy absorption response of
straight and tapered rectangular tubes having two (double-tapered), three (triple-
tapered) and four (frusta) tapered sides, were compared. The tube geometries are
illustrated in Figure 3.2-1. In the case of the triple-tapered tube, the longitudinal axis
refers to the line perpendicular to the flat surface of the base or top of the tube, as
shown in Figure 3.2-1. Any loading that is not normal to this surface is treated as
oblique.

θ
θ

Straight Double taper Triple taper Frusta


(four tapered sides)

Figure 3.2-1 Geometries of the straight and tapered rectangular tubes.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 71
Figure 3.2-2 shows the mesh and loading arrangement used to model each tube. Due
to symmetry of the geometry, loading and deformation modes, only one quarter of
each of the straight tubes, double-tapered tubes and frusta were modelled by applying
suitable boundary conditions along each unloaded edge of the quarter model. Such a
modelling technique serves to reduce computation time and has been used in
previous studies on the axial loading of thin-walled square columns (Karagiozova &
Jones 2004; Langseth et al. 1999). In the case of the double- and triple-tapered tubes
the taper angle was obtained by varying the base width, in accordance with the
existing models used for validation.

Fully fixed
rigid body

y
Symmetry boundary
conditions applied along z x
each unloaded edge for
the frusta, straight and
double-tapered tubes

Rigid body and


tube translate in
y direction

Base width varied to


Base of tube fully fixed obtain taper angle
to rigid body for double- and
triple-tapered tubes

Quarter model of Quarter model of


straight tube double-tapered tube

Figure 3.2-2 Finite element mesh and loading arrangement for straight and
tapered tubes.

Each tube was modelled using shell elements of designation S4R, which is
interpreted as a quadrilateral element with four nodes, suitable for large strain
analyses. Five integration points were used through the shell thickness to model
bending. The tubes were modelled with rounded rather than perpendicular corners
since this resulted in closer correlation with the existing models, and after
convergence an element size of 5 mm was found to produce suitable correlation.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 72
Since each tube deformed via the symmetric collapse mode required by the existing
models, no initial imperfections were required to perturb the mesh.

3.2.2 Loading, interaction and boundary conditions

For both quasi-static and dynamic loading the base of each tube (the large end for
tapered tubes) was fully fixed to a supporting rigid body using a tied constraint. A
tied constraint is a multi-point constraint (MPC) which makes the translational
motion as well as all other active degrees of freedom for the nodes on the base edge
of the tube the same as that of the supporting rigid body. This rigid body was
modelled as an analytical rigid surface. Analytical rigid surfaces are efficient
computation-wise for contact simulations since they are not element based rigid
surfaces and do not require the calculation of stresses and strains. The supporting
rigid body was constrained to translate vertically such that the free end of the tube
was axially crushed onto a fully fixed rigid body which was also modelled as an
analytical rigid surface. A reference node was defined for each rigid body, and the
motion of the rigid body as a whole could then be controlled by assigning the
boundary conditions to this reference node.

A FREQUENCY linear perturbation analysis step, provided in ABAQUS/Standard,


was used to determine the step time for the quasi-static nonlinear crushing analysis.
This is recommended when using the ABAQUS/Explicit code to perform quasi-static
nonlinear analyses (Hibbitt, Karlsson & Sorensen, Inc. 2002a). A step time which is
too small causes the simulation to occur too quickly resulting in localised rather than
structural deformation. On the other hand a step time which is too large results in
unacceptably large CPU time. Since in a quasi-static analysis the lowest mode of the
structure usually dominates the response, the frequency analysis allowed the
frequency of this mode to be obtained for each tube. The corresponding period
determines the lower limit on the step time, which was made 10 times this period as
recommended in the ABAQUS/Explicit manual, to ensure structural deformation
was modelled, which is characteristic of a quasi-static analysis. The step time for all
quasi-static and dynamic analyses was 30 ms. In an explicit analysis the step time for
dynamic impact simulations is a measure of the “real time” of the impact event. The

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 73
step time for the impact simulations conducted in this thesis was found to coincide
with the impact duration of straight and tapered tubes of similar dimensions reported
by Reid and Reddy (1986).

Quasi-static loading was simulated by constraining the supporting rigid body to


translate over a pre-defined displacement. The motion of the rigid body was defined
using an amplitude curve via the AMPLITUDE option in ABAQUS/Explicit, which
moved the rigid body over a time duration equal to the total step time for the
nonlinear analysis. To ensure an accurate and efficient quasi-static analysis the
motion of this rigid body was controlled using the SMOOTH STEP sub-option in
ABAQUS/Explicit. This ensures the rigid body has a smooth motion rather than
sudden, jerky movement which can induce noisy and inaccurate solutions (Hibbitt,
Karlsson & Sorensen, Inc. 2002a). Figure 3.2-3 shows the amplitude curve
describing the motion of the rigid body, with zero gradient at the start and finish of
the step time to ensure smooth motion.

1
Displacement amplitude

Time Step time

Figure 3.2-3 Displacement curve for the supporting rigid body.

Dynamic loading was simulated by applying a point mass element and initial
velocity, V to the supporting rigid body such that it translated to axially impact the
tube into the fully fixed rigid body. Such a modelling approach was adopted to

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 74
simulate experimental impact loading of straight and tapered tubes presented in one
of the studies used for validation of the finite element model (Reid & Reddy 1986).
Furthermore, it was found that impacting a fixed tube with a moving rigid body had
no significant effect on the response.

For both quasi-static and dynamic loading, contact within each tube as the walls
buckle under an applied load was modelled using the 3D self-contact interaction
option available in ABAQUS/Explicit. Self-contact between the tube walls during
collapse, and surface-to-surface contact between each rigid surface and the tube were
defined using the finite sliding “penalty” based contact algorithm with contact pairs
and “hard” contact. All surface-to-surface contact was of the “master-slave” type. In
accordance with the existing models, all surface contact in the finite element models
for the straight and double-tapered tubes was treated as frictionless, while the
Coulomb friction coefficient for all contact in the model of the frusta was 0.2.

To ensure a suitable quasi-static and dynamic response is obtained when using


explicit time integration, the finite element model must meet certain criteria (Hibbit,
Karlsson & Sorensen, Inc. 2002a). In the first place, for the quasi-static analyses the
kinetic energy of the model should be no greater than a few percent of the internal
energy to indicate an acceptable quasi-static solution. The reason for this is the
loading rigid body has no mass and so the only kinetic energy is due to the motion of
the tube. The maximum kinetic energy was measured in a trial simulation and was
found to be sufficiently small compared with the internal energy. Furthermore, since
the quasi-static loading is smooth, the internal energy- and kinetic energy-time
profiles should also be smooth, since highly oscillatory behaviour is an indication of
significant plasticity which could affect the results since the model contains a
plasticity material model. Both the kinetic and internal energy-time profiles were
observed to be smooth. For both the quasi-static and dynamic analyses performed in
ABAQUS/Explicit, the ratio of artificial strain energy to internal energy was less
than 5 %, indicating that hourglassing was not a problem in the model (Hibbitt,
Karlsson & Sorensen, Inc. 2002a).

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 75
3.2.3 Material models

The finite element model of the frusta used the same material model as that in the
existing numerical model. For the finite element model of the straight and double-
tapered tubes, however, there was no existing material model and thus one was
developed for use in this thesis. A tensile test was carried out in accordance with
Australian Standard AS1391-1991 (Standards Australia 1991) on 2 mm thick
specimens of mild steel of a similar grade to that used in the existing theoretical
models for the straight and double-tapered tubes. From this test an isotropic plasticity
material model was developed to allow the post-yield behaviour of the steel to be
accurately defined in the finite element model. Figure 3.2-4 shows the apparatus used
for the tensile test, which was carried out on a universal testing machine.

Figure 3.2-4 Tensile test apparatus.

The true static stress-strain curve of the steel was obtained (Figure 3.2-5), from
which approximated true stress-plastic strain data points were used in the finite
element models, as shown in Table 3.2-1. The steel had a yield strength of σy = 304
MPa, Young’s modulus, E = 207 GPa and Poisson ratio, υ = 0.3. The steel was
assumed to have only isotropic strain hardening, and strain rate effects on the yield
strength of the steel were neglected due to the relatively low overall average strain
rate used in the tensile tests.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 76
500
450
400
350
True stress (MPa) 300
250
200
150
100
50
0
0 0.05 0.1 0.15 0.2
True strain

Figure 3.2-5 True static stress-strain curve for mild steel.

Table 3.2-1 True stress-plastic strain data points used for steel in the FE model.

σt (N/mm2) 304.6 344.19 385.51 424.88 450.39 470.28


εp 0 0.0244 0.0485 0.0951 0.1384 0.1910

For dynamic loading the effect of strain rate on the steel was included in the finite
element model using the Cowper-Symonds constitutive equation given by:

q
σ 
ε& p = D d − 1 for σ d ≥ σ s (3.2-1)
σs 

where σ d is the dynamic flow stress at a uniaxial plastic strain rate ε& p , σ s is the

associated static flow stress, and the constants D and q are material parameters.
Equation 3.2-1 is an overstress power law and was incorporated into the finite
element model using the RATE DEPENDENT option. The parameter values were D
= 6844 s-1 and q = 3.91, as used in previous studies for the dynamic axial crushing of
mild steel tubes (Abramowicz & Jones 1984; Abramowicz & Jones 1986; Reid et al.
1986; Reid & Reddy 1986). These particular parameter values have been adopted to
fit the experimental data for the ultimate stress of steel specimens (Campbell &
Cooper 1966), and thus have been found appropriate for use in crashworthiness

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 77
design calculations which involve large strains, such as the present problem.
Furthermore, finite element simulations conducted as part of this thesis showed that
the strains during deformation of the tubes were in the vicinity of those
corresponding to the static ultimate tensile stress for the grade of mild steel used.
Therefore, the above-mentioned values of D and q were used in this thesis for the
simulation of dynamic impact loading.

3.3 Existing models for the axial crushing of straight and


tapered tubes

3.3.1 Theoretical model for straight and double-tapered tubes

The finite element model for the axial loading of straight and double-tapered
rectangular tubes was validated using existing theoretical models which predict the
axial mean load-deflection response for these tubes. The principles used to develop
these models were described in Chapter 2. The quasi-static mean load, Pm for straight
rectangular tubes was obtained using the expression

Pm / M 0 = 52.22(c / h )
1/ 3
(3.3-1)

proposed by Abramowicz and Jones (1986) for a rigid-plastic column of side length c
and wall thickness h, undergoing the symmetric collapse mode, i.e. Type I as defined
in Reid et al. (1986) and Abramowicz and Jones (1984), and described in Chapter 2.
Here, M0 = σ0h2/4 is the fully-plastic bending moment per unit length for the sheet
metal, σ0 being the yield stress of the material. The effects of strain hardening can be
accounted for by using the average of the yield and ultimate stresses as the value for
σ0 (Langseth & Hopperstad 1996; Reyes et al. 2002). Strictly speaking, Equation 3.3-
1 is only applicable to square tubes, however it has also been found to produce
reasonable results for rectangular tubes (Reid & Reddy 1986). Equation 3.3-1 takes
into account the effective crushing distance in which, due to material strain
hardening and the finite thickness of the tube wall, the folds in a straight tube do not

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 78
flatten completely. This leads to a reduction in the maximum crush distance and a
subsequent increase in the mean load. The effective crushing distance was
approximated by assuming the axial deformation, δ of a fold to be only 73% of the
axial length, 2H of a folding unit, in accordance with Abramowicz and Jones (1986).

The finite element model for the axial crushing of rectangular double-tapered tubes
was validated using a theoretical model which predicts the mean load for tapered
tubes during their progressive buckling (or folding) phase (Reid & Reddy 1986).
Based on an analysis for the crushing of straight tubes and verified using
experiments, the theoretical model treats the behaviour of a tapered tube as a series
of uniform rectangular sectioned tubes whose peripheral lengths increase from one
fold mechanism to the next down the length of the tube. The tube is assumed to
deform via the symmetric collapse mode, with adjacent sides undergoing lateral
deformations which are out of phase. Based on the principle of virtual work, the
mean load and compression for a rectangular tapered tube under axial quasi-static
loading can be respectively expressed by

Pmi = 4M 0 [9.63 ci −1 / h + π tan θ ] (3.3-2)

and

δ i = 2 H mi (3.3-3)

where M0 is the fully-plastic bending moment per unit length defined earlier. In
Equation 3.3-2 and Equation 3.3-3, ci-1 is the peripheral length of a tube quadrant
(sum of half width and half breadth for a rectangular section), h is the wall thickness,
θ is the taper angle, δi is the deformation of a single fold in the tube, and 2Hmi is the
axial length of a single fold. The above analysis does not take account of the
effective crushing distance. Nevertheless, it has been shown that this correction is not
necessary in the case of tapered tubes because almost complete flattening of the folds
can be accommodated due to the presence of the taper (Reid & Reddy 1986).

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 79
The theoretical dynamic mean crush load-deflection response for straight and
double-tapered rectangular tubes was obtained by considering the effects of strain
rate as outlined in Reid and Reddy (1986). This theoretical response was used to
validate the finite element model of the dynamic axial loading of these tubes. The
ratio of the dynamic to quasi-static mean crush loads for identical, straight tubes of
mild steel can be expressed as

Pmd / Pms = 1 + (ε& / 6844)1 / 3.91 (3.3-4)

where 3.91 and 6844 s-1 are material constants. The effective strain rate, ε& , in a fold
unit deforming at a velocity V was calculated following Abramowicz and Jones
(1984), taking into account the effective crushing distance in the case of straight
tubes, however assuming complete flattening for tapered tubes. The velocity of
deformation of a particular fold was assumed constant at its initial value during
deformation of the fold, in accordance with Reid and Reddy (1986).

3.3.2 Numerical model for frusta

The finite element model for frusta was validated by comparing the predicted
response with that of Mamalis et al. (2001), in which the explicit finite element code
LS-DYNA was used to simulate the axial collapse of steel thin-walled square frusta
under impact loading. In the cited study four frusta specimens were analysed, with
variations in impact velocity, wall thickness and taper angle. The latter was adjusted
by varying the cross-section dimensions of the top (small end) and base (large end)
of the frusta. The impact mass was 60 kg, while the impact velocity ranged from 6.05
to 9.25 m/s. Since no previous results were available to validate the model of the
triple-tapered tube, it was modelled using the same material, mesh and contact and
boundary conditions as the straight and double-tapered tubes.

3.4 Validation of the finite element model for axial loading

The finite element model was validated by comparing the predicted tube response
with that of the existing models for a range of input parameters under both quasi-
static and dynamic loading. The input parameters were wall thickness, h, taper angle,

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 80
θ, impact velocity, V and impact mass, M. Each geometry parameter is illustrated in
Figure 3.4-1 and the values used are summarised in Table 3.4-1. All the taper angles
are equal and are denoted by θ, for the double- and triple-tapered tubes and frusta.

Wt
Bt y

z x

L θ

Bb
Wb

Figure 3.4-1 Model geometry for the (l) straight and (r) tapered tubes.

Table 3.4-1 Tube dimensions used for validating the finite element model.

Geometry θ L Wb Bb Wt Bt h
(deg) (mm) (mm) (mm) (mm) (mm) (mm)
Straight 300 50 100 50 100 1.5, 2, 2.5
Double taper 5 300 102.49 100 50 100 1.5, 2, 2.5
10 300 155.8 100 50 100 1.5, 2, 2.5
15 300 210.77 100 50 100 1.5, 2, 2.5
20 300 268.38 100 50 100 1.5, 2, 2.5
Triple taper 5 300 76.25 152.49 50 100 1.5, 2, 2.5
10 300 102.9 205.8 50 100 1.5, 2, 2.5
15 300 130.38 260.77 50 100 1.5, 2, 2.5
20 300 159.19 318.38 50 100 1.5, 2, 2.5
Frusta 5 127 51.9 50 35.6 34.5 0.97
7.5 127 59.1 58.5 36.4 35.7 1.47
10 127 57.2 55.8 27.5 26.5 1.6
14 127 56.5 56.8 11.6 11.7 1.52

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 81
3.4.1 Validation results for the straight tube

Of interest for validation were the load-deflection and mean load-deflection


response, and the deformation mode. Figure 3.4-2 shows the quasi-static load-
deflection (Fs-d) and mean load-deflection (Fms-d) curves for the straight tube with a
wall thickness of 1.5 mm. The finite element response consists of an initial peak in
the mean load at the onset of crushing, after which the load quickly decreases to
reach a steady value. This is due to the effect of the initial peak crush load predicted
by the finite element model.

160
140
120
100
Load (kN)

80
60
40
20
0
0 25 50 75 100 125 150 175 200 225 250
Deflection (mm)

Finite element model Theory

(a) Load-deflection response


100
90
80
70
Mean load (kN)

60
50
40
30
20
10
0
0 25 50 75 100 125 150 175 200 225 250
Deflection (mm)

Finite element model Theory

(b) Mean-load deflection response


Figure 3.4-2 Comparison of results for the straight tube under quasi-static axial
loading (here h = 1.5 mm).

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 82
The finite element load-deflection curve is characterised by oscillating local
maximum and minimum loads as the folds are formed. Towards the end of the crush
cycle the load increases abruptly as the first (folding) phase terminates, and the
deformation proceeds into the second phase consisting of fold consolidation. Close
correlation of the results was also obtained for other values of wall thickness.

Figure 3.4-3 compares the results for the straight tube under dynamic loading,
showing the dynamic load-deflection (Fd-d) and mean load-deflection (Fmd-d) curves.

200
180
160
Dynamic load (kN)

140
120
100
80
60
40
20
0
0 25 50 75 100 125 150 175 200 225
Deflection (mm)

Finite element model Theory

(a) Dynamic load-deflection response


120
110
Dynamic mean load (kN)

100
90
80
70
60
50
40
30
20
10
0
0 25 50 75 100 125 150 175 200 225
Deflection (mm)

Finite element model Theory

(b) Dynamic mean load-deflection response


Figure 3.4-3 Comparison of results for the straight tube under dynamic axial
loading (here h = 1.5 mm, V = 15 m/s and M = 90 kg).

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 83
The response in Figure 3.4-3 corresponds to an impact velocity, V of 15 m/s, an
impact mass, M of 90 kg, and a wall thickness of 1.5 mm. The loading parameters
were chosen such that they would provide sufficient kinetic energy to crush the tube.
Close correlation can be observed, as was also obtained for other values of wall
thickness, impact velocity and impact mass.

By comparing the quasi-static and dynamic results presented in Figure 3.4-2 and
Figure 3.4-3, it is evident that the loads predicted by the finite element model are
higher under dynamic loading than under quasi-static loading. One reason for this is
due to strain rate effects, however it may also be due to inertia effects and the
associated higher input (kinetic) energy. The dynamic response under axial loading
will be further investigated in the parametric analysis presented in Chapter 6.
Furthermore, it can be seen from Figure 3.4-2 and Figure 3.4-3 that the theoretical
quasi-static mean load is constant with deflection, whereas the theoretical dynamic
mean load decreases with deflection. This is due to the diminishing influence of
strain rate as the velocity of the tube reduces (Reid & Reddy 1986), and this effect is
generally well predicted by the finite element model.

3.4.2 Validation results for the double-tapered tube

Figure 3.4-4 shows the quasi-static load-deflection and mean load-deflection curves
for the double-tapered tube with a wall thickness of 1.5 mm and a taper angle of 100.
As for the straight tube, the initial peak crush load predicted by the finite element
model causes an initial peak in the mean load, however this effect is not as great as
for the straight tube. This may be due to the presence of the taper, and will be further
investigated in Chapter 5 as part of the parametric analysis pertaining to quasi-static
loading of the tubes. As for straight tubes, Figure 3.4-4 shows that the load-
deflection curve predicted by the finite element model is characterised by oscillating
local maximum and minimum loads as the lobes are formed. Towards the end of the
crush cycle the load increases abruptly as the first (folding) phase terminates, and the
deformation proceeds into the second phase consisting of fold consolidation.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 84
100
90
80
70

Load (kN)
60
50
40
30
20
10
0
0 25 50 75 100 125 150 175 200 225 250
Deflection (mm)

Finite element model Theory

(a) Load-deflection response

40
35
30
Mean load (kN)

25
20
15
10
5
0
0 25 50 75 100 125 150 175 200 225 250
Deflection (mm)

Finite element model Theory

(b) Mean load-deflection response


Figure 3.4-4 Comparison of results for the double-tapered tube under quasi-
static axial loading (here h = 1.5 mm and θ = 100).

For quasi-static loading, whereas for the straight tube the theoretical mean load is
constant throughout the crush cycle, in the case of tapered tubes the mean load
increases with deflection. This is clearly due to the increasing width down the length
of the tube, and hence more material is available per fold for deformation and
subsequent energy absorption. Thus, as the taper angle increases, the mean load
increases at a greater rate with deflection and this was predicted by the finite element
model.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 85
The dynamic response predicted by the finite element model for the double-tapered
tube is shown in Figure 3.4-5 for the same geometry as before, and with an impact
velocity of 15 m/s and an impact mass of 90 kg. Close correlation can be observed,
as wall also the case for other values of each geometry and loading parameter.

140
120
Dynamic load (kN)

100
80
60
40
20
0
0 25 50 75 100 125 150 175 200 225
Deflection (mm)

Finite element model Theory

(a) Dynamic load-deflection response

80
70
Dynamic mean load (kN)

60
50
40
30
20
10
0
0 25 50 75 100 125 150 175 200 225
Deflection (mm)

Finite element model Theory

(b) Dynamic mean load-deflection response


Figure 3.4-5 Comparison of results for the double-tapered tube under dynamic
axial loading (here h = 1.5 mm, V = 15 m/s and M = 90 kg).

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 86
3.4.3 Validation results for the frusta

The available data to validate the finite element model of the frusta were the
maximum crush distance and total energy absorption under dynamic loading. Table
3.4-2 provides a comparison of these parameters obtained by the finite element
model and by the analysis presented in Mamalis et al. (2001). On average the
difference is within ± 10 %.

Table 3.4-2 Comparison of results for the frusta finite element model with those
of Mamalis et al. (2001).

Method of Max. crush distance (mm) Total energy absorption (kJ)


solution (taper angle, deg) (taper angle, deg)
(5 ) (7.5) (10) (14) (5) (7.5) (10) (14)
FE model 87.97 88.30 84.30 98.56 1.08 2.42 2.40 2.29
Mamalis et al. 89.50 83.50 89.00 90.50 1.04 2.67 2.50 2.07
(2001)

3.4.4 Deformation modes for the straight and tapered tubes

The deformation modes predicted by the finite element model for each of the straight
and double-tapered tubes were similar under both quasi-static and dynamic loading
for the range of impact velocities considered, as required by the theoretical model.
All tubes deformed via the global symmetric (Type I) collapse mode with buckling
initiating at the impacted end (small end in the case of tapered tubes), as was the case
for the existing models. The behaviour was generally compact for all tubes, which,
by definition indicates that the folds were not separated by relatively undeformed
portions of the side-walls. The plastic fold length increased down the length of the
tapered tubes yet was constant in the case of straight tubes, which was consistent
with the theoretical and numerical results. Overall, the good correlation of results
obtained with the validated FE model placed confidence in the subsequent analyses.
The deformation modes of straight and tapered tubes under quasi-static and dynamic
loading are further studied in Chapters 5 and 6.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 87
It should be noted that the existing theoretical and numerical models with which the
finite element model was validated were based on a progressive, symmetric collapse
mode. Thus, the tube dimensions analysed must be such that they produce this
collapse mode under quasi-static loading. This is particularly the case for the straight
tubes, double-tapered tubes and frusta, since they were modelled as quarter-tubes
based on the assumption that deformation is symmetric.

3.5 Summary

As highlighted in the introduction, a finite element model must first be validated for
it to be sufficiently accurate. This chapter detailed the development and validation of
a finite element model for the axial crushing of straight and tapered thin-walled
rectangular tubes under both quasi-static and dynamic loading conditions. The model
showed good correlation with the predictions of existing theoretical and numerical
models. The finite element model can predict the response of the tubes with
sufficient accuracy for variations in their wall thickness, taper angle, number of
tapered sides, and impact mass and velocity. Thus, a robust model has been produced
which can be used to perform a parametric analysis on the axial loading of straight
and tapered rectangular tubes having a variety of geometries, and for different
loading conditions. Such a parametric analysis can then be used as a design tool for
energy absorbers, as will be demonstrated in Chapters 5 and 6.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 88
4 Experimental Testing and Finite Element
Model Validation for Oblique Loading

4.1 Introduction and background

As identified in Chapter 2, loading in real-world crashworthiness applications is


seldom purely axial or bending, rather comprises off-axis or oblique loads. Such
loading causes the energy absorber to deform via a combination of both axial and
global bending (Euler) collapse modes. As such it is important to understand how the
energy absorber responds under oblique loading. For instance, automotive industries
often require that the bumper system of vehicles must endure a load up to a certain
angle to the longitudinal axis, typically 300 (Reyes 2002). The basis for this is that
during an oblique impact the crush boxes are subjected to both axial forces and
moments. The energy absorption is reduced if the crush boxes undergo global
bending as opposed to axial collapse, thus leading to increased bending and axial
forces being transmitted to the rest of the vehicle. As another case, it has been
identified that the main axially loaded beams in vehicle bodies (such as the front
longitudinal beams) are not straight in reality and are subject to oblique impacts, and
hence can be difficult to control against Euler-type buckling (Kecman 1997). Finally,
protection structures such as Roll Over Protection (ROP) systems, and vehicles such
as buses and coaches are largely comprised of tubing sections. Under impact loading
such tubes usually undergo bending collapse via local buckling and can potentially
intrude into the survival space (Kecman 1983). As such the design of energy
absorbing structures for global bending failure as well as progressive axial collapse is
important in many crashworthiness applications.

As discussed in Chapter 2, information in the open literature on the oblique loading


of thin-walled energy absorbers is limited. Nevertheless some useful studies have
been carried out, as described in Section 2.2.2.1 of Chapter 2. Studies on the oblique
loading response of tapered tubes are further limited. However, experimental

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 89
investigations have been carried out on the dynamic oblique impact of single- and
double-tapered thin-walled rectangular tubes (Reid & Reddy 1986), as discussed in
Chapter 2. Furthermore, the energy absorption characteristics of circular frusta made
of composite material has been investigated experimentally under axial and off-axis
loading over a range of 5 to 350 (Karbhari & Chaoling 2003). The tests showed that
the angle of load orientation can significantly affect the crush energy absorption
characteristics of circular frusta, especially as the degree of off-axis load orientation
increases.

In the light of the literature review, it appears that limited studies have been
conducted on the energy absorption response of tapered rectangular tubes under
oblique loading. Furthermore, limited research has been conducted to compare the
response of straight and tapered tubes under oblique loading. This thesis sought to
further investigate and compare the behaviour of straight and tapered thin-walled
rectangular tubes under oblique loads. Specifically, the thesis sought to investigate
the following aspects of the tube behaviour:

(1) The critical load angle(s) at which straight and tapered tubes make the
transition from the axial collapse mode to the bending collapse mode.
(2) How the response (mean load and absorbed energy) of each tube geometry is
affected by the load angle, impact velocity and geometry parameters of the
tubes, such as wall thickness, taper angle, number of tapers, breadth and
length.
(3) How the deformation modes of each tube geometry depend on the load angle.

As for axial loading, a validated finite element model was used to carry out these
investigations. However, a suitable theoretical or numerical model for the oblique
loading of straight and tapered rectangular tubes was not available in the open
literature with which to validate the finite element model. Thus, a comprehensive set
of oblique loading experiments was conducted as part of this thesis, to validate the
finite element model of the straight and tapered tubes. This chapter describes the
testing procedure and results, followed by the development and validation of the
finite element model for oblique loading. The result is a validated finite element
model which can be used in a parametric analysis to further study and compare the

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 90
response of straight and tapered tubes under oblique loading conditions, for
variations in the loading and geometry parameters. The findings of this chapter are
also presented in Nagel and Thambiratnam (2005a).

4.2 Experimental testing of straight and tapered


rectangular tubes under oblique loading

4.2.1 Specimen details and experimental program

It is difficult to perform precision dynamic tests of obliquely loaded thin-walled


tubes (Reyes et al. 2002). As such quasi-static testing was used to validate the FE
model for obliquely loaded straight and tapered rectangular tubes. Furthermore, it
was found in a previous study that ABAQUS/Explicit suitably takes account of the
strain rate and inertia effects associated with the impact loading of such structures
(Nagel & Thambiratnam 2004a).

Quasi-static compression tests can be conveniently performed using a standard


universal testing machine. Quasi-static tests are convenient for analysing energy
absorber response since they allow continuous monitoring of load, displacement and
deformation mode. These results are important since they are necessary for
validation of the finite element models. Furthermore, together, the load and
deflection provide a quantitative measure of the mean load, which is also necessary
for model validation. A screw-type Tinius Olsen universal testing machine (Figure
4.2-1) was used for all quasi-static compression tests conducted in this thesis. For
details on other testing techniques commonly used to investigate energy absorber
response, refer to Lu and Yu (2003).

Both straight and tapered tubes were studied experimentally. To reduce the number
of experiments, tapered tubes having only one tapered side (single-tapered tubes),
and one taper angle, were tested. Tubes with various dimensions could later be
analysed using the validated finite element model.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 91
Figure 4.2-1 Tinius Olsen universal testing machine.

The length of each test specimen was constant at 199 mm, while the average section
dimensions of all specimens for each tube are shown in Table 4.2-1. The dimensions
were chosen based on the experimental constraints (maximum load capacity of the
testing machine, practical size of the test rig), and were typical of those used in
crashworthiness applications such as automobiles. Also, it became difficult to
distinguish between axial and global bending collapse for smaller lengths than those
used. The wall thickness of each specimen was measured prior to testing, and the
variation in thickness relative to the average value was less than 4 %. Initially a
numerical model was used to determine the critical load angle at which the tubes
underwent the transition from axial collapse to bending collapse. Based on the
analysis, load angles of 100 and 300 were chosen for the experiments to achieve
respective axial and global bending collapse modes. To ensure repeatability and
accuracy, three tubes were tested for each combination of geometry and load angle,
giving a total of 12 tests.

Table 4.2-1 Average section dimensions of all specimens used in the oblique
loading experiments.
Wt

Geometry ș L B, Wb, Wt h
ș L = 99 mm (deg) (mm) (mm) (mm)
Straight tube 0 199 48.31, 79.88, 79.88 1.93
Single taper 10 199 48.15, 80, 45.28 1.93

B Wb

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 92
Each specimen was open at both ends, and was manufactured from 2 mm thick hot
rolled mild steel of yield strength, ıy = 304.6 MPa, Young’s modulus, E = 205 GPa,
Poisson ratio, Ȟ = 0.3 and density, ȡ = 7700 kg/m3. This was the same grade of mild
steel used in the finite element model for axial loading described in Chapter 3.
Furthermore, hot rolled mild steel was found to possess suitable ductility and
strength for energy absorption, and was also readily available. The geometry was cut
from sheet steel and bent such that each tube consisted of two sections joined using
continuous welds along the narrow sides. Each specimen was welded at one end
(large end for the tapered tube) to a flat steel plate which was bolted to the upper
cross-head of the universal testing machine via a special rig designed to achieve the
required oblique loading angle. This ensured each specimen was fully fixed at one
end and free at the other before load was applied. Two rigs were used for each of 100
and 300 loading, the details of which are shown in Figure 4.2-2. Each rig consisted of
a horizontal plate which was clamped flush with the underside of the cross-head on
the testing machine. A length of RHS was welded between this plate and a second
plate oriented at the desired load angle, to which the flat steel plate supporting each
specimen was bolted.

Applied load
Cross-head of testing machine

Test rig d

Base
plate

Į Į is tested at 100 and 300


Specimen welded to
base plate
(arrow points to
compression flange,
web faces out of the
page) Load-bearing
surface

Figure 4.2-2 Test rig used for oblique loading.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 93
Before fabrication of the test rig, numerical modelling was used to determine a
suitable size for the RHS to ensure its stiffness was much larger than that of the
specimens. The experimental test procedure was such that each rig could be attached
to and removed from the testing machine, and each specimen bolted to the rigs, to
allow consecutive tests to be easily carried out. For validation of the finite element
model, the load-deflection response of each specimen as it was crushed was obtained
using load recordings from the testing machine, and by measuring the deflection of
the cross-head using a Linear Variable Displacement Transducer (LVDT). Data
acquisition, monitoring and recording was accomplished using the Labtec Notebook
software package. Loading was achieved by lowering the cross-head such that the
specimen was crushed at its free end onto the load-bearing surface of the testing
machine. The specimen was also free to slide over the load-bearing surface. All tests
were conducted at a crosshead speed of 10 mm/min. Figure 4.2-3 shows typical
specimens and the test apparatus used.

(a)

(b) (c)

Figure 4.2-3 (a) Straight and tapered tube specimens, (b) and (c) oblique loading
test apparatus.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 94
4.2.2 Observed deformation modes

All the specimens in the experiments subjected to a load angle of 100 deformed by
progressive buckling initiating at the loaded end, as observed for the quasi-static
axial crushing of single- and double-tapered rectangular mild steel tubes (Reid &
Reddy 1986), and square frusta (Mamalis et al. 2001). Folds in both the straight and
tapered tubes formed via the global symmetric (Type I) collapse mode, with each
group of lobes consisting of two outward forming lobes on opposite sides of the tube,
and two inward forming lobes on the other two opposite sides. The subsequent group
of lobes which formed down the length of the tube were arranged at a direction of
approximately 900 to the previous group. All specimens showed a tendency to slide
sideways slightly in a direction towards the compression flange due to the oblique
nature of the loading, however the deformation generally remained compact.

For all except one of the straight tube specimens, an inward forming lobe developed
at the loaded end of the compression flange at the onset of buckling for each straight
and tapered tube, leading to three lobes being formed down the length of the
compression flange. However, one of the straight tube specimens developed an
outward forming lobe on the compression flange at the onset of buckling, such that
only two lobes formed on this flange.

Figure 4.2-4 shows typical deformed profiles of the straight and tapered tubes when
subjected to a load angle of 100. The profiles correspond to cross-head deflections of
140 mm and 120 mm, respectively. The direction of applied load is towards the
bottom-right of the page. Similar deformation modes were displayed by the other
tubes tested for a 100 load angle.

All the specimens subjected to a 300 load angle underwent global bending collapse.
Initially the tension flange buckled at the loaded end, after which a plastic hinge or
lobe formed on the compression flange close to the fixed end of the tube due to the
transition to a global collapse mode, as shown in Figure 4.2-5 for a deflection of 75
mm.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 95
(a) (b)
Figure 4.2-4 Typical final deformed profiles for the (a) straight and (b) tapered
tube, for a 100 load angle.

(a) (b)

Figure 4.2-5 Typical final deformed profiles for the (a) straight and (b) tapered
tube, for a 300 load angle.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 96
Similar deformed shapes were displayed by the other tube specimens tested for this
load angle. Such response is similar to that observed by Han and Park (1999) for the
oblique impact loading of thin-walled square columns, and by Reid and Reddy
(1986) for the axial quasi-static crushing of straight rectangular tubes. For the present
tests, a single bending mode was exhibited for all specimens, in which the
compression flange buckled inward while the two narrow side-walls buckled
outward to form the lobe, as shown in Figure 4.2-6. A similar mode has been
observed in previous studies on the bending collapse of straight rectangular tubes
(Kecman 1983), and the oblique loading of square tubes (Reyes et al. 2002). For
straight and tapered tubes examined in the present study, the extent of the buckle was
observed to be the same since the same material was used for all specimens and
hence all exhibited similar amounts of strain hardening. The distance from the
middle of one of the outward forming hinges on the narrow side-wall, to the fixed
end of each specimen, was measured and was found to be on average 4 mm larger for
the tapered tubes. This appears to be an effect of the tapered side. Finally, it was
observed that two of the tapered tube specimens slid in a direction away from the
tapered face as they were crushed due to lack of symmetry in the loading.

Compression
flange

TOP VIEW

Figure 4.2-6 Lobe formed near fixed base of each specimen for a 300 load angle.

4.2.3 Energy absorption response

Figure 4.2-7 shows the load-deflection curves for each tube under a 100 load angle.
For this figure and all subsequent figures in this section, all three test repetitions are
shown, while the following identification system is used: S/T-Į-i, where “S” and “T”
stand for straight tube and tapered tube, respectively, Į is the load angle, and i is the
specimen number.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 97
90
80
70
60
Load (kN)
50
40
30
20
10
0
0 20 40 60 80 100 120 140 160
Deflection (mm)

S-10-1 S-10-2 S-10-3

(a)

70
60
50
Load (kN)

40
30
20
10
0
0 20 40 60 80 100 120 140
Deflection (mm)

T-10-1 T-10-2 T-10-3

(b)
Figure 4.2-7 Load-deflection curve for the (a) straight and (b) tapered tubes, for
a 100 load angle (results for all three test specimens are shown).

According to Figure 4.2-7, for a 100 load angle the load for each tube initially
increases rapidly, however unlike for axial loading it reaches a local maximum value
before decreasing and then rapidly increasing to an absolute maximum value. This is
due to the initial buckling of the tension flange followed by the subsequent collapse
of the other sides of the specimen. After the absolute maximum value is reached, the
load drops and fluctuates about a steady state value as the specimen undergoes
progressive buckling. It is evident in Figure 4.2-7(a) that one of the specimens
exhibits a different load-deflection curve to the others, and this was due to the
compression flange initially buckling outward for this specimen as mentioned earlier.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 98
Figure 4.2-8 shows the load-deflection curves for each tube under a 300 load angle.
For each tube, the load initially increases to a local maximum value before
decreasing, corresponding to the initial buckling of the tension flange at the loaded
end. Thereafter the load increases once more as the plastic hinge forms near the base
of the specimen, promoting increased energy absorption.

40
35
30
25
Load (kN)

20
15
10
5
0
0 10 20 30 40 50 60 70 80
Deflection (mm)

S-30-1 S-30-2 S-30-3

(a)

45
40
35
30
Load (kN)

25
20
15
10
5
0
0 10 20 30 40 50 60 70 80
Deflection (mm)

T-30-1 T-30-2 T-30-3

(b)
Figure 4.2-8 Load-deflection curves for the (a) straight and (b) tapered tubes,
for a 300 load angle.

After reaching a maximum value, the load subsequently decreases as the specimen
undergoes global bending collapse, which reduces the energy absorption capacity of
the specimen.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 99
Figure 4.2-9 and Figure 4.2-10 show the mean load-deflection response of each tube
under 100 and 300 load angles, respectively. For each tube the mean load reduces
slightly with deflection when the applied load angle is 100. However, this reduction
is more significant for a 300 load angle. This coincides with the reduced energy
absorption capacity when each tube undergoes global bending collapse.

70
60
50
Mean load (kN)

40
30
20
10
0
0 20 40 60 80 100 120 140
Deflection (mm)

S-10-1 S-10-2 S-10-3

(a)

60

50
Mean load (kN)

40

30

20

10

0
0 20 40 60 80 100 120 140
Deflection (mm)

T-10-1 T-10-2 T-10-3

(b)
Figure 4.2-9 Mean load-deflection curves for the (a) straight and (b) tapered
tubes, for a 100 load angle.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 100
30

25

Mean load (kN)


20

15

10

0
0 10 20 30 40 50 60 70 80
Deflection (mm)

S-30-1 S-30-2 S-30-3

(a)

35
30
Mean load (kN)

25
20
15
10
5
0
0 10 20 30 40 50 60 70 80
Deflection (mm)

T-30-1 T-30-2 T-30-3

(b)
Figure 4.2-10 Mean load-deflection curves for the (a) straight and (b) tapered
tubes, for a 300 load angle.

Figure 4.2-11 summarises the response by showing the initial peak load, Fbs and
mean load, Fms for the straight and tapered tubes as functions of load angle. The
mean load is calculated up to a deflection of 75 mm for each case. From the results
shown in Figure 4.2-7 to Figure 4.2-11, it is obvious that the initial peak and mean
loads decrease significantly with increasing load angle, which will be investigated
further in Chapter 7. Furthermore, according to Figure 4.2-11, for a 100 load angle
the peak load decreases with the introduction of a taper. This will be further
investigated in Chapter 5.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 101
90
80
70
60
Fbs , Fms (kN)

50
40
30
20
10
0
0 10 20 30 40
Load angle, α (deg)
Straight tube (peak load) Tapered tube (peak load)
Straight tube (mean load) Tapered tube (mean load)

Figure 4.2-11 Initial peak load, Fbs and mean load, Fms for each tube geometry
as functions of load angle.

For a 300 load angle, however, the peak load is greater for the tapered tube than for
the straight tube by approximately 17 %, while the mean load is also greater for the
tapered tube by approximately 20 % compared with the straight tube. This may be
due to the different deformation mechanisms between a 100 load angle and a 300 load
angle. For a 100 load angle the deformation mode of each tube is predominantly
axial, whereas for a 300 load angle it is predominantly bending. The mean load
enhancement may also be due to the taper tending to increase the resistance of the
tapered tube to further bending collapse under a higher load angle. Tapered tubes
may thus prove advantageous when energy absorption under oblique loading must be
maximised, as commonly encountered in crashworthiness applications. It also
appears from Figure 4.2-11 that the initial peak and mean load are less sensitive to
load angle for tapered tubes than for straight tubes, even though each tube had the
same cross-sectional dimensions at the fixed end. Hence, tapered tubes may prove
advantageous for energy absorption applications involving oblique loading. This
behaviour will be further examined in Chapter 7.

Overall, the peak load reduces by approximately 58 % and 40 % for the straight and
tapered tubes, respectively, when the load angle is increased from 100 to 300. The

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 102
reduction in mean crush load calculated up to a deflection of 75 mm is approximately
62 % and 52 % for the straight and tapered tubes, respectively. In terms of
repeatability, the mean load for each of the three tests in each combination of
geometry and load angle was within 10 %.

The initial peak load, Fbs and mean load, Fms from all the experiments are
summarised in Table 4.2-2. The mean load was calculated up to a deflection of 75
mm by dividing the energy absorbed by the crush distance. The energy absorbed was
obtained by integrating the crush load with respect to crush distance using the
CUMTRAPZ function in Matlab version 6.5. For the specimens labelled in the first
column of Table 4.2-2, the following identification system is used: S/T-Į-i, where
“S” and “T” stand for straight tube and tapered tube, respectively, Į is the load angle,
and i is the specimen number. The results for the three test repetitions of each
combination of tube geometry and load angle are shown, and display close
repeatability. Therefore, the tests sufficiently represent the oblique loading response
of the tubes, and can be subsequently used for validation of a finite element model.

Table 4.2-2 Summary of experimental results.

Specimen Test h Initial peak load, Fbs Mean load, Fms


number (mm) (kN) (at d = 75 mm)
(kN)
S-10-1 1 1.93 78.22 50.29
S-10-2 2 1.96 79.98 48.04
S-10-3 3 1.96 78.22 49.06
T-10-1 1 1.89 65.19 47.68
T-10-2 2 1.89 65.04 45.98
T-10-3 3 1.98 64.31 45.74
S-30-1 1 1.93 31.79 18.3
S-30-2 2 1.97 33.69 18.81
S-30-3 3 1.90 34.5 18.39
T-30-1 1 1.93 38.82 21.74
T-30-2 2 1.89 39.99 22.48
T-30-3 3 1.90 38.31 22.64

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 103
4.3 Development of the finite element model for oblique
loading

In this section the development of the finite element model used for oblique loading
is described by addressing key aspects of the model, namely geometry, material
properties, finite element mesh, and finally the interactions, loading and boundary
conditions.

4.3.1 Model geometry and material properties

The dimensions used to model the straight and tapered tubes are shown in Table
4.3-1. The average section dimensions of all test specimens for each tube geometry
were rounded to the nearest mm to ensure a more consistent mesh. It was deemed
that this would not significantly affect the results since the maximum variation for
any given dimension relative to its average value was only 0.65 %, corresponding to
dimension W of the straight tube. The length of each tube model was 199 mm, as in
the experiments.

Table 4.3-1 Dimensions used to model the straight and tapered tubes under
oblique loading.

Wt

Tube ș L B, Wb, Wt
ș L geometry (deg) (mm) (mm)
Straight tube 0 199 48, 80, 80
Single taper 10 199 48, 80, 45
B Wb

The wall thickness can have significant influence on the energy absorption response,
as will be studied in Chapter 7, and its value used in the finite element model was
carefully chosen. The wall thickness used in the model for each tube was the average

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 104
measured for all experimental specimens in each combination of tube geometry and
load angle, as shown in Table 4.3-2. Since the experimental specimens were loaded
quasi-statically, the isotropic plasticity material model used in the finite element
model was the same as that defined in Chapter 3 for axial quasi-static loading.

Table 4.3-2 Wall thickness used in the finite element model for oblique loading.

Tube geometry and load angle combination h (average of three specimens)


(mm)
S-10 1.95
T-10 1.92
S-30 1.95
T-30 1.92

4.3.2 Element mesh and initial imperfections

Each specimen was modelled using S4R shell elements to suitably model the
buckling response of the tubes, while 5 integration points were used through the
thickness of each element. Simulations were carried out to determine the proper
mesh density in terms of reduced computation time and correlation with
experimental results. The tube subjected to a 100 load angle was analysed with
several element sizes, and the effect on the initial peak load and the load-deflection
response was recorded. The optimal element size was 5 × 5 mm in terms of reduced
solution time and correlation with the experimental results, and thus was chosen for
the finite element model for oblique loading of the straight and tapered tubes. The
corresponding total number of elements was 2000 and 1760, respectively. This
element size has also been used in previous studies of thin-walled energy absorbers
(Kim & Wierzbicki 2001). The element mesh for each of the straight and tapered
tubes is shown in Figure 4.3-1.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 105
(a) (b)

Figure 4.3-1 Element mesh for the (a) straight tube and (b) tapered tube.

It was observed that the experimental specimens had rounded corners due to the
bending operation used to form the specimens. Analyses were carried out to
determine the influence of wall corner radius on the accuracy of the models
compared with the experimental results. The corners were modelled using a 1-
element wide strip of elements along each edge, diagonal to the four sides of the
tube, as shown in Figure 4.3-2. This allowed the width of the elements along each
strip to be controlled to simulate the corner radius.

Figure 4.3-2 Corner detail in the finite element mesh.

It was found that introducing a rounded corner brings the numerical peak and mean
loads closer to the experimental results. As expected, the mean load does not change
significantly with the introduction of a corner radius, however the initial peak load
does, such that when no radius is used, the peak load is considerably high compared
with the experimental result. Simulations showed that a corner radius of 3 mm was
suitable for use in the model. This value was similar to that measured on the
experimental specimens and gave good correlation with the experimental results

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 106
when used in conjunction with appropriate mesh imperfections, as will be introduced
in the next section.

It was observed that the tapered tube specimens used in the experiments had slight
warping along their tapered side due to fabrication. The depth of the warp was
measured and included in the finite element model, as shown in Figure 4.3-3. This
served to bring the predicted load-deflection response closer to the experimental
response.

1 mm depth down centre


length of tapered side

Figure 4.3-3 Fabrication warp used in finite element mesh for the tapered tube.

Existing studies on the axial and oblique crushing of thin-walled tubes have shown
that the initial peak and mean loads are sensitive to initial geometry imperfections in
the tube wall (Langseth et al. 1999; Reyes et al. 2002). Such studies have achieved
model validation by assuming a sine curve for the variation of the imperfection
magnitude (or “out-of-flatness”) either down the length and width of the column, or
as a single trigger at a discreet location on the tube. For the present tubes, initial
imperfections were introduced into the finite element mesh for each tube modelled
under 100 loading to ensure they deformed in the same modes as in the experiments.
Numerical analyses for the straight tube without initial imperfections resulted in the
same deformation mode as in the experiments, however imperfections were
introduced to improve correlation of the load-deflection response. Analyses of the
tapered tube without imperfections did not produce the experimental deformation
mode, rather resulted in all four sides buckling inward at the start of the crush
process. Since it was difficult to measure the distribution of imperfections on each
specimen, eigenmode imperfections were introduced using an eigenvalue buckling

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 107
analysis in ABAQUS/Standard before the nonlinear crushing analysis performed in
ABAQUS/Explicit, as recommended in the ABAQUS user’s manual for such
analyses (Hibbitt, Karlsson & Sorensen, Inc. 2002a). The buckling analysis, executed
using the BUCKLE linear perturbation step, also allows a smooth postbuckling
response to be obtained for each tube during the nonlinear analysis. Close correlation
was obtained with the experimental deformation modes for 100 loading when the first
eigenmode was used to perturb the mesh for the straight tube, and the second
eigenmode was used for the tapered tube. These eigenmodes are shown in Figure
4.3-4.

(a) (b)

(a) (b)
0
Figure 4.3-4 Imperfection eigenmodes for 10 loading: (a) first eigenmode for
straight tube and (b) second eigenmode for tapered tube.

The IMPERFECTION keyword option was used in the input file for the nonlinear
crushing analysis, to read the nodal positions of each eigenmode profile into the
unseeded tube mesh. The nodal positions were introduced into each mesh as

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 108
imperfections which were a percentage of the shell thickness, being small enough so
as not to unrealistically deform the mesh. Furthermore, the peak and mean loads
were sensitive to imperfection amplitude, such that the amplitude was adjusted to
obtain correlation with the experimental results.

For 300 loading, the introduced eigenmodes had no significant influence on the
nonlinear deformation mode since the specimens underwent global collapse as
opposed to progressive deformation. It was found that the experimental global
collapse mode could be closely simulated by introducing a collapse trigger near the
fixed base on the compression flange of each modelled tube, in the absence of any
eigenmode imperfections. The trigger was located a similar distance from the fixed
end of each tube as the buckle observed in the experiments, while the trigger
geometry is shown in Figure 4.3-5.

(a) (b)

(c)

Figure 4.3-5 Imperfection trigger geometry for 300 loading: (a) straight tube, (b)
tapered tube and (c) trigger detail.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 109
The trigger, which is in the form of a chamfer, was formed by creating a perfect
mesh in the ABAQUS/CAE pre-processor, compiling the input file and adjusting the
coordinates of nodes directly in the input file. The particular trigger geometry shown
in Figure 4.3-5 was chosen to achieve close correlation between the predicted and
experimental peak and mean loads, and deformation profile. It was found that
simulations without the trigger over-predicted the peak and mean loads for the
straight and tapered tubes. Hence, various trigger depths were analysed to determine
their effect on the peak and mean loads, and hence achieve correlation with the
experimental results. Five sets of trigger dimensions were analysed, as shown in
Table 4.3-3.

Table 4.3-3 Trigger dimensions.

Analysis Trigger dimensions


a (mm)
a b
b
1 0 0
2 0.25 0.5
3 0.5 1.0
4 0.75 1.5
5 1.0 2.0

The finite element model of the straight tube without the trigger produced an outward
forming lobe near the fixed based, as opposed to an inward forming lobe observed in
the experiments. Also, the lobe formed about 17 mm further up the tube than the lobe
in the experiments. An inward forming lobe appeared for the analysis of the tapered
tube without a trigger, however the lobe was located approximately 10 mm above the
lobe in the experiments. When a trigger was incorporated in the model, an inward
lobe formed for all trigger depths for both the straight and tapered tubes. Controlling
the position of the trigger along the length of the tube allowed the inward lobe in the
model to form in a similar location as that in the experiments. As expected, the peak
and mean loads reduce when a trigger is introduced, as shown in Figure 4.3-6.
However, both response parameters, and therefore energy absorption, are relatively
independent of trigger depth.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 110
50
45
40

Initial peak load (kN)


35
30
25
20
15
10
5
0
0.0 0.5 1.0 1.5 2.0 2.5
Trigger depth, b (mm)

Straight tube Tapered tube

(a)

25

20
Mean load (kN)

15

10

0
0.0 0.5 1.0 1.5 2.0 2.5
Trigger depth, b (mm)

Straight tube Tapered tube

(b)
Figure 4.3-6 Effect of trigger depth on (a) peak load and (b) mean load.

Furthermore, for each trigger depth, the load for the tapered tube is much higher than
that for the straight tube. This was also observed in the experiments (refer to Section
4.2.3), indicating that the trigger does not reverse this trend. The higher initial peak
and mean loads for the tapered tube compared with the straight tube will be further
investigated in Chapter 7.

The reduction in peak load when a trigger is introduced is evident in the load-
deflection curve for each tube (Figure 4.3-7), however each curve remains relatively
unchanged as trigger depth increases above 0.5 mm.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 111
40
35

Load (kN) 30
25
20
15
10
5
0
0 10 20 30 40 50 60 70 80
Deflection (mm)

b = 0.0 mm b = 0.5 mm b = 1.0 mm b = 1.5 mm b = 2.0 mm

(a)

50
45
40
35
Load (kN)

30
25
20
15
10
5
0
0 10 20 30 40 50 60 70 80
Deflection (mm)

b = 0.0 mm b = 0.5 mm b = 1.0 mm b = 1.5 mm b = 2.0 mm

(b)
Figure 4.3-7 Load-deflection curves for various trigger depths: (a) straight tube,
(b) tapered tube.

In the validated finite element model, a trigger depth, b = 1.0 mm was found to
produce suitable correlation with the experimental results for both straight and
tapered tubes, in terms of peak load, mean load and deformation profile.

4.3.3 Interactions, loading and boundary conditions

The oblique loading arrangement used for the finite element model is shown in
Figure 4.3-8.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 112
F Rigid body translates in
y direction
y Į

x
.
z
d
Nodes along bottom edges
of specimen are “tied” to
Fully fixed rigid the rigid body
body

Figure 4.3-8 Oblique loading model.

The flat steel plate supporting each specimen in the experiments was modelled as an
analytical rigid surface since its properties do not contribute significantly to the
analysis. This rigid body was constrained in all degrees of freedom to support the
tube, while a tied constraint was found sufficient to model the weld fixing each
specimen to its supporting plate. Load was applied to the tube by an analytical rigid
surface. This essentially modelled the load-bearing surface of the testing machine,
and was constrained to translate over a pre-defined displacement using the
AMPLITUDE option, with the SMOOTH STEP sub-option to ensure smooth motion
as in the quasi-static test. To achieve oblique loading, the specimen and its
supporting rigid body was oriented at the required loading angle as in the
experiments.

The step time was determined using a FREQUENCY linear perturbation analysis
step. The period corresponding to the lowest mode of the straight tube was obtained
as a lower limit on the step time, which was made 10 times this period to ensure
structural deformation was modelled, characteristic of a quasi-static analysis. The
analyses were performed with a step time of 0.03 s. However, changing the step time
was found to alter the load-deflection curve. This effect was briefly studied for the
straight tube under an applied 100 load angle, with dimensions as shown in Table
4.3-1, and initial imperfections in the mesh as shown in Figure 4.3-4(a). Figure 4.3-9
shows the response when the step time, T is increased by one order of magnitude.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 113
90
80
70
60
Load (kN)

50
40
30
20
10
0
0 20 40 60 80 100 120 140
Deflection (mm)

Step time = 0.03 s Step time = 0.3 s

(a)
60

50
Mean load (kN)

40

30

20

10

0
0 20 40 60 80 100 120 140
Deflection (mm)

Step time = 0.03 s Step time = 0.3 s

(b)
Figure 4.3-9 Effect of step time on the (a) load-deflection response and (b) mean
load-deflection response.

It can be seen that the load-deflection curve changes when the step time is increased
from 0.03 to 0.3 s, and this is due to an observed change in the deformation mode.
However, the peak and mean crush loads (and hence energy absorption) remain
relatively unchanged, whereas increasing the step time led to an 11-fold increase in
CPU time. Since close correlation with experimental results was achieved with a step
time of 0.03 s as well as an acceptable CPU time, this step time was chosen for all
subsequent quasi-static oblique loading simulations of the straight and tapered tubes.

In this study the step time was reduced to speed up the analysis. Another technique
commonly used to speed up quasi-static analyses performed with explicit time

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 114
integration is mass scaling, which involves increasing or decreasing the mass density
of the elements in the thin-walled absorber (Hibbitt, Karlsson & Sorensen, Inc.
2002a). This technique has been used for modelling the axial crushing of square
aluminium extrusions under impact loads (Langseth et al. 1999), in which it was
found that reducing the step time and absorber density using mass scaling led to a
significant reduction in the mean load over the displacement range considered. In the
present study, however, it was found that adjusting the step time without mass
scaling has no significant effect on the initial peak or mean load. This observed
phenomenon can be attributed to the inertia effects influencing the crush load due to
changes in the density of the absorber when using mass scaling. In this study it was
observed, however, that adjusting the step time without mass scaling changes the
nonlinear deformation mode when eigenmode imperfections are included in the
mesh, and a constant maximum crush distance is used in the nonlinear analysis.
Hence it can be concluded that the deformation mode is dependent on loading rate
when simulating a quasi-static problem with explicit time integration, whereas the
crush load is independent of loading rate when mass scaling is not used in the
analysis. The study highlights the fact that the use of explicit time integration for
modelling quasi-static problems must be made with great care, as has been identified
in other studies dealing with post-buckling phenomena (Langseth et al. 1999).

As well as comparing the energy absorption response, the suitability of the model to
simulate a quasi-static problem was controlled by ensuring the ratio of kinetic energy
to internal energy was small (less than a few percent), since the loading rigid body
has no mass such that the only kinetic energy in the model is due to motion of the
tube. The calculated ratio for the nonlinear analysis was 0.18 %, which is acceptable.
Furthermore, as addressed in Chapter 3, the ratio of artificial strain energy to internal
energy should be no more than 5 % to indicate that hourglassing is not a problem in
the model (Hibbitt, Karlsson & Sorensen, Inc. 2002a). The calculated ratio was 3.43
%, which once again is acceptable. Finally, since the quasi-static loading is smooth,
the internal energy and kinetic energy-time profiles should also be smooth, as highly
oscillatory behavior is an indication of significant plasticity which could affect the
results given the model contains a plasticity material model. Both the kinetic and
internal energy-time profiles were observed to be smooth. Hence, based on these
criteria the finite element model suitably simulates quasi-static loading.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 115
Correct modelling of the contact between the walls of the specimens as they crush is
essential to accurately model their post-buckling response. This contact was
modelled using the 3D self-contact interaction option available in
ABAQUS/Explicit. Self-contact between the tube walls during collapse, and surface-
to-surface contact between each rigid surface and the tube were defined using the
finite sliding “penalty” based contact algorithm with contact pairs and “hard”
contact. All surface-to-surface contact was of the “master-slave” type. The friction
coefficient associated with contact between the tubes and rigid bodies significantly
influenced the magnitude of the crush load. Since this parameter is difficult to
determine accurately under experimental conditions, it was treated as a variable in
the numerical model to achieve correlation with experimental results. Simulations
determined that a friction coefficient of 0.3 was suitable for 100 loading of the
straight and tapered tubes, while for 300 loading values of 0.1 and 0.2 were used for
the straight and tapered tubes, respectively. The lower coefficients for the higher load
angle were indicative of the lower contact forces between the load surface and
specimen during the experiment. The friction coefficient for self-contact within the
specimens was found to have no significant influence on the energy absorption
response, such that a coefficient of 0.1 was used for all combinations of tube
geometry and load angle.

4.4 Validation of the finite element model

As for axial loading, the finite element model for oblique loading was validated by
comparing the predicted load-deflection response, mean load-deflection response and
deformation modes, with the experimental results. The mean load was obtained by
dividing the energy absorbed by the crush distance. The energy absorbed was
calculated by integrating the crush load obtained from the numerical analysis with
respect to crush distance using the CUMTRAPZ function in Matlab version 6.5. This
method provided close approximation of the experimental mean load, and as such
was used for the oblique loading validation. Furthermore, this method has been used
in a previous study on the oblique loading collapse of thin-walled columns (Han &
Park 1999).

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 116
4.4.1 Load-deflection and mean load-deflection response

The load-deflection curves for the validated model of each tube are shown in Figure
4.4-1 and Figure 4.4-2, respectively. In each case the model closely predicts the
experimental peak load however the curves are slightly out of phase, particularly for
the 100 load angle. This may be due to the use of shell elements to model the tubes,
as observed in a previous study on the numerical simulation of thin-walled square
frusta undergoing axial collapse (Mamalis et al. 2001). The model generally under-
predicted the load for the straight tube with a 300 load angle and the tapered tube
with a 100 load angle, resulting in a lower mean load than in the experiments.

90
80
70
60
Load (kN)

50
40
30
20
10
0
0 20 40 60 80 100 120 140
Deflection (mm)

S-10-2 S-10-3 Finite element model

(a)
40
35
30
Load (kN)

25
20
15
10
5
0
0 10 20 30 40 50 60 70 80
Deflection (mm)

S-30-1 S-30-2 S-30-3 Finite element model

(b)
Figure 4.4-1 Experimental vs. numerical load-deflection response for the
straight tube: (a) 100 load angle and (b) 300 load angle.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 117
80
70
Load (kN) 60
50
40
30
20
10
0
0 20 40 60 80 100 120 140
Deflection (mm)

T-10-1 T-10-2 T-10-3 Finite element model

(a)

45
40
35
30
Load (kN)

25
20
15
10
5
0
0 10 20 30 40 50 60 70 80
Deflection (mm)

T-30-1 T-30-2 T-30-3 Finite element model

(b)
Figure 4.4-2 Experimental vs. numerical load-deflection response for the
tapered tube: (a) 100 load angle and (b) 300 load angle.

Figure 4.4-3 and Figure 4.4-4 show the mean load-deflection response predicted by
the model for the straight and tapered tubes, respectively. The straight tube specimen
(S-10-1) which exhibited a differing deformation mode and hence energy absorption
response under 100 loading, as mentioned earlier, was not included in the validation
since it was not duplicated in the experiments.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 118
70
60

Mean load (kN)


50
40
30
20
10
0
0 20 40 60 80 100 120 140
Deflection (mm)

S-10-2 S-10-3 Finite element model

(a)
30

25
Mean load (kN)

20

15

10

0
0 10 20 30 40 50 60 70 80
Deflection (mm)

S-30-1 S-30-2 S-30-3 Finite element model

(b)
Figure 4.4-3 Experimental vs. numerical mean load-deflection response for the
straight tube: (a) 100 load angle and (b) 300 load angle.

The initial peak load is also an important measure of a tube’s deformation response.
Figure 4.4-5 compares the initial peak load predicted by the finite element model for
each tube, with the corresponding experimental results. The experimental results are
the average for all specimens in each combination of tube geometry and load angle.
Overall, the model predicts the initial peak load with sufficient accuracy.
Furthermore, Figure 4.4-5 shows that the initial peak load decreases more for the
straight tube than for the tapered tube as the load angle increases from 10 to 300. This
further indicates the decreased sensitivity of tapered rectangular tubes to the applied
load angle, and hence their advantage for energy absorption applications involving
oblique loading.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 119
60

50
Mean load (kN)
40

30

20

10

0
0 20 40 60 80 100 120 140
Deflection (mm)

T-10-1 T-10-2 T-10-3 Finite element model

(a)
35
30
Mean load (kN)

25
20
15
10
5
0
0 10 20 30 40 50 60 70 80
Deflection (mm)

T-30-1 T-30-2 T-30-3 Finite element model

(b)
Figure 4.4-4 Experimental vs. numerical mean load-deflection response for the
tapered tube: (a) 100 load angle and (b) 300 load angle.

90
80
Initial peak load (kN)

70
60
50
40
30
20
10
0
0 10 20 30 40
Load angle (deg)
Straight tube (Experiment) Straight tube (Finite element model)
Tapered tube (Experiment) Tapered tube (Finite element model)

Figure 4.4-5 Comparison of the predicted and experimental initial peak loads.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 120
Differences arising between the results of the finite element model and the
experiments may partly be due to the presence of the fillet welds in the experimental
specimens. These welds ran down the full length of the narrow sides of each
specimen and were found to influence the deformation profile, particularly under a
300 load angle. However, since the influence of the welds on the mean crush load and
subsequent energy absorption response was not significant, these welds were not
included in the validated models. Furthermore, these welds were neglected due to the
difficulty of accurately quantifying the associated variables such as variations in fillet
height and width as well as material effects in the region of the Heat Affected Zone
(HAZ). Nevertheless, these affects associated with the welds attaching each
specimen to its supporting base plate had no significant influence on the energy
absorption and deformation response of each tube. Another source of variance
between the predicted and experimental results would be residual stresses present in
the walls of the tubes due to the bending and welding operations used to fabricate the
specimens. Since the magnitude and distribution of these stresses are unknown and
difficult to determine, they were not included in the validated model. Despite these
differences, the validated model provides a sufficiently accurate prediction of the
response of the straight and tapered tubes under oblique loading, over the range of
load angles considered in this study.

4.4.2 Deformation modes

The deformation modes predicted by the FE model were also compared with those of
the experimental specimens. Close correlation was achieved using the imperfections
described earlier. Figure 4.4-6 shows the experimental and numerical deformed
profiles for each tube under 100 loading. The formation and hence number of lobes
was accurately predicted in each case, however it can be seen that the predicted
initial lobe formation for the tapered tube differs from in the experiment. This is due
to the presence of the weld in the test specimens, which acts as a stiffening member,
thus causing the top edge on the tapered side of the tube to buckle inwards. Since the
weld was absent in the model, this top edge does not have the stiffening effect due to
the weld, and hence folds in on itself in the model (Figure 4.4-6). This phenomenon
did not have any significant effect on the energy absorption response.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 121
Straight tube (crush distance = 143 mm)

Effect of weld
on initial
deformation

Tapered tube (crush distance = 120 mm)

Figure 4.4-6 Experimental and numerical deformed profiles under 100 loading.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 122
The experimental and numerical deformed profiles for the straight and tapered tubes
under 300 loading are shown in Figure 4.4-7, at a crush distance of 76 mm. The
model accurately predicted the formation of the buckle close to the fixed end.
However, the presence of the weld in the test specimens restricted the progressive
deformation at the loaded end. Since the weld was not included in the finite element
model, the extent of this deformation was slightly over-predicted, as can be seen in
Figure 4.4-8, though this did not significantly affect the energy absorption response.
All the straight and tapered tube specimens subjected to 300 loading experienced
negative curvature of the tension flange close to the fixed end, where the flange
curves inwards to relieve tensile and shear strain as the specimen undergoes global
bending collapse. This response was displayed in the finite element model for each
tube, as shown in Figure 4.4-9 for the straight tube.

Straight tube (crush distance = 76 mm)

Tapered tube (crush distance = 76 mm)

Figure 4.4-7 Experimental and numerical deformed profiles under 300 loading.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 123
Straight tube

Tapered tube

Figure 4.4-8 Progressive deformation at the loaded end for each tube.

Negative
curvature

Experiment Finite element model

Figure 4.4-9 Negative curvature of the straight tube under 300 loading:
comparison of experimental and predicted result.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 124
4.5 Summary

This chapter detailed the development and validation of a finite element model to
simulate the response of straight and tapered thin-walled rectangular tubes under
quasi-static oblique loading. A comprehensive set of experiments has been carried
out to validate the model, for loads applied at 10 and 300 to the axis of the tube.
Three experiments were conducted for each geometry and load angle combination,
and repeatability for the mean load was within 10 %. Close correlation was observed
between the experimental and predicted load-deflection and mean load-deflection
response, and the model was able to simulate the deformation modes with sufficient
accuracy. Overall, the good correlation of results allows the FE model to be used in a
parametric analysis to further study and compare the response of straight and tapered
rectangular tubes under oblique loading conditions, for variations in the loading and
geometry parameters. In turn, this will enable control of the energy absorption
response for the design of such energy absorbers under oblique loading.

Several practical modelling techniques arose during development and validation of


the FE model for oblique loading. For instance, closer correlation with the
experimental results was achieved by modelling the tubes with rounded rather than
perpendicular corners. The initial peak load in particular was affected by the
presence of a corner radius. Therefore, for numerical studies on thin-walled tubes
based on experimental tests, the author recommends that the effect of rounding the
tube corners be studied, particularly for oblique loading studies.

The experimental results show that the initial peak and mean load are less sensitive
to load angle for tapered tubes than for straight tubes. Also, the mean load of the
tapered tubes is greater than for straight tubes of equivalent cross-section
dimensions, when the response is dominated by global collapse. Thus, tapering a
rectangular tube enhances the energy absorption capacity under oblique loading.
Such structures may therefore be advantageous in applications where oblique loading
conditions are expected, such as vehicle crashworthiness. Using the validated FE
model developed in this chapter, the oblique loading response of straight and tapered
rectangular tubes will be further investigated in Chapter 7.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 125
5 Quasi-Static Response of Straight and
Tapered Rectangular Tubes under Axial
Loading

5.1 Introduction

The response of energy absorbers can be heavily influenced by their geometry


parameters, such as cross-section dimensions (width and breadth), length and wall
thickness. Investigating these parameters by means of laboratory tests can be quite
expensive and time-consuming, as such it was desirable to utilise the advantages of
finite element techniques to carry out a parametric study of the straight and tapered
tubes. This chapter and the following three chapters compare the quasi-static and
dynamic energy absorption response of straight and tapered thin-walled rectangular
tubes under both axial and oblique loading. These regimes typify the type of loading
such energy absorbers experience in real-world applications. The calibrated finite
element models presented in Chapters 3 and 4 are used to determine how the
geometry parameters influence the response of the tubes. Such information is useful
in the design of energy absorbers where a desired deformation mode or energy
absorption response is required.

This chapter presents the quasi-static axial behaviour of straight and tapered
rectangular tubes. A quasi-static analysis is commonly used as a starting point for
energy absorber design, as it allows the influence of geometry parameters on the
response to be compared in the absence of inertia and strain rate effects associated
with dynamic loading. Of interest in the present study is the individual and combined
influence of the wall thickness, taper angle and number of tapers on the response of
straight and tapered rectangular tubes. The results of this chapter show that these
parameters influence the quasi-static energy absorption response and deformation
modes of the tubes by varying amounts. Having determined the quasi-static
behaviour of the straight and tapered tubes, their response under axial impact loading

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 126
is investigated in Chapter 6. Oblique loading of straight and tapered rectangular tubes
is then examined in Chapter 7. The main findings of the present Chapter can also be
found in Nagel and Thambiratnam (2005b).

5.2 Response variables and tube dimensions

The quasi-static behaviour of the straight and tapered tubes was compared using a
parametric analysis by varying their geometry parameters. For axial loading these
parameters were the wall thickness, h and the taper angle, θ. Tube response was
described with commonly used response parameters, namely the initial peak load,
Fbs, the mean crush load, Fms and the energy dissipation, Es. These latter two
parameters were calculated up to a given deflection, d. From these response
parameters other descriptors of energy absorber behaviour were obtained, namely the
energy absorbed per unit crush length, Ecl, and the energy absorbed per unit mass,
Em. These parameters are commonly used such as in crashworthiness applications,
and were defined in Chapter 2. The tube geometries studied were the straight tube,
and tapered tubes having two tapered sides (double taper), three tapered sides (triple
taper) and four tapered sides (frusta). To allow meaningful comparison of the results,
the dimensions of the frusta were changed such that the height of all tubes was 300
mm, while the cross-section dimensions for the straight tube and small end of the
tapered tubes were 100 × 50 mm. The loading and boundary conditions as well as
contact and material properties were the same as those for the validated straight and
double-tapered tubes presented in Chapter 3. Figure 5.2-1 and Table 5.2-1 show the
geometry and dimensions used to model the tubes in the quasi-static parametric
analysis. The different parameters are chosen independently to investigate their effect
on the tube response, while the optimum dimensions will depend on the requirements
and constraints of the intended application. These geometrical parameters and
measurements were chosen in general since they are typical tube sizes as used for
energy absorbers in automotive structures, and suit the application studied in Chapter
8. Also, similar dimensions have been used by others (e.g. Reid & Reddy 1986) for
straight and tapered rectangular tubes, and produced the global symmetric (Type I)
collapse mode required by the existing theoretical and numerical models used for
calibration of the finite element model for axial loading in this thesis.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 127
Wt
Bt
y

z x

300 mm θ

Bb
Wb

Figure 5.2-1 Model geometry for the straight and tapered tubes.

Table 5.2-1 Model dimensions used in the quasi-static parametric analysis.

Tube θ L Wb Bb Wt Bt h
geometry (deg) (mm) (mm) (mm) (mm) (mm) (mm)
Straight 300 50 100 50 100 1.5, 2, 2.5
Double taper 5 300 102.49 100 50 100 1.5, 2, 2.5
10 300 155.8 100 50 100 1.5, 2, 2.5
15 300 210.77 100 50 100 1.5, 2, 2.5
Triple taper 5 300 76.25 152.49 50 100 1.5, 2, 2.5
10 300 102.9 205.8 50 100 1.5, 2, 2.5
15 300 130.38 260.77 50 100 1.5, 2, 2.5
Frusta 5 300 102.49 152.49 50 100 1.5, 2, 2.5
10 300 155.80 205.80 50 100 1.5, 2, 2.5
15 300 210.77 260.77 50 100 1.5, 2, 2.5

5.3 Load-deflection response for straight and tapered


rectangular tubes

Initially the load-deflection response of the straight and tapered rectangular tubes
was compared, as shown in Figure 5.3-1 for a wall thickness of 1.5 mm.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 128
140

120

100

Load (kN)
80

60

40

20

0
0 25 50 75 100 125 150 175 200 225 250
Deflection (mm)

Straight tube
80
70
60
50
Load (kN)

40
30
20
10
0
0 25 50 75 100 125 150 175 200 225 250
Deflection (mm)

Taper angle = 5 deg Taper angle = 10 deg Taper angle = 15 deg

Double taper
80
70
60
50
Load (kN)

40
30
20
10
0
0 25 50 75 100 125 150 175 200 225 250
Deflection (mm)

Taper angle = 5 deg Taper angle = 10 deg Taper angle = 15 deg

Triple taper
Figure 5.3-1 Load-deflection response for straight and tapered rectangular
tubes.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 129
It can be seen in Figure 5.3-1 that the initial peak load is significantly less for the
tapered tubes than for the straight tubes of equivalent cross-section dimensions. This
phenomenon will be investigated further in this chapter. Furthermore, for a given
number of tapers, it appears that the phase of the load-deflection response appears to
be delayed with deflection as the taper angle increases. This may be due to an
observed increase in the fold length as the taper angle increases. Nevertheless, it can
be seen from Figure 5.3-1 that the maximum crush distance at which fold
consolidation begins is approximately the same for each taper angle and number of
tapers. Similar results were also displayed by the frusta. From Figure 5.3-1 it appears
that the initial peak load is significantly affected by tapering the tubes, whereas the
mean load is not. However, the mean load may depend on interaction effects between
the number of tapers and other important geometry parameters, as will be further
investigated in this chapter.

5.4 Parametric study of straight and tapered tubes under


quasi-static loading

5.4.1 Effect of wall thickness

Figure 5.4-1 shows the quasi-static mean load-deflection response for the straight
and double-tapered tubes, as the wall thickness is increased for a fixed taper angle of
100. It can be seen that an increase in wall thickness causes an increase in the mean
load up to a given deflection, as also found for the triple-tapered tube and frusta. This
trend is obviously due to the increased amount of material available for plastic
deformation and subsequent energy absorption, as demonstrated in Figure 5.4-2 for
the straight and double-tapered tubes. Clearly, the energy absorbed up to a given
deflection increases as the wall thickness increases, as was also observed for the
triple-tapered tube and frusta. Also, it can be seen in Figure 5.4-1 and Figure 5.4-2
that the mean load and energy absorption increase abruptly after 250 mm deflection.
This is due to the compaction of the folds at the end of the folding phase, as
identified in Chapter 3.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 130
140
120
100
Mean load (kN)

80
60
40

20
0
0 50 100 150 200 250 300
Deflection (mm)

h = 1.5 mm h = 2 mm h = 2.5 mm

Straight tube

120

100
Mean load (kN)

80

60

40

20

0
0 50 100 150 200 250 300
Deflection (mm)

h = 1.5 mm h = 2 mm h = 2.5 mm

Double taper
Figure 5.4-1 Effect of wall thickness on the quasi-static mean load (here θ = 100).

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 131
35
30
Absorbed energy (kJ)

25
20
15
10
5
0
0 50 100 150 200 250 300
Deflection (mm)

h = 1.5 mm h = 2 mm h = 2.5 mm

Straight tube
35
30
Absorbed energy (kJ)

25
20
15
10

5
0
0 50 100 150 200 250 300
Deflection (mm)

h = 1.5 mm h = 2 mm h = 2.5 mm

Double taper
Figure 5.4-2 Effect of wall thickness on the absorbed energy (here θ = 100).

Increasing the wall thickness also leads to an increase in the initial peak load, as
shown in Figure 5.4-3 for each tube with θ = 100. The increase in peak load with
increasing wall thickness is due to the greater amount of material across the section
of the tube, which effectively increases the axial stiffness of the tube and hence the
load required to initiate buckling.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 132
250
225
200

Initial peak load (kN)


175
150
125
100
75
50
25
0
1.5 2.0 2.5
Wall thickness (mm)

Straight Double taper Triple taper Frusta

Figure 5.4-3 Effect of wall thickness on the initial peak load (here θ = 100).

Finally, Figure 5.4-4 shows the ratio of mean load to initial peak load as a function of
wall thickness for each tube with θ = 100, and the mean load calculated up to a
deflection, d = 200 mm. It can be seen for each tube that the mean load converges to
the initial peak load as the wall thickness increases.

1.0
Mean load / Initial peak load

0.8

0.6

0.4

0.2

0.0
1.5 2.0 2.5
Wall thickness (mm)

Straight Double taper Triple taper Frusta

Figure 5.4-4 Ratio of mean load to initial peak load vs. wall thickness
(here θ = 100).

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 133
5.4.2 Effect of taper angle

Tapering one or more of the sides of a rectangular tube provides an extra geometry
variable, namely the taper angle, which might influence the energy absorption
response. The effects of taper angle were studied for the double-tapered tube, triple-
tapered tube and frusta, for a constant wall thickness of 1.5 mm. The mean load-
deflection response for the double-tapered tube and frusta is shown in Figure 5.4-5 as
a function of the taper angle.

50
45
40
Mean load (kN)

35
30
25
20
15
10
5
0
0 50 100 150 200 250 300
Deflection (mm)

Taper angle = 5 deg Taper angle = 10 deg Taper angle = 15 deg

Double taper

60

50
Mean load (kN)

40

30

20

10

0
0 50 100 150 200 250 300
Deflection (mm)

Taper angle = 5 deg Taper angle = 10 deg Taper angle = 15 deg

Frusta
Figure 5.4-5 Effect of taper angle on the quasi-static mean load
(here h = 1.5 mm).

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 134
For the double-tapered tube it appears that the taper angle has no significant effect on
the mean load-deflection response for lower taper angles. However, increasing the
taper angle to 150 causes an increase in the mean load towards the start of the crush
cycle. This leads to a slightly higher mean load as the deflection increases. For the
triple-tapered tube the taper angle has no significant effect on the response. The
frusta, on the other hand, appears to show a small overall increase in the mean load
with increasing taper angle, for deflections above 100 mm. These results differ for
those of thin-walled circular cylinders and frusta subjected to quasi-static axial
crushing, in which the mean load decreases as the taper angle increases (Mamalis &
Johnson 1983). In the cited study the taper angle was increased by reducing the small
end dimensions and keeping the large end dimensions constant, which may account
for the different mean load response compared with the present study.

The observed effect of taper angle on the tapered rectangular tubes is also reflected
in the energy absorption response, shown in Figure 5.4-6. For the double-tapered
tube the energy absorbed up to a given deflection is approximately the same for taper
angles of 5 and 100, whereas increasing the taper angle to 150 slightly increases the
absorbed energy. The energy absorption response for the triple-tapered tube (not
shown) is not influenced by a change in taper angle, whereas the frusta shows an
overall increase in the energy absorption as the taper angle increases. From these
observations, the mean load and energy absorption response of the tapered tubes is
more influenced by the wall thickness than the taper angle. Nevertheless, it appears
that the response of the double-tapered tube and frusta displays certain dependence
on the taper angle. Hence, it may be possible to use the taper angle to “fine tune” the
response of these absorbers for design purposes.

The effect of the taper angle on the initial peak load for each of the tapered tubes is
shown in Figure 5.4-7. The initial peak load for the straight tube is also shown for
comparison. It can be seen that the initial peak load decreases significantly when a
taper is introduced, and continues to decrease with increasing taper angle. This
phenomenon has been observed for circular frusta (Mamalis & Johnson 1983;
Mamalis et al. 1984; 1986a; b), however no theoretical representation of the
relationship between the initial peak load and taper angle for tapered rectangular
tubes was found in the open literature.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 135
12

10
Absorbed energy (kJ)

0
0 50 100 150 200 250 300
Deflection (mm)

Taper angle = 5 deg Taper angle = 10 deg Taper angle = 15 deg

Double taper
16
14
Absorbed energy (kJ)

12
10
8
6
4
2
0
0 50 100 150 200 250 300
Deflection (mm)

Taper angle = 5 deg Taper angle = 10 deg Taper angle = 15 deg

Frusta
Figure 5.4-6 Effect of taper angle on the absorbed energy (here h = 1.5 mm).

150

125
Initial peak load (kN)

100

75

50

25

0
0 5 10 15
Taper angle (deg)

Straight Double taper Triple taper Frusta

Figure 5.4-7 Effect of taper angle on the initial peak load.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 136
The classical buckling load for circular frusta is given as (Spagnoli &
Chryssanthopoulos 1999):

2πEh 2 cos 2 θ
Pcr = (5.4-1)
3(1 − υ 2 )

where θ is the taper angle, h is the wall thickness, E is the Young’s modulus and v is
the Poisson ratio. This formula predicts that the buckling load decreases with
increasing taper angle, as was the case for the initial peak load examined for the
tapered rectangular tubes in the present study. This phenomenon may be beneficial
for energy absorber design such as in crashworthiness applications, where it is often
desirable to reduce the initial peak loads and decelerations associated with impact
loading. Another possible advantage of using tapered rectangular tubes as energy
absorbers is as follows. As shown in Chapter 2, the frontal structure of a vehicle
usually consists of an energy absorbing “crush box” supported on the end of the
chassis rail. In order for the “crush box” to progressively collapse and effectively
dissipate impact energy, the initial peak load required to crush it must be less than
the maximum load carrying capacity of the supporting chassis rail. Thus, controlling
the taper angle of a tapered “crush box” may be an effective means of reducing its
initial peak load.

Finally, Figure 5.4-8 shows the ratio of mean load to initial peak load as a function of
taper angle for each tube with h = 1.5 mm, and the mean load calculated up to a
deflection, d = 200 mm. It can be seen that the initial peak load converges to the
mean load as the taper angle increases, as was also found for circular frusta (Mamalis
& Johnson 1983). The ratio of mean load to initial peak load is often referred to as
the “form factor” or “crush force efficiency” (see Section 2.2.1.5 of Chapter 2), and
it is often desirable to maximise the ratio to improve the stability of the load-
deflection response. Figure 5.4-8 shows that the efficiency increases with increasing
taper angle and number of tapers. Furthermore, the number of tapers has more
influence on the efficiency as the taper angle increases.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 137
1.2

Mean load / Initial peak load


1.0

0.8

0.6

0.4

0.2

0.0
0 5 10 15
Taper angle (deg)

Straight Double taper Triple taper Frusta

Figure 5.4-8 Ratio of mean load to initial peak load vs. taper angle.

The results thus far indicate that the initial peak load is influenced by the wall
thickness and taper angle. For design purposes, it is of interest to know the relative
effect of each parameter on the initial peak load. In other words, is the initial peak
load more sensitive to changes in the wall thickness or taper angle, for the range of
each parameter studied? This was investigated using a simple factorial study
investigating the main and interaction effects of the wall thickness and taper angle on
the initial peak load. Of interest for the present study are the effects of increasing the
wall thickness from 1.5 to 2.5 mm, and increasing the taper angle from 5 to 150.

For explanation of factorial design and the procedure used to obtain the results
presented in this section, the reader is referred to Section 2.3.3 of Chapter 2. If there
are n input parameters, the study requires 2n tests to be conducted by varying each
parameter between a high and a low value. Two input parameters were varied: wall
thickness and taper angle, giving 4 tests to be performed. The simulation tests
conducted for the different combinations of high and low values of each input
parameter are shown in Table 5.4-1.

The results of the factorial study are summarised in Figure 5.4-9 for each tapered
tube, where each effect is shown as a percentage of the initial peak load. Clearly, the
initial peak load is more influenced by the wall thickness than by the taper angle.
Furthermore, both the wall thickness and taper angle have less influence on the initial

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 138
peak load for the double-tapered tube than for the other tapered tubes. Indeed, the
taper angle has more influence on the initial peak load as the number of tapers
increases. Finally, the h-θ interaction shows that, for each tapered tube, the wall
thickness has less influence on the initial peak load as the taper angle increases.

Table 5.4-1 Levels used in the factorial study.

Level h θ
(mm) (deg)
Low 1.5 5
High 2.5 15

80

60
% of average initial peak load

40

20

0
h θ h-θ
-20

-40

-60
Geometry parameters

Double taper Triple taper Frusta

Figure 5.4-9 Relative effect of wall thickness and taper angle on the initial peak
load.

Also of interest for design purposes is the relative effect of wall thickness and taper
angle on the ratio of mean load to initial peak load (crush force efficiency). The same
factorial study was carried out, the results of which are summarised in Figure 5.4-10.
Here the mean load is calculated up to a deflection of 200 mm. It can be seen that the
wall thickness has more influence than the taper angle on the load ratio for the
double-tapered tube, but not for the triple-tapered tube or frusta. As a practical
outcome, if it is desirable to maximise the crush force efficiency of the response for
the double-tapered tube, it would be more beneficial to vary the wall thickness than
the taper angle.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 139
70

% of average crush force efficiency


60
50
40
30
20
10
0
h θ h-θ
Geometry parameters

Double taper Triple taper Frusta

Figure 5.4-10 Relative effect of wall thickness and taper angle on the ratio of
mean load to initial peak load.

It can also be seen from Figure 5.4-10 that, as the number of tapers increases, the
influence of wall thickness on the crush force efficiency generally decreases while
the influence of taper angle increases. This highlights the significant effect which the
number of tapered sides may have on the response, and this geometry parameter will
be further investigated in the next section. Finally, the h-θ interaction shows that the
wall thickness has more influence on the crush force efficiency as the taper angle
increases. Practically, therefore, the wall thickness is more effective for controlling
the crush force efficiency for larger taper angles.

5.4.3 Effect of the number of tapers

The results thus far indicate that the response of the tubes is influenced by the wall
thickness and the taper angle. It also appears that the response depends on the
number of tapered sides, and this will be further investigated. Figure 5.4-11 shows
the effect of the number of tapered sides on the initial peak load for a taper angle of
100. Clearly, introducing a taper leads to a significant reduction in the initial peak
load, as was observed earlier.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 140
250
225
200

Initial peak load (kN)


175
150
125
100
75
50
25
0
1.5 2.0 2.5
Wall thickness (mm)

Straight Double taper Triple taper Frusta

Figure 5.4-11 Effect of the number of tapered sides on the initial peak load
(here θ = 100).

Furthermore, according to Figure 5.4-11, for a given wall thickness, the triple-
tapered tube generally has a higher initial peak load than the double-tapered tube and
frusta, as was also found for other taper angles. This may be due to the unsymmetric
cross-section of the triple-tapered tube.

The influence of the number of tapered sides on the mean load is shown in Figure
5.4-12 and Figure 5.4-13 for a constant taper angle and wall thickness, respectively.
The mean load is calculated up to a deflection, d = 200 mm.

120

100
Mean load (kN)

80

60

40

20

0
1.5 2.0 2.5
Wall thickness (mm)

Straight Double taper Triple taper Frusta

Figure 5.4-12 Effect of the number of tapered sides on the mean load
(here θ = 100).

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 141
40

Mean load (kN) 38

35

33

30
0 5 10 15
Taper angle (deg)

Straight Double taper Triple taper Frusta

Figure 5.4-13 Effect of the number of tapered sides on the mean load
(here h = 1.5 mm).

According to Figure 5.4-12 and Figure 5.4-13, the triple-tapered tube generally has
the highest mean load for a given wall thickness and taper angles of 5 and 100. This
may be due to the unsymmetrical cross-section of the triple-tapered tube restricting
the limit of compression of each fold, thus increasing the mean load. However, when
the taper angle is increased to 150 the frusta has a higher mean load than the other
tapered tubes, which may be due to its greater cross-section at a given deflection, and
hence more material available for energy absorption. From Figure 5.4-12 it appears
that the number of tapers has more influence on the mean load as the wall thickness
increases. Furthermore, it can be seen from Figure 5.4-12 and Figure 5.4-13 that the
mean load for the frusta is generally larger than that for double-tapered and straight
tubes at a given wall thickness and taper angle, and may once again be due to the
larger cross-section of the frusta at a given deflection compared with these tubes.

These same trends are also displayed by the energy absorption response of each tube,
as shown in Figure 5.4-14 and Figure 5.4-15. Overall the results indicate that the tube
response can also be controlled by varying the number of tapered sides.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 142
23
20

Absorbed energy (kJ)


18
15
13
10
8
5
3
0
1.5 2.0 2.5
Wall thickness (mm)

Straight Double taper Triple taper Frusta

Figure 5.4-14 Effect of the number of tapers on the absorbed energy


(here θ = 100).

8.0
Absorbed energy (kJ)

7.5

7.0

6.5

6.0
0 5 10 15
Taper angle (deg)

Straight Double taper Triple taper Frusta

Figure 5.4-15 Effect of the number of tapers on the absorbed energy


(here h = 1.5 mm).

5.4.4 Effect of wall thickness, taper angle and number of tapers on


other design parameters

Thus far, the results of this chapter indicate that the load-deflection, mean load-
deflection and energy absorption response of straight and tapered rectangular tubes
are influenced by the wall thickness, taper angle and the number of tapered sides.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 143
Also of interest, particularly from the viewpoint of energy absorber design, is the
effect of these parameters on various other important measures of tube response,
namely the energy absorbed per unit crush length, Ecl, and the energy absorbed per
unit mass, Em.

5.4.4.1 Energy absorbed per unit crush length, Ecl

As described in Chapter 2, the energy absorbed per unit length is important when the
available crush zone is limited. It was observed that the maximum crush distance
under quasi-static loading was approximately the same for each tube geometry over
the range of wall thickness and taper angle studied. Thus, to provide a valid
comparison of the response for each tube, Ecl was calculated using the energy
absorbed up to a deflection of 200 mm. The results are summarised in Figure 5.4-16
for a taper angle of 50. It was found that the taper angle does not significantly
influence Ecl compared with the wall thickness. From Figure 5.4-16 it is obvious
however, that Ecl is an increasing function with wall thickness, as was also found for
other taper angles. This is obviously due to the increased energy absorption
associated with a higher wall thickness. Finally, for each taper angle and wall
thickness, the triple-tapered tube appears to have the highest energy absorption per
unit crush length, followed by the frusta, while the response for the straight tube and
double-tapered tube is virtually identical.
Energy absorbed per unit crush length

100
90
80
70
60
(kJ / m)

50
40
30
20
10
0
1.5 2.0 2.5
Wall thickness (mm)

Straight Double taper Triple taper Frusta

Figure 5.4-16 Effect of wall thickness and number of tapers on Ecl (θ = 50).

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 144
The following practical conclusions can be drawn from these observations:

(1) For a given number of tapers and taper angle, increasing the wall thickness
maximises the energy absorbed within in a given crush distance.
(2) For a given wall thickness and number of tapers, increasing the taper angle
does not significantly increase the energy absorbed within a given crush
distance.
(3) For a given wall thickness and taper angle, the triple-tapered tube and frusta
absorb more energy within a given crush distance than the straight and
double-tapered tubes.

5.4.4.2 Energy absorbed per unit mass, Em

Increasing the wall thickness, taper angle and number of tapers increases the
undeformed tube’s mass, as shown in Table 5.4-2.

Table 5.4-2 Increase in tube mass with wall thickness and taper angle.

Geometry θ Undeformed tube mass, m


(mm) (kg)
h = 1.5 mm h = 2.0 mm h = 2.5 mm
Straight 1.02 1.35 1.69
Double taper 5 1.20 1.60 2.00
10 1.39 1.86 2.32
15 1.60 2.13 2.66
Triple taper 5 1.29 1.72 2.15
10 1.58 2.11 2.64
15 1.90 2.53 3.16
Frusta 5 1.38 1.85 2.31
10 1.78 2.37 2.96
15 2.21 2.94 3.68

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 145
This increase in tube mass has a significant effect on the energy absorbed per unit
mass, as shown in Figure 5.4-17 for a taper angle of 50. Here the absorbed energy has
been calculated up to a deflection of 200 mm.

10
Energy absorbed per unit mass

9
8
7
6
(kJ / kg)

5
4
3
2
1
0
1.5 2.0 2.5
Wall thickness (mm)

Straight Double taper Triple taper Frusta

Figure 5.4-17 Effect of wall thickness and number of tapers on Em (θ = 50).

Clearly, Em increases as the wall thickness increases for a given taper angle and
number of tapers. In other words, each tube can absorb more energy for a given mass
as the wall thickness increases. Furthermore, for each taper angle and wall thickness
it can be seen that the straight tube has a higher Em than any of the tapered tubes,
while the frusta has the lowest Em. This is indicative of the increased mass of the
absorber as the number of tapers is increased, despite the associated increase in
absorbed energy. This was also found for other taper angles. On the other hand, the
triple-tapered tube had a higher Em than the double-tapered tube for taper angles of 5
and 100, but not for 150.

For each tapered tube, the energy absorbed per unit mass decreases as the taper angle
increases, as shown in Figure 5.4-18 for the double-tapered tube. A factorial study
was carried out to determine the relative influence of wall thickness and taper angle
on Em for each tapered tube. This involved comparing the effects of increasing the
wall thickness from 1.5 to 2.5 mm, and increasing the taper angle from 5 to 150.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 146
9

Energy absorbed per unit mass


8
7
6

(kJ / kg)
5
4
3
2
1
0
1.5 2.0 2.5
Wall thickness (mm)

Taper angle = 5 deg Taper angle = 10 deg Taper angle = 15 deg

Figure 5.4-18 Effect of wall thickness and taper angle on Em for the double-
tapered tube.

The results of the factorial study, summarised in Figure 5.4-19, indicate that except
for the double-tapered tube, both the wall thickness and taper angle have virtually the
same influence on Em for a given number of tapers. However, Em increases with
increasing wall thickness, yet decreases with increasing taper angle. Furthermore, for
each tapered tube, the h-θ interaction shows that the wall thickness has less influence
on Em as the taper angle increases.

60
% of average energy absorbed

40

20
per unit mass

0
h θ h-θ
-20

-40

-60
Geometry parameters

Double taper Triple taper Frusta

Figure 5.4-19 Relative effect of wall thickness and taper angle on Em.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 147
In the light of these observations, the following practical conclusions can be drawn:

(1) For a given number of tapers and taper angle, increasing the wall thickness
maximises the energy absorbed per unit mass. This indicates that increasing
the wall thickness would be a favourable means of increasing the energy
absorption capability for straight and tapered rectangular tubes, when mass is
an important consideration.
(2) For a given wall thickness and number of tapers, increasing the taper angle
decreases the energy absorbed per unit mass. Thus it would be preferable to
minimise the taper angle when mass is an important consideration.
(3) For the triple-tapered tube and frusta, both the wall thickness and taper angle
have approximately the same influence on the energy absorbed per unit mass.
(4) For a given wall thickness and taper angle, the straight tube absorbs the most
energy for a given absorber mass, while the frusta absorbs the least.
Generally, therefore, it would be desirable to minimise the number of tapers
when comparing tubes of equivalent cross-section where mass is an important
consideration.
(5) Finally, for each tapered tube, the wall thickness is not as effective for
controlling Em as the taper angle increases.

5.5 Developing an empirical relation between the geometry


and energy absorption response of tapered tubes

For the designer who intends to use a tapered thin-walled rectangular tube for impact
energy absorption, it would be useful to have a simple method of determining the
quasi-static crush force efficiency, in other words, the ratio of peak load to mean
load, Fbs / Fms. This would allow initial assessment of the energy absorption response
of tapered tubes, from which a suitable geometry could be selected and further
examined under simulated or experimental impact loading conditions.

With this in mind, the results of the validated numerical model were used to develop
a simple empirical model which could be used to predict the quasi-static crush force
efficiency. The double-tapered tube was used as an example.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 148
5.5.1 Developing the empirical relation

Figure 5.5-1 shows the geometry of the double-tapered rectangular tube, and the
associated geometry parameters.
Wt
Bt

Figure 5.5-1 Double-tapered rectangular tube geometry.

The geometry parameters which typically control the energy absorption response are
the taper angle, θ, wall thickness, h, the cross-section dimensions of the small end,
Wt and Bt, and the deflection, d. For convenience of design purposes, these
parameters can be combined into the following dimensionless groups:

h / Wt (slenderness ratio)

Wt / Bt (aspect ratio)

d/L (relative crush distance)

The slenderness ratio and aspect ratio are commonly used to define the geometry of
thin-walled columns used as impact energy absorbers, and the relative crush distance
takes account of the increasing mean load as the tapered tube is crushed.1 The aim is
to develop a relation between these dimensionless groups and the energy absorption
response, represented by the crush force efficiency, of a double-tapered tube having a
given taper angle. This relation would be empirical in form, based on results for the
1
For this particular application, the slenderness ratio is defined as the wall thickness over the width,
as has also been adopted in previous studies on circular frusta (Mamalis & Johnson 1983; Mamalis et
al. 1984; 1986a; b). Although in other applications it is also defined as the tube length over a cross-
sectional dimension.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 149
crush force efficiency over a range of values for each dimensionless group. Empirical
modelling was described in Section 2.3.2 of Chapter 2. The validated finite element
model of the double-tapered tube under axial quasi-static loading is used. For
simplicity, the analysis presented only applies to a double-tapered tube of 150 taper
angle, however the analysis can also be applied to other taper angles and number of
tapers.

The empirical relation takes the form:

α β γ
Fms W   h  d 
= C  t      (5.5-1)
Fbs  Bt   Wt  L

which is a form commonly used for dimensional analysis. The constant C and power
coefficients α, β and γ are determined using an available data set containing results of
the crush force efficiency for a range of aspect ratios, slenderness ratios and relative
crush distances. The results for an aspect ratio of 1 are shown in Figure 5.5-2 for a
taper angle of 150.

1.2
Crush force efficiency, F ms / Fbs

1.0

0.8

0.6

0.4

0.2

0.0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Relative crush distance, d / L

h / Wt = 0.02 h / Wt = 0.03 h / Wt = 0.04 h / Wt = 0.05


h / Wt = 0.06 h / Wt = 0.07

Figure 5.5-2 Variation of crush force efficiency with relative crush distance and
slenderness ratio for the double-tapered tube (here θ = 150 and the aspect ratio,
Wt / Bt = 1).

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 150
For Figure 5.5-2, The mean load was calculated using the theory presented in
Chapter 3, while the initial peak load was obtained using the validated finite element
model. As can be seen, the crush force efficiency increases as the relative crush
distance increases due to the corresponding increase in mean load. Furthermore, the
crush force efficiency increases with increasing slenderness ratio. Similar results
were also found for other aspect ratios, and it was found that the crush force
efficiency is relatively independent of aspect ratio.

Using the least squares theorem with linear programming and applying the available
results to Equation 5.5-1, the following values were obtained for the constant and
power coefficients:

C = 5.51
α = -0.04
β = 0.61
γ = 0.11

These values can be used to obtain the equation that provides the “best fit” to the
finite element results. Thus, the empirical relation which determines the quasi-static
crush force efficiency for a double-tapered rectangular tube of taper angle 150, based
on the aspect ratio, slenderness ratio and relative crush distance, is:

−0.04 0.61 0.11


Fms W   h  d 
= 5.51 t      (5.5-2)
Fbs  Bt   Wt  L

The exponent α is relatively small compared with the corresponding values for β and
γ since the load ratio is relatively insensitive to aspect ratio. The values of β and γ are
relatively larger and positive since the load ratio increases as both the slenderness
ratio and relative crush distance increase.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 151
5.5.2 Testing the empirical relation for various aspect ratios,
slenderness ratios and relative crush distances

Equation 5.5-2 was tested for various aspect ratios, slenderness ratios and relative
crush distances, for a double-tapered rectangular tube with a taper angle of 150 and a
length of 300 mm. The material used was the same as that used in the validated finite
element model presented in Chapter 3. Three cases were studied, the results of which
are summarised in Figure 5.5-3. The peak load has been calculated using the
empirical model and the theoretical prediction of the mean load. This predicted peak
load has then been compared with the peak load obtained by the finite element
model. As can be seen, reasonable correlation is obtained. Thus, a useful tool has
been developed which can provide the designer with an initial estimate of the crush
loads and hence energy absorption characteristics for a double-tapered tube of given
dimensions. Furthermore, the relationship provides a guide as to which parameters
have most influence on the crush loads and energy absorption.

From Figure 5.5-3 it is obvious that the accuracy of the empirical relation depends on
the range of each geometry parameter used in Equation 5.5-2. For instance, for an
aspect ratio = 1 and slenderness ratio = 0.035, the initial peak load is over-predicted
by the empirical relation. A tube length of 200 mm was modelled for this case,
whereas the empirical relation in Equation 5.5-2 was developed using a tube length
of 300 mm. Thus, greatest accuracy is obtained when using Equation 5.5-2 for a tube
length of 300 mm. Developing the empirical equation using results for other tube
lengths would improve the equation’s accuracy over a range of tube lengths.
Furthermore, Equation 5.5-2 was developed for wall thicknesses between 1 and 3.5
mm, a small end width, Wt of 50 mm, and a small end breadth, Bt between 50 and
120 mm. Thus, Equation 5.5-2 is most accurate for values of each parameter within
these ranges. Furthermore, it was found that a wall thickness less than 1 mm may
lead to a non-compact tube response (refer to Figure 2.2-5 of Chapter 2), for which
the theoretical model used to validate the FE model was not valid. Thus, care should
be taken when using Equation 5.5-2 for slenderness ratios (h / Wt) less than 0.02,
since this was the minimum ratio used to develop Equation 5.5-2.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 152
60
55
50
45
40

Fbs (kN)
35
30
25
20
15
10
5
0
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 1.1 1.2 1.3
Relative crush distance (d / L )

Empirical Finite element model

Aspect ratio = 1, slenderness ratio = 0.035


130
120
110
100
90
80
Fbs (kN)

70
60
50
40
30
20
10
0
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
Relative crush distance (d / L )

Empirical Finite element model

Aspect ratio = 0.8, slenderness ratio = 0.028


100
90
80
70
60
Fbs (kN)

50
40
30
20
10
0
0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6
Relative crush distance (d / L )

Empirical Finite element model

Aspect ratio = 0.83, slenderness ratio = 0.015


Figure 5.5-3 Prediction of the initial peak load using Equation 5.5-2 and the
finite element model.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 153
Finally, it should be noted that since geometry alone is not sufficient in determining
the crush force efficiency, Equation 5.5-2 may only work for the mild steel grade
used and not other material properties.

Similar relations for taper angles of 5 and 100 have been developed in Nagel and
Thambiratnam (2005b), for which correlations within 6 % were obtained.

5.6 Deformation modes of straight and tapered rectangular


tubes under quasi-static axial loading

The axial deformation mode of a thin-walled tube can be influenced by its geometry
parameters. This becomes particularly important for energy absorber design, when
the deformation mode producing maximised energy absorption may be required. For
instance, it is often desirable to select the wall thickness and section dimensions of
the tube such that progressive collapse is obtained under axial loading, as opposed to
the global (Euler) type buckling mode which limits the energy absorption. This
section investigates the influence of the number of tapers, taper angle and wall
thickness on the deformation modes of the straight and tapered rectangular tubes
examined in the present study.

5.6.1 Effect of the number of tapers on the deformation mode

Figure 5.6-1 illustrates the collapse sequence of each of the straight and tapered
rectangular tubes for a wall thickness of 1.5 mm and a taper angle of 100. The quarter
models are shown for the straight tube, double-tapered tube and frusta, while the full
model is shown for the triple-tapered tube. For each tube, collapse starts at the loaded
end and progresses down the length of the tube. The plastic fold length increases
down the length of the tapered tubes yet is constant for the straight tube.
Furthermore, it appears that the fold length increases with increasing number of
tapers.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 154
Undeformed d = 61 mm d = 108 mm d = 158 mm d = 239 mm

Straight

Double taper

Triple taper

Frusta

Figure 5.6-1 Collapse sequence of straight and tapered rectangular tubes under
a quasi-static axial load (here h = 1.5 mm and θ = 100).

5.6.2 Effect of the taper angle on the deformation mode

Also of interest is the dependence of the deformation mode on the taper angle for
each tapered tube. For the double-tapered tube and frusta it was found that the fold

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 155
length increases with increasing taper angle, while the deformation mode remains
symmetric, as found for the axial impact loading of square frusta (Mamalis et al.
2001). This is illustrated in Figure 5.6-2 for the frusta at a deflection of 158 mm, and
a wall thickness of 1.5 mm. Also shown is the length of the large fold formed for the
double-tapered tube and frusta at the above deflection, as the taper angle increases.
As the taper angle increases for the double-tapered tube, the same number of lobes
forms on the tapered side up to the maximum crush distance. However, for the frusta
the number of lobes on each tapered side decreases as the taper angle increases.

δ
δ Geometry δ (mm)
(θ, deg)
0
(5 ) (100) (150)

y y Double taper 21 27 40
Frusta 20 62 96
x z
Frusta Double taper

z
x

Frusta (taper angle = 50) Frusta (taper angle = 100) Frusta (taper angle = 150)

Figure 5.6-2 Dependence of fold length on taper angle for the double-tapered
tube and frusta (h = 1.5 mm).

The deformation mode of the triple-tapered tube is not symmetric as the taper angle
increases, nevertheless the folds remained inextensional in a global sense. This is
illustrated in Figure 5.6-3, which shows respective isometric, side and top views of
the triple-tapered tube at a deflection of 223 mm, for a wall thickness of 1.5 mm. An
increase in taper angle leads to a decrease in the number of lobes formed on the
larger tapered side.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 156
Isometric view Side view Top view

Triple taper (taper angle = 50)

Triple taper (taper angle = 100)

Triple taper (taper angle = 150)

Figure 5.6-3 Dependence of deformation mode on the taper angle for the triple-
tapered tube (h = 1.5 mm).

As the tube is crushed for taper angles of 5 and 150, it tends to slide in a direction
towards the vertical face, as well as twist about the longitudinal axis. This may be
due to the unsymmetric cross-section creating a resultant force which acts oblique to
the longitudinal axis of the tube, as illustrated in Figure 5.6-4. Such a deformation
mode may be impractical from a design viewpoint. For instance, the absorber may
slide outside of its available crush zone, or the twisting behaviour may induce
undesirable stresses in the mounting system. For a taper angle of 100 the tube
deformed in an axial manner with no sliding or twisting. It appears that this may be
due to the first lobe “sinking into” the second lobe, thus decreasing the reaction force
that leads to sliding deformation.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 157
Fr

Figure 5.6-4 Resultant force created in the triple-tapered tube due to its
unsymmetric cross-section.

5.6.3 Effect of wall thickness on the deformation mode

Thus far the deformation response of the straight and tapered tubes has pertained to a
wall thickness of 1.5 mm. Increasing the wall thickness to 2.5 mm does not change
the number of lobes formed nor the fold length for the straight tube. However,
increasing the wall thickness to 2.5 mm appears to decrease the fold length for the
double-tapered tube and frusta. For instance, at a deflection of 158 mm and for a
taper angle of 150, the length of the large fold for the double-tapered tube and frusta
was 32 mm (recall 40 and 96 mm with h = 1.5 mm for the double-tapered tube and
frusta, respectively). Nevertheless, the number of lobes formed for each taper angle
for the double-tapered tube remained the same when the wall thickness was
increased from 1.5 to 2.5 mm. For the frusta the number of lobes decreased as the
wall thickness increased for a taper angle of 100, whereas the number of lobes
increased for taper angles of 50 and 150. For the triple-tapered tube with a taper angle
of 150, increasing the wall thickness to 2.5 mm prevented the tube from moving
rearward, however the tube still rotated about its longitudinal axis. Furthermore,
buckling initiated at the large end of the tube beyond a deflection of 100 mm. For a
taper angle of 50, the triple-tapered tube did not exhibit any twisting, while twisting
and sliding in a direction towards the vertical side was observed for a taper angle of
100. Thus, it appears that the deformation mode of the triple-tapered tube is sensitive
to wall thickness.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 158
5.7 Conclusions and design guidelines

This chapter studied the quasi-static axial behaviour of straight and tapered
rectangular tubes. The parameters investigated were the tube wall thickness, taper
angle, number of tapers and cross-section dimensions. The main conclusions from
the study are summarised below. Though they are based on quasi-static analyses, the
conclusions provide design information regarding the relative effect of each
geometry parameter on the tube response. This information could then be used as an
initial guide for implementing tapered rectangular tubes in applications where impact
energy must be absorbed. Overall, the results highlight the advantages of using
tapered rectangular tubes for absorbing energy under quasi-static axial loading.

(1) The initial peak load decreases with the introduction of a taper, and as the
taper angle increases. This would be beneficial when it is necessary to reduce
the peak impact loads and decelerations transmitted to the structure under
protection.

(2) The mean load and energy absorption response of tapered rectangular tubes
subjected to axial quasi-static loading is more influenced by the wall
thickness than the taper angle and number of tapers. Nevertheless, the number
of tapers has more influence on the mean load as the wall thickness increases.
Thus, as the wall thicknesses increases, the triple-tapered tube appears to
have the highest mean load and hence energy absorption.

(3) For tapered rectangular tubes, the initial peak load is more influenced by the
wall thickness than the taper angle. However, the influence of wall thickness
on the initial peak load diminishes as the taper angle increases. Practically,
therefore, the wall thickness is more effective for controlling the initial peak
load for smaller taper angles. Furthermore, the taper angle has more influence
on the initial peak load as the number of tapers increases.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 159
(4) The crush force efficiency (ratio of mean crush load to initial peak crush
load) increases with increasing taper angle and number of tapers.
Furthermore, the number of tapers has more influence on the efficiency as the
taper angle increases. The wall thickness is more effective for controlling the
crush force efficiency for larger taper angles. Furthermore, as the number of
tapers increases, the influence of wall thickness on the crush force efficiency
generally decreases while the influence of taper angle increases.

(5) For a given wall thickness and taper angle, the triple-tapered tube generally
has a higher initial peak load than the double-tapered tube and frusta. Thus,
when using the triple-tapered tube for impact energy absorption, care must be
taken to avoid excessive initial peak crush loads.

(6) The energy absorbed within a given crush distance for a rectangular tube can
be maximised by:

a. Increasing the wall thickness, and/ or


b. Using a tube with three or four tapered sides.

(7) However, the taper angle has no significant influence on the energy absorbed
within a given crush distance.

(8) The energy absorbed for a given absorber mass can be increased by:

a. Increasing the wall thickness, and/ or


b. Decreasing the taper angle, and/ or
c. Decreasing the number of tapers.

(9) For each tapered tube, the wall thickness is not as effective for controlling the
energy absorbed per unit mass as the taper angle increases.

(10) As the triple-tapered tube is crushed it tends to slide in a direction


towards the vertical face, as well as twist about its longitudinal axis. Such a
deformation mode may be impractical from a design viewpoint, since the

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 160
absorber may slide outside of its available crush zone, or the twisting
behaviour may induce undesirable stresses in the mounting system. Such
behaviour may be avoided by appropriate selection of the wall thickness and
taper angle.

(11) A simple empirical relationship has been derived for calculating the
crush force efficiency of double-tapered rectangular tubes, using the aspect
ratio (ratio of small end width to breadth), slenderness ratio (ratio of wall
thickness to small end width), and relative crush distance (ratio of crush
distance to tube length). The crush force efficiency increases with increasing
slenderness ratio and relative crush distance. However, the crush force
efficiency is relatively independent of aspect ratio. The empirical relationship
can be used for preliminary design and selection of double-tapered tubes for
energy absorption.

When designing energy absorbing structures it is often convenient to start with a


quasi-static analysis, which, as has been shown already, is relatively simple
compared with a dynamic impact analysis due to the absence of strain rate and inertia
effects. However, loading in real-world energy absorption applications is primarily
dynamic or impulsive in nature. Thus the response of straight and tapered rectangular
tubes under dynamic impact loading is of particular interest, and this will be the
focus of the next chapter.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 161
6 Dynamic Response of Straight and Tapered
Rectangular Tubes under Axial Impact
Loading

6.1 Introduction

As mentioned in Chapter 5, quasi-static analyses provide a convenient means of


comparing the influence of geometry parameters on the energy absorber response.
However, thin-walled tubular energy absorbers used in crashworthiness applications
are typically subjected to impact energies which are significantly higher than the
corresponding energies under quasi-static loading. Thus, it is important to examine
the dynamic response of such absorbers so as to be able to predict their energy
absorbing characteristics and estimate the associated crushing forces. Furthermore,
for impact velocities above typically tens of metres per second, the inertia forces in a
tube become important (Jones 1989). This, together with material strain rate effects
as outlined in Chapter 2, needs to be taken into account for the analysis and design of
energy absorbers which are subjected to higher impact velocities.

As identified in Chapter 2, there appears to be relatively few studies conducted on


the energy absorption behaviour of tapered thin-walled rectangular tubes under
dynamic loading, in particular with the use of numerical methods. Thus far the thesis
has highlighted the advantages of tapered rectangular tubes for use as energy
absorbers, in particular the ability to control their response via the taper angle and
number of tapers. An understanding of the behaviour of these tubes under dynamic
loading is therefore warranted, which will be the focus of this chapter.

The aim of this chapter is to study the dynamic energy absorption response of
straight and tapered rectangular tubes under axial impact loading. Simulations are
conducted using the finite element model validated for dynamic loading, as described
in Chapter 3. Dynamic loading introduces two loading parameters: the impact mass

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 162
and the impact velocity. These have been found to influence the energy absorption
response of thin-walled circular tubes and square columns under axial impact loading
(see for example Karagiozova & Jones 2004; Karagiozova et al. 2000; Langseth et al.
1999). The effects of impact mass and velocity on the response of straight and
tapered rectangular tubes are studied in this chapter. Also, the effects of geometry
parameters such as wall thickness, taper angle and number of tapers under dynamic
loading are compared with the corresponding quasi-static response. The results of
such investigations allow comparison of the inertia sensitivity of straight and tapered
rectangular tubes, as well as determination of the relative influence of the various
geometry parameters on the dynamic response. The practical application of the
results is the development of guidelines to facilitate the design of straight and tapered
rectangular tubes under axial impact loading.

Overall, the results show that inertia effects influence the dynamic impact response
of straight and tapered rectangular tubes, yet these effects are less significant for the
latter. This would be beneficial for impact energy absorption applications where
“strengthening” of the absorber due to inertia effects needs to be minimised.
Furthermore, the properties of the impact surface also influence the energy
absorption response of straight and tapered rectangular tubes. For other studies
carried out by the author on the response of tapered rectangular tubes under axial
impact loading, refer to Nagel and Thambiratnam (2004a; b).

6.2 Tube dimensions and loading variables

The same tube dimensions were used as for the quasi-static analysis, and are shown
in Figure 6.2-1 and Table 6.2-1 for ease of reference. Material strain rate sensitivity
is taken into account using the Cowper-Symonds constitutive equation and the
associated material parameters as described in Chapter 3. As described in Chapter 3,
dynamic loading was simulated by applying a point mass, M and initial velocity, V to
the rigid body supporting each tube such that the body translated to axially impact
the tube into a fully fixed rigid body. The loading arrangement is shown in Figure
6.2-2. The step time was 30 ms, as established in Section 3.2.2 of Chapter 3.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 163
Wt
Bt y

z x

300 mm θ

Bb
Wb

Figure 6.2-1 Model geometry for the (l) straight and (r) tapered tubes.

Table 6.2-1 Model dimensions used in the parametric analysis.

Geometry θ L Wb Bb Wt Bt h
(deg) (mm) (mm) (mm) (mm) (mm) (mm)
Straight 300 50 100 50 100 1.5, 2, 2.5
Double taper 5 300 102.49 100 50 100 1.5, 2, 2.5
10 300 155.8 100 50 100 1.5, 2, 2.5
15 300 210.77 100 50 100 1.5, 2, 2.5
Triple taper 5 300 76.25 152.49 50 100 1.5, 2, 2.5
10 300 102.9 205.8 50 100 1.5, 2, 2.5
15 300 130.38 260.77 50 100 1.5, 2, 2.5
Frusta 5 300 102.49 152.49 50 100 1.5, 2, 2.5
10 300 155.80 205.80 50 100 1.5, 2, 2.5
15 300 210.77 260.77 50 100 1.5, 2, 2.5

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 164
Fully fixed
rigid body

Rigid body and


y tube translate in
V y direction
z x

M
Base of tube fully fixed
to rigid body

Figure 6.2-2 Arrangement for axial impact loading (straight tube quarter model
shown).

6.3 Effect of impact velocity and impact mass

Firstly, the effect of impact velocity and mass on the response of the straight and
tapered rectangular tubes was examined and compared with the corresponding quasi-
static response, to determine the dynamic effects for each tube. As for quasi-static
loading presented in Chapter 5, this was accomplished by examining the load-
deflection, mean load-deflection, and energy absorption-deflection response.

6.3.1 Effect of impact velocity

Figure 6.3-1 shows the effect of impact velocity, V, on the load-deflection response
for the straight and tapered rectangular tubes. The impact velocity range was chosen
to cover the typical range of speeds encountered in many crashworthiness
applications such as automobiles. Here, the impact mass is 80 kg, while the wall
thickness is 1.5mm and taper angle is 100. This impact mass was chosen since it
provided sufficient kinetic energy to crush the tubes.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 165
250

200
Dynamic load (kN)
150

100

50

0
0 25 50 75 100 125 150 175 200 225 250 275
Deflection (mm)

Quasi-static V = 5 m/s V = 10 m/s V = 15 m/s V = 20 m/s

Straight tube
160
140
Dynamic load (kN)

120
100
80
60
40
20
0
0 25 50 75 100 125 150 175 200 225 250 275
Deflection (mm)

Quasi-static V = 5 m/s V = 10 m/s V = 15 m/s V = 20 m/s

Double taper
160
140
Dynamic load (kN)

120
100
80
60
40
20
0
0 25 50 75 100 125 150 175 200 225 250 275
Deflection (mm)

Quasi-static V = 5 m/s V = 10 m/s V = 15 m/s V = 20 m/s

Triple taper
Figure 6.3-1 Effect of impact velocity on the load-deflection response
(here h = 1.5 mm and θ = 100).

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 166
Each curve in Figure 6.3-1 is shown up to the maximum deflection, defined herein as
the crush distance at which the tube is brought to rest. For each tube, an increased
impact velocity causes the load to increase during the first part of the impact. In
particular, the initial peak load is increased as the impact velocity increases. A
similar effect has been observed for the axial impact loading of square tubes made of
strain rate-insensitive aluminium (Langseth & Hopperstad 1996; Langseth et al.
1999), in which the observed difference in loads was associated with inertia effects
set up at the instant of impact due to the lateral movement of the sidewalls in order to
initiate the folding process. Since the tubes considered in the present study are made
from strain rate-sensitive mild steel, the increase in load is also due to strain rate
effects causing an increase of the flow stress for higher impact velocities.

The increase in dynamic load with velocity shown in Figure 6.3-1 causes a
subsequent increase in the mean load and energy absorption up to a given deflection
for each tube, as shown in Figure 6.3-2 and Figure 6.3-3 for the straight and double-
tapered tubes. A similar response was also found for axially impacted aluminium
alloy square tubes (Langseth et al. 1999), in which more energy was needed to start
the folding process when increasing the impact velocity. Thus, for the straight and
tapered rectangular tubes in the present study, inertia effects appear to increase the
crush loads and energy absorbed under dynamic loading compared with quasi-static
loading. This has been observed for the axial impact loading of numerous energy
absorbing structures, such as aluminium honeycombs (Harrigan et al. 1999), circular
frusta (Singace et al. 2001), circular tubes (Karagiozova & Jones 2001; Karagiozova
et al. 2000), and square tubes (Karagiozova & Jones 2004).

Several further important observations can be made from the dynamic response
presented in Figure 6.3-2 and Figure 6.3-3. Firstly, it can be seen in Figure 6.3-2 that,
for the straight tube, the axial deflection before buckling (i.e. before the initial peak
load is reached) increases as the impact velocity increases. This was also found for
axially impacted aluminium alloy square tubes (Langseth et al. 1999), and was
attributed to lateral inertia forces in the tube. These forces tended to “support” the
side-walls in the early phase of the response and consequently delayed the initial
buckling.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 167
140
120
Dynamic mean load (kN) 100
80
60
40
20
0
0 50 100 150 200 250 300
Deflection (mm)

Quasi-static V = 5 m/s V = 10 m/s V = 15 m/s V = 20 m/s

Straight tube

120
Dynamic mean load (kN)

100

80

60

40

20

0
0 50 100 150 200 250 300
Deflection (mm)

Quasi-static V = 5 m/s V = 10 m/s V = 15 m/s V = 20 m/s

Double taper
Figure 6.3-2 Effect of impact velocity on the dynamic mean load
(here h = 1.5 mm and θ = 100).

Thus, lateral inertia forces tend to be influencing the response of the straight tubes in
the present study. However, it can be seen from the results for the double-tapered
tube in Figure 6.3-2 that the axial deflection at the initial peak load is not as sensitive
to impact velocity. This was also found for the triple-tapered tube and frusta.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 168
16

Dynamic absorbed energy (kJ)


14
12
10
8
6
4
2
0
0 50 100 150 200 250 300
Deflection (mm)

Quasi-static V = 5 m/s V = 10 m/s V = 15 m/s V = 20 m/s

Straight tube

16
Dynamic absorbed energy (kJ)

14
12
10
8
6
4
2
0
0 50 100 150 200 250 300
Deflection (mm)

Quasi-static V = 5 m/s V = 10 m/s V = 15 m/s V = 20 m/s

Double taper
Figure 6.3-3 Effect of impact velocity on the dynamic absorbed energy
(here h = 1.5 mm and θ = 100).

Thus, it appears that the presence of a taper may be influencing the inertia effects on
the tube response. This will be further investigated in subsequent sections of this
chapter. Secondly, both Figure 6.3-2 and Figure 6.3-3 show the results up to the
maximum deflection at which the tube is brought to rest. Clearly, the maximum
deflection for each tube increases as the impact velocity increases, and this is
obviously due to the increased impact (kinetic) energy of the tube and loading rigid
body.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 169
Overall, the main conclusion that can be drawn from Figure 6.3-1 to Figure 6.3-3 is
that increasing the impact velocity increases the initial peak (buckling) load and the
associated mean load and energy absorption response for both straight and tapered
rectangular tubes made of a strain rate-sensitive material, in this case mild steel.
Thus, as long as the main difference between the dynamic and static response of a
tube is related to the first part of the impact, it appears that the straight and tapered
rectangular tubes are behaving much like Type II structures. As explained in Section
2.1.1.1 of Chapter 2, the load-deflection response of these structures consists of an
initial peak load whose magnitude is influenced by inertia effects. However, further
study is required of the strain rate and inertia effects of the tubes under impact
loading, as will be investigated in this chapter.

To determine the relative influence of strain rate and inertia effects on the dynamic
response of the straight and tapered tubes, simulations were conducted without the
strain rate-dependent material option in the finite element model. This was
accomplished for each tube having the same dimensions and loading conditions as
before. In Figure 6.3-4 the initial peak load is compared with the corresponding
strain rate-sensitive case as a ratio of the quasi-static initial peak load. Results are
shown for the straight and double-tapered tubes. For each tube it can be seen that
using a strain rate-dependent material causes a larger increase in the initial peak load
for a given impact velocity compared with a strain rate-independent material, even
for relatively low impact velocities. Such was also found for the triple-tapered tube
and frusta. Thus, the combined effects of strain rate and inertia appear to have an
enhancement effect on the initial peak load. This was also found for the Type II
structure defined in Chapter 2 (Su et al. 1995a; b). Thus, strain rate effects can have
significant influence on the dynamic response of structures which plastically deform
to absorb energy. Nevertheless, Figure 6.3-4 shows that even without strain rate-
dependency in the material model for the present tubes, inertia effects may have a
significant effect on the response for the range of impact velocities considered. Since
the tubes in the present study were modelled with a strain rate-dependent material,
namely hot rolled mild steel, the effects of strain rate are retained in the subsequent
analyses.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 170
1.6
1.5

Dynamic / quasi-static
1.4

initial peak load


1.3
1.2
1.1
1.0
0.9
0 5 10 15 20
Impact velocity (m/s)

Strain rate-dependent Strain rate-independent

Straight tube

1.8
1.7
Dynamic / quasi-static

1.6
initial peak load

1.5
1.4
1.3
1.2
1.1
1.0
0.9
0 5 10 15 20
Impact velocity (m/s)

Strain rate-dependent Strain rate-independent

Double taper
Figure 6.3-4 Relative effect of strain rate and inertia effects on the dynamic
response (here h = 1.5 mm, θ = 100 and M = 80 kg).

6.3.2 Effect of impact mass

The second loading variable of interest is the impact mass, M. The effect of impact
mass on the dynamic response of the straight and tapered tubes is summarised in
Figure 6.3-5 for the straight and double-tapered tubes, with h = 1.5 mm, θ = 100 and
V = 15 m/s. A mass range of 60 to 100 kg was chosen since this provided sufficient
kinetic energy to crush the tubes.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 171
200
175
Dynamic load (kN) 150
125
100
75
50
25
0
0 25 50 75 100 125 150 175 200 225 250
Deflection (mm)

M = 60 kg M = 80 kg M = 100 kg

Straight tube

125

100
Dynamic load (kN)

75

50

25

0
0 25 50 75 100 125 150 175 200 225 250 275
Deflection (mm)

M = 60 kg M = 80 kg M = 100 kg

Double taper
Figure 6.3-5 Effect of impact mass on the dynamic load-deflection response
(here h = 1.5 mm, θ = 100 and V = 15 m/s).

It can be seen that there is no significant change in the load-deflection response as


the mass is increased from 60 to 100 kg. The only difference observed is an increase
in maximum deflection associated with the increased impact (kinetic) energy of the
tube and loading rigid body as the impact mass increases. These results were also
observed for axially impacted aluminium alloy square tubes for the same impact
velocity and a similar range of impact masses (Langseth et al. 1999). The same
trends were also observed when typical vehicle masses are analysed (Bignell 2004).
Thus it appears that the dynamic or non-linear stiffness of both the straight and

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 172
tapered tubes is independent of the impact mass, such that the impact mass has no
influence on the internal forces set up in the tubes to resist the projectile when the
impact velocity is kept constant for the mass range considered.

6.3.3 Determining the presence of axial inertia effects

Thus far the influence of lateral inertia effects on the dynamic response of straight
and tapered rectangular tubes has been identified. Such effects arise due to the lateral
movement of the sidewalls at the instant of impact. Also of interest is the presence of
axial inertia affects, which can be identified by comparing the crush loads at the
impacted and fixed ends of the tube over time. The load at the fixed end is the
reaction load, while the load at the impacted end is the product of the acceleration of
the moving rigid body and its mass. The response is compared for a low impact
velocity (20 m/s) and a high impact velocity (60 m/s), for all tubes with a wall
thickness of 1.5 mm and a taper angle of 100. Though these velocities may be
relatively large for certain crashworthiness applications, they were chosen to capture
the effects of inertia on the tube response. Also, such values have been used by
others to study inertia effects on straight square tubes made of aluminium alloy
(Karagiozova & Jones 2004). The initial impact energy for both velocities was kept
constant at 16 kJ, by varying the impact mass.

The dynamic load response for the straight tube is shown in Figure 6.3-6 as a
function of time (measured after the tube initially contacts the impact surface). It
should be noted that all the load-deflection curves shown thus far and in subsequent
sections correspond to the magnitude of the load. In this section, however, the
nominal load is compared to show the direction of the acting force. For instance, a
negative load is one which acts down the tube’s axis, in a direction away from the
impacted end. Figure 6.3-6 shows that, for a relatively low impact velocity, the axial
loads at each end of the straight tube are virtually the same as the tube crushes,
except at the onset of impact where there is significant “retardation”. This indicates
that axial inertia effects are not significant for the large part of the response duration,
as was also found for axially impacted aluminium alloy square tubes over an impact
range of 5-25 m/s (Langseth et al. 1999). However, for a relatively high impact

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 173
velocity, in this case 60 m/s, Figure 6.3-6 shows that the load at each end of the tube
begin to differ slightly, and this may be due to the propagation and reflection of
elastic-plastic stress waves along the tube (Karagiozova & Jones 2004),
corresponding to axial inertia effects.

The load-time curves for the double-tapered tube at the same velocities are shown in
Figure 6.3-7.

50

0
0 0.005 0.01 0.015 0.02 0.025
Dynamic load (kN)

-50

-100

-150

-200

-250
Time (sec)

Fixed end Impacted end

V = 20 m/s

50

0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014
Dynamic load (kN)

-50

-100

-150

-200

-250
Time (sec)

Fixed end Impacted end

V = 60 m/s
Figure 6.3-6 Load-time curves for the straight tube (here h = 1.5 mm).

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 174
100
50
0

Dynamic load (kN)


-50 0 0.005 0.01 0.015 0.02 0.025

-100
-150
-200
-250
-300
Time (sec)

Fixed end Impacted end

V = 20 m/s

150
100
50
Dynamic load (kN)

0
-50 0 0.002 0.004 0.006 0.008 0.01 0.012
-100
-150
-200
-250
-300
Time (sec)

Fixed end Impacted end

V = 60 m/s
Figure 6.3-7 Load-time curves for the double-tapered tube
(here h = 1.5 mm and θ = 100).

Apart from at the initial impact, the load-time curves in Figure 6.3-7 are virtually the
same for a relatively low impact velocity. However, increasing the impact velocity to
60 m/s causes the loads to differ for most of the impact duration. Moreover,
comparing the straight and double-tapered tubes for an impact velocity of 60 m/s, it
appears that the initial peak load is the same at each end of the straight tube, whereas
for the double-tapered tube the peak load is higher at the impacted end. This has also
been found for other taper angles (Nagel & Thambiratnam 2004a). It could be, in the
case of tapered tubes, that wave propagation energy is being dissipated before the

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 175
reaction load at the fixed end is induced. The response of the triple-tapered tube and
frusta displayed similar trends as for the double-tapered tube for each impact
velocity. Thus, the results of this section show that axial inertia effects are not a
significant factor to account for in the design of straight and tapered rectangular
tubes at relatively low impact velocities, i.e. on the order of 20 m/s. For relatively
high impact velocities (around 60 m/s), the effects of axial inertia becomes more
significant and should not be neglected. Furthermore, the higher peak load at the
impacted end for tapered tubes should be estimated and used in design, due to the
effects which the corresponding deceleration may have on the protected structure and
its occupants.

The results of this section show that the dynamic response of straight and tapered
rectangular tubes under axial impact loading is significantly affected by the impact
velocity but not the impact mass, for the range of each loading parameter considered.
Both strain rate and inertia effects are present, which tend to increase the dynamic
crush loads and energy absorption of the tubes as the impact velocity increases. Thus,
these effects become important when controlling the energy absorption response of
straight and tapered rectangular tubes under axial impact loading, as will be further
discussed in the next section.

6.4 Effect of geometry parameters on the dynamic response


of straight and tapered rectangular tubes

Of interest is how the dynamic response of the straight and tapered rectangular tubes
is influenced by the various geometry parameters. This is examined for variations in
wall thickness, taper angle and number of tapers. For consistency the impact velocity
and mass are kept constant at 20 m/s and 80 kg, respectively. Such information
provides knowledge of how the tube response can be controlled using the geometry.

6.4.1 Effect of wall thickness

Figure 6.4-1 shows the dynamic mean load-deflection response for the straight and
double-tapered tubes, as the wall thickness is increased for a fixed taper angle of 100.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 176
250
225

Dynamic mean load (kN)


200
175
150
125
100
75
50
25
0
0 25 50 75 100 125 150 175 200 225 250 275
Deflection (mm)

h = 1.5 mm h = 2 mm h = 2.5 mm

Straight tube

225
200
Dynamic mean load (kN)

175
150
125
100
75
50
25
0
0 25 50 75 100 125 150 175 200 225 250 275 300
Deflection (mm)

h = 1.5 mm h = 2 mm h = 2.5 mm

Double taper
Figure 6.4-1 Effect of wall thickness on the dynamic mean load (here θ = 100).

It can be seen that an increase in wall thickness causes an increase in the mean load
up to a given deflection, which was also found for the triple-tapered tube and frusta.
As for quasi-static loading, this is due to the increased amount of material available
for plastic deformation and subsequent energy absorption. However, it may also be
due to larger lateral inertia forces in the tube walls as the wall thickness increases.
Figure 6.4-2 shows that increasing the wall thickness increases the dynamic absorbed
energy, as expected.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 177
16

Dynamic absorbed energy (kJ)


14
12
10
8
6
4
2
0
0 25 50 75 100 125 150 175 200 225 250 275
Deflection (mm)

h = 1.5 mm h = 2 mm h = 2.5 mm

Straight tube

16
Dynamic absorbed energy (kJ)

14
12
10
8
6
4
2
0
0 25 50 75 100 125 150 175 200 225 250 275 300
Deflection (mm)

h = 1.5 mm h = 2 mm h = 2.5 mm

Double taper
Figure 6.4-2 Effect of wall thickness on the dynamic absorbed energy
(here θ = 100).

The response in Figure 6.4-1 and Figure 6.4-2 is shown up to the maximum
deflection, dmax, at which the tube was brought to rest. It can be seen that dmax
decreases with increasing wall thickness, and this is due to the increasing energy
absorption and decreasing limit of compression of each fold as the wall thickness
increases. Figure 6.4-3 shows that the maximum deflection for a given wall thickness
also depends on the number of tapers and taper angle.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 178
300

Maximum deflection (mm)


250

200

150

100

50

0
1.5 2.0 2.5
Wall thickness (mm)

Straight Double taper Triple taper Frusta

Taper angle = 50
300
Maximum deflection (mm)

250

200

150

100

50

0
1.5 2.0 2.5
Wall thickness (mm)

Straight Double taper Triple taper Frusta


0
Taper angle = 10
300
Maximum deflection (mm)

250

200

150

100

50

0
1.5 2.0 2.5
Wall thickness (mm)

Straight Double taper Triple taper Frusta

Taper angle = 150


Figure 6.4-3 Effect of wall thickness on the maximum crush distance.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 179
Though the number of tapers and taper angle have less influence on the maximum
deflection than the wall thickness, it appears that dmax is greatest for the double-
tapered tube for a given wall thickness and taper angle. Such information provides a
guide as to how much crush zone is required to absorb a given amount of energy, for
straight and tapered tubes of equivalent cross-section dimensions.

6.4.2 Effect of taper angle

The effect of taper angle was studied for the double-tapered tube, triple-tapered tube
and frusta, for a constant wall thickness of 1.5 mm. The mean load-deflection
response as a function of the taper angle is shown in Figure 6.4-4 for the double-
tapered tube. The taper angle has no significant influence on the dynamic mean load
and hence energy absorption response for each tapered tube. However, increasing the
taper angle leads to a small reduction in the maximum crush distance.

120
Dynamic mean load (kN)

100

80

60

40

20

0
0 25 50 75 100 125 150 175 200 225 250 275 300
Deflection (mm)

Taper angle = 5 deg Taper angle = 10 deg Taper angle = 15 deg

Figure 6.4-4 Effect of taper angle on the dynamic mean load for the double-
tapered tube (here h = 1.5 mm).

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 180
6.4.3 Combined effects of wall thickness, taper angle and number of
tapers on the absorbed energy

So far the individual effect of wall thickness, taper angle and number of tapers on the
tube response has been determined. Figure 6.4-5 shows the effect on the dynamic
absorbed energy for combinations of wall thickness, taper angle and number of
tapers. The impact velocity is 20 m/s and the impact mass 80 kg, while the energy
absorbed up to 100 mm is shown.

The results show that the absorbed energy is influenced by the number of tapers, and
this influence appears to increase with increasing wall thickness. The results also
show that, for each wall thickness, the double-tapered tube absorbs the least energy
up to the given deflection, as was found for other deflections and impact velocities
(Nagel & Thambiratnam 2004b). For a wall thickness of 1.5 and 2 mm, the straight
tube, triple-tapered tube and frusta appear to absorb similar amounts of energy, while
the straight tube absorbs the most energy for a wall thickness of 2.5 mm.
Furthermore, as shown already, it can be seen that the taper angle has no significant
effect on the energy absorption response for a given number of tapers. Wall
thickness, on the other hand, has a significant effect on the response. Thus, Figure
6.4-5 provides a summary of the results of this section, and illustrates the relative
effect of each geometry parameter on the dynamic energy absorption. The results are
used in subsequent sections to determine the effects of these parameters on other
response parameters.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 181
6

Dynamic absorbed energy (kJ)


5

0
0 5 10 15
Taper angle (deg)

Straight Double taper Triple taper Frusta

(a) h = 1.5 mm
10
Dynamic absorbed energy (kJ)

9
8
7
6
5
4
3
2
1
0
0 5 10 15
Taper angle (deg)

Straight Double taper Triple taper Frusta

(b) h = 2 mm
16
Dynamic absorbed energy (kJ)

14
12
10
8
6
4
2
0
0 5 10 15
Taper angle (deg)

Straight Double taper Triple taper Frusta

(c) h = 2.5 mm
Figure 6.4-5 Dynamic absorbed energy vs. taper angle and number of tapers
(calculated at d = 100 mm).

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 182
6.4.4 Effect of wall thickness, taper angle and number of tapers on
other design parameters

Thus far the effect of the loading and geometry parameters on the load-deflection and
energy absorption-deflection response has been quantified. Also of interest is the
collective effect of these parameters on other measures of energy absorption
response, namely the:

(1) Dynamic initial peak load, Fbd


(2) Ratio of dynamic mean load to initial peak load, Fmd / Fbd
(3) Energy absorbed per unit crush length, Ecl, and
(4) Energy absorbed per unit mass, Em

These parameters have been introduced for quasi-static loading in Chapter 5 and are
adopted here for dynamic impact loading.

6.4.4.1 Dynamic initial peak load, Fbd

The effect of each geometry parameter on the dynamic initial peak load for an impact
velocity of 20 m/s is shown in Figure 6.4-6. As for quasi-static loading, the
introduction of a taper greatly reduces the initial peak load. Furthermore, the peak
load decreases as the taper angle increases for the double- and triple-tapered tubes,
however the load is relatively independent of taper angle for the frusta. For a given
wall thickness and taper angle, the initial peak load is generally highest for the triple-
tapered tube compared with tubes having two and four tapers. This may be due to the
triple-tapered tube’s unsymmetrical cross-section providing greater resistance to
initial buckling. On the other hand, the frusta has the lowest initial peak load for the
range of wall thickness and taper angle considered. Finally, as expected, the initial
peak load increases with increasing wall thickness. Overall, Figure 6.4-6 shows that
the initial peak load is more influenced by the wall thickness and number of tapers
than the taper angle. Nevertheless, it appears that all three parameters can be used to
control the initial peak load under dynamic impact loading.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 183
250

200
Initial peak load (kN)
150

100

50

0
0 5 10 15
Taper angle (deg)

Straight Double taper Triple taper Frusta

(a) h = 1.5 mm
300

250
Initial peak load (kN)

200

150

100

50

0
0 5 10 15
Taper angle (deg)

Straight Double taper Triple taper Frusta

(b) h = 2 mm
400
350
Initial peak load (kN)

300
250
200
150
100
50
0
0 5 10 15
Taper angle (deg)

Straight Double taper Triple taper Frusta

(c) h = 2.5 mm
Figure 6.4-6 Effect of each geometry parameter on the dynamic initial peak load
(here V = 20 m/s and M = 80 kg).

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 184
6.4.4.2 Ratio of dynamic mean load to initial peak load, Fmd / Fbd

Figure 6.4-7 shows the effect of each parameter on the ratio of dynamic mean load to
initial peak load, where the mean load has been calculated up to a deflection of 100
mm. This ratio is also referred to as the crush force efficiency, and is shown as a
percentage in Figure 6.4-7. It is desirable to maximise this ratio, particularly from a
crashworthiness point of view, since peak loads during an impact event correspond to
peak decelerations which may be detrimental to the structure being protected or to
the safety of the occupants it contains.

An important observation from Figure 6.4-7 is that the ratio is higher for the tapered
tubes than for the straight tube, thus tapered tubes would be preferable for use as
energy absorbers in regard to maximising the crush force efficiency. Furthermore,
the ratio is also a function of the number of tapers, and thus it would appear that
frusta generally have a higher crush force efficiency compared with the double- and
triple-tapered tubes. This may be due to the frusta having the lowest initial peak load
under dynamic loading, as was shown in Figure 6.4-6. Double-tapered tubes have the
lowest efficiency, since they absorb the least energy for a given deflection compared
with the other tubes, as was shown in Figure 6.4-5. Finally, it appears that the crush
force efficiency generally increases as the wall thickness increases.

6.4.4.3 Energy absorbed per unit crush length, Ecl

The energy absorbed per unit crush length, Ecl is useful when the crush zone is
restricted. Ecl was calculated using the impact energy (16 kJ) and maximum
deflection for each combination of geometry parameters, as shown in Figure 6.4-8.
The impact velocity is 20 m/s and the impact mass is 80 kg. Evidently, Ecl increases
with increasing wall thickness for both straight and tapered tubes. This is due to an
observed decrease in the maximum crush distance as the wall thickness increases,
which is associated with the decreasing limit of compression of each fold (Nagel &
Thambiratnam 2004a). Practically, it would appear that increasing the wall thickness
would be an effective means of maximising energy absorption when the available
crush zone is limited. It can also be seen that Ecl is influenced by the taper angle and
number of tapers, the effect of the latter being more significant.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 185
Dynamic mean load / Initial peak load
80
70
60
(%) 50
40
30
20
10
0
0 5 10 15
Taper angle (deg)

Straight Double taper Triple taper Frusta

(a) h = 1.5 mm
Dynamic mean load / initial peak load

80
70
60
50
(%)

40
30
20
10
0
0 5 10 15
Taper angle (deg)

Straight Double taper Triple taper Frusta

(b) h = 2 mm
Dynamic mean load / initial peak load

80
70
60
50
(%)

40
30
20
10
0
0 5 10 15
Taper angle (deg)

Straight Double taper Triple taper Frusta

(c) h = 2.5 mm
Figure 6.4-7 Effect of each geometry parameter on the dynamic crush force
efficiency (here V = 20 m/s and M = 80 kg).

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 186
Energy absorbed per unit crush length
62
61
61

(kJ / m)
60
60
59
59
58
0 5 10 15
Taper angle (deg)

Straight Double taper Triple taper Frusta

(a) h = 1.5 mm
Energy absorbed per unit crush length

88
86
84
82
(kJ / m)

80
78
76
74
72
70
0 5 10 15
Taper angle (deg)

Straight Double taper Triple taper Frusta

(b) h = 2 mm
Energy absorbed per unit crush length

150
145
140
135
130
(kJ / m)

125
120
115
110
105
100
0 5 10 15
Taper angle (deg)

Straight Double taper Triple taper Frusta

(c) h = 2.5 mm
Figure 6.4-8 Energy absorbed per unit crush length vs. taper angle, number of
tapers and wall thickness.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 187
According to Figure 6.4-8, for each wall thickness, the double-tapered tube generally
absorbs the least energy per unit crush length since the tube requires a greater
distance to be brought to rest. The triple-tapered tube appears to have the highest Ecl
for wall thicknesses of 2 and 2.5 mm, since it generally requires the lowest crush
distance to absorb the impact energy. All the tubes “bottomed out” for a wall
thickness of 1.5 mm, such that the trends of Ecl for this wall thickness are somewhat
different to those for the other two thicknesses.

The following practical conclusions can be drawn from these observations:

(1) For a given number of tapers and taper angle, increasing the wall thickness
maximises the dynamic energy absorbed within a given crush distance.
(2) For a given wall thickness, the double-tapered tube generally absorbs the
least energy within a given crush distance, while the triple-tapered tube
generally absorbs the most (when “bottoming out” does not occur).

6.4.4.4 Energy absorbed per unit mass, Em

As described in Chapter 2, the energy absorbed per unit mass is a convenient


measure when the weight of the energy absorbing system is important. The
dependence of Em on the various geometry parameters is shown in Figure 6.4-9. Em is
calculated based on the energy absorbed up to a deflection of 100 mm, and the
original mass of each undeformed tube. The impact velocity is 20 m/s and the impact
mass is 80 kg.

Since the tapered tubes have larger cross-section dimensions that the straight tube,
increasing the number of tapers and taper angle increases the tube mass. Thus,
straight rectangular tubes absorb more energy per unit mass than tapered rectangular
tubes, as shown in Figure 6.4-9. For each tube, Em increases with increasing wall
thickness yet decreases with increasing taper angle. Finally, the double- and triple-
tapered tubes absorb more energy per unit mass than the frusta, for a given wall
thickness and taper angle.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 188
10

Energy absorbed per unit mass


9
8
7
6

(kJ / kg)
5
4
3
2
1
0
1.5 2.0 2.5
Wall thickness (mm)

Straight Double taper Triple taper Frusta


0
(a) θ = 5
10
Energy absorbed per unit mass

9
8
7
6
(kJ / kg)

5
4
3
2
1
0
1.5 2.0 2.5
Wall thickness (mm)

Straight Double taper Triple taper Frusta

(b) θ = 100
10
Energy absorbed per unit mass

9
8
7
6
(kJ / kg)

5
4
3
2
1
0
1.5 2.0 2.5
Wall thickness (mm)

Straight Double taper Triple taper Frusta

(c) θ = 150
Figure 6.4-9 Energy absorbed per unit mass vs. wall thickness, number of
tapers and taper angle.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 189
6.5 Comparison between static and dynamic response

Thus far the effect of the various loading and geometry parameters on the dynamic
response has been examined. Also of interest is how these parameters influence the
dynamic response compared with the static response. Such is examined in this
section using the Dynamic Amplification Factor, or DAF. As described in Chapter 2,
the DAF is the ratio of energy absorbed under dynamic loading to the energy
absorbed under quasi-static loading. This measure is useful for design purposes
because it allows the effect of various parameters on the dynamic response to be
quantified with respect to the quasi-static response.

The effect of impact velocity, wall thickness, taper angle and number of tapers on the
DAF was studied for a constant impact mass of 200 kg. This represents an increased
mass from before so as to allow sufficient impact energy to crush most of the tube
for the range of geometry parameters used in this section.

6.5.1 Effect of impact velocity on the DAF

Figure 6.5-1 shows the effect of impact velocity on the DAF for the straight and
double-tapered tubes, with h = 1.5 mm and θ = 100. The same scale has been used in
each graph to illustrate the relative effect for each tube. For each tube, including the
triple-tapered tube and frusta, an increase in impact velocity causes an increase in the
DAF up to a given deflection. This was also found for axially impacted aluminium
alloy square tubes (Langseth et al. 1999), and was attributed to the increased lateral
inertia forces set up in the sidewalls as the velocity increased. For the tubes in the
present study, the increased DAF with velocity may therefore be due to inertia
effects, as well as strain rate and increased input energy. Practically, therefore,
increasing the impact velocity of straight and tapered rectangular tubes of equivalent
dimensions increases the amplification of absorbed energy. This observation has
been verified with other geometry and loading parameters as well (Nagel &
Thambiratnam 2004b).

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 190
2.25

Dynamic / quasi-static
absorbed energy
1.75

1.5

1.25

1
0 25 50 75 100 125 150 175 200 225
Deflection (mm)

V = 10 m/s V = 20 m/s

Straight tube

2.25

2
Dynamic / quasi-static
absorbed energy

1.75

1.5

1.25

1
0 25 50 75 100 125 150 175 200 225
Deflection (mm)

V = 10 m/s V = 20 m/s

Double taper
Figure 6.5-1 Effect of impact velocity on the DAF (here h = 1.5 mm and θ = 100).

The DAF shown in Figure 6.5-1 has been plotted versus axial deflection. For the
straight tube the DAF is a decreasing function with increasing deflection. Once
again, this was also found for axially impacted aluminium alloy square tubes
(Langseth et al. 1999). This is due to the diminishing influence of inertia and strain
rate effects as the tube is brought to rest, such that the dynamic and static responses
converge as the crushing process slows down (Nagel & Thambiratnam 2004a).
However, for the double-tapered tube it can be seen that the DAF is almost constant
with respect to axial deflection, and the DAF is reduced compared with the straight

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 191
tube. This was also found for the triple-tapered tube and frusta. Thus, it appears that
the influence of lateral inertia in tapered tubes is less than in straight tubes of
equivalent cross-section dimensions. A similar response was also found when
introducing an initial geometric imperfection into an initially straight tube (Langseth
& Hopperstad 1996). Pre-buckling initially straight square aluminium tubes lead to
an almost constant ratio between the dynamic and quasi-static mean loads with
respect to axial deflection, and the mean load ratio was reduced compared with an
initially straight tube. In the same study, pre-buckling also reduced the initial peak
load and increased the maximum axial deflection compared with an initially straight
tube. These results are similar to those for the Type II structure studied by Su et al.
(1995b), in which the initial peak load was found to decrease as the initial kink angle
of the structure increased (refer to Section 2.1.1.1 of Chapter 2. Thus, as identified in
Section 6.3.1, it appears that the response of the straight and tapered rectangular
tubes is similar to that of a Type II structure.

As a practical implication of these results, the taper in a tube face may be viewed as
an inherent imperfection which reduces the influence of inertia effects. This
coincides with the reduction in initial peak load when a taper is introduced, as shown
earlier in this chapter. That the inertia effects are less significant for tapered tubes
than for straight tubes would be beneficial for impact energy absorption applications
where the “strengthening” behaviour of the absorber due to inertia effects needs to be
minimised.

6.5.2 Effect of wall thickness on the DAF

Figure 6.5-2 shows the effect of wall thickness on the DAF for the straight and
double-tapered tubes, with θ = 100 and V = 20 m/s. For the straight tube, the DAF
clearly decreases as the wall thickness increases. Thus, the amplification of absorbed
energy appears to be greater for “thin” straight tubes than for “thick” straight tubes.
This effect has been qualitatively explained using the ratio of dynamic mean load to
quasi-static mean load (Langseth & Hopperstad 1996), in which it was found that
this ratio is proportional to h −2 / 3 . Thus, keeping all other parameters constant, an
increase in wall thickness will reduce the mean load ratio, which is effectively the
DAF.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 192
2.25

Dynamic / quasi-static
absorbed energy
1.75

1.5

1.25

1
0 25 50 75 100 125 150 175 200 225
Deflection (mm)

h = 1.5 mm h = 2 mm

Straight tube

2.25

2
Dynamic / quasi-static
absorbed energy

1.75

1.5

1.25

1
0 25 50 75 100 125 150 175 200 225
Deflection (mm)

h = 1.5 mm h = 2 mm

Double taper
Figure 6.5-2 Effect of wall thickness on the DAF (here θ = 100 and V = 20 m/s).

However, it can be seen from Figure 6.5-2 that the DAF for a given wall thickness
and deflection is less for the double-tapered tube than for the straight tube, while the
DAF is less dependent on wall thickness for the double-tapered tube. This is also the
case for the triple-tapered tube and frusta, and may once again be due to the presence
of a taper influencing the inertia effects.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 193
6.5.3 Effect of taper angle on the DAF

Figure 6.5-3 shows the effect of taper angle on the DAF for each tube, with h = 1.5
mm and V = 20 m/s. Generally, it is difficult to determine the effect that increasing
the taper angle has on the DAF. However, it generally appears that the DAF reduces
slightly when the taper angle is increased, and this coincides with the effect of
increasing the wall thickness.

6.5.4 Effect of the number of tapers on the DAF

The effect of the number of tapered sides on the DAF is illustrated in Figure 6.5-4
and Figure 6.5-5. For a constant wall thickness and taper angle, the DAF is less for
the tapered tubes than for the straight tube, as shown in Figure 6.5-4 for impact
velocities of 10 and 20 m/s. The tapered tubes generally lie in the same DAF range
for each velocity. Increasing the wall thickness, however, reduces the DAF for the
straight tube relative to that of the tapered tubes (Figure 6.5-5). Finally, it was found
that the DAF for the tapered tubes is approximately the same, irrespective of the
taper angle or number of tapers.

In conclusion, it appears that the amplification of absorbed energy under dynamic


loading is most influenced by the impact velocity and introducing a taper. In
particular, therefore, tapering one or more of the sides of a rectangular tube can be
used to control the amount of energy absorbed under dynamic loading relative to that
under quasi-static loading. Also, since the DAF for tapered tubes is less sensitive to
the various loading and geometry parameters than for straight tubes, it may be
preferable to use tapered tubes when inertia effects and the associated load
enhancement need to be minimised. Further study of the DAF for straight and
tapered rectangular tubes is provided in Nagel and Thambiratnam (2004b).

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 194
2.25

Dynamic / quasi-static
absorbed energy
1.75

1.5

1.25

1
0 25 50 75 100 125 150 175 200 225
Deflection (mm)

a = 5 deg a = 15 deg

Double taper
2.25

2
Dynamic / quasi-static
absorbed energy

1.75

1.5

1.25

1
0 25 50 75 100 125 150 175 200 225
Deflection (mm)

a = 5 deg a = 15 deg

Triple taper
2.25

2
Dynamic / quasi-static
absorbed energy

1.75

1.5

1.25

1
0 25 50 75 100 125 150 175 200 225
Deflection (mm)

a = 5 deg a = 15 deg

Frusta
Figure 6.5-3 Effect of taper angle on the DAF (here h = 1.5 mm and V = 20 m/s).

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 195
1.75
Dynamic / quasi-static
absorbed energy

1.5

1.25

1
0 25 50 75 100 125 150 175 200 225
Deflection (mm)

Straight Double taper Triple taper Frusta

V = 10 m/s
2.25

2
Dynamic / quasi-static
absorbed energy

1.75

1.5

1.25

1
0 25 50 75 100 125 150 175 200 225
Deflection (mm)

Straight Double taper Triple taper Frusta

V = 20 m/s
Figure 6.5-4 Effect of the number of tapers on the DAF
(here h = 1.5 mm and θ = 100).

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 196
2.25

Dynamic / quasi-static
absorbed energy
1.75

1.5

1.25

1
0 25 50 75 100 125 150 175 200 225
Deflection (mm)

Straight Double taper Triple taper Frusta

h = 1.5 mm

2.25

2
Dynamic / quasi-static
absorbed energy

1.75

1.5

1.25

1
0 25 50 75 100 125 150 175 200 225
Deflection (mm)

Straight Double taper Triple taper Frusta

h = 2 mm
Figure 6.5-5 Effect of the number of tapers on the DAF
(here θ = 100 and V = 20 m/s).

6.6 Use of a factorial study for energy absorber design

The analyses thus far in this chapter have been based on a one-factor-at-a-time
methodology, where the effect of one parameter on the tube response is investigated
by varying this single parameter and keeping all remaining parameters fixed. Hence,
the method provides an estimate of the effect of a single variable at selected fixed
values of the other variables. However, for such an estimate to be generally

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 197
meaningful it must be assumed that the effect is the same at other values of the other
variables. If, in reality, the effect is not the same, then the variables are said to
“interact”, and such interaction cannot be detected and estimated using the one-
factor-at-a-time approach to analysing energy absorber response.

A factorial study was used to investigate the main and interaction effects between
pertinent geometry and loading parameters on the impact response of the straight and
tapered tubes. If there are n input parameters, the study requires 2n tests to be
conducted by varying each parameter between a high and a low value. Three input
parameters were varied: wall thickness, taper angle and impact velocity, giving 8
tests to be performed. This was carried out for both straight tubes and tubes having
two, three and four tapered sides. The impact mass was increased to 275 kg to allow
sufficient deflection to be reached, while the impact mass was not varied as a
parameter since it has no significant influence on the response up to a given
deflection for the range of parameter values studied. The initial peak load and mean
load were the response parameters, the latter calculated at a deflection of 100 mm.

The simulation tests conducted for the different combinations of high and low values
of each input parameter are shown in Table 6.6-1.

Table 6.6-1 Levels used in the factorial study.

Level h θ V
(mm) (deg) (m/s)
Low 1.5 5 10
High 2.5 15 20

The relative effect of each parameter on the initial peak and mean loads is shown in
Figure 6.6-1 for each tube. Results are shown as a percentage of the average force of
all combinations of high and low values for each tube. Obviously, the wall thickness
has the most influence on the initial peak and mean load for both straight and tapered
tubes. Furthermore, the influence of the wall thickness on the initial peak load
becomes more significant as the number of tapers increases.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 198
100

80

% of average dynamic
60

initial peak load


40

20

0
h θ V h-θ h-V θ-V h-θ-V
-20

-40
Geometry parameters

Straight Double taper Triple taper Frusta

(a)

120
% of average dynamic mean load

100
80
60
40
20
0
-20 h θ V h-θ h-V θ-V h-θ-V
-40
-60
Geometry parameters

Straight Double taper Triple taper Frusta

(b)
Figure 6.6-1 Relative effect of wall thickness, taper angle and impact velocity on
the (a) initial peak load and (b) mean load for each tube.

According to Figure 6.6-1, the initial peak and mean load for tapered tubes are
generally more sensitive to impact velocity than to taper angle. Furthermore, the
influence of taper angle on the initial peak load is not significantly dependent on the
number of tapers. On the other hand, the influence of taper angle on the mean load is
significantly dependent on the number of tapers. Overall, it appears that the mean
load is most sensitive to the input parameters for the triple-tapered tube followed by
the double-tapered tube. Positively, this implies that the response of the double- and
triple-tapered tubes can be more effectively controlled when optimising the design

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 199
using the available parameters. On the other hand, it implies that tighter tolerances
may be required when manufacturing these tubes for impact energy absorption, so as
to achieve the desired mean load response.

The interaction effects are only significant for the double- and triple-tapered tubes,
particularly for the mean load. For instance, for these two tubes, the effect of wall
thickness on the mean load becomes more significant as the taper angle and impact
velocity increase, and the effect of taper angle becomes more significant as the
impact velocity increases.

The results of the factorial study provide a convenient way of determining the
relative influence of each geometry and loading parameter on the tube response, and
in effect summarises the pertinent findings of this chapter. For example, it was
shown earlier in this chapter that the dynamic initial peak load increases with
increasing wall thickness and velocity, yet decreases with increasing taper angle.
This is illustrated in Figure 6.6-1 by the positive effect of wall thickness and velocity
on the initial peak load, and negative effect of taper angle.

6.7 Deformation modes of straight and tapered rectangular


tubes under axial dynamic impact loading

As highlighted in Chapter 5, the deformation mode of thin-walled tubes is important


since it can significantly influence the amount of energy absorbed. This is
particularly the case under dynamic impact loading, since impact mass and velocity
can influence the deformation mode and the final deformed profile of the tube. This
section examines the deformation modes of the straight and tapered rectangular tubes
under axial impact loading. In the first case the effect of increasing the impact energy
is studied, while in the second case the effect of impact velocity is studied for
constant impact energy.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 200
6.7.1 Effect of impact energy on the deformation mode

Firstly the deformation modes of the straight and tapered rectangular tubes were
studied under high and low impact energies. The low impact energy selected was 10
kJ (corresponding to V = 20 m/s and M = 50 kg), and the high impact energy was
20.25 kJ (corresponding to V = 90 m/s and M = 5 kg). The mass values were chosen
such that there was sufficient kinetic energy to crush the tubes without bottoming-out
occurring. Similar velocity values have been used by others to determine the
influence of inertia effects on the deformation response of straight square aluminium
alloy tubes under axial impact loads (Karagiozova & Jones 2004). For the present
study, the straight and tapered tubes were compared for a wall thickness of 1.5 mm
and a taper angle of 100.

Figure 6.7-1 and Figure 6.7-2 show the deformation profiles of the straight and
tapered tubes at various deflections. For an impact velocity of 20 m/s, the straight
and tapered tubes deform via dynamic progressive buckling with sequential fold
formation starting at the impacted end. However, increasing the velocity to 90 m/s
causes a buckle to form near the fixed end of the straight tube, such that buckling
occurs at both ends of the tube. For each of the tapered tubes a wrinkle has formed
near the fixed end of each tapered face, however a higher impact velocity was
required to cause this wrinkle to localise into a buckle. This observation may be
explained in the light of the inertia characteristics of the tubes. The larger width at
the fixed end of the tapered tube may cause larger lateral inertia forces which tend to
support the unbuckled shape, thus preventing buckling from occurring at this end.
Thus, a higher impact energy (and hence force) is required to cause buckling near the
base of the tapered tube, and so tapered tubes prove advantageous in that they
promote sequential buckling to occur from the impacted end. This would be desirable
in a practical sense since it provides a more controlled deformation mode.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 201
d = 128 mm dmax = 203 mm d = 112 mm dmax = 232 mm
Initial
V = 20 m/s V = 90 m/s

d = 135 mm dmax = 231 mm d = 107 mm dmax = 221 mm


Initial
V = 20 m/s V = 90 m/s

d = 148 mm dmax = 204 mm d = 101 mm dmax = 280 mm


Initial
V = 20 m/s V = 90 m/s

Figure 6.7-1 Deformation profiles for the straight tube, double-tapered tube and
frusta at high and low impact velocities (h = 1.5 mm and θ = 100).

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 202
d = 145 mm dmax =191 mm
Initial
V = 20 m/s

d = 102 mm dmax = 187 mm

V = 90 m/s

Figure 6.7-2 Deformation profiles for the triple-tapered tube at high and low
impact velocities (h = 1.5 mm and θ = 100).

6.7.2 Effect of velocity on the deformation mode- constant impact


energy

In previous studies it has been shown that the impact velocity significantly influences
the maximum crush distance caused by similar impact energies for axially impacted
square and circular tubes (Karagiozova & Jones 2004; 2001). The deformation
behaviour of the straight and tapered rectangular tubes in the present study was
investigated by increasing the impact velocity while keeping the impact energy
constant at 9 kJ. The wall thickness of the tubes was 1.5 mm, while the taper angle
was 100. The final deformed profiles for the straight tube, double-tapered tube and
frusta are shown in Figure 6.7-3.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 203
Initial V = 15 m/s V = 30 m/s V = 60 m/s V = 90 m/s

Initial V = 15 m/s V = 30 m/s V = 60 m/s V = 90 m/s

Initial V = 15 m/s V = 30 m/s V = 60 m/s V = 90 m/s

Figure 6.7-3 Final deformed profiles for the straight tube, double-tapered tube
and frusta for a constant impact energy (h = 1.5 mm and θ = 100).

Comparing the deformation profiles of the straight tube in Figure 6.7-3, it can be
seen that the profile corresponding to a velocity of 90 m/s consists of small wrinkles
down the undeformed length of the tube, in addition to the localised lobes at the
impacted end. This is a phenomenon called dynamic plastic buckling (Jones 1989,
Section 2.1.1.1 of Chapter 2), and arises due to inertia effects associated with the

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 204
higher impact velocity. Such behaviour has also been observed for axially impacted
aluminium alloy square tubes (Karagiozova & Jones 2004; Langseth et al. 1999) and
circular tubes (Karagiozova & Jones 2001; Karagiozova et al. 2000). However, there
appear to be no such wrinkles on either the double-tapered tube, the frusta or the
triple-tapered tube, which is shown in Figure 6.7-4. This may be due to the larger
width of the tapered tubes causing larger lateral inertia forces which tend to prevent a
rapid development of lateral displacements for all impact velocities.

Initial V = 15 m/s V = 30 m/s

V = 60 m/s V = 90 m/s

Figure 6.7-4 Final deformed profiles for the triple-tapered tube for a constant
impact energy (h = 1.5 mm and θ = 100).

From Figure 6.7-3 and Figure 6.7-4 it can be seen that the impact velocity
significantly influences the maximum crush distance caused by the same impact
energy, for both straight and tapered tubes. This is due to two factors. Firstly, a
higher impact velocity invokes larger lateral inertia forces which support the
unbuckled shape for a longer time. Secondly, strain rate effects for higher impact

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 205
velocities cause an increase of the flow stress of the shell, thus stiffening the
structure. As a result, a larger impact energy would be required to cause bottoming-
out of each tube when subjected to a higher impact velocity. In other words, the tubes
can absorb more impact energy when this energy is applied at a higher velocity. This
is illustrated in Figure 6.7-5, which shows the energy absorbed as a function of
deflection and impact velocity for the straight and double-tapered tubes. Results are
shown up to the maximum deflection in each case.

9
8
Dynamic absorbed energy (kJ)

7
6
5
4
3
2
1
0
0 25 50 75 100 125 150 175 200 225
Deflection (mm)

V = 15 m/s V = 30 m/s V = 60 m/s V = 90 m/s

Straight tube

9
8
Dynamic absorbed energy (kJ)

7
6
5
4
3
2
1
0
0 25 50 75 100 125 150 175 200 225
Deflection (mm)

V = 15 m/s V = 30 m/s V = 60 m/s V = 90 m/s

Double taper
Figure 6.7-5 Influence of impact velocity on the energy absorbed up to a given
deflection (here h = 1.5 mm, θ = 100 and impact energy = 9 kJ).

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 206
For each tube the amount of energy absorbed up to a given deflection increases as the
impact velocity increases, as was also found for the triple-tapered tube and frusta.
Furthermore, the maximum crush distance decreases as the impact velocity increases.

6.8 Impact pulse loading

6.8.1 Introduction

Full scale dynamic testing of vehicle crashworthiness is frequently supplemented


using mathematical models and laboratory tests (Huang 2002). Such tests are used to
predict changes in overall safety performance as vehicle structural and occupant
restraint parameters are varied. This objective can be achieved using crash pulse
characterisation, in which the crash pulse time history is represented in a simplified
manner while retaining the necessary response parameters to allow parametric study
of the crash performance. The crash pulse is defined using response parameters
which describe the physical events during the crash such as velocity change and
impact duration. The half-sine curve is commonly used as a crash pulse
approximation, and can be defined by the following acceleration-time curve (Chan
2000):

π∆V  πt 
a (t ) =  sin  (6.8-1)
2T  T

where ∆V is the speed change or area under the curve, T is the pulse duration, and 0 ≤
t ≤ T. In the case of tube crushing considered here, the tube is brought to rest such
that the speed change in Equation 6.8-1 can be replaced by the initial velocity, V.
Equation 6.8-1 can be used to study the effect of speed change and pulse duration on
the impact response. The pulse magnitude depends on the speed change and impact
duration, while the pulse duration depends on the mass and stiffness of the impact
surface (Bignell 2004). For instance, a rigid barrier has a relatively high stiffness and
leads to a short impact duration, thus causing a larger pulse magnitude. A deformable
barrier, on the other hand, has a relatively low stiffness and causes an impact of
longer duration and a lower pulse magnitude. Thus, the effect of different surfaces on

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 207
the impact response can be studied by varying the impulse duration. The finite
element model used for dynamic loading was modified by constraining the loading
rigid body to translate with an acceleration-time curve defined by Equation 6.8-1,
during the tube crushing process.

6.8.2 Effect of impact surface properties

Using a constant impact mass it is possible to study the effect of different impact
surfaces, by varying the pulse duration. This was carried out for impact velocities of
10, 20 and 30 m/s, as used in the dynamic impact analyses, while the wall thickness
was 1.5 mm and the taper angle was 100. Three pulse durations were used: 20, 30 and
40 ms, which are typical of impact tests on thin-walled tubes for the given loading
conditions and tube geometries. The study sought to determine whether straight or
tapered tubes were more sensitive to different impact surfaces, and whether the
number of tapers affected this sensitivity. Comparisons were made using the energy
absorbed under pulse loading, Ep, as the response parameter, which is shown
calculated up to 100 mm in the following graphs. The effect of strain rate was
included as in the previous impact analyses. Figure 6.8-1 shows the input
acceleration-time curves for pulse durations of 20, 30 and 40 ms. It can be seen that
the peak acceleration increases with increasing velocity, yet decreases with
increasing pulse duration.

Figure 6.8-2 shows that, except for the triple-tapered tube with V = 10 m/s, the
energy absorbed up to a given deflection decreases with increasing pulse duration,
and increases with increasing impact velocity. This was also found for the frusta.
Thus, the energy absorbed by the tubes generally decreases as the stiffness of the
impact surface decreases. Once again, strain rate and inertia effects appear to account
for the increasing energy absorption as the velocity increases.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 208
2500

Acceleration, a(t) (m/s )


2000

2
1500

1000

500

0
0 5 10 15 20
Time, t (msec)

V = 10 m/s V = 20 m/s V = 30 m/s

T = 20 ms
1800
1600
Acceleration, a(t) (m/s )
2

1400
1200
1000
800
600
400
200
0
0 5 10 15 20 25 30
Time, t (msec)

V = 10 m/s V = 20 m/s V = 30 m/s

T = 30 ms
1400
1200
Acceleration, a(t) (m/s )
2

1000
800
600
400
200
0
0 5 10 15 20 25 30 35 40
Time, t (msec)

V = 10 m/s V = 20 m/s V = 30 m/s

T = 40 ms
Figure 6.8-1 Input acceleration-time curves for pulse durations of 20, 30
and 40 ms.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 209
Energy absorbed under pulse loading
1.06
1.05
1.04
1.03
1.02
(kJ)

1.01
1.00
0.99
0.98
0.97
20 30 40
Pulse duration, T (msec)

V = 10 m/s V = 20 m/s V = 30 m/s

Straight tube
Energy absorbed under pulse loading

0.96
0.94
0.92
0.90
0.88
(kJ)

0.86
0.84
0.82
0.80
0.78
20 30 40
Pulse duration, T (msec)

V = 10 m/s V = 20 m/s V = 30 m/s

Double taper
Energy absorbed under pulse loading

1.12
1.11
1.10
1.09
1.08
(kJ)

1.07
1.06
1.05
1.04
1.03
1.02
20 30 40
Pulse duration, T (msec)

V = 10 m/s V = 20 m/s V = 30 m/s

Triple taper
Figure 6.8-2 Energy absorbed under pulse loading(here h = 1.5 mm and θ =
100).

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 210
The effect of the impact surface properties on the energy absorption response of the
different tube geometries was compared by analysing the reduction of absorbed
energy as the pulse duration is increased from 20 to 40 ms, as shown in Figure 6.8-3.
Clearly, the double-tapered tube has the highest sensitivity to the properties of the
impact surface, since it has the most reduction in absorbed energy for each velocity.
Furthermore, the impact surface properties have more effect on the response of each
tube as the impact velocity increases. Such results provide a guide as to how each
tube behaves when impacted by different surfaces.

90
Reduction of absorbed energy (J)

80
70
60
50
40
30
20
10
0
10 20 30
Impact velocity, V (m/s)

Straight Double taper Triple taper Frusta

Figure 6.8-3 Effect of surface properties (shown as reduction of absorbed


energy when the pulse duration increases from 20 to 40 ms).

6.9 Conclusions and design guidelines

This chapter has investigated the response of straight and tapered mild steel
rectangular tubes subjected to axial dynamic impact loading. The geometry
parameters investigated were the tube wall thickness, taper angle and number of
tapers. The loading parameters were the impact velocity, mass and duration. The
pertinent findings of the chapter and the design guidelines which can be drawn from
them are summarised below.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 211
(1) Both strain rate and inertia effects influence the response of straight and
tapered rectangular tubes under axial impact loading. Furthermore, these
effects become more significant as the impact velocity increases, leading to
an increase in the initial peak load, mean load and energy absorption
compared with that under quasi-static loading. Thus, while quasi-static
analyses may be used as an initial tool to compare the response of different
tube geometries and dimensions, this should be followed by dynamic
analyses for the impact energies expected in service. This way the dynamic
amplifications can be accounted for.

(2) The influence of lateral inertia in tapered rectangular tubes is less than in
straight rectangular tubes of equivalent cross-section dimensions. Thus,
tapered tubes would be beneficial for impact energy absorption applications
where the dynamic energy amplification and associated increase in crush
loads due to inertia effects needs to be minimised.

(3) Axial inertia effects (corresponding to elastic-plastic stress wave propagation)


are not significant in either straight or tapered rectangular tubes at relatively
low impact velocities (i.e. those on the order of 20 m/s). For relatively high
impact velocities (around 60 m/s), the effect of axial inertia becomes more
significant and should not be neglected. Furthermore, for tapered tubes the
peak load is highest at the impacted end for both relatively low and high
impact velocities. Therefore, in design calculations the load at the impacted
end should be used in preference to the load at the fixed end so as to avoid
under-prediction of the expected service loads.

(4) The energy absorption response of straight and tapered rectangular tubes
under axial impact loading is influenced by the wall thickness, taper angle
and number of tapered sides. Thus, these geometry parameters may be used
to control the response. However, the wall thickness has the most influence
on the initial peak and mean load for both straight and tapered rectangular
tubes, while the triple-tapered tube generally absorbs the most energy within
a given crush distance.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 212
(5) As for quasi-static loading, tapered tubes have a higher crush force efficiency
than straight tubes under axial impact loading, and this response parameter
generally increases with increasing wall thickness. Thus tapered tubes would
be preferable for use as energy absorbers when the initial peak loads and
corresponding decelerations must be minimised, such as in crashworthiness
applications.

(6) When the impact energy absorbed within in a given crush distance needs to
be maximised, the following guidelines should be followed:

a. For a given number of tapers and taper angle, increasing the wall
thickness maximises the dynamic energy absorbed within in a given
crush distance.
b. For a given number of tapers and wall thickness, the taper angle has
no significant effect on the energy absorption response.
c. For tapered tubes of a given wall thickness and taper angle, the
double-tapered tube generally absorbs the least energy within a given
crush distance, while the triple-tapered tube generally absorbs the
most.

(7) When the impact energy absorption needs to be maximised for a given
absorber mass, the following guidelines should be followed:

a. Straight rectangular tubes absorb more energy per unit mass than
tapered rectangular tubes.
b. For each tube, the energy absorption per unit mass increases with
increasing wall thickness yet decreases with increasing taper angle.
c. Double- and triple-tapered tubes absorb more energy per unit mass
than frusta, for a given wall thickness and taper angle.

(8) Increasing the impact velocity of straight and tapered rectangular tubes of
equivalent cross-section dimensions increases the Dynamic Amplification
Factor (DAF), due to inertia effects, strain rate sensitivity and increased input
energy. However, the amplification of absorbed energy is less for tapered

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 213
tubes. Thus, tapering a thin-walled tube may be a more practical means of
reducing the amplification of absorbed energy under impact loads, compared
with the more common techniques of pre-buckling and adding crush triggers.
Furthermore, the amplification of absorbed energy appears to be greater for
“thin” straight tubes than for “thick” straight tubes. However, the DAF is less
dependent on the wall thickness for tapered rectangular tubes compared with
straight rectangular tubes. Also, the DAF only reduces slightly when the taper
angle is increased, while the DAF is not significantly influenced by the
number of tapers for tapered tubes.

(9) Finally, the properties of the impact surface have most influence on the
energy absorption response of tubes with two tapered sides compared with
the other tubes investigated. Thus, particular consideration of the impact
surface may be required when using the double-tapered tube for absorbing
impact energy. Furthermore, the impact surface properties have more effect
on the response of each tube as the impact velocity increases. Such results
provide a guide as to how each tube behaves when impacted by different
surfaces.

In Chapter 2 it was identified that, in the context of vehicle crashworthiness, off-axis


or oblique loads are unavoidable. Furthermore, it was identified that there is no
published knowledge on the energy absorption response of tapered rectangular tubes
under oblique loading. This is clearly important for impact energy absorber design
such as in crashworthiness applications, and will be addressed in the next chapter.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 214
7 Dynamic Response of Straight and Tapered
Rectangular Tubes under Oblique Impact
Loading

7.1 Introduction

Real-world crashworthiness applications frequently involve loads which act oblique


to the axis of the energy absorber. Such loading causes the absorber to deform via a
combination of both axial and global bending (Euler) collapse modes. The latter
mode is generally unstable with an associated reduction in the energy absorption
capacity of the tube. As such it is important to understand how the energy absorber
responds under oblique loading, in order to maximise energy absorption under such
conditions. This thesis sought to further investigate and compare the behaviour of
straight and tapered thin-walled rectangular tubes under oblique loads. Specifically,
the thesis sought to investigate the following aspects of the tube behaviour:

(1) The critical load angle(s) at which straight and tapered tubes make the
transition from the axial collapse mode to the bending collapse mode.
(2) How the response (mean load and absorbed energy) of each tube geometry is
affected by the load angle, impact velocity and geometry parameters of the
tubes, such as wall thickness, taper angle, number of tapers, breadth and
length.
(3) How the deformation modes of each tube geometry depend on the load angle.

Using the validated finite element model for oblique loading presented in Chapter 4,
this chapter extends the axial loading analyses of Chapters 5 and 6 by investigating
the dynamic impact response of straight and tapered tubes under oblique loading
conditions. It was found that the mean load and energy absorption decrease as the
load angle increases for both straight and tapered rectangular tubes. However,
tapered tubes display a more stable decrease in mean load with increasing load angle,

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 215
and their energy absorption capacity is more maintained when global collapse
dominates the response. Overall, the results show that tapered tubes generally absorb
energy more effectively than straight tubes under oblique loading conditions. The
findings of this chapter are also presented in Nagel and Thambiratnam (2005a).

7.2 Loading arrangement and description of the finite


element model

The oblique loading arrangement and load cases (range of load angle, α ) used for
the parametric study are shown in Figure 7.2-1.

Fully fixed rigid body

α
V
y
α range is from 0 to 400
x
.
z

Rigid body of mass M Base of tube fully fixed


to rigid body

Figure 7.2-1 Model arrangement for oblique loading.

As for the finite element model used to simulate axial loading, the base of each tube
(the large end for tapered tubes) was fully fixed to a supporting rigid body using a
tied constraint. This supporting rigid body was constrained to translate such that the
free end of the tube was axially crushed onto a fully fixed rigid body. Both rigid
bodies were modelled as analytical rigid surfaces. Oblique loading conditions were
achieved by orienting the fully fixed rigid body at a pre-defined load angle, α, to the
horizontal plane in Figure 7.2-1. Such a loading arrangement has been adopted by
others to simulate oblique loading of thin-walled tubes (Han & Park 1999), and
follows the arrangement used to simulate axial impact loading as presented in

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 216
Chapter 6 of this thesis. For the present study, impact loading was simulated by
prescribing the moving rigid body with a point mass, M and initial velocity, V, such
that it translated to axially impact the tube into the fully fixed rigid body. To avoid
spurious deformation modes, a friction coefficient of 0.2 was used for contact
between the sides of the tube and the rigid bodies. Such a value has been used in
previous studies simulating the axial crushing of square tapered tubes under impact
loads (Mamalis et al. 2001). The friction coefficient for self-contact within each tube
was found to have no significant influence on the energy absorption response, such
that a coefficient of 0.1 was used for all analyses.

In the validation study, geometry imperfections were included due to fabrication of


the specimens. However, for simplicity, initial imperfections were not included for
the parametric study, and since it was found that the trigger depth has minor effect on
the energy absorption response under global collapse (see Figure 4.3-6 of Chapter 4).
Furthermore, it has been argued in a similar study (Reyes et al. 2002) that, even if the
FE model for the parametric study is not able to exactly describe the experimental
load-deflection response, the main effects are present in the model and thus it is
suitable for parametric studies. Finally, the effects of the welds on the tube
deformation mode and energy absorption response were not significant, and thus
were neglected for the parametric study.

The step time was 30 ms, as established in Section 3.2.2 of Chapter 2. The effects of
strain rate were included using the Cowper-Symonds equation with parameter values
D = 6844 s-1 and q = 3.91, as detailed in Chapter 3. It was shown in Chapter 6 that
the impact velocity significantly affects the energy absorption response of the tubes
under impact loading. An impact velocity range of 10-20 m/s was chosen for the
oblique loading parametric study, which is a range typically encountered in
automobile crashworthiness applications. Values within this range have been used in
previous studies on the oblique loading of thin-walled tubes (Han & Park 1999;
Reyes et al. 2002). An impact mass of 400 kg was used for all analyses, as this
provided the minimum impact energy required to fully crush the tubes for the given
impact velocities. Finally, the full tube was modelled in all cases since the loading is
unsymmetric.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 217
7.3 Tube geometries

The main purpose of the oblique loading parametric study was to compare the
response of straight and tapered tubes. Single- and double-tapered tubes were
compared since they have the same breadth, B as straight tubes. Triple-tapered tubes
and frusta were not considered since their breadth is larger than that of straight tubes,
and thus would not provide a meaningful comparison. Figure 7.3-1 shows the tube
geometries compared in the oblique loading parametric analysis.

Wt

L
θ

z x B Wb

Straight Single taper Double taper

Figure 7.3-1 Tube geometries compared for oblique loading.

Of interest in this study was the effect of the various geometry parameters on the
response of the tubes under oblique loading. Such knowledge would provide a guide
as to the design of such structures in applications involving impact loads which act
oblique to the longitudinal axis of the tube. The main geometry parameters varied
were the tube wall thickness, h, length, L, web breadth, B and taper angle, θ. The
taper angle provides the means of comparison between straight and tapered tubes.
The range of each parameter studied is shown in Table 7.3-1, and the dimensions
were chosen to cover the typical range of tube sizes as used in automotive structures
(Han & Park 1999). The top width, Wt of the straight tube and small end of the
tapered tubes was maintained constant at 100 mm, such that the taper angle was
obtained by varying the base width, Wb.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 218
Table 7.3-1 Range of each geometry parameter used in the oblique loading
parametric analysis.

Geometry h θ L B
(mm) (deg) (mm) (mm)
Straight tube 1.5, 2, 2.5 5, 10, 15 300, 350, 400 60, 80, 100
Single taper
Double taper

7.4 Effect of load angle on tube response

Initially the effect of the load angle on the energy absorption response of the straight
and tapered tubes was compared. Also of particular interest was the critical load
angle at which there is a transition from progressive axial to global bending collapse,
for a given geometry. As in previous chapters of this thesis, response is measured in
terms of the crush load, the mean crush load and the energy absorbed with deflection
of the tube. Results in this section are compared for a baseline tube geometry with h
= 2 mm, L = 350 mm, B = 80 mm and θ = 100, while the impact velocity is 15 m/s.

The effect of the load angle is very significant for both straight and tapered tubes, as
can be seen from the load-deflection curves shown in Figure 7.4-1. As has been
illustrated in earlier chapters of this thesis, thin-walled columns undergoing pure
axial deformation display a characteristic initial peak load as the first lobe forms,
followed by a region where the load oscillates between local minimum and
maximum values as subsequent lobes form. For each tube described in Figure 7.4-1,
the load oscillates over the full crush deflection as in pure axial deformation, up to a
load angle of 150. When the load angle is increased to 200, however, the load
oscillations cease at approximately 75 mm deflection for the straight tube, 85 mm for
the single-tapered tube and 100 mm for the double-tapered tube. Thus, each tube
undergoes progressive collapse up to these respective deflections, beyond which the
load decreases for the remainder of the crush cycle. However, for each tube the load
then increases slightly towards the end of the crush cycle, indicating that each tube is
still undergoing progressive collapse. This was verified by observing the deformation
history of each tube simulated by the finite element model.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 219
350
300

Dynamic load (kN)


250
200
150
100
50
0
0 25 50 75 100 125 150 175 200
Deflection (mm)

Axial loading Load angle = 10 deg load angle = 15 deg


Load angle = 20 deg Load angle = 25 deg Load angle = 30 deg
Load angle = 35 deg

Straight tube
350
300
Dynamic load (kN)

250
200
150
100
50
0
0 25 50 75 100 125 150 175 200
Deflection (mm)

Axial loading Load angle = 10 deg Load angle = 15 deg


Load angle = 20 deg Load angle = 25 deg Load angle = 30 deg
Load angle = 35 deg

Single taper
400
350
Dynamic load (kN)

300
250
200
150
100
50
0
0 25 50 75 100 125 150 175 200
Deflection (mm)

Axial loading Load angle = 10 deg Load angle = 15 deg


Load angle = 20 deg Load angle = 25 deg Load angle = 30 deg
Load angle = 35 deg

Double taper
Figure 7.4-1 Effect of load angle on the load-deflection response.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 220
When the load angle is increased to 250, it is clear from Figure 7.4-1 that for both
straight and tapered tubes the load does not oscillate, rather it increases steadily up to
a certain deflection, before decreasing continuously for the remainder of the crush
cycle.

The above observed phenomena can be explained in terms of the deformation


profiles of the tubes, which are shown in Figure 7.4-2. The deformed profiles are
shown up to a deflection, d = 200 mm. For a load angle of α = 100 each tube fails via
progressive collapse, corresponding to the load oscillations which were observed up
to a load angle of 200 in Figure 7.4-1. However, bending collapse has occurred for a
load angle of 300 after the formation of a lobe at the impacted end. This lobe
formation is characterised by the initial increase in load for α ≥ 250, while the global
collapse corresponds to the subsequent decrease in load for these higher values of α.
Such behaviour was also observed for a taper angle of θ = 250, such that the critical
load angle at which each tube makes the transition from progressive to global
bending collapse lies between α = 250 and 300. This critical load angle may depend
on the various geometry parameters, as will be further investigated in subsequent
sections of this chapter. It can be seen from Figure 7.4-2 that bending collapse occurs
due to the formation of a buckle towards the fixed end of each tube. It was found that
the distance from this buckle to the fixed end of each tube increased as the number of
tapers increased, as was also found in the experiments (refer to Section 4.2.2 of
Chapter 4).

The effect of the load angle can be clearly seen from the mean load-deflection and
energy absorption-deflection curves for each tube, as shown in Figure 7.4-3 and
Figure 7.4-4, respectively. The mean load and energy absorption up to a given
deflection decrease as the load angle increases, and this may be due to the associated
decrease in the initial peak load, as depicted in Figure 7.4-1. Furthermore, increasing
the load angle when progressive collapse dominates causes the lobes formed to be
pushed laterally away from the sides of the tube, thus reducing their amount of axial
compression and hence energy absorption. It is clear from Figure 7.4-3 and Figure
7.4-4 that, as the loading goes from axial to oblique, the reduction in mean load and
energy absorption becomes less significant as the number of tapers increases.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 221
y

z x

Undeformed α = 100 α = 300

Undeformed α = 100 α = 300

Undeformed α = 100 α = 300

Figure 7.4-2 Typical deformation modes for straight and tapered rectangular
tubes under oblique loading (deformed profiles shown at d = 200 mm). Straight
tube (top), single-tapered tube (middle), and double-tapered tube (bottom).

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 222
250

Dynamic mean load (kN)


200

150

100

50

0
0 25 50 75 100 125 150 175 200
Deflection (mm)

Axial loading Load angle = 10 deg load angle = 15 deg


Load angle = 20 deg Load angle = 25 deg Load angle = 30 deg
Load angle = 35 deg

Straight tube
250
Dynamic mean load (kN)

200

150

100

50

0
0 25 50 75 100 125 150 175 200
Deflection (mm)

Axial loading Load angle = 10 deg Load angle = 15 deg


Load angle = 20 deg Load angle = 25 deg Load angle = 30 deg
Load angle = 35 deg

Single taper
250
Dynamic mean load (kN)

200

150

100

50

0
0 25 50 75 100 125 150 175 200
Deflection (mm)

Axial loading Load angle = 10 deg Load angle = 15 deg


Load angle = 20 deg Load angle = 25 deg Load angle = 30 deg
Load angle = 35 deg

Double taper
Figure 7.4-3 Effect of load angle on the mean load-deflection response.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 223
Dynamic absorbed energy (kJ)
20

15

10

0
0 25 50 75 100 125 150 175 200
Deflection (mm)

Axial loading Load angle = 10 deg load angle = 15 deg


Load angle = 20 deg Load angle = 25 deg Load angle = 30 deg
Load angle = 35 deg

Straight tube
Dynamic absorbed energy (kJ)

20

15

10

0
0 25 50 75 100 125 150 175 200
Deflection (mm)

Axial loading Load angle = 10 deg Load angle = 15 deg


Load angle = 20 deg Load angle = 25 deg Load angle = 30 deg
Load angle = 35 deg

Single taper
Dynamic absorbed energy (kJ)

20

15

10

0
0 25 50 75 100 125 150 175 200
Deflection (mm)

Axial loading Load angle = 10 deg Load angle = 15 deg


Load angle = 20 deg Load angle = 25 deg Load angle = 30 deg
Load angle = 35 deg

Double taper
Figure 7.4-4 Effect of load angle on the energy absorption-deflection response.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 224
This important observation will be the focus of discussion for the rest of this section.
It can also be seen from Figure 7.4-4 that, for each tube, the absorbed energy at d ≥
175 mm decreases significantly when the load angle increases from 150 to 250. In
comparison, the reduction in absorbed energy is not as significant as the load angle
increases beyond 250.

To compare the relative effect of the load angle for the straight and tapered tubes, the
mean load was calculated up to a deflection of 200 mm for each tube, as shown in
Figure 7.4-5. Each tube had undergone sufficient axial compression or global
rotation at this deflection, thus it was chosen for adequate comparison of results.

100
90
Dynamic mean load (kN)

80
Global collapse dominates
70
60
50
40
30
20 Progressive collapse dominates Transition region
10
0
0 5 10 15 20 25 30 35 40
Load angle, α (deg)

Straight Single taper Double taper

Figure 7.4-5 Effect of load angle on the mean load under dynamic impact
loading (calculated up to d = 200 mm).

Once again, Figure 7.4-5 clearly shows that the dynamic mean load decreases as the
load angle is increased for each tube. Furthermore, three regions arise from the
relationship between the mean load and the load angle: a region where progressive
collapse dominates, a region where there is a transition from progressive to global
bending collapse, and finally a region where global bending collapse dominates the
response. This relationship has also been found for the oblique loading of mild steel
straight square columns (Han & Park 1999). Furthermore, it can be seen from Figure
7.4-5 that the mean load of each tube is less sensitive to load angle when global
collapse dominates the response. For instance, the mean load decreases by

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 225
approximately 40 kN for each tube when the load angle increases from 0 to 200,
whereas it decreases by approximately only 20 kN for 200 < α < 400.

In Figure 7.4-5, additional results are shown between α = 20 and 250, to determine
the critical load angle. It was observed that global buckling occurs for all tubes when
the load angle reaches 230, indicating that this is the critical load angle for the tubes
of the given geometries. Figure 7.4-5 shows that for the straight tube the mean load
drops noticeably (~ 29 %), as the load angle increases from 22 to 230. On the other
hand, the tapered tubes show a gradual decrease in mean load over the transition
region. This may be because, at a load angle of 230, the deformation at the impacted
end before global buckling occurred was greater for the tapered tubes than for the
straight tube. Furthermore, the larger cross-section of the tapered tubes in the region
of the plastic hinge provides greater resistance to rotation, thus increasing the energy
absorption. As the load angle increases beyond 250, the mean load for a given load
angle increases with increasing number of tapers. Once again, this is predominantly
due to the larger cross-section of the tapered tubes in the region of the buckle formed
during global collapse, thus increasing the energy absorption and hence mean load.

These results highlight several advantages of using tapered rectangular tubes for
energy absorption under oblique loading conditions. Firstly, the gradual decrease in
mean load they display over the transition region compared with the straight tubes
represents a more stable response for the tapered tubes. Secondly, the enhanced mean
load for tapered tubes compared with straight tubes in the global collapse region
indicates that the energy absorption capacity of tapered tubes is more maintained
under oblique loading. This is verified by Figure 7.4-6, which shows the energy
absorbed, Ed by each tube up to a deflection of 200 mm, as a percentage of the
energy absorbed under axial loading, Edaxial, up to the same deflection. The reduction
in absorbed energy is less as the number of tapers increases, particularly in the global
collapse region. These results imply that tapered tubes may be preferable to straight
tubes for applications involving oblique loading conditions. Finally, Figure 7.4-6
clearly shows the significant reduction in mean load for each tube over the transition
region between load angles of 20 and 250, however this reduction is most significant
for straight tubes.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 226
100

80

axial 60
Ed /Ed
(%) 40

20

0
0 10 15 20 25 30 35 40
Load angle, α (deg)

Straight tube Single-tapered tube Double-tapered tube

Figure 7.4-6 Reduction of absorbed energy with increasing load angle.

The plastic hinge causing global collapse formed according to several bending modes
for each tube. Three modes were observed, as shown for a selection of load angles
and tube geometries in Figure 7.4-7. Mode A consisted of an outward forming lobe
on the compression flange, while the narrow sides of the tube were drawn inwards.
This mode occurred for the single-tapered tube with load angles of 30 and 350, and
for the double-tapered tube with load angles of 35 and 400. Mode C consisted of an
inward forming lobe on the compression flange with the narrow sides moving
outwards. A similar mode has been observed in previous studies on the bending
collapse of straight rectangular tubes (Kecman 1983), and the oblique loading of
straight square tubes (Reyes et al. 2002), as well as for the experimental specimens
described in Chapter 4 of this thesis. Mode C was exhibited for the straight tube with
load angles of 35 and 400, for the single-tapered tube with load angles of 25 and 400,
and for the double-tapered tube with load angles of 25 and 300. Finally, Mode B
consisted of a first lobe which formed according to Mode A at the onset of global
collapse. However, as the tube was further deformed, underneath this lobe a second
lobe began to form according to Mode C. The first lobe ceased to grow as the tube
was further loaded, whereas the second lobe continued to grow. Mode B occurred for
the straight tube at load angles of 25 and 300. For all three modes the tension flange
moved inwards to relieve tensile and shear strain. According to these observations, it
appears that the global buckling mode changes for each tube as the load angle is
increased. This may be due to the elastic and plastic half-waves interacting in
different ways as the load angle increases.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 227
Mode B

First lobe

Mode C

Second lobe

Straight tube (α = 300) Top view Straight tube (α = 400)

Mode C
Mode A

Single taper (α = 300) Single taper (α = 400)

Mode C
Mode A

Double taper (α = 300) Double taper (α = 400)

Figure 7.4-7 Deformation modes of the plastic hinge formed during global
collapse.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 228
7.5 Effect of wall thickness on the tube response under
oblique loading

It has been shown in previous chapters of this thesis that the wall thickness has a
significant effect on the energy absorption response of straight and tapered
rectangular tubes under axial loads. To examine the effect of wall thickness under
oblique loading, the baseline dimensions of the straight and tapered tubes were
adopted at three values of wall thickness: 1.5, 2 and 2.5 mm. The results for the
straight and single-tapered tubes are shown in Figure 7.5-1.

140
120
Dynamic mean load (kN)

100
80
60
40
20
0
0 5 10 15 20 25 30 35 40
Load angle, α (deg)

h = 1.5 mm h = 2 mm h = 2.5 mm

Straight tube
140
120
Dynamic mean load (kN)

100
80
60
40
20
0
0 5 10 15 20 25 30 35 40
Load angle, α (deg)

h = 1.5 mm h = 2 mm h = 2.5 mm

Single taper
Figure 7.5-1 Effect of wall thickness on the mean load with increasing load
angle.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 229
Similar trends were displayed for the double-tapered tube. The critical load angle lies
within the transition range of 220 ≤ α ≤ 240 for the straight tube and 230 ≤ α ≤ 300 for
the single-tapered tube, for each wall thickness. Thus, increasing the wall thickness
has no significant effect on the critical load angle range for both straight and tapered
rectangular tubes, as has also been found for straight square tubes (Han & Park
1999). For both straight and tapered tubes it appears that the mean load is more
sensitive to load angle as the wall thickness increases, particularly in the region of
progressive collapse. This was also found for obliquely loaded square columns
(Reyes et al. 2002). Furthermore, whereas the wall thickness has a significant effect
on the mean load in the region of progressive collapse, the opposite seems to be true
when global collapse dominates the response. Thus, the sensitivity of the mean load
to wall thickness is less when the tube response is dominated by global collapse. This
result was also found for straight square tubes (Han & Park 1999), and the present
analysis shows that this also holds true for straight and tapered rectangular tubes.

It was shown in Chapter 6 that, under axial impact loading, the energy absorbed per
unit mass, Em is sensitive to the wall thickness. However, in comparison, the wall
thickness has less effect on Em when straight and tapered tubes respond via global
collapse, as illustrated in Figure 7.5-2. Here the energy absorbed up to d = 200 mm
has been used in the calculations, to include the effect of load angle. Nevertheless, it
appears that the influence of wall thickness under global collapse increases slightly
as the number of tapers increases. Furthermore, Figure 7.5-3 shows that the number
of tapers has no significant effect on Em under global collapse, whereas the opposite
is the case when axial collapse dominates the response. The overall practical
implications of these results is that the wall thickness and number of tapers have less
influence on the energy absorption capacity, in terms of absorber mass, of straight
and tapered rectangular tubes when the tubes respond via global collapse as opposed
to progressive, axial collapse. Thus, when weight reduction is important, increasing
the wall thickness or number of tapers provides no significant improvement to the
energy absorption capacity when global collapse dominates the response, which can
happen when there is significant oblique impact.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 230
12

Energy absorbed per unit mass


10

(kJ / kg)
6

0
0 5 10 15 20 25 30 35 40
Load angle, α (deg)

h = 1.5 mm h = 2 mm h = 2.5 mm

Straight tube
9
Energy absorbed per unit mass

8
7
6
(kJ / kg)

5
4
3
2
1
0
0 5 10 15 20 25 30 35 40
Load angle, α (deg)

h = 1.5 mm h = 2 mm h = 2.5 mm

Single taper
8
Energy absorbed per unit mass

7
6
5
(kJ / kg)

4
3
2
1
0
0 5 10 15 20 25 30 35 40
Load angle, α (deg)

h = 1.5 mm h = 2 mm h = 2.5 mm

Double taper
Figure 7.5-2 Energy absorbed per unit mass vs. load angle and wall thickness.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 231
10

Energy absorbed per unit mass


9
8
7
6
(kJ / kg)
5
4
3
2
1
0
0 5 10 15 20 25 30 35 40
Load angle, α (deg)

Straight Single taper Double taper

Figure 7.5-3 Energy absorbed per unit mass vs. load angle and number of
tapers (here h = 2 mm).

7.6 Effect of taper angle on the tube response under oblique


loading

It was found that, as for axial impact loading, the taper angle has no significant
influence on the dynamic response for each tube as the load angle is increased. Thus,
as for the wall thickness, an increase in the taper angle provides no significant
improvement to the energy absorption capacity when global collapse dominates the
response, or when progressive collapse dominates. Furthermore, the energy absorbed
per unit mass is relatively independent of taper angle when global collapse dominates
the response. This is in contrast to the findings of Chapter 6, where it was found that
Em is dependent on taper angle under axial impact loading conditions. Furthermore,
the taper angle has no significant influence on the critical load angle.

Finally, Figure 7.6-1 shows the effect of taper angle on the energy absorbed under
oblique loading relative to that under axial loading. It is obvious that the relative
energy absorbed decreases with increasing load angle. Furthermore, for both single-
and double-tapered tubes it appears that increasing the taper angle increases the
relative energy absorption when global collapse dominates the response. This implies
that it would be desirable to maximise the taper angle to minimise the relative
reduction of absorbed energy as the load angle increases.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 232
100

80

axial 60
Ed /Ed
(%) 40

20

0
0 10 15 20 25 30 35 40
Load angle, α (deg)

Taper angle = 5 deg Taper angle = 10 deg Taper angle = 15 deg

Single taper

100

80

axial 60
Ed /Ed
(%) 40

20

0
0 10 15 20 25 30 35 40
Load angle, α (deg)

Taper angle = 5 deg Taper angle = 10 deg Taper angle = 15 deg

Double taper
Figure 7.6-1 Effect of taper angle and load angle on the percentage of energy
absorbed under axial loading.

7.7 Effect of breadth on the tube response under oblique


loading

Figure 7.7-1 depicts the mean load response of each tube versus load angle with
varying breadth, B (refer to Figure 7.3-1).

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 233
120

Dynamic mean load (kN)


100

80

60

40

20

0
0 5 10 15 20 25 30 35 40
Load angle, α (deg)

B = 60 mm B = 80 mm B = 100 mm

Straight tube
90
80
Dynamic mean load (kN)

70
60
50
40
30
20
10
0
0 5 10 15 20 25 30 35 40
Load angle, α (deg)

B = 60 mm B = 80 mm B = 100 mm

Single taper
90
80
Dynamic mean load (kN)

70
60
50
40
30
20
10
0
0 5 10 15 20 25 30 35 40
Load angle, α (deg)

B = 60 mm B = 80 mm B = 100 mm

Double taper
Figure 7.7-1 Effect of breadth on the mean load with increasing load angle.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 234
The results show that the mean load is more sensitive to the breadth when global
collapse dominates the response as opposed to when progressive collapse dominates.
This is in contrast to the case for wall thickness. Furthermore, it is interesting to note
that the reduction in mean load as the load angle increases is less significant when
the breadth is larger. This is also in contrast to the case with increasing wall
thickness shown earlier.

From the results shown thus far it appears that the wall thickness and taper angle
have no significant effect on the critical load angle range over which the tubes make
the transition from progressive collapse to global bending collapse. However, Figure
7.7-1 shows that the critical load angle increases as the web breadth increases. For
instance, in the case of the straight tube, it was found that the critical load angle lies
between 10 and 150 when the breadth is 60 mm, between 20 and 250 for a breadth of
80 mm, and between 25 and 300 when the breadth is 100 mm. The critical load angle
also increases with increasing breadth for the single-tapered tube however not for the
double-tapered tube. This is because it was observed that a progressive buckle tends
to form on the lower half of the compression flange of the double-tapered tube,
which then localises to promote global buckling and hence limiting the critical load
angle.

Figure 7.7-2 shows the effect of web breadth on the percentage of energy absorbed
under axial loading. When global collapse dominates the response, a higher
percentage of energy is absorbed for each load angle when the breadth increases.
Also, for a given breadth it can be seen that tapered tubes absorb a higher percentage
of energy for each load angle when the response is governed by global collapse.
Finally, under global collapse, the percentage of absorbed energy is more sensitive to
breadth for a given load angle as the number of tapers increases. These results further
highlight the advantages of tapering a rectangular tube to maximise its energy
absorbed under oblique loading.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 235
100

80

axial 60
Ed /Ed
(%) 40

20

0
0 10 15 20 25 30 35 40
Load angle, α (deg)

B = 60 mm B = 80 mm B = 100 mm

Straight tube

100

80

axial 60
Ed /Ed
(%) 40

20

0
0 10 15 20 25 30 35 40
Load angle, α (deg)

B = 60 mm B = 80 mm B = 100 mm

Single taper

100

80

axial 60
Ed /Ed
(%) 40

20

0
0 10 15 20 25 30 35 40
Load angle, α (deg)

B = 60 mm B = 80 mm B = 100 mm

Double taper
Figure 7.7-2 Reduction of absorbed energy with increasing load angle.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 236
The following important practical conclusions can be drawn from these observations:

(1) Increasing the web breadth increases the range of load angles over which both
straight and tapered rectangular tubes effectively absorb energy without
global collapse occurring.
(2) Increasing the web breadth increases the energy absorption capacity for both
straight and tapered rectangular tubes when global collapse dominates the
response.
(3) For a given web breadth and load angle, increasing the number of tapers
increases the energy absorption capacity of rectangular tubes when the
response is governed by global collapse.

12
Energy absorbed per unit mass

10

8
(kJ / kg)

0
0 5 10 15 20 25 30 35 40
Load angle, α (deg)

B = 60 mm B = 80 mm B = 100 mm

Straight tube
9
Energy absorbed per unit mass

8
7
6
(kJ / kg)

5
4
3
2
1
0
0 5 10 15 20 25 30 35 40
Load angle, α (deg)

B = 60 mm B = 80 mm B = 100 mm

Single taper
Figure 7.7-3 Energy absorbed per unit mass vs. load angle and breadth.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 237
Finally, the energy absorbed per unit mass increases with increasing breadth when
the tubes fail via global bending collapse, as shown in Figure 7.7-3 for the straight
and single-tapered tubes.

7.8 Effect of length on the tube response under oblique


loading

The effect of tube length on the mean load response under oblique loading is shown
in Figure 7.8-1. Clearly, the tube length has no significant effect on the response for
both straight and tapered tubes which fail via progressive collapse. This is in
accordance with pure axial collapse theory for straight (Abramowicz & Jones 1986)
and tapered (Reid & Reddy 1986) rectangular tubes. However, under global bending
collapse, an increase in the length causes a reduction in the mean load (maximum
reduction at α = 300 is ~27 % for the straight and single-tapered tubes and 22 % for
the double-tapered tube). Since all other parameters remain constant, an increase in
tube length increases the available moment arm, thus reducing the force required to
bend each tube over a given deflection. Such results were also found for straight
square tubes made of aluminium (Reyes et al. 2002) and mild steel (Han & Park
1999), under oblique impact loading.

Overall it was found that the critical load angle is not as sensitive to the tube length
as it is to the tube breadth. This may be due to friction creating an opposing force at
the sliding interface between the tube and impacted rigid surface, thus promoting
progressive collapse, even when the tube length increases. When global collapse
dominates, increasing the length of each tube reduces the percentage of energy
absorbed under axial loading, as shown in Figure 7.8-2. However, this reduction
becomes less as the number of tapers increases. Thus, the absorbed energy is more
maintained for a tapered rectangular tube under oblique impact loads. Finally, Figure
7.8-3 shows that, for each tube, the length has virtually the same influence on the
energy absorbed per unit mass under both axial and global collapse response.
Furthermore, as the length increases for each tube, the associated increase in mass
causes a reduction in Em at each load angle.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 238
100
90

Dynamic mean load (kN)


80
70
60
50
40
30
20
10
0
0 5 10 15 20 25 30 35 40
Load angle, α (deg)

L = 300 mm L = 350 mm L = 400 mm

Straight tube
90
80
Dynamic mean load (kN)

70
60
50
40
30
20
10
0
0 5 10 15 20 25 30 35 40
Load angle, α (deg)

L = 300 mm L = 350 mm L = 400 mm

Single taper
90
80
Dynamic mean load (kN)

70
60
50
40
30
20
10
0
0 5 10 15 20 25 30 35 40
Load angle, α (deg)

L = 300 mm L = 350 mm L = 400 mm

Double taper
Figure 7.8-1 Effect of length on the mean load with increasing load angle.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 239
100

80

axial 60
Ed /Ed
(%) 40

20

0
0 10 15 20 25 30 35 40
Load angle, α (deg)

L = 300 mm L = 350 mm L = 400 mmm

Straight tube

100

80

axial 60
Ed /Ed
(%) 40

20

0
0 10 15 20 25 30 35 40
Load angle, α (deg)

L = 300 mm L = 350 mm L = 400 mmm

Single taper

100

80

axial 60
Ed /Ed
(%) 40

20

0
0 10 15 20 25 30 35 40
Load angle, α (deg)

L = 300 mm L = 350 mm L = 400 mmm

Double taper
Figure 7.8-2 Reduction of absorbed energy with increasing load angle.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 240
12

Energy absorbed per unit mass


10

(kJ / kg)
6

0
0 5 10 15 20 25 30 35 40
Load angle, α (deg)

L = 300 mm L = 350 mm L = 400 mm

Straight tube
9
Energy absorbed per unit mass

8
7
6
(kJ / kg)

5
4
3
2
1
0
0 5 10 15 20 25 30 35 40
Load angle, α (deg)

L = 300mm L = 350 mm L = 400 mm

Single taper
8
Energy absorbed per unit mass

7
6
5
(kJ / kg)

4
3
2
1
0
0 5 10 15 20 25 30 35 40
Load angle, α (deg)

L = 300 mm L = 350 mm L = 400 mm

Double taper
Figure 7.8-3 Energy absorbed per unit mass vs. load angle and length.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 241
7.9 Effect of impact velocity on the tube response under
oblique loading

Thus far the effect of the various geometry parameters on the oblique loading
response of the straight and tapered tubes has been determined. The impact velocity
is an important loading parameter since it can have significant effect on the energy
absorption response of thin-walled tubes, as revealed in previous chapters of this
thesis. An impact velocity range of 10-20 m/s was chosen for the oblique loading
parametric study, while the baseline tube dimensions were used. This represents a
typical range of velocities encountered in automotive crashworthiness applications.
Values within this range have also been used in previous studies on the oblique
loading of thin-walled tubes (Han & Park 1999; Reyes et al. 2002).

Chapter 6 compared the dynamic and quasi-static response of the straight and tapered
rectangular tubes under axial impact loading, using the Dynamic Amplification
Factor, or DAF. The DAF for straight and tapered tubes under oblique loading is
shown in Figure 7.9-1 for various impact velocities. Here the DAF has been
calculated up to a deflection of 200 mm. The factor increases with increasing impact
velocity for each tube and load angle. This may be due to strain rate and inertia
effects, as identified for axial impact loading in Chapter 6. It is interesting to observe
that the amplification of absorbed energy seems to generally decrease as the load
angle increases for each tube. Thus, inertia and strain rate effects may be less
significant as the load angle increases. Nevertheless, the reduction in the Dynamic
Amplification Factor with increasing load angle is less significant as the impact
velocity and number of tapers increase. Practically, therefore, the number of tapers
can be used to control the amplification of absorbed energy under oblique impact
loading. Finally, the impact velocity has no significant effect on the critical load
angle range. However, the range was reduced to 15-200 for the double-tapered tube
with V = 10 and 20 m/s, compared with 20-250 for V = 15 m/s. This was due to an
inward forming buckle on the compression flange towards the base of the tube, thus
promoting global buckling at lower load angles.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 242
1.8

Dynamic Amplification Factor


1.6
1.4
1.2
1.0
0.8
0.6
0.4
0.2
0.0
0 5 10 15 20 25 30 35 40
Load angle, α (deg)

V = 10 m/s V = 15 m/s V = 20 m/s

Straight tube
1.6
Dynamic Amplification Factor

1.4
1.2
1.0
0.8
0.6
0.4
0.2
0.0
0 5 10 15 20 25 30 35 40
Load angle, α (deg)

V = 10 m/s V = 15 m/s V = 20 m/s

Single taper
1.6
Dynamic Amplification Factor

1.4
1.2
1.0
0.8
0.6
0.4
0.2
0.0
0 5 10 15 20 25 30 35 40
Load angle, α (deg)

V = 10 m/s V = 15 m/s V = 20 m/s

Double taper
Figure 7.9-1 Effect of load angle on the DAF.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 243
As for dynamic loading, the quasi-static energy absorption capacity of the tapered
tubes is more maintained under oblique loading compared with straight tubes. This is
illustrated in Figure 7.9-2 and Figure 7.9-3. Figure 7.9-2 shows that the quasi-static
mean load for the tapered tubes is generally greater than that for straight tubes of
equivalent cross-section dimensions. Furthermore, the mean load for the straight tube
drops quite dramatically over the transition region, whereas it decreases more
gradually for the tapered tubes. Figure 7.9-3 shows the ratio of quasi-static energy
absorbed under oblique loading to that absorbed under axial loading. Over the
transition region the ratio drops most dramatically for the straight tube. Furthermore,
overall, the reduction in energy absorption is less for the tapered tubes than for the
straight tubes over the range of load angles considered, particularly in the global
collapse region. Finally, it was observed that the critical load angle for each tube was
less under quasi-static loading compared with dynamic loading. However, this may
be due to the higher coefficient of friction (µ = 0.3) used to model the contact
between the tube and stationary rigid body under quasi-static loading, in accordance
with a previous study on the axial loading of square frusta (Mamalis et al. 2001).

70
60
Quasi-static mean load

50
40
(kN)

30
20
10
0
0 5 10 15 20 25 30 35 40
Load angle, α (deg)

Straight Single taper Double taper

Figure 7.9-2 Effect of load angle on the quasi-static mean load.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 244
100

80

axial 60
Es /Es
(%) 40

20

0
0 10 15 20 25 30 35 40
Load angle, α (deg)

Straight Single taper Double taper

Figure 7.9-3 Reduction of absorbed energy with increasing load angle under
quasi-static loading.

7.10 Conclusions and design guidelines

Real-world crashworthiness applications often involve oblique loads, causing the


energy absorber to fail via a combination of progressive axial and global bending
collapse modes. This chapter has examined the energy absorption response of
straight and tapered rectangular tubes under oblique loading, in which the angle of
applied load was varied from axial loading to 400 to the tube’s longitudinal axis. The
loading parameters investigated were the load angle and impact velocity, while the
geometry parameters were the tube wall thickness, taper angle, number of tapers,
web breadth and length. The main conclusions from the study are summarised below.

(1) The mean load and energy absorption decrease significantly as the angle of
applied load increases, due to a change in the failure mode.

(2) Tapered rectangular tubes display several key advantages over straight
rectangular tubes for absorbing energy under oblique impact loading
conditions, as follows:

a. Tapered tubes display a more stable and gradual decrease in mean


load over the transition region.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 245
b. The energy absorption capacity of tapered tubes is greater than for
straight tubes of equivalent cross-section dimensions, when the
response is dominated by global bending collapse.
c. The initial peak load, mean load and energy absorption are less
sensitive to load angle for tapered tubes than for straight tubes.

Tapered rectangular tubes may therefore be advantageous in applications


where oblique loading conditions are expected, such as vehicle
crashworthiness.

(3) For both straight and tapered rectangular tubes, the mean load is more
sensitive to load angle as the wall thickness increases and as the web breadth
decreases. The wall thickness has a significant effect on the energy
absorption response in the region of progressive collapse, whereas the taper
angle, web breadth and length do not.

(4) Increasing the web breadth increases the range of load angles over which
both straight and tapered rectangular tubes effectively absorb energy without
global collapse occurring.

(5) When global collapse cannot be avoided in a particular application, the


energy absorption capacity under oblique impact loading can be increased by:

a. Increasing the number of tapers, and/ or


b. Increasing the web breadth, and/ or
c. Decreasing the length, and/ or
d. Increasing the wall thickness

(6) Increasing the load angle appears to decrease the amplification of energy
absorbed under oblique impact loading relative to the energy absorbed under
oblique quasi-static loading. Nevertheless, the number of tapers can be used
to control this amplification.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 246
(7) Finally, the energy absorbed per unit mass can be increased for a global
collapse response by increasing the web breadth and decreasing the tube
length.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 247
8 Application: Dynamic Impact Simulation
and Energy Absorption of a Vehicle Bumper
System

8.1 Introduction

In many applications energy absorbers are frequently incorporated with other


structural components in the system design. It is therefore of particular interest how
individual energy absorbers behave when they are incorporated into an energy
absorbing system. Furthermore, the absorbers are commonly mounted to other
structural members. Thus, the absorber response may differ to when the absorbers
are analysed on their own, as in previous chapters of this thesis.

Energy absorbing systems have been developed for a host of applications particularly
in the fields of structural safety and crashworthiness. A common example is the front
longitudinal columns in automobiles, which, as well as providing energy absorption
during frontal impacts are also used for mounting vehicle auxiliary equipment such
as the bumper beam. Bumper beams are one of the main structures of passenger cars
that protect them from front and rear collisions (Hosseinzadeh et al. 2004). Hanssen
et al. (2000c) studied the dynamic axial impact loading of a car bumper system
consisting of crash boxes filled with aluminium foam. Experiments were used to
validate a finite element model which simulated symmetric and asymmetric loading
of the system. The finite element program LS-DYNA was used to model the bumper
system with existing constitutive models, and it was able to represent the complete
crushing process with sufficient accuracy. Other studies on automotive bumper
systems have focused on the use of composite materials to reduce weight compared
with traditional steel and aluminium bumper systems, without sacrificing strength
(Cheon S et al. 1995; Hosseinzadeh et al. 2004).

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 248
No research information was found in the open literature relating to the use of
tapered rectangular tubes in energy absorbing systems. As such, one of the primary
objectives of this thesis was the development of such a system. This chapter
describes the design and analysis of the energy absorbing system. Experimental tests
were conducted on a prototype, and these tests were used to validate a finite element
model of the system. The model was then used to analyse the response of the system
under quasi-static and dynamic impact loading. The response was compared with
that of a similar system consisting of straight rectangular tubes, and a parametric
analysis was conducted to determine the most influential parameters on the system’s
energy absorption response. The findings were used to develop design guidelines and
recommendations for the implementation of tapered rectangular tubes in energy
absorbing systems. To this end, the system was conceptual in form such that it could
be adopted for a variety of applications. Nevertheless, for convenience, the approach
in this chapter is to treat the system as a demonstrator car bumper system used to
absorb impact energy during minor frontal collisions.

While maximising the strength to weight ratio is a primary deign objective of vehicle
bumpers nowadays, this aspect was not considered for the present system. Rather, the
focus involved comparing the energy absorption performance of bumper systems
involving straight and tapered tubes, in lieu of the primary objectives of this thesis.
Furthermore, the intention concerning the bumper system studied in this chapter was
that it could be relevant to other energy absorbing applications, such as highway
guardrails, where weight reduction is not as much of a concern. Other examples of
energy absorbing systems which could incorporate tapered rectangular tubes include
shipping dock yard fenders and “survival cells” at the front of trains, as discussed in
Chapter 2.

Overall, the findings of this chapter highlight the importance of analysing the
response of an energy absorber as part of the energy absorbing system. The
advantages of using tapered rectangular tubes in such a system are also shown.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 249
8.2 Description of the energy absorbing system

Figure 8.2-1 shows the assembly and components of the bumper system. The
components are joined to each other using fillet and stitch welds. The vehicle
longitudinal axis indicates the orientation of the system should it be installed on a
vehicle. Two energy absorbers are located at each end of a channel or cross beam
which acts to transfer an applied load to each absorber. Each absorber itself is a
tapered thin-walled rectangular tube having one tapered side and a closed face at the
small end.

Channel brace

Channel
Trigger

Energy absorber
Vehicle
longitudinal axis Mounting plate

Figure 8.2-1 Energy absorbing system consisting of tapered rectangular tubes


(green component labelled “energy absorber”).

As identified in Chapter 2, when thin-walled tubes and columns are axially crushed
there is an associated initial peak load. In the case of the bumper system in an impact
situation this would lead to peak decelerations which may be above those which the
protected vehicle and its occupants can withstand. Furthermore, these decelerations
may be sufficiently high and prolonged to trigger and deploy the vehicle’s SRS
(Supplementary Restraint System) airbag in an otherwise “no deploy” event (Bignell
2004). The trigger component, a concept conceived and developed by the author, is
designed to reduce this peak load and hence provide a relatively stable load response
for the absorber as it is crushed. The vanes of the trigger act as stress raisers as they

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 250
indent the closed small end face of the tapered tube. This interaction initiates
localised deformation and yielding rather than deformation across the whole of the
face, and thus avoiding the characteristic peak loads. Figure 8.2-2 shows the
simulated deformed profile of the tube under an axially applied quasi-static load,
with the stress concentrations due to the trigger vanes clearly evident in the end face
of the tube.

Trigger vanes cause stress


concentrations which
reduce the initial peak load

Figure 8.2-2 Trigger/ energy absorber deformation mode under an axial load.

The effectiveness of the trigger is shown in Figure 8.2-3, which compares the
simulated load-deflection response of the system under an axial quasi-static load both
with and without the trigger installed. Thus it can be seen that the trigger could
potentially be used in bumper system design as a means of reducing and controlling
the peak loads transmitted to the vehicle.

160
140
120
100
Load (kN)

80
60
40
20
0
0 10 20 30 40 50 60
Deflection (mm)

Trigger No trigger

Figure 8.2-3 Comparison of absorber response when incorporating the trigger.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 251
The channel brace supports and stiffens the channel in the region of the absorber
attachment, and eases assembly of the channel onto the attachment. The mounting
plate provides a fixed load-bearing support for the absorber, and can be used to
attach the whole system to an external fixture such as a vehicle chassis rail.

8.3 Experimental testing

8.3.1 Quasi-static test setup

Quasi-static testing was carried out in order to validate the finite element model of
the bumper system. The specimen used for quasi-static testing is shown in Figure
8.3-1. Each component of the specimen was fabricated from mild steel plate with the
following thicknesses: 3 mm for the channel centre, 4 mm for the channel brace and
trigger, 1.6 mm for the energy absorber and 8 mm for the mounting plate. Each
component was welded together using fillet welds. The bumper system was then
fully fixed to a rigid test rig to support the specimen, using fillet welds around the
base edge of each energy absorber, as shown in Figure 8.3-1.

Figure 8.3-1 Test specimen of the bumper system welded to its supporting
test rig.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 252
Validation of the finite element model involves comparing the load-deflection
response of the bumper system, and the stresses and strains at various regions of the
system. Strain was measured using rosette strain gauges placed at regions where
stresses were highest, and these regions were first determined using a preliminary
finite element model of the system. A rosette was placed on a side wall of each
energy absorber, and on each trigger and channel brace, giving a total of six rosettes
(Figure 8.3-2). Locating rosettes on both the left-hand-side and right-hand-side
components was so as to account for any lack of symmetry in manufacture or loading
between each absorber assembly. Gauge A was placed on the outer face of the
energy absorber, near the location of the buckle predicted using the finite element
model. Gauge B was placed on the lower vertical vane of the trigger to assess strains
in the region of the connection between the channel and the energy absorber, while
gauge C was placed on the channel brace, below the horizontal trigger vane.

C
A
B

Figure 8.3-2 Strain gauge location.

The deflection response of the channel was measured using a Linear Variable
Displacement Transducer (LVDT) fixed to the cross-head of the testing machine, as
shown in Figure 8.3-3. This deflection would also provide a measure of the
deflection (in mm) of the energy absorbers as the load was applied.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 253
Cross-head

Cross-head extension

Figure 8.3-3 LVDT (circled) to measure deflection of the energy absorbers.

Figure 8.3-4 shows the overall quasi-static test setup. The test rig supporting the
bumper system was positioned on the load-bearing surface of the testing machine.
All measured data, i.e. strain from the strain gauges, channel centre deflection from
the LVDT and applied load from the testing machine, was received by a Remote
Strain Unit (RSU) and processed using a PC. Data acquisition, monitoring and
recording was accomplished using the Labtec Notebook software package operating
with a Visual Basic user interface.

Figure 8.3-4 Quasi-static test setup for the bumper system.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 254
8.3.2 Experimental results

The load-deflection and mean load-deflection response of the tested bumper system
is shown in Figure 8.3-5. It can be seen that the load initially rises rapidly until a
peak value is reached. This is due to the yielding of each energy absorber as it was
indented by the trigger vanes. However, this initial peak load is also due to the
formation of an outward moving buckle near the small end of each absorber, as
illustrated in Figure 8.3-6. The load then decreases as the buckle collapses, until a
deflection of approximately 27 mm at which point the load begins to increase
rapidly. This corresponds to the formation of a second, inward forming buckle
towards the large end of the absorber (Figure 8.3-6). The load decreases as this
buckle collapses, until approximately 50 mm deflection at which point the channel
centre contacted the test rig. Further applied load effectively by-passed each energy
absorber and the test was stopped. The final deformed profile of the bumper system
showed that the two energy absorbers were the most deformed components, while
each displayed identical deformed profiles. The mean load ratio shown in Figure
8.3-5 is relatively stable, which is a desirable response for energy absorbers used in
crashworthiness applications.

140
120
100
Load (kN)

80
60
40
20
0
0 10 20 30 40 50 60
Deflection (mm)

Load Mean load

Figure 8.3-5 Load-deflection and mean load-deflection response of the bumper


system.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 255
Outward forming
buckle

Inward forming
buckle

Figure 8.3-6 Deformation mode of the energy absorber.

The principal strains recorded on the right-hand-side components are shown in


Figure 8.3-7, up to the maximum load of approximately 120 kN. It is evident that the
strains are greatest in the energy absorber (gauge A), and least in the channel brace
(gauge C). This is expected, since the absorber is considerably less stiff and therefore
undergoes much larger deformation. It can also be seen from Figure 8.3-7 that the
principal strains measured on the channel brace increase at a relatively low rate up to
120 kN. This also occurs with the strains in the trigger (gauge B) until a load of
approximately 80 kN is reached, at which point the strains begin to increase more
rapidly. These observations indicate that the channel brace, at the point of measured
strain, has not yielded and plastically (permanently) deformed once the maximum
load of 120 kN is reached, whereas the trigger has possibly yielded at an applied load
of approximately 80 kN. On the other hand, the strains in the absorber increase at a
relatively larger rate until the peak load of 120 kN, indicating significant deformation
of the absorber is occurring before the peak load is reached.

140
120
100
Load (kN)

80
60
40
20
0
-1000 -500 0 500 1000 1500 2000
Strain

Energy absorber- Princ A Energy absorber- Princ B Trigger- Princ A


Trigger- Princ B Channel brace- Princ A Channel brace- Princ B

Figure 8.3-7 Principal strains of the right-hand-side components.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 256
8.4 Development and validation of the finite element model

8.4.1 Description of the finite element model

A model of the bumper system was developed using the finite element code
ABAQUS/Explicit version 6.3. The modelling technique used was similar to that for
the individual energy absorbers studied in previous chapters of this thesis. Figure
8.4-1 shows the mesh, geometry and loading arrangement of the model for the
bumper system. Since the thickness of each component was sufficiently small, all
components were modelled using shell elements of designation S4R, while five
integration points were used through the shell thickness to model bending.

Rigid body translates along


vehicle’s longitudinal axis to
axially crush bumper system

Detail of absorber assembly

Each mounting plate is


supported by a fixed
rigid body (not shown)

Vehicle’s
longitudinal axis

Figure 8.4-1 Finite element model of the bumper system.

The welded connections between each component were simulated using tied
constraints. The base of each mounting plate was fully fixed to a supporting rigid
body using a tied constraint. This rigid body was modelled as an analytical rigid
surface and constrained in all degrees of freedom to support the bumper system. In
this case the rigid body was used to simulate the test rig supporting the bumper
system during quasi-static testing.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 257
Quasi-static loading was simulated using an analytical rigid surface which was
positioned in front of the channel and constrained to translate over a pre-defined
displacement such that it axially crushed the bumper system. The rigid body
essentially simulates the loading cross-head of the testing machine used in the quasi-
static test. The motion of the rigid body was defined using an amplitude curve via the
AMPLITUDE option in ABAQUS/Explicit, which moved the rigid body over a time
duration equal to the total step time for the nonlinear analysis. To ensure an accurate
and efficient quasi-static analysis the SMOOTH STEP sub-option was used. This
option ensures the loading rigid body has smooth motion and thus avoids any
sudden, jerky movement which can induce noisy and inaccurate solutions (Hibbitt,
Karlsson & Sorensen, Inc. 2002a). Figure 8.4-2 shows the amplitude curve
describing the motion of the rigid body, with zero gradient at the start and finish of
the step time to ensure smooth motion.

1
Displacement amplitude

0
Time Step time

Figure 8.4-2 Displacement curve for the loading rigid body.

A FREQUENCY linear perturbation analysis step was used to determine the step
time for the nonlinear crushing analysis, as recommended when using the
ABAQUS/Explicit code to perform quasi-static nonlinear analyses (Hibbitt, Karlsson
& Sorensen, Inc. 2002a). As for the finite element model developed for the
individual energy absorbers of earlier chapters, the suitable step time chosen for the
nonlinear crushing analysis was 30 ms (refer to Section 3.2.2 of Chapter 3).

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 258
Contact within each energy absorber as the tube walls buckle under the applied load
was modelled using the 3D self-contact interaction option available in
ABAQUS/Explicit. Self-contact between the absorber walls during collapse, and
surface-to-surface contact between each component were defined using the finite
sliding “penalty” based contact algorithm with contact pairs and “hard” contact.
Respective Coulomb friction coefficients of 0.1 and 0.25 for self-contact and all
other surface-to-surface contact were found to produce suitable results. Since the
same mild steel grade was used as for the individual energy absorbers examined in
earlier chapters of this thesis, the isotropic plasticity material model used in the finite
element model was the same as that defined in Chapter 3.

8.4.2 Comparison between the finite element and experimental


results

Figure 8.4-3 shows the experimental and predicted deformed profiles of the bumper
system at a cross-head deflection of 59 mm. Overall, the lobe formation has been
well predicted.

Outward forming lobe

Figure 8.4-3 Experimental and predicted deformed profiles of the bumper


system.

The experimental and predicted load-deflection and mean-load deflection response is


compared in Figure 8.4-4. It can be seen that the elastic portion of the experimental
load-deflection curve is less stiff than for the predicted curve. This is due to the
elastic deformation of the cross-head extension used in the experiment (Figure
8.3-3), which was modelled as a rigid body. Due to this discrepancy the curves are
“out of phase”. Nevertheless, the initial peak load, as an important energy absorption
parameter, is well predicted by the FE model, as is the mean load-deflection
response.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 259
140
120
Load (kN) 100
80
60
40
20
0
0 10 20 30 40 50 60
Deflection (mm)

Finite element model Experiment

Load-deflection response
120

100
Mean load (kN)

80

60

40

20

0
0 10 20 30 40 50 60
Deflection (mm)

Finite element model Experiment

Mean load-deflection response


Figure 8.4-4 Comparison of experimental and predicted quasi-static response
for the bumper system.

8.5 Parametric study

The aim of the parametric study was to compare the dynamic impact response of
straight and tapered rectangular tubes when mounted in an energy absorbing system.
The validated finite element model of the bumper system was used as the basis for
the study, while the absorber dimensions were modified to include a range of
absorbers as shown in Table 8.5-1.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 260
Table 8.5-1 Absorber dimensions used in the parametric study.

W Absorber θ L W B h
Geometry (deg) (mm) (mm) (mm) (mm)
L
Straight 200 70 50 2
B
Tapered 10 200 70 50 2
W

The length of the absorber was increased to 200 mm, which is typical for crush cans
used at the front of vehicles. The wall thickness, h was increased to 2 mm to increase
the energy absorption capacity of the tubes, and hence allow a wider range of impact
energies to be studied. The same modelling technique was used to simulate dynamic
impact loading as for the individual tubes examined in earlier chapters of this thesis.
This involved axially impacting the bumper system onto a rigid body which was
fully fixed in space, while the rigid body to which the bumper system was attached
was assigned a point mass and initial velocity. A step time of 30 ms was used as for
the numerical analyses presented thus far in this thesis. The effects of strain rate were
only included for the absorber, since this component underwent the most severe
deformations and strains. The effects of strain rate were included using the Cowper-
Symonds equation with parameter values D = 6844 s-1 and q = 3.91, as detailed in
Chapter 3. As for the simulations presented in Chapter 7, the dynamic friction
coefficients were 0.1 for self-contact within each tube, and 0.2 for all other contact
between the various components.

As identified in earlier chapters, real-world energy absorbing systems are subjected


to combinations of axial and oblique loading. Thus, the response of the bumper
system was assessed for two types of loading conditions: symmetric loading and
oblique loading, as illustrated in Figure 8.5-1. Symmetric loading is where the
bumper system axially impacts the fully fixed rigid body, and this rigid body extends
across the full width of the channel. The oblique loading condition involves axially

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 261
impacting the bumper system into the fixed rigid body which is oriented at an angle,
α to the horizontal plane in Figure 8.5-1. The impact velocity, V and impact mass, M
were varied to account for a range of typical vehicle speeds and masses. The analysis
and results for each loading condition will be presented and discussed in turn.

Fully fixed
impact surface
α
V V

Supporting rigid body of


mass M

Symmetric loading Oblique loading


Figure 8.5-1 Dynamic loading conditions for the bumper system (top view).

8.5.1 Symmetric loading

For symmetric loading the response of the bumper system was compared when either
a straight or tapered tube was used as the energy absorbing component. Also of
interest was the effect of impact mass and velocity, to account for different vehicle
types and collision speeds. Three impact masses were used: 1000, 1500 and 2000 kg,
to cover the typical range of current passenger vehicle masses. The impact velocity
was varied from 5 to 20 km/hr, in 5 km/hr increments, to represent the range of low
speed impacts for which vehicle front crush cans are typically designed. Since the
loading and geometry were symmetric, only half the bumper system was modelled
(i.e. one set of absorber components and half the channel), by applying suitable
boundary conditions mid-way along the length of the channel.

Figure 8.5-2 shows the velocity of the moving rigid body as a function of its
deflection, for impact velocities of 10 and 20 km/hr. Here, a straight tube has been
used as the energy absorber. Similar trends as those reported here were also found
when incorporating a tapered tube. It can be seen that the velocity of the rigid body
remains relatively constant for the first 3 mm of deflection. This is due to the initial
gap between the channel and the impact surface. Figure 8.5-2 also shows that the

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 262
velocity-deflection curve is relatively smooth as the tubes crush. This was also found
when using tapered tubes, and highlights the advantage of using such thin-walled
structures in energy absorbing systems, when stable, controlled retardation is
required during an impact event.

12
Velocity of the rigid body (km/hr)

10

0
0 5 10 15 20 25 30 35 40 45 50 55
Deflection of the rigid body (mm)

M = 1000 kg M = 1500 kg M = 2000 kg

Impact velocity, V = 10 km/hr

25
Velocity of the rigid body (km/hr)

20

15

10

0
0 25 50 75 100 125 150 175
Deflection of the rigid body (mm)

M = 1000 kg M = 1500 kg M = 2000 kg

Impact velocity, V = 20 km/hr


Figure 8.5-2 Velocity-deflection response of the bumper system with a straight
tube as the energy absorber.

Each curve in Figure 8.5-2 is shown up to the maximum deflection, i.e. at which the
bumper system was either brought to rest or when “bottoming-out” of the absorber
occurred. Clearly, the maximum crush distance increases as the impact mass and

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 263
velocity increase, due to the associated increase in impact energy. For the straight
tube, bottoming-out occurred for an impact velocity of 20 km/hr and impact masses
of 1500 and 2000 kg. Thus, the maximum impact energy which the bumper system
can absorb without excessive forces being transmitted to the vehicle (via bottoming
out) can be determined for a given set of absorber dimensions. In another light, it can
be said that the velocity change of the bumper system up to a given deflection is less
as the impact mass and velocity increase. In other words, the effect that the absorbers
have on diminishing the velocity of the bumper system decreases. This was also
found for Vehicle Frontal Protection Systems (Bignell 2004), and is also a measure
of the “effectiveness” of the energy absorbing system to absorb sufficient impact
energy within a given crush distance, and hence protect the vehicle structure. By way
of illustration, Figure 8.5-3 shows the change in velocity of the moving rigid body up
to a deflection of 20 mm for various impact masses and velocities, for the present
bumper system with straight tube energy absorbers.

6
Change in velocity of the rigid

5
body (km/hr)

0
1000 1500 2000
Impact mass, M (kg)

V = 10 km/hr V = 15 km/hr V = 20 km/hr

Figure 8.5-3 Change in velocity of the bumper system up to a deflection


of 20 mm.

Thus far the dynamics of the bumper system have been investigated, and the results
show the significant effect of impact mass and velocity on the system’s response.
Also of interest is the effect of these loading parameters on the energy absorption
response of the system. Figure 8.5-4 and Figure 8.5-5 show the load-deflection

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 264
response for both straight and tapered tubes, respectively, for various impact masses
and velocities. As found in previous chapters, increasing the impact mass for a given
impact velocity has no significant effect on the crush load up to a given deflection.
However, the maximum deflection increases due to the associated increase in impact
energy, as identified earlier.

120

100
Dynamic load (kN)

80

60

40

20

0
0 2 4 6 8 10 12 14
Deflection of the rigid body (mm)

M = 1000 kg M = 1500 kg M = 2000 kg

V = 5 km/hr

160
140
Dynamic load (kN)

120
100
80
60
40
20
0
0 20 40 60 80 100 120 140 160
Deflection of the rigid body (mm)

M = 1000 kg M = 1500 kg M = 2000 kg

V = 20 km/hr
Figure 8.5-4 Load-deflection response of the bumper system with a straight tube
as the energy absorber.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 265
120

100
Dynamic load (kN)
80

60

40

20

0
0 2 4 6 8 10 12 14
Deflection of the rigid body (mm)

M = 1000 kg M = 1500 kg M = 2000 kg

V = 5 km/hr

160
140
Dynamic load (kN)

120
100
80
60
40
20
0
0 20 40 60 80 100 120 140 160
Deflection of the rigid body (mm)

M = 1000 kg M = 1500 kg M = 2000 kg

V = 20 km/hr
Figure 8.5-5 Load-deflection response of the bumper system with a tapered tube
as the energy absorber.

The different load-deflection response for each velocity is due to the different
degrees of deformation of the energy absorbers. For an impact velocity of 5 km/hr
the absorber had not deflected far enough to contact the inside of the channel brace,
whereas for an impact velocity of 20 km/hr the absorber had undergone significant
progressive deformation, as evidenced by the load fluctuations. Nevertheless, it
appears that the load-deflection response for the tapered tube is more stable than for
the straight tube, which is desirable for crashworthiness. The deformation sequence
of the bumper system will be addressed in a later discussion.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 266
Chapter 6 demonstrated that the Dynamic Amplification Factor or DAF of
individually crushed tapered tubes is less than that of straight tubes, indicating that
the dynamic response of the former is less influenced by inertia effects. Of interest is
the DAF for the bumper system when either a straight or tapered tube is used as the
energy absorber. Figure 8.5-6 shows the DAF as function of impact velocity at
various rigid body deflections, for an impact mass of 2000 kg. For low deflections,
where the effect of the trigger on the response is more prevalent, the impact velocity
does not have a significant effect on the DAF. However as the deflection increases,
the impact velocity has more effect, while the dynamic amplification of the bumper
system with the tapered tube is less influenced by velocity compared with using a
straight tube as the energy absorber. Thus, it appears that using a tapered tube
reduces the inertia effects on the dynamic response of the bumper system.

The most important practical conclusions which can be drawn from the symmetric
loading analysis are as follows:

(1) The trigger mechanism serves to reduce the dynamic effects on the energy
absorption of the bumper system for low deflections. Nevertheless, it appears
that the load-deflection response for the tapered tube is more stable than for
the straight tube, which is desirable for crashworthiness. Thus a tapered tube
provides a means of controlling the response.

(2) For impacts of higher energy, the trigger mechanism has less influence on the
dynamic response of the system for higher deflections. Furthermore,
incorporating a tapered tube reduces the dynamic amplification of the energy
absorption for these higher deflections.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 267
1.05

Dynamic Amplification Factor


1.04
1.03
1.02
1.01
1
0.99
0.98
10 12 14 16 18 20
Impact velocity, V (km/hr)

Straight Tapered

Rigid body deflection = 20 mm


1.14
Dynamic Amplification Factor

1.13
1.12
1.11
1.1
1.09
1.08
1.07
1.06
10 12 14 16 18 20
Impact velocity, V (km/hr)

Straight Tapered

Rigid body deflection = 35 mm


1.13
1.12
Dynamic Amplification Factor

1.11
1.1
1.09
1.08
1.07
1.06
1.05
1.04
1.03
1.02
10 12 14 16 18 20
Impact velocity, V (km/hr)

Straight Tapered

Rigid body deflection = 50 mm


Figure 8.5-6 DAF at various deflections with either a straight or tapered tube as
the energy absorber.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 268
Figure 8.5-7 shows the crush sequence of the bumper system under an axial impact
load, in which a straight tube is used as the energy absorber. The impact event
corresponds to an impact mass of 1000 kg and an impact velocity of 15 km/hr. The
numerical simulations show that the channel impacts the fixed rigid surface first, and
thereafter the channel remains in contact with the surface. The rest of the bumper
system then “catches up” with the channel until the crush cans impact the inside
surface of the channel brace. The tubes then begin to crush progressively, thus
slowing down the bumper system.

Moving rigid body Channel impacts Tube impacts


rigid surface inside of
Impact surface
channel brace

Rigid body deflection,


d = 14 mm d = 25 mm
d = 0 mm

d = 30 mm d = 35 mm d = 42 mm

d = 48 mm d = 55 mm dmax = 59 mm

Figure 8.5-7 Crush sequence of the bumper system under an axial impact load.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 269
8.5.2 Oblique loading

The effects of load angle on the mean load and energy absorption of the bumper
system were investigated. In particular, the ability of the system to maintain its
energy absorption capacity under increasing load angles was of interest from a
practical point of view. Furthermore, the structural properties of the channel centre
become important under oblique loading conditions, because this member becomes a
load path between each energy absorber. Thus, the influence of the channel centre’s
stiffness on the energy absorption response of the bumper system was examined by
controlling the channel’s wall thickness. Initially a 6 mm wall thickness was ascribed
to the channel, while the remaining geometry parameters of the bumper system were
the same as for symmetric loading. The full finite element model of the system (i.e.
two sets of absorber components and a full channel) was used to take account of the
unsymmetric loading conditions. The impact mass and velocity were 1500 kg and 20
km/hr, respectively.

Figure 8.5-8 shows the mean load of the bumper system as a function of deflection
and load angle, when either a straight or tapered tube is used as the energy absorber.
Clearly, for each type of absorber, increasing the load angle decreases the mean load
up to a given deflection. In Chapter 7 it was identified that such a phenomenon for
individual tubes may partly be due to a reduction in the initial peak load. However, it
is evident from Figure 8.5-8 that the presence of the trigger mechanism in the
bumper system serves to increase the initial peak load as the load angle increases
from 00 (axial loading). Furthermore, the initial peak load remains relatively
constant as the taper angle increases from 100. Thus, the reduction in mean load with
increasing load angle must be due to another explanation. It is most likely that this
response is due to a phenomenon identified in Chapter 7, in which the lobes formed
in the tube are pushed laterally away from the sides of the tube, thus reducing their
amount of axial compression and hence energy absorption. The corresponding
reduction in energy absorption with increasing load angle is depicted in Figure 8.5-9.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 270
80

Dynamic mean load (kN)


70
60
50
40
30
20
10
0
0 10 20 30 40 50 60 70 80 90 100 110 120
Rigid body deflection (mm)

Axial loading Load angle = 10 deg Load angle = 15 deg


Load angle = 20 deg Load angle = 25 deg Load angle = 30 deg
Load angle = 35 deg Load angle = 40 deg

With a straight tube as the energy absorber

80
Dynamic mean load (kN)

70
60
50
40
30
20
10
0
0 10 20 30 40 50 60 70 80 90 100 110 120
Rigid body deflection (mm)

Axial loading Load angle = 10 deg Load angle = 15 deg


Load angle = 20 deg Load angle = 25 deg Load angle = 30 deg
Load angle = 35 deg Load angle = 40 deg

With a tapered tube as the energy absorber


Figure 8.5-8 Effect of load angle on the mean load-deflection response of the
bumper system.

As identified in Chapter 7, it is important to determine the critical load angle at


which the energy absorbers in the bumper system make the transition from
progressive axial collapse to global bending collapse. This is depicted in Figure
8.5-10, which shows the mean crush load calculated up to a moving rigid body
deflection, d of 120 mm for each load angle. Also compared is the response when
either a straight or tapered rectangular tube is used as the energy absorber.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 271
Dynamic absorbed energy (kJ)
8
7
6
5
4
3
2
1
0
0 10 20 30 40 50 60 70 80 90 100 110 120
Rigid body deflection (mm)

Axial loading Load angle = 10 deg Load angle = 15 deg


Load angle = 20 deg Load angle = 25 deg Load angle = 30 deg
Load angle = 35 deg Load angle = 40 deg

With a straight tube as the energy absorber


Dynamic absorbed energy (kJ)

8
7
6
5
4
3
2
1
0
0 10 20 30 40 50 60 70 80 90 100 110 120
Rigid body deflection (mm)

Axial loading Load angle = 10 deg Load angle = 15 deg


Load angle = 20 deg Load angle = 25 deg Load angle = 30 deg
Load angle = 35 deg Load angle = 40 deg

With a tapered tube as the energy absorber


Figure 8.5-9 Effect of load angle on the energy absorption-deflection response of
the bumper system.

As for individual energy absorbers, the response of the bumper system under oblique
loading conditions can be divided into three regions: a region where progressive
axial collapse dominates, a region where there is a transition from progressive to
global bending collapse, and finally a region where global bending collapse
dominates the response. From Figure 8.5-10 it appears that the transition region, and
hence the critical load angle, lies between load angles of 300 and 350. This was
verified from the simulated deformed profile of the bumper system, as will be
examined shortly.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 272
70
60

Dynamic mean load (kN)


50
40
30
20
10
0
0 10 20 30 40 50
Load angle, α (deg)

Straight energy absorber Tapered energy absorber

Figure 8.5-10 Effect of load angle on the mean load of the bumper system under
dynamic impact loading (calculated up to d = 120 mm).

Several important observations and conclusions can be drawn from Figure 8.5-10 in
terms of the response of the bumper system. Firstly, and perhaps most obviously, it is
interesting to note that the mean load response either side of the transition region is
approximately the same when either a straight or tapered tube is used as the energy
absorber. This is in contrast to the findings of Chapter 7, where it was found that the
mean load and hence energy absorption capacity of tapered tubes decreases to a
lesser extent than for straight tubes of equivalent cross-section dimensions. Thus it
appears that the trigger mechanism reduces this relative effect between straight and
tapered rectangular tubes. Furthermore, the front end of each absorber (i.e. the end
closest to the impact surface) on the left-hand-side components underwent
progressive deformation inside the channel centre for load angles up to 300. Thus the
similar contact conditions between each absorber and the channel and other
components reduce the effect of tapering the energy absorber. The second important
observation from Figure 8.5-10 is that the mean load for each tube remains relatively
constant over the region in which the absorbers undergo progressive collapse (α <
300). This is in contrast to the response for the individual straight and tapered tubes
examined in Chapter 7, for which the mean load dropped by approximately 50 %
between axial loading and the transition region (refer to Figure 7.4-5 of chapter 7).
This may be due to load-sharing between the two tubes of each bumper system. It
may also be due to the contact between the absorbers and channel components

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 273
(channel brace, trigger and trigger vanes), in particular the friction between these
components, such that the mean load is maintained as the load angle increases. This
is a desirable characteristic when the energy absorption capacity of the bumper
system must be maintained under oblique loading. Nevertheless, Figure 8.5-10 shows
that the mean load over the transition region is more maintained when a tapered tube
is used as the energy absorber.

The previous observations are further evident in Figure 8.5-11, which shows the
energy absorbed by each tube up to a moving rigid body deflection of 120 mm, as a
percentage of the energy absorbed under axial loading. Clearly, the energy
absorption response for load angles less than the critical angle is relatively the same
when either a straight or tapered tube is used as the energy absorber. Furthermore,
the transition region can be clearly seen between α = 30 and 350. Nevertheless, the
energy absorption during global collapse is more maintained for using tapered tubes.

120

100

80
axial
Ed /Ed
60
(%)
40
20
0
0 10 15
20 25 30
35 40 45 50
Load angle, α (deg)

Straight energy absorber Tapered energy absorber

Figure 8.5-11 Reduction of absorbed energy with increasing load angle.

The effect of the channel’s stiffness on the energy absorption response of the bumper
system under oblique loading was examined by reducing the wall thickness of the
channel to 3 mm. The results when incorporating either a straight or tapered tube
energy absorber are shown in Figure 8.5-12. It appears that the channel’s wall
thickness has no significant effect on the mean load of the bumper system when
either progressive collapse or global bending collapse dominates the response.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 274
80
70

Dynamic mean load (kN)


60
50
40
30
20
10
0
0 10 20 30 40 50
Load angle, α (deg)

Channel wall thickness = 6 mm Channel wall thickness = 3 mm

With a straight tube as the energy absorber

70
60
Dynamic mean load (kN)

50
40
30
20
10
0
0 10 20 30 40 50
Load angle, α (deg)

Channel wall thickness = 6 mm Channel wall thickness = 3 mm

With a tapered tube as the energy absorber


Figure 8.5-12 Effect of channel wall thickness on the response of the bumper
system under oblique loading.

However, it is evident that decreasing the channel’s wall thickness reduces the mean
crush load just before the transition region (30-350), nevertheless, this reduction is
less for the tapered tube. Furthermore, the load angle range containing the critical
load angle did not change when reducing the channel’s wall thickness.

Thus far the energy absorption response of the bumper system has been quantified
for oblique loading conditions. Also important is the deformation response, since this
allows unstable collapse modes to be identified. Figure 8.5-13 and Figure 8.5-14

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 275
show the deformed profile of the bumper system when a straight or tapered tube is
used as the energy absorber, respectively. The channel’s wall thickness is 6 mm. The
profiles are shown from a plan view, while the channel is transparent to show the
deformed energy absorbers. The profiles correspond to a deflection of 150 mm, close
to the maximum or “bottoming out” deflection of the system. The bumper system
moves from right to left across the page.

100 load angle 250 load angle 300 load angle 350 load angle

Figure 8.5-13 Deformed profile of the bumper system with a straight tube
energy absorber (top view shown).

The deformation response of the bumper system under oblique loading is similar
when either a straight or tapered tube is used as the energy absorber. For each load
angle the impact surface was oriented such that the lower energy absorber shown in
each figure was the first to impact the surface, which is shown by the heavy line in
Figure 8.5-13. Up to a load angle of 300 this energy absorber undergoes primarily
progressive axial deformation, as well as some lateral deformation.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 276
V

100 load angle 250 load angle 300 load angle 350 load angle

Figure 8.5-14 Deformed profile of the bumper system with a tapered tube
energy absorber (top view shown).

The top energy absorber shown in each figure undergoes slight progressive
deformation at its front end up to a load angle of 300, such that most of the impact
energy is absorbed by the lower absorber. This causes the channel to rotate anti-
clockwise about the top absorber. However, at a load angle of 300 this top absorber
begins to buckle globally near its base, causing a slight reduction in mean load, as is
evident from Figure 8.5-10. The whole bumper system maintains its energy
absorption capacity since the lower absorber still buckles progressively, though with
large lateral deformation. Increasing the load angle to 350 results in global
deformation of both absorbers, thus causing a significant reduction in the energy
absorption capacity of the bumper system (Figure 8.5-11). For such global collapse,
the entire bumper system tends to displace laterally towards the top of the page, as
indicated by the feint arrow in Figure 8.5-13 and Figure 8.5-14.

Reducing the wall thickness of the channel to 3 mm causes the channel to buckle for
several load angles, as shown in Figure 8.5-15 and Figure 8.5-16 when either a
straight or tapered tube is used as the energy absorber, respectively.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 277
V

100 load angle 250 load angle 300 load angle 350 load angle

Figure 8.5-15 oblique loading response of the bumper system when the
channel’s wall thickness is reduced to 3 mm (with a straight tube as the energy
absorber).

100 load angle 250 load angle 300 load angle 350 load angle

Figure 8.5-16 oblique loading response of the bumper system when the
channel’s wall thickness is reduced to 3 mm (with a tapered tube as the energy
absorber).

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 278
Clearly, the buckling of the channel is most significant for load angles of 25 and 300,
i.e. as the load angle approaches the transition region. This affects the buckling mode
of the lower straight tube energy absorber, which appears to be buckling globally for
a load angle of 300 in Figure 8.5-15. By comparison, this same absorber buckled
progressively for a load angle of 300 when the channel’s wall thickness was higher,
as was shown in Figure 8.5-13. This causes a subsequent reduction in the mean load
and hence energy absorption capacity of the bumper system, as evidenced by Figure
8.5-12. However, the response of the bumper system with a tapered tube as the
energy absorber is less sensitive to the channel’s wall thickness. Figure 8.5-16 shows
that the lower tapered tube energy absorber still undergoes progressive collapse for a
load angle of 300, even when the channel’s wall thickness is reduced to 3 mm. Thus,
the mean load of the bumper system is more maintained over this load angle range
when using a tapered tube rather than a straight tube for the energy absorber, as is
clear from Figure 8.5-12. The practical outcome of these observations is that using a
tapered tube as the primary energy absorbing component in the bumper system is
more likely to maintain the energy absorption capacity of the system under oblique
loading compared with using a straight tube as the energy absorber.

Figure 8.5-17 compares the deformation of the channel for a load angle of 300 with a
straight tube as the energy absorber, for channel wall thicknesses of 6 and 3 mm.

Bm

Channel wall thickness = 6 mm Channel wall thickness = 3 mm

Figure 8.5-17 Dependence of the bumper system’s deformation mode on the


channel’s wall thickness under oblique loading (here α = 300).

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 279
The energy absorber on the lower left of each figure is the one which is closest to the
impact surface. For a channel wall thickness of 6 mm this energy absorber undergoes
primarily progressive collapse, whereas decreasing the channel’s wall thickness to 3
mm causes this absorber to buckle globally. To explain this phenomenon, it appears
that the energy absorber which is situated furthest from the impact surface provides a
bending moment, designated Bm in Figure 8.5-17, which opposes the bending force
acting on this absorber due to the oblique loading condition. The bending moment,
Bm, together with the stiffness of the channel, provides a support for the energy
absorber closest to the impact surface, such that this energy absorber deforms
progressively. Reducing the channel’s stiffness by decreasing its wall thickness
effectively removes some of the support for the energy absorber, such that it buckles
globally. When this energy absorber is a tapered tube, however, it still deforms
progressively due to the higher mean force required to globally buckle the tube, as
was shown in Chapter 7.

In summary, the results of this section show that the stiffness of the channel can have
a significant effect on the energy absorption response of the bumper system for load
angles approaching and close to the critical load angle. Practically, therefore, even
though reducing the wall thickness of the channel favours reduced vehicle weight,
care must be taken to ensure the channel’s stiffness remains sufficient to promote
progressive buckling of each energy absorber, and hence sustained energy
absorption.

8.6 Conclusions and design guidelines

This chapter has investigated the impact energy absorption response of a bumper
system consisting of straight and tapered thin-walled rectangular tubes. A finite
element model of the system was validated using experimental techniques, and the
energy absorption response was quantified for both axial and oblique dynamic
impact loads.

The conclusions and design guidelines which can be drawn from this chapter are
general since the applications are diverse in which tapered rectangular tubes may be

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 280
used in energy absorbing systems. Nevertheless, the following important
observations and recommendations arise from the results of this chapter. Overall, the
results highlight the importance of analysing thin-walled energy absorbers as part of
an energy absorbing system, since the response of the absorbers may be different to
when they are treated on their own.

(1) Straight and tapered thin-walled rectangular tubes are advantageous for use as
energy absorbers since they provide a relatively smooth velocity-deflection
response which promotes stable, controlled retardation as required during an
impact event.

(2) The velocity change of an energy absorbing system up to a given deflection is


less as the impact mass and velocity increase. This phenomenon provides
measurement of the “effectiveness” of the system to absorb sufficient impact
energy within a given crush distance, and hence protect the vehicle structure.

(3) The axial load-deflection response of the bumper system is more stable when
a tapered rectangular tube is used as the energy absorber compared with a
straight rectangular tube. This is desirable for crashworthiness applications
when controlled retardation of the vehicle structure is required.

(4) The presence of the trigger mechanism serves to reduce the dynamic effects
on the energy absorption of the bumper system for low deflections. Thus, for
impacts of relatively low energy which produce small deflections, it makes
no significant difference to include either a straight or tapered tube as the
energy absorber. Thus, the type of absorber can be chosen based on other
factors such as weight reduction and installation.

(5) For impacts of higher energy, the trigger mechanism has less influence on the
dynamic response of the system for higher deflections. Furthermore, using a
tapered tube reduces the inertia and hence “strengthening” effects on the
dynamic response of the bumper system for these higher deflections.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 281
(6) The mean crush load of the energy absorbing system decreases under oblique
loading. As such, it is important to analyse the response of the energy
absorbing system under the load angles expected in service.

(7) For the energy absorbing system examined in this chapter, the mean load for
each tube remains relatively constant over the load angle range in which the
absorbers undergo progressive collapse. This is in contrast to the response for
the individual straight and tapered tubes, and is a desirable characteristic
when the energy absorption capacity of the bumper system must be
maintained under oblique loading. Nevertheless, the mean load over the
transition region is more maintained when a tapered tube is used as the
energy absorber.

(8) The stiffness of the load-transferring member (or bumper beam in


automobiles) between the energy absorbing tubes can have a significant
effect on the energy absorption response of the energy absorbing system for
load angles approaching and close to the critical load angle. A decrease in
beam stiffness reduces the energy absorption capacity of the tubes near the
critical load angle. Practically, therefore, even though reducing the wall
thickness of the beam favours reduced vehicle weight, care must be taken to
ensure the beam’s stiffness remains sufficient to promote progressive
buckling of each energy absorber, and thus maintaining the energy absorption
capacity of the system. Nevertheless, the response of the energy absorbing
system with a tapered rectangular tube as the energy absorber is less sensitive
to the beam’s wall thickness.

(9) From these observations, it is first important to establish the load angle range
expected in service. Then, for a typical range of tube dimensions and
geometries, the load angles which fall within the transition region for these
geometries can be determined. If these load angles are likely to be expected
in service, then decisions can be made whether to tailor the tube dimensions
such that the transition region is adjusted, or such that the energy absorption
capacity over the transition region is enhanced.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 282
(10) The practical outcome of these conclusions is that using a tapered
rectangular tube as opposed to a straight rectangular tube as the primary
energy absorbing component, is more likely to maintain the energy
absorption capacity of the system under oblique loading.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 283
9 Conclusions

9.1 Contributions of the thesis

The primary aim of this thesis was to generate research information on the impact
and energy absorption of tapered thin-walled rectangular tubes, to facilitate their
application in energy absorbing systems.

In accordance with the objectives outlined in Chapters 1 and 2, this thesis has
examined the impact and energy absorption of straight and tapered rectangular tubes
under axial and oblique loading. The geometry parameters varied were the tube wall
thickness, taper angle, number of tapers and cross-section dimensions. The loading
parameters varied were the impact mass, velocity, duration and angle of applied load.
To investigate the response of straight and tapered rectangular tubes in an energy
absorbing system, they were incorporated with a vehicle bumper which may be used
in crashworthiness applications. Finite element models were developed to simulate
the response of the individual tubes and the vehicle bumper system. These models
were validated using experiments conducted by the author, and existing theoretical
and numerical models. Extensive finite element analyses were then conducted to
simulate the response of both the individual tubes and the system. Thus, these
validated finite element models are one of the main contributions of this thesis since
they provide a cost-effective means of performing parametric studies on the tubes as
part of a “virtual” design cycle, for example.

One other main contribution of this thesis is a better understanding of the response of
tapered rectangular tubes under both axial and oblique loading. Based on the relative
effects of each geometry and loading parameter, design guidelines have been
developed to facilitate the use of such structures as energy absorbers in impact
applications.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 284
Overall, the results highlight the advantages of using tapered rectangular tubes for
impact energy absorbers such as in crashworthiness applications. The most
significant advantages of these tubes arising from the thesis are listed below.

(1) Tapering a rectangular tube provides two extra geometry parameters; the
taper angle and number of tapered sides, which can be used to control the
response. For instance, the peak load required to crush the tubes decreases
with the introduction of a taper, and as the taper angle increases. This is
desirable for minimising the impact loads transmitted to the protected
structure.

(2) The dynamic initial peak and mean loads of straight and tapered rectangular
tubes are greater than the corresponding quasi-static loads due to strain rate
effects, and the likely presence of axial wave propagation and inertia effects.
Nevertheless, tapered tubes are less influenced by lateral inertia effects than
straight tubes. This would be beneficial for impact energy absorption
applications where “strengthening” of the absorber due to inertia effects
needs to be minimised.

(3) The energy absorption response of straight and tapered rectangular tubes
under axial impact loading is influenced by the wall thickness, taper angle
and number of tapered sides. Thus, these geometry parameters may be used
to control the response. However, the wall thickness has the most influence
on the initial peak and mean load for both straight and tapered rectangular
tubes, while the triple-tapered tube generally absorbs the most energy within
a given crush distance. Tapered rectangular tubes have a higher dynamic
crush force efficiency than straight rectangular tubes, and this efficiency is
generally highest for tubes with four tapered sides. However, the properties of
the impact surface have most influence on the energy absorption response of
tubes with two tapered sides compared with the other tubes investigated.

(4) The mean load and energy absorption of both straight and tapered rectangular
tubes decrease significantly as the angle of applied load increases.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 285
Nevertheless, tapering a rectangular tube enhances its energy absorption
capacity under oblique loading.

(5) The axial load-deflection response of the bumper system is more stable when
a tapered rectangular tube is used as the energy absorber compared with a
straight rectangular tube. This is desirable for crashworthiness applications
when controlled retardation of the vehicle structure is required. Furthermore,
using a tapered tube reduces the inertia and hence “strengthening” effects on
the dynamic response of the bumper system. Under oblique impact loading,
the mean load and hence energy absorption capacity of the bumper system is
more maintained when using a tapered rectangular tube as opposed to a
straight rectangular tube as the primary energy absorbing component.

9.2 Recommendations for future work

From the results of this thesis, several recommendations arise for future work, as
follows.

(1) It was shown in Section 3.2.1 of Chapter 3 and Section 4.3.2 of Chapter 4 that
modelling the tubes with rounder corners resulted in closer correlation with
the experimental and existing theoretical results. This was particularly the
case for oblique loading, since the experimental initial peak and mean loads
were over-predicted when rounded corners were not included in the FE
model. Since the majority of numerical studies in the published literature
model thin-walled tubes without corner radii, the author recommends that the
effect of this geometry parameter be further studied for a variety of tube
sections, particularly under oblique loading, since this may have a noticeable
improvement on subsequent FE model validation.

(2) The energy absorbing vehicle bumper system studied in Chapter 8 proves to
be promising as a concept for use in crashworthiness applications. The author
recommends further studies by carried out on similar systems incorporating
tapered tubes, in particular with the use of dynamic impact testing. Such

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 286
would be possible with the use of a pendulum testing facility, suitable for
low-speed impact tests. Such tests could be implemented with the FE
modelling techniques adopted in Chapter 8 to develop other concepts for
vehicle bumper systems. Furthermore, the advantages of tapered rectangular
tubes as impact energy absorbers warrants investigation of their usage in
other energy absorbing systems, notably road-side or highway barriers.

(3) The empirical formula developed in Section 5.5 of Chapter 5 may be made
more accurate and take account of a wider range of tapered tube geometries
by including other geometry parameters such as the corner radius, taper angle
and number of tapers. Furthermore, the particular formula presented in this
thesis has been developed for tube lengths of 300 mm. Developing the
formula for other tube lengths and materials would increase its range of
applicability. Finally, while the formula developed in this thesis is ideal for
quasi-static design, developing future formulas on dynamic results would also
provide an invaluable design tool.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 287
References

Abramowicz, W. 2003. Thin-walled structures as impact energy absorbers. Thin-


Walled Structures. 41: 91-107.

Abramowicz, W. and Jones, N. 1984. Dynamic axial crushing of square tubes.


International Journal of Impact Engineering. 2 (2): 179-208.

Abramowicz, W. and Jones, N. 1986. Dynamic progressive buckling of circular and


square tubes. International Journal of Impact Engineering. 4 (4): 243-270.

Alghamdi, A. 2001. Collapsible impact energy absorbers: an overview. Thin-Walled


Structures. 39: 189-213.

Alghamdi, A. A. A. 2002. Reinversion of aluminium frustra. Thin-Walled Structures.


40 (12): 1037-1049.

Alghamdi, A. A. A. 2002. Folding-crumpling of thin-walled aluminium frusta.


International Journal of Crashworthiness. 7 (1): 67-78.

Alghamdi., A. A. A., Aljawi, A. A. N. and Abu-Mansour, T. M.-N. 2002. Modes of


axial collapse of unconstrained capped frusta. International Journal of Mechanical
Sciences. 44: 1145-1161.

Aljawi, A. A. N. 2000. Finite element and experimental analysis of axially


compressed plastic tubes. European Journal of Mechanical and Environmental
Engineering. 45 (1): 3-10.

Arnaud, P. and Hamelin, P. 1998. Dynamic characterization of structures: a study of


energy absorption in composite tubes. Composites Science and Technology. 58: 709-
715.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 288
Ashby, M. F., Evans, A., Fleck, N. A., Gibson, L. J., Hutchinson, J. W. and Wadley,
H. N. G. 2000. Metal foams: A Design Guide. Boston, MA, USA: Butterworth-
Heinemann.

Beard, S. J. and Chang, F.-K. 2000. Energy absorption of braided composite tubes. In
Chirwa, E. C. and Otte, D. (eds), Proc. ICrash 2000- International Crashworthiness
Conference, London, 6-8 October 2000. London: The Royal Aeronautical Society.

Bignell, P. 2004. Evaluation of the performance and testing techniques of Vehicle


Frontal Protection Systems. Ph.D. Thesis, School of Civil Engineering, Queensland
University of Technology.

Blazynski, T. Z. (ed). 1987. Materials at High Strain Rates. London: Elsevier


Applied Science Publishers Ltd.

Borvik, T., Hopperstad, O. S., Reyes, A., Langseth, M., Solomos, G. and Dyngeland,
T. 2003. Empty and foam-filled circular aluminium tubes subjected to axial and
oblique quasi-static loading. International Journal of Crashworthiness. 8 (5): 481-
494.

Box, G. E. P., Hunter, W. G. and Hunter, J. S. 1978. Statistics for experimenters.


New York: Wiley.

Calladine, C. R. and English, R. W. 1984. Strain rate and inertia effects in the
collapse of two types of energy-absorbing structure. International Journal of
Mechanical Sciences. 26: 689-701.

Campbell, J. D. and Cooper, R. H. 1966. Yield and flow of low-carbon steel at


medium strain rates. In Proc. Conf on the Physical Basis of Yield and Fracture, Inst.
of Physics and Physical Soc., 77-87.

Chan, C. 2000. Fundamentals of crash sensing in automotive air bag systems.


Warrendale, PA, USA: Society of Automotive Engineers, Inc.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 289
Chen, W. 2001. Experimental and numerical study on bending collapse of aluminium
foam-filled hat profiles. International Journal of Solids and Structures. 38: 7919-
7944.

Chen, W. and Wierzbicki, T. 2001. Relative merits of single-cell, multi-cell and


foam-filled thin-walled structures in energy absorption. Thin-Walled Structures. 39:
287-306.

Chen, W., Wierzbicki, T., Breuer, O. and Kristiansen, K. 2001. Torsional crushing of
foam-filled thin-walled square columns. International Journal of Mechanical
Sciences. 43: 2297-2317.

Cheon, S. S., Choi, J. H. and Lee, D. G. 1995. Development of the composite


bumper beam for passenger cars. Composite Structures. 32 (1-4): 491-499.

Chirwa, E. C. 1993. Theoretical analysis of tapered thin-walled metal inverbucktube.


International Journal of Mechanical Sciences. 35 (3/4): 325-351.

El-Sobky, H. and Singace, A. A. 1999. Profiled polymer pipes as re-usable energy


absorption elements. International Journal of Mechanical Sciences. 41: 1385-1400.

Gibson, L. J. and Ashby, M. F. 1997. Cellular Solids, Structure and Properties. 2nd
edn. Cambridge, UK: Cambridge University Press.

Gupta, N. K. (ed). 1993. Plasticity and Impact Mechanics. New Delhi: Wiley Eastern
Limited.

Han, D. C. and Park, S. H. 1999. Collapse behavior of square thin-walled columns


subjected to oblique loads. Thin-Walled Structures. 35: 167-184.

Hanssen, A. G., Langseth, M. and Hopperstad, O. S. 2000. Static and dynamic


crushing of square aluminium extrusions with aluminium foam filler. International
Journal of Impact Engineering. 24 (4): 347-383.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 290
Hanssen, A. G., Langseth, M. and Hopperstad, O. S. 2000. Static and dynamic
crushing of circular aluminium extrusions with aluminium foam filler. International
Journal of Impact Engineering. 24 (5): 475-507.

Hanssen, A. G., Langseth, M. and Hopperstad, O. S. 2001. Optimum design for


energy absorption of square aluminium columns with aluminium foam filler.
International Journal of Mechanical Sciences. 43: 153-176.

Hanssen, A. G., Lorenzi, L., Berger, K. K., Hopperstad, O. S. and Langseth, M.


2000. A demonstrator bumper system based on aluminium foam filled crash boxes.
International Journal of Crashworthiness. 5 (4): 381-392.

Harrigan, J. J., Reid, S. R. and Peng, C. 1999. Inertia effects in impact energy
absorbing materials and structures. International Journal of Impact Engineering. 22:
955-979.

Haug, E. and De Rouvray, A. 1993. Crash response of composite structures. In


Jones, N. and Wierzbicki, T. (eds), Structural Crashworthiness and Failure. Essex:
Elsevier Applied Science Ltd.

Hibbitt, Karlsson and Sorensen, Inc. 2002. Getting Started with ABAQUS/Explicit
Version 6.3. Pawtucket, USA: HKS.

Hibbitt, Karlsson and Sorensen, Inc. 2002. Getting Started with ABAQUS/Standard
Version 6.3. Pawtucket, USA: HKS.

Hosseinipour, S. J. and Daneshi, G. H. 2003. Energy absorption and mean crushing


load of thin-walled grooved tubes under axial compression. Thin-Walled Structures.
41: 31-46.

Hosseinzadeh, R., Mahmood, M. S. and Lessard, L. B. 2004. Parametric study of


automotive composite bumper beams subjected to low-velocity impacts. Composite
Structures. In press.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 291
Huang, M. 2002. Vehicle crash mechanics. New York, USA: CRC Press.

Johnson, W. 1972. Impact Strength of Materials. London, UK: Edward Arnold


(Publishers) Limited.

Johnson, W. and Mamalis, A. G. 1978. Crashworthiness of vehicles. London:


Mechanical Engineering Publications Ltd.

Johnson, W. and Reid, S. R. 1978. Metallic energy dissipating systems. Applied


Mechanics Review. 31 (3): 277–288.

Jones, N. 1989. Structural Impact. Cambridge, UK: Cambridge University Press.

Jones, N. 1999. Some phenomena in the structural crashworthiness field.


International Journal of Crashworthiness. 4 (4): 335-350.

Jones, N. 2000. Dynamic material properties and inelastic failure in structural


crashworthiness. In Chirwa, E. C. and Otte, D. (eds), Proc. ICrash 2000-
International Crashworthiness Conference, London, 6-8 October 2000. London: The
Royal Aeronautical Society.

Jones, N. and Wierzbicki, T. (eds). 1993. Structural Crashworthiness and Failure.


Essex, UK: Elsevier Science Publishers Ltd.

Karagiozova, D. and Jones, N. 2004. Dynamic buckling of elastic-plastic square


tubes under axial impact-II: structural response. International Journal of Impact
Engineering. 30 (2): 167-192.

Karagiozova, D. and Jones, N. 2001. Dynamic effects on buckling and energy


absorption of cylindrical shells under axial impact. Thin-Walled Structures. 39: 583-
610.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 292
Karagiozova, D., Alves, M. and Jones, N. 2000. Inertia effects in axisymmetrically
deformed cylindrical shells under axial impact. International Journal of Impact
Engineering. 24 (10): 1083-1115.

Karbhari, V. M. and Chaoling, X. 2003. Energy absorbing characteristics of circular


frustra. International Journal of Crashworthiness. 8 (5): 471-479.

Kecman, D. 1983. Bending collapse of rectangular and square section tubes.


International Journal of Mechanical Sciences. 25 (9/10): 623-636.

Kecman, D. 1997. An engineering approach to crashworthiness of thin-walled beams


and joints in vehicle structures. Thin-Walled Structures. 28 (3/4): 309-320.

Kelly, D. J. and Rucker, J. R. 1997. New polyurethane polymers for energy


absorbing automotive applications. In SAE International Congress & Exposition,
Detroit, Michigan, 24-27 February 1997. SAE Paper No 970168.

Kim, H.-S. and Wierzbicki, T. 2001. Crush behavior of thin-walled prismatic


columns under combined bending and compression. Computers and Structures. 79:
1417-1432.

Kindervater, C. and Georgi, H. 1993. Composite strength and energy absorption as


an aspect of structural crash resistance. In Jones, N. and Wierzbicki, T. (eds),
Structural Crashworthiness and Failure. Essex: Elsevier Applied Science Ltd.

Kinslow, R. (ed). 1970. High-Velocity Impact Phenomena. New York: Academic


Press, Inc.

Lin, K. and Mase, G. 1990. An assessment of add-on energy absorbing devices for
vehicle crashworthiness. Journal of Engineering Materials and Technology. 112:
406-411.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 293
Langseth, M. and Hopperstad, O. S. 1996. Static and dynamic axial crushing of
square thin-walled aluminium extrusions. International Journal of Impact
Engineering. 18 (7/8): 949-968.

Langseth, M., Hopperstad, O. S. and Berstad, T. 1999. Crashworthiness of


aluminium extrusions: validation of numerical simulation, effect of mass ratio and
impact velocity. International Journal of Impact Engineering. 22: 829-854.

Lorenzi, L., Fuganti, A., Todaro, E. and Fossat, E. 1997. Aluminium foam
applications for impact energy absorbing structures. In SAE International Congress
& Exposition, Detroit, Michigan, 24-27 February 1997. SAE Paper No 970015.

Lorenzo, L., Burr, S. and Fennessy- Ketola, K. 1997. Integrated inner door panel/
energy absorber designs for side impact occupation protection. In SAE International
Congress & Exposition, Detroit, Michigan, 27 February-2 March 1995. SAE Paper
No 960151.

Lu, G. and Yu, T. 2003. Energy absorption of structures and materials. Cambridge,
UK: Woodhead Publishing Limited.

Mamalis, A. G. and Johnson, W. 1983. The quasi-static crumpling of thin-walled


circular cylinders and frusta under axial compression. International Journal of
Mechanical Sciences. 25 (9/10): 713-732.

Mamalis, A. G., Johnson, W. and Viegelahn, G. L. 1984. The crumpling of steel


thin-walled tubes and frusta under axial compression at elevated strain rates: some
experimental results. International Journal of Mechanical Sciences. 26: 537-548.

Mamalis, A. G., Manolakos, D. E., Saigal, S., Viegelahn, G. I. and Johnson, W.


1986. Extensible plastic collapse of thin-wall frusta as energy absorbers.
International Journal of Mechanical Sciences. 28: 219-229.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 294
Mamalis, A. G., Manolakos, D. E., Viegelahn, G. I., Vaxevanidis, N. M. and
Johnson, W. 1986. On the inextensional axial collapse of thin PVC conical shells.
International Journal of Mechanical Sciences. 28: 323-335.

Mamalis, A. G., Manolakos, D. E., Ioannidis, M. B., Kostazos, P. K. and Hassiotis,


G. 2001. Finite element simulation of the axial collapse of thin-wall square frusta.
International Journal of Crashworthiness. 6 (2): 155-164.

Mamalis, A. G., Manolakos, D. E., Ioannidis, M. B. and Papapostolou, D. P. 2004.


Crashworthy characteristics of axially statically compressed thin-walled square
CFRP composite tubes: experimental. Composite Structures. 63: 347-360.

Mamalis, A. G., Manolakos, D. E. and Viegelahn, G. L. 1989. The axial crushing of


thin PVC tubes and frusta of square cross-section. International Journal of Impact
Engineering. 8 (3): 241- 264.

Mamalis, A. G., Manolakos, D. E., Ioannidis, M. B., Kostazos, P. K. and Dimitriou,


C. 2003. Finite element simulation of the axial collapse of metallic thin-walled tubes
with octagonal cross-section. Thin-Walled Structures. 41: 891-900.

Mamalis, A. G., Manolakos, D. E., Demosthenous, G. A. and Ioannidis, M. B. 1997.


Analytical modelling of the static and dynamic axial collapse of thin-walled
fibreglass composite conical shells. International Journal of Impact Engineering. 19
(5/6): 477-492.

Miscow F., P. C. and Al-Qureshi, H. A. 1997. Mechanics of static and dynamic


inversion processes. International Journal of Mechanical Sciences. 39 (2): 147-161.

Murray, N. 1994. When it comes to the crunch- the mechanics of car collisions.
Singapore: World Scientific Publishing Co. Pte. Ltd.

Nagel, G. M. and Thambiratnam, D. P. 2004. A numerical study on the impact


response and energy absorption of tapered thin-walled tubes. International Journal of
Mechanical Sciences. 46 (2): 201-216.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 295
Nagel, G. M. and Thambiratnam, D. P. 2004. Dynamic simulation and energy
absorption of tapered tubes under impact loading. International Journal of
Crashworthiness. 9 (4): 389-399.

Nagel, G. M. and Thambiratnam, D. P. 2005. Dynamic simulation and energy


absorption of tapered thin-walled tubes under oblique impact loading. International
Journal of Impact Engineering. (Accepted and in press).

Nagel, G. M. and Thambiratnam, D. P. 2005. Computer simulation and energy


absorption of tapered thin-walled rectangular tubes. Thin-Walled Structures.
(Submitted).

Rechnitzer, G., Powell, C. and Seyer, K. 1996. Development and testing of energy
absorbing rear underrun barriers for heavy vehicles. In Proc. Fifteenth International
Technical Conference on the Enhanced Safety of Vehicles, Melbourne, Australia, 13-
16 May 1996. USA: U.S. Department of Transportation- National Highway Traffic
Safety Administration.

Reddy, T. Y. and Wall, R. J. 1988. Axial compression of foam-filled thin-walled


circular tubes. International Journal of Impact Engineering. 7 (2): 151-166.

Reddy, T. Y., Yu, T. X. and El-Dufani, A. I. 1996. Plastic collapse of circular tubes
under combined bending and torsion. In The Third Asia-Pacific Symposium on
Advances in Engineering Plasticity and its Applications (AEPA ’96), Hiroshima,
Japan, August 1996.

Reid, S. R. 1993. Plastic deformation mechanisms in axially compressed metal tubes


used as impact energy absorbers. International Journal of Mechanical Sciences. 35
(12): 1035–1052.

Reid, S. R. and Reddy, T. Y. 1986. Static and dynamic crushing of tapered sheet
metal tubes of rectangular cross-section. International Journal of Mechanical
Sciences. 28 (9): 623-637.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 296
Reid, S. R., Reddy, T. Y. and Gray, M. D. 1986. Static and dynamic axial crushing of
foam-filled sheet metal tubes. International Journal of Mechanical Sciences. 28 (5):
295-322.

Reyes, A., Langseth, M. and Hopperstad, O. S. 2002. Crashworthiness of aluminium


extrusions subjected to oblique loading: experiments and numerical analyses.
International Journal of Mechanical Sciences. 44: 1965-1984.

Reyes, A., Hopperstad, O. S. and Langseth, M. 2004. Aluminum foam-filled


extrusions subjected to oblique loading: experimental and numerical study.
International Journal of Solids and Structures. 41: 1645-1675.

Royal Automobile Club of Queensland (RACQ). 2004. The Road Ahead- June/July
2004. The Road Ahead Publishing Co Pty Ltd.

Santosa, S. and Wierzbicki, T. 1998. Crash behavior of box column filled with
aluminium honeycomb or foam. Computers and Structures. 68 (4): 343-368.

Santosa, S., Wierzbicki, T., Hanssen, A. G. and Langseth, M. 2000. Experimental


and numerical studies of foam-filled sections. International Journal of Impact
Engineering. 24: 509-534.

Singace, A. A. and El-Sobky, H. 1997. Behaviour of axially crushed corrugated


tubes. International Journal of Mechanical Sciences. 39 (3): 249-268.

Singace, A. A., El-Sobky, H. and Petsios, M. 2001. Influence of end constraints on


the collapse of axially impacted frusta. Thin-Walled Structures. 39 (5): 415-428.

Spagnoli, A. and Chryssanthopoulos, M. K. 1999. Elastic buckling and postbuckling


behaviour of widely-stiffened conical shells under axial compression. Engineering
Structures. 21: 845-855.

Standards Australia. 1991. AS1391-1991 Methods for tensile testing of metals.


Sydney, Australia: Standards Australia.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 297
Stronge, W. J. 2000. Impact Mechanics. Cambridge: Cambridge University Press.

Su, X. Y., Yu, T. X. and Reid, S. R. 1995. Inertia-sensitive impact energy-absorbing


structures part I: effects of inertia and elasticity. International Journal of Impact
Engineering. 16 (4): 651-672.

Su, X. Y., Yu, T. X. and Reid, S. R. 1995. Inertia-sensitive impact energy-absorbing


structures part II: effect of strain rate. International Journal of Impact Engineering.
16 (4): 673-689.

Tam, L. L. 1990. Strain rate and inertia effects in the collapse of energy absorbing
structures. Ph.D. Thesis, University of Cambridge.

Tam, L. L. and Calladine, C. R. 1991. Inertia and strain rate effects in a simple plate-
structure under impact loading. International Journal of Impact Engineering. 11:
349-377.

Watson, P. 1974. Some methods of absorbing the energy of motor vehicles and their
occupants. In Proc. Fifth International Technical Conference on Experimental Safety
Vehicles, London, 4-7 June 1974. Washington: U.S. Department of Transportation-
National Highway Traffic Safety Administration.

White, M. and Jones, N. 1999. Experimental quasi-static axial crushing of top-hat


and double-hat thin-walled sections. International Journal of Mechanical Sciences.
41: 179-208.

Wierzbicki, T. 1983. Crushing analysis of metal honeycombs. International Journal


of Impact Engineering. 1 (2): 157-174.

Yasui, Y. 2000. Dynamic axial crushing of multi-layer honeycomb panels and


impact tensile behavior of the component members. International Journal of Impact
Engineering. 24: 659-671.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 298
Yu-Hallada, L. C., Kuczynski, E. T. and Weierstall, M. 1998. Polyurethane: the
material of choice for occupant protection and energy management. Journal of
Cellular Plastics. 34: 272-282.

Zhang, J. and Ashby, M. F. 1994. Mechanical selection of foams and honeycombs


used for packaging and energy absorption. Journal of Materials Science. 29: 157-
163.

Zhao, H. and Gary, G. 1998. Crushing behaviour of aluminium honeycombs under


impact loading. International Journal of Impact Engineering. 21 (10): 827-836.

G. Nagel: Impact and Energy Absorption of Straight and Tapered Rectangular Tubes 299

You might also like