You are on page 1of 176

Cross-sectional instability of

aluminium extrusions with complex


cross-sectional shapes

Natalia Kutanova
Cross-sectional instability of
aluminium extrusions with complex
cross-sectional shapes

PROEFSCHRIFT

ter verkrijging van de graad van doctor aan de


Technische Universiteit Eindhoven, op gezag van de
rector magnificus, prof.dr.ir. C.J. van Duijn, voor een
commissie aangewezen door het College voor
Promoties in het openbaar te verdedigen
op dinsdag 23 juni 2009 om 16.00 uur

door

Natalia Kutanova

geboren te Kavalerovo, Rusland


Dit proefschrift is goedgekeurd door de promotoren:

prof.ir. F. Soetens
en
prof.ir. H.H. Snijder
Samenstelling van de Promotiecommissie:

prof. ir. J. Westra TU Eindhoven (voorzitter)


prof. ir. F. Soetens TU Eindhoven
prof. ir. H.H. Snijder TU Eindhoven
prof. ir. F. van Herwijnen TU Eindhoven
prof. ir. F.S.K. Bijlaard TU Delft
prof. dr. T. Peköz Cornell University, USA
dr. ir. J. Mennink TNO
ir. B.W.E.M. van Hove TU Eindhoven

ISBN 978-90-77172-47-6
First printing May 2009
Keywords: cross-sectional stability, aluminium, local buckling, distortional
buckling

This thesis was prepared in LATEX by the author and printed by Print Partners
Ipskamp, Enschede

Cover design: N.Kutanova


Cover photography: Eva&Ed

c
Copyright °2009 by N.Kutanova, The Hague, The Netherlands

All rights reserved. No part of this publication may be reproduced, stored in a


retrieval system, or transmitted in any form or by any means, electronic,
mechanical, photocopying, recording or otherwise, without prior written
permission of the copyright holder.
This research was carried out under the project number MC1.02146 in the
framework of the Research Program of the Materials innovation institute (M2i)
(www.m2i.nl)
Acknowledgments

This thesis is the result of an extensive work, which often seemed to me as a never-
ending period of time. However, the work is finished and that would not have been
possible without people around me who have been very helpful during the time it
took me to write this thesis.
First, I would like to show my utmost gratitude to my first promotor, Frans
Soetens for inviting me to the Netherlands and giving me the opportunity to be-
come a PhD researcher. Frans, thank you very much for sharing your expertise, for
the support at any moment and for immense kindness. I especially appreciate your
patience to correct hundreds of missing articles in my English writing. My grati-
tude also extends to my second promotor Bert Snijder for careful reading any piece
of my work, for critical suggestions and challenging questions. I am deeply grateful
to Jeroen Mennink without whom this work would not have started and, moreover,
successfully ended. My thanks and appreciation goes to my thesis committee mem-
bers, Frans van Herwijnen, Frans Bijlaard, Dianne van Hove and Theoman Peköz. I
had very fruitful discussions with my reading committee and the questions I have
received were a great help.
This thesis would not have been accomplished without funds of the Materials In-
novation Institute. I would like to acknowledge the help and organization support
of people at the M2i head office. The research was carried out at the University of
Eindhoven and also at TNO. I am indebted to many of my TNO and Eindhoven col-
leagues who supported me and provided me assistance in numerous ways. Many
thanks to Sander, Paul, Roel, Ernst and Edwin for informal chats and for the great
working atmosphere. My closest colleague Johan Maljaars was a great source of help
especially during the first year of my research. He has made available his support
with respect to the DIANA program. Johan also inspired me to bike everyday from
the Hague to Delft and was able to bear my company on every Thursday train trip
from Delft to Eindhoven. It would have been an impossible task for me to execute
experiments without the help of people in Pieter van Musschenbroek laboratory:
Theo van de Loo, Erik Wijen, Hans Lamers and Martien Ceelen. It is a pleasure
to thank my student Jeroen Berkmortel for performing a part of the experimental
program.
Further, I would like to thank those closest to me, whose direct or indirect pres-
ence helped making the completion of my work possible. I have been fortunate to

i
meet many good friends without whom my life would be bleak. I thank my friends
in the Netherlands and France. I would like to thank Eva for helping me with a
cover design. I would love to name all of your here, but the list might be too long.
However, some of my Russian friends who are more like a family to me, I have to
mention here: thank you Daria, Andrey, Denis, Andrew, Alexey, Katya and last but
not least, Gena.
It is difficult to express how grateful am I to my mother for her everyday emails
helping me feeling that I am not away from my country and my family. Special
thanks to my sister Katya and my brother Syoma for visiting me in the Netherlands
and making me laugh even in my thoughts about you. Many thanks go to my French
family in law for their kindness and affection.
Finally, I would like to thank my husband Ugo Lafont for an enormous amount
of faith in me, for mental support and lots of wise advices. He encouraged me to
concentrate on completing this thesis more than anyone, while keeping me away
from other responsibilities. Nothing would have been possible without him.

Natalia Kutanova
Eindhoven
May 14, 2009
Summary

Aluminium extrusions are of great interest for different industrial fields such as
structural applications and transport. The extrusion process allows one to optimize
structural elements according to the design requirements with a relative ease. Op-
timization of the shape of the aluminium elements often results in the use of thin-
walled cross-sectional shapes, which increases the complexity of the cross-section.
As a result of thin-walled elements cross-sectional instability - in particular local and
distortional buckling - has a substantial effect on the structural behaviour.
In classification of cross-sectional instability local buckling implies changes in
geometry without any lateral displacement or twist, while for distortional buck-
ling lateral displacement and twist take place with changes in the cross-sectional
geometry. The current design rules used by engineers are limited to local buckling
of simple and symmetrical cross-sections. Hence, these design rules do not pro-
vide an accurate description of distortional buckling behaviour and can not be used
for more complex shapes. Extensive research into cross-sectional instability of alu-
minium structural elements concerning distortional buckling is required, which is
the main subject of this thesis.
An experimental program has been executed in order to predict the ultimate
resistance of aluminium structural elements due to cross-sectional instability. This
program consists of extruded and cold-formed aluminium specimens with Z-shaped,
Angle-shaped and C-shaped sections. Test specimens were subjected to uniform
axial compression. Extensive measurements have been performed on initial im-
perfections of extruded and cold-formed specimens. The influence of a gradual
increase of the complexity of the geometry and material influence on the buckling
behaviour have been investigated. The experimental program has resulted in a set
of test data on the cross-sectional instability of aluminium structural members with
various cross-sectional shapes, including local and distortional buckling, as well as
interaction of modes. These data have been used for the numerical model valida-
tion.
A finite element (FE) model has been developed and validated based on the re-
sults of the experimental program. In this respect, experiments are simulated using
the actual geometry, imperfections and material. The comparison between the FE-

iii
model and test results indicates the accuracy of the numerical prediction. It has been
shown that the FE-model is a useful tool for the prediction of structural behaviour
of uniformly compressed aluminium members with various cross-sectional shapes.
The validated FE-model has been used for a detailed investigation of the actual dis-
tortional buckling behaviour and local-distortional (distortional-local) interaction.
A substantial number of analyses have been carried out to study the distortional
buckling effect using the validated FE-model. Based on the actual distortional buck-
ling behaviour of C-shaped sections, a prediction model is proposed for the calcula-
tion of the ultimate resistance of compressed C-sections subjected to cross-sectional
instability. As a result, the existing model for local buckling prediction and the
newly developed model for distortional buckling prediction including mode inter-
action are able to describe the cross-sectional instability behaviour of aluminium
C-sections. Furthermore, these models are an important step in the investigation of
cross-sectional instability of complex cross-sectional shapes.
List of Figures

1.1 Aluminium profiles application . . . . . . . . . . . . . . . . . . . . . . 1


1.2 Instability types . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.3 Research approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

2.1 Plate geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8


2.2 Buckling coefficient for a simply supported plate. . . . . . . . . . . . 9
2.3 Buckling coefficients kcr of uniformly compressed rectangular plates
with various boundary conditions. . . . . . . . . . . . . . . . . . . . . 10
2.4 Stages of stress distribution in simply supported compressed plates. 10
2.5 Effective width . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.6 Stress-strain diagram based on the Ramberg-Osgood law. . . . . . . . 13
2.7 Relationship between slenderness parameter and axial resistance. . . 16
2.8 Classification of cross-sections . . . . . . . . . . . . . . . . . . . . . . . 17
2.9 Buckling modes of lipped channel in compression. . . . . . . . . . . . 20

3.1 Examples of cross-sections selected for investigation in the current


project . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
3.2 1st subprogram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.3 2nd subprogram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.4 Specimen notations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.5 Z-shaped profile cut to measure radiuses. . . . . . . . . . . . . . . . . 29
3.6 Measured heart-to-heart values of cross-sectional dimensions. . . . . 29
3.7 Schematization of the measuring process. . . . . . . . . . . . . . . . . 30
3.8 Imperfection test set-up. . . . . . . . . . . . . . . . . . . . . . . . . . . 31
3.9 Imperfection measurements for web and flange. . . . . . . . . . . . . 31
3.10 Test set-up for long profile. . . . . . . . . . . . . . . . . . . . . . . . . . 32
3.11 Fitting plane for measured points, using the Least Square method. . . 32
3.12 Schematization of the measured points. . . . . . . . . . . . . . . . . . 33
3.13 Imperfection amplitude calculation (case 1). . . . . . . . . . . . . . . . 33
3.14 Imperfection amplitude calculation (case 2). . . . . . . . . . . . . . . . 34
3.15 Initial deflections of the surface. . . . . . . . . . . . . . . . . . . . . . . 35

v
3.16 Tensile coupon . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
3.17 Tensile coupons for cold-formed program. . . . . . . . . . . . . . . . . 37
3.18 Test set-up for tensile test in 250 kN test bench. . . . . . . . . . . . . . 37
3.19 Tensile test results: stress-strain relation. . . . . . . . . . . . . . . . . . 38
3.20 Tensile test results: stiffness-strain relation. . . . . . . . . . . . . . . . 39
3.21 Compression test set-up for specimens . . . . . . . . . . . . . . . . . . 40
3.22 Gap between the specimen and support. . . . . . . . . . . . . . . . . . 41
3.23 CUFSM result for the selected Z-shaped profile. . . . . . . . . . . . . 41
3.24 Experimental result for selected Z-shaped profile. . . . . . . . . . . . 42
3.25 Compression test results ”Extruded” subprogram. . . . . . . . . . . . 43
3.26 Compression test results ”Cold-formed” subprogram. . . . . . . . . . 44
3.27 Compression test results: gradually increased complexity of the cross-
sectional geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
3.28 Comparison of local and distortional buckling tests for the same profile. 46
3.29 Load-displacement curves for the batch of specimens . . . . . . . . . 47
3.30 Compression test results: material influence. . . . . . . . . . . . . . . 47

4.1 Material stress-strain curves. . . . . . . . . . . . . . . . . . . . . . . . . 51


4.2 Mesh applied for the FE-model. . . . . . . . . . . . . . . . . . . . . . . 52
4.3 Material models for applied springs. . . . . . . . . . . . . . . . . . . . 53
4.4 Gap during the test and modelled gap. . . . . . . . . . . . . . . . . . . 54
4.5 FEM results ”Extruded” subprogram. . . . . . . . . . . . . . . . . . . 55
4.6 Comparison of FEM results with and without gap modelling (3C80E7) 55
4.7 Load-displacement and load-deflection curves (specimen 3C80E7). . 56
4.8 Comparison of tangential stiffness of the experiment (LVDT’s and
strain gauges), tensile test and numerical analysis (3C80E7). . . . . . 57
4.9 FEM results ”Cold-formed” subprogram. . . . . . . . . . . . . . . . . 58
4.10 Comparison of test and FEM results for the specimen 2Z100F11A. . . 59
4.11 Mesh density. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
4.12 Load-displacement diagram varying mesh density. . . . . . . . . . . 62
4.13 Tensile test results for two coupons in rolling and perpendicular to
rolling directions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
4.14 Load-displacement diagram varying material characteristics. . . . . . 64
4.15 Comparison FE-model and test results, applying enhanced material
properties for the rounded corners. . . . . . . . . . . . . . . . . . . . . 65
4.16 Euler buckling analysis results for specimen 3C80E7. . . . . . . . . . 66

5.1 Z-shaped specimen dimensions. . . . . . . . . . . . . . . . . . . . . . 70


5.2 Buckling behaviour of cross-sectional plates and critical limits. . . . . 71
5.3 Cross-sectional deformation at different load limits. . . . . . . . . . . 71
5.4 Initial and secondary buckling stress determination for local buckling. 73
5.5 Proposed inelastic buckling approximation. . . . . . . . . . . . . . . . 77
5.6 Illustration of the local and distortional buckling behaviour of the
cross-section, according to the model. . . . . . . . . . . . . . . . . . . 80

vi
5.7 Considered cross-sectional shapes. . . . . . . . . . . . . . . . . . . . . 83

6.1 Cross-section selected for parameter study. . . . . . . . . . . . . . . . 88


6.2 CUFSM results for selected C-profile with defined buckling shapes
for local and distortional modes. . . . . . . . . . . . . . . . . . . . . . 89
6.3 CUFSM results for selected C-profile with defined buckling curves
and critical points for local and distortional modes. . . . . . . . . . . 89
6.4 Applied springs for C-shaped profile. . . . . . . . . . . . . . . . . . . 90
6.5 CUFSM result for C-shaped profile with additional springs. . . . . . 90
6.6 Spring application at two points on a critical length distance from
both edges. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
6.7 CUFSM results for specimen 5(6)C5. . . . . . . . . . . . . . . . . . . . 95
6.8 Initial deformed shape according to Euler analysis (mode 1). . . . . . 95
6.9 FEM deformed shapes for specimen 5(6)C5 according to non-linear
analysis. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
6.10 Load-displacement (left) and load-deflection (right) plots for speci-
men 5(6)C5. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
6.11 FE-results for plate elements of specimen 5(6)C5. . . . . . . . . . . . . 97
6.12 Tangential stiffness for plate elements of specimen 5(6)C5. . . . . . . 97
6.13 CUFSM results for specimen 2.5(3)C5. . . . . . . . . . . . . . . . . . . 98
6.14 FEM deformed shapes for specimen 2.5(3)C5 according to non-linear
analysis. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
6.15 Load-displacement (left) and load-deflection (right) plots for speci-
men 2.5(3)C5. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
6.16 FE-results for plate elements of specimen 2.5(3)C5. . . . . . . . . . . . 100
6.17 Tangential stiffness for plate elements of specimen 2.5(3)C5. . . . . . 100
6.18 CUFSM results for specimen 1(2)C10. . . . . . . . . . . . . . . . . . . 101
6.19 FEM deformed shapes for specimen 1(2)C10 according to non-linear
analysis. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
6.20 Load-displacement (left) and load-deflection (right) plots for speci-
men 1(2)C10. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
6.21 FE-results for plate elements of specimen 1(2)C10. . . . . . . . . . . . 103
6.22 Tangential stiffness for plate elements of specimen 1(2)C10. . . . . . . 103
6.23 FEM deformed shapes for specimen 1(2)C5 according to non-linear
analysis. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
6.24 Load-displacement (left) and load-deflection (right) plots for speci-
men 1(2)C5. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
6.25 FE-results for plate elements of specimen 1(2)C5. . . . . . . . . . . . . 105
6.26 Tangential stiffness for plate elements of specimen 1(2)C5. . . . . . . 106
6.27 Initial buckling (outstanding plates group I) and secondary buckling
(internal plates group II). . . . . . . . . . . . . . . . . . . . . . . . . . . 107
6.28 Initial buckling (outstanding plates group I) and secondary buckling
(internal plates group II). . . . . . . . . . . . . . . . . . . . . . . . . . . 108
6.29 Post-buckling behaviour. Distortional-local interaction. . . . . . . . . 108

vii
6.30 Local-distortional interaction. . . . . . . . . . . . . . . . . . . . . . . . 109
6.31 Investigation of the scatter in distortional buckling behaviour of the
outstanding plate elements. . . . . . . . . . . . . . . . . . . . . . . . . 110
6.32 Investigation of the scatter in distortional buckling behaviour of the
internal plate elements. . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
6.33 Comparison of the inelastic distortional buckling (6082–T6 alloy) with
the predicted elastic model at εp (squares) and inelastic model σcr;T
(dots). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
6.34 Comparison of the inelastic distortional buckling (6060–T66 alloy)
with the predicted elastic model at εp (squares) and inelastic model
σcr;T (dots). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
6.35 Comparison of the inelastic distortional buckling (5083–H111 alloy)
with the predicted elastic model at εp (squares) and inelastic model
σcr;T (dots). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117

C.1 Buckling curves for three aluminium alloys. . . . . . . . . . . . . . . . 138


C.2 Piecewise idealization of the curves. . . . . . . . . . . . . . . . . . . . 140

D.1 Compressed specimens 6082 aluminium alloy. . . . . . . . . . . . . . 145


D.2 Compressed specimens 6082 aluminium alloy. . . . . . . . . . . . . . 145
D.3 Compressed specimens 6060 aluminium alloy. . . . . . . . . . . . . . 146
D.4 Compressed specimens 6060 aluminium alloy. . . . . . . . . . . . . . 146
D.5 Compressed specimens 5083 aluminium alloy. . . . . . . . . . . . . . 147
D.6 Compressed specimens 5083 aluminium alloy. . . . . . . . . . . . . . 147

viii
List of Tables

2.1 Values of kcr for determining the critical buckling stress . . . . . . . . 10

3.1 Specification for ”Extruded” subprogram. . . . . . . . . . . . . . . . . 27


3.2 Specification for ”Cold-formed” subprogram 6082-T6 aluminium alloy 27
3.3 Specification for ”Cold-formed” subprogram 5083-H111 aluminium
alloy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.4 Compression test results for ”Extruded” subprogram. . . . . . . . . . 43
3.5 Compression test results for ”Cold-formed” subprogram. . . . . . . . 44
3.6 Repeatability of tests. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

4.1 Comparison of experimental and FE-results for extruded profiles. . . 58


4.2 Comparison of experimental and FE-results for cold-formed profiles. 60
4.3 Sensitivity analysis results for the specimen 3C80E7. . . . . . . . . . . 66

5.1 Material characteristics used for parameter study . . . . . . . . . . . 76


5.2 Comparison of the FE-results with Mennink’s prediction model re-
sults. ”Extruded” profiles, 6060-T66 aluminium alloy. . . . . . . . . . 83
5.3 Comparison of the FE-results with Mennink’s prediction model re-
sults. ”Cold-formed” profiles, 6082-T6 aluminium alloy. . . . . . . . . 83
5.4 Comparison of the FE-results with Mennink’s prediction model re-
sults. ”Cold-formed profiles”, 5083-H111 aluminium alloy. . . . . . . 83

6.1 Specimen specification for parameter study. . . . . . . . . . . . . . . . 88


6.2 Parameter study C–shaped specimens specification. . . . . . . . . . . 94
6.3 CUFSM results for C–shaped specimens of the parameter study (6082-
T6 aluminium alloy). . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
6.4 Equations for the average post-buckling stiffness of outstanding com-
ponents. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
6.5 Equations for the average post-buckling stiffness of internal compo-
nents. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
6.6 FEM results for C-profiles selected for the parameter study (6082-T6). 115

ix
6.7 FEM results for C-specimens (6060-T66). . . . . . . . . . . . . . . . . . 116
6.8 FEM results for C-shaped specimens (5083-H111). . . . . . . . . . . . 117

A.1 Measured dimensions and characteristics, ”Extruded” subprogram. . 129


A.2 Measured dimensions and characteristics, ”Cold-formed” subprogram 130
A.3 Measured imperfections, ”Extruded” subprogram. . . . . . . . . . . . 131
A.4 Measured imperfections, ”Cold-formed” subprogram . . . . . . . . . 132
A.5 Material characteristics for aluminium alloy EN AW-6060 T66. . . . . 133
A.6 Material characteristics for aluminium alloy EN AW-6060 T66. . . . . 133
A.7 Material characteristics for aluminium alloy EN AW-6060 T66. . . . . 133
A.8 Material characteristics for aluminium alloy EN AW-6082 T6. . . . . . 134
A.9 Material characteristics for aluminium alloy EN AW-6082 T6. . . . . . 134
A.10 Material characteristics for aluminium alloy EN AW-5083 H111. . . . 134
A.11 Material characteristics for aluminium alloy EN AW-5083 H111. . . . 134

B.1 Dimensions used in FE simulations for extruded specimens. . . . . . 135


B.2 Dimensions used in FE simulations (in mm) for cold-formed specimens.135
B.3 Imperfections used in FE simulations (in mm) for extruded specimens. 136
B.4 Imperfections used in FE simulations (in mm) for cold-formed speci-
mens. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136

D.1 Comparison of the FSM and Eigenvalue analysis results for C-shaped
specimens of the parameter study. . . . . . . . . . . . . . . . . . . . . 143
D.2 FEM results for C-shaped specimens of the parameter study (6082-T6). 143
D.3 FEM results for C-shaped specimens of the parameter study (6060-T66).144
D.4 FEM results for C-shaped specimens of the parameter study (5083-
H111). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144

E.1 Prediction model results for C-profiles of the parameter study (6082-T6).149
E.2 Prediction model results for C-profiles of the parameter study (6060-
T66). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
E.3 Prediction model results for C-profiles of the parameter study (5083-
H111). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150

x
List of symbols

Ncr;E Euler load of a column, [N] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7


E Modulus of elasticity / Young’s modulus, [N/mm2 ] . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
I Second moment of inertia, [mm4 ] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
L Axial length of the specimen, [mm] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
t Plate thickness, [mm] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
σcr Elastic critical stress, [N/mm2 ] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
D Plate flexural rigidity, [N/mm2 ] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
kcr Plate buckling coefficient, [-] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
ϕ Plate slenderness parameter, [-] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
b Plate width, [mm] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
fy Yield stress, [N/mm2 ] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
bef f Effective plate width, [mm] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
σmax Maximum axial stress at the unloaded plate edges, [N/mm2 ] . . . . . . . . . . . . . . . . . . 11
C Effective width parameter, [-] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
f0.1 0.1% proof strength of the material characteristic, [N/mm2 ] . . . . . . . . . . . . . . . . . . . 13
f0.2 0.2% proof strength of the material characteristic, [N/mm2 ] . . . . . . . . . . . . . . . . . . . 12
fu Ultimate stress of the material characteristic, [N/mm2 ] . . . . . . . . . . . . . . . . . . . . . . . 13
n Ramberg-Osgood strain-hardening parameter, [-] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
σ Axial stress, [N/mm2 ] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
ε Axial strain, [-] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
ν Poisson’s ratio, [-] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
λ Slenderness parameter, [-] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
N Axial resistance, [N] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
A Cross-sectional area [mm2 ] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
N0.2 Axial squash load, [N] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
Ncr Critical axial load at initial buckling, [N] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
Nu Ultimate or failure load, [N] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
u Axial displacement of the specimen, [mm] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
w Out-of-plane deflection, [mm] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
σpl Plastic stress, [N/mm2 ] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
εpl Plastic strain, [-] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

xi
εel Elastic strain, [-] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .51
σ∗ Non-dimensionalised axial stress at initial buckling , [-] . . . . . . . . . . . . . . . . . . . . . . . 70
ε∗ Non-dimensionalised axial strain with respect to initial buckling , [-] . . . . . . . . . . 70
fp Proportional limit of the material, [kN] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
Lcr Critical buckling length, [mm] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
εp Elastic strain according to the proportional limit of the material, [-] . . . . . . . . . . . . 72
σcr;2 Elastic critical stress at secondary buckling, [N/mm2 ] . . . . . . . . . . . . . . . . . . . . . . . . . 75
εcr;2 Elastic critical strain at secondary buckling, [-] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
Nu;pm Predicted ultimate load, [kN] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
σcr;T Inelastic critical axial stress at initial buckling, [N/mm2 ] . . . . . . . . . . . . . . . . . . . . . . 76
χT Inelastic buckling coefficient, [-] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
λ̄ Relative slenderness, [-] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137

xii
Contents

Acknowledgements i

Summary iii

List of figures v

List of tables ix

List of symbols xi

Table of contents xiii

1 Introduction 1
1.1 Complex cross-sectional shapes . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Problem statement and research goals . . . . . . . . . . . . . . . . . . 3
1.3 Research approach and thesis outline . . . . . . . . . . . . . . . . . . . 3

2 State of the art: cross-sectional instability 7


2.1 Column Buckling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.2 Buckling Theory of Plates . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.2.1 Elastic critical buckling . . . . . . . . . . . . . . . . . . . . . . . 7
2.2.2 Postbuckling Strength . . . . . . . . . . . . . . . . . . . . . . . 10
2.2.3 Effective Width Approach . . . . . . . . . . . . . . . . . . . . . 10
2.2.4 Effective Thickness Approach . . . . . . . . . . . . . . . . . . . 12
2.3 Stability: influence of material and imperfections . . . . . . . . . . . . 12
2.3.1 Material characteristics . . . . . . . . . . . . . . . . . . . . . . . 12
2.3.2 Geometrical Imperfections . . . . . . . . . . . . . . . . . . . . 14
2.4 Local Buckling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.4.1 Slenderness parameter . . . . . . . . . . . . . . . . . . . . . . . 15
2.4.2 Classification of cross-sections . . . . . . . . . . . . . . . . . . 16
2.5 Distortional buckling . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.6 Numerical Investigations . . . . . . . . . . . . . . . . . . . . . . . . . . 19

xiii
2.6.1 Finite Element Method . . . . . . . . . . . . . . . . . . . . . . . 19
2.6.2 Finite Strip Method . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.6.3 Generalized Beam Theory . . . . . . . . . . . . . . . . . . . . . 21
2.7 Direct Strength Method . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.8 Mennink’s prediction model . . . . . . . . . . . . . . . . . . . . . . . . 23
2.9 Literature Study Evaluation . . . . . . . . . . . . . . . . . . . . . . . . 24

3 Experimental Investigation 25
3.1 Program Description . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
3.2 Measuring Dimensions . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.3 Measuring Imperfections . . . . . . . . . . . . . . . . . . . . . . . . . . 29
3.3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
3.3.2 Measuring concept . . . . . . . . . . . . . . . . . . . . . . . . . 30
3.3.3 Test set-up . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
3.3.4 Test results processing . . . . . . . . . . . . . . . . . . . . . . . 32
3.3.5 Imperfections results . . . . . . . . . . . . . . . . . . . . . . . . 34
3.4 Material characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
3.4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
3.4.2 Tensile coupon . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
3.4.3 Test set-up . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
3.4.4 Tensile test results . . . . . . . . . . . . . . . . . . . . . . . . . 38
3.4.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
3.5 Test set-up for compression test . . . . . . . . . . . . . . . . . . . . . . 40
3.6 Compression test results . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.6.1 Agreement with FSM-prediction . . . . . . . . . . . . . . . . . 41
3.6.2 Extruded experimental subprogram . . . . . . . . . . . . . . . 42
3.6.3 Cold-formed experimental subprogram . . . . . . . . . . . . . 44
3.6.4 Results of the geometry influence . . . . . . . . . . . . . . . . 44
3.6.5 Local vs distortional buckling . . . . . . . . . . . . . . . . . . . 46
3.6.6 Test repeatability . . . . . . . . . . . . . . . . . . . . . . . . . . 46
3.6.7 Results of the material study . . . . . . . . . . . . . . . . . . . 47
3.7 Conclusions and summary of observations . . . . . . . . . . . . . . . 48

4 Finite Element Model Validation 49


4.1 Scope of the chapter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
4.2 Numerical program . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
4.2.1 Applied imperfections . . . . . . . . . . . . . . . . . . . . . . . 50
4.2.2 Material input . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
4.3 FE-model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
4.3.1 Element types . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
4.3.2 Mesh . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.3.3 Supports, loading and modelled gap . . . . . . . . . . . . . . . 52
4.3.4 Analyses types . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
4.4 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

xiv
4.4.1 Extruded subprogram . . . . . . . . . . . . . . . . . . . . . . . 54
4.4.2 Cold-formed subprogram . . . . . . . . . . . . . . . . . . . . . 58
4.5 Influence of the modelling parameters . . . . . . . . . . . . . . . . . . 60
4.5.1 General . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
4.5.2 Mesh density . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
4.5.3 Material anisotropy . . . . . . . . . . . . . . . . . . . . . . . . . 62
4.5.4 Material model with enhanced corner properties . . . . . . . . 64
4.5.5 Imperfection sensitivity . . . . . . . . . . . . . . . . . . . . . . 65
4.6 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

5 Assessment of the existing model with respect to distortional buckling 69


5.1 Scope of the chapter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
5.2 Specimen example for model evaluation . . . . . . . . . . . . . . . . . 69
5.3 Mennink’s model for local buckling . . . . . . . . . . . . . . . . . . . 72
5.3.1 General . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
5.3.2 Initial and secondary buckling determination . . . . . . . . . 72
5.3.3 Elastic local buckling prediction: (σcr < fp ) . . . . . . . . . . . 74
5.3.4 Inelastic buckling prediction: (fp ≤ σcr ) . . . . . . . . . . . . . 76
5.3.5 Inelastic buckling coefficient . . . . . . . . . . . . . . . . . . . 76
5.4 Model extension for distortional buckling . . . . . . . . . . . . . . . . 79
5.4.1 Description of the behaviour . . . . . . . . . . . . . . . . . . . 79
5.4.2 Post-buckling resistance . . . . . . . . . . . . . . . . . . . . . . 80
5.4.3 Model proposals . . . . . . . . . . . . . . . . . . . . . . . . . . 81
5.5 Application of the extended prediction model . . . . . . . . . . . . . 81
5.6 Summary and conclusions . . . . . . . . . . . . . . . . . . . . . . . . . 85

6 Prediction model for distortional buckling 87


6.1 Scope of the chapter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
6.2 Selection of the cross-sectional shape for distortional buckling . . . . 88
6.3 FEM set-up for the parameter study . . . . . . . . . . . . . . . . . . . 91
6.4 Parameter study definition . . . . . . . . . . . . . . . . . . . . . . . . . 93
6.5 Parameter study results: elastic buckling . . . . . . . . . . . . . . . . . 94
6.5.1 Pure distortional buckling . . . . . . . . . . . . . . . . . . . . . 94
6.5.2 Secondary buckling and distortional-local interaction . . . . . 98
6.5.3 Secondary buckling and local-distortional interaction . . . . . 101
6.5.4 Elastic buckling of the reference C-shape . . . . . . . . . . . . 104
6.5.5 Evaluation of elastic buckling results . . . . . . . . . . . . . . . 106
6.6 Development of a prediction model for distortional buckling of C-
shaped specimens . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
6.6.1 Initial buckling . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
6.6.2 Secondary buckling . . . . . . . . . . . . . . . . . . . . . . . . 111
6.7 Prediction model for distortional buckling behaviour of C-shaped alu-
minium structural elements . . . . . . . . . . . . . . . . . . . . . . . . 112
6.8 Inelastic distortional buckling . . . . . . . . . . . . . . . . . . . . . . . 114

xv
6.9 Chapter conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118

7 Conclusions and recommendations 119


7.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
7.1.1 Distortional buckling in the current design rules . . . . . . . . 119
7.1.2 Experimental research . . . . . . . . . . . . . . . . . . . . . . . 119
7.1.3 Numerical research . . . . . . . . . . . . . . . . . . . . . . . . . 120
7.1.4 Prediction model for distortional buckling . . . . . . . . . . . 120
7.2 Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122

Bibliography 123

Appendixes 129

A Experimental results 129


A.1 Measured dimensions . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
A.2 Measured imperfections . . . . . . . . . . . . . . . . . . . . . . . . . . 131
A.3 Measured material characteristics . . . . . . . . . . . . . . . . . . . . . 133

B Numerical analysis 135


B.1 Applied dimensions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
B.2 Applied imperfections . . . . . . . . . . . . . . . . . . . . . . . . . . . 136

C Inelasticity coefficient derivation 137


C.1 Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
C.2 Buckling curves for three aluminium alloys . . . . . . . . . . . . . . . 138
C.3 Mennink’s approximation . . . . . . . . . . . . . . . . . . . . . . . . . 140

D Parameter Study Results 143

E Prediction model results 149

Samenvatting 151

Curriculum Vitae 153

xvi
Chapter 1
Introduction

Abstract

This chapter outlines the major aspects of the thesis. A general survey of the research
background is given and the objectives and the structure of the thesis are introduced.
Cross-sectional instability is one of the important design issues for aluminium struc-
tural members with complex cross-sectional shapes. Therefore, the current investigation
is focused on cross-sectional instability regarding various cross-sectional shapes and re-
sulting in the development of a prediction model.

1.1 Complex cross-sectional shapes

T he last decades structural applications of aluminium have experienced a fast


growth due to the outstanding properties of aluminium, namely high strength
to weight ratios, good corrosion resistance and ease of maintenance Mazzolani [2002].
Aluminium’s flexibility is an important quality for structural design, as it allows one
to meet the requirements of different industrial situations (see Figure 1.1).

Figure 1.1: Application of aluminium structural elements with complex cross-sectional


shapes in greenhouses.

Extrusion allows one to optimize the cross-section of structural elements accord-


ing to the design requirements with relative ease. Extrusion is a technological pro-
cess used to form sections with different shapes by pushing hot metal through an

1
opening called a die. The resulted cross-section corresponds to the shape of such
a die. So, even when the cross-sectional shape is complex, the desired aluminium
structural element can be simply manufactured by preparing the die accordingly.
Aluminium characteristics enable sections of incredible complexity to be produced
for a certain application.
One of the important aspects of the design process is to satisfy the practical needs
while optimally utilizing the material. This often results in thin-walled sections. En-
gineers are searching for new structural forms of aluminium extrusions to enhance
practical efficiency. Further optimizing is possible if some additional functions are
added to the cross-section like stiffeners or weld backings, all of which leads to the
application of complex cross-sectional shapes (see Figure 1.1). Thus, improvements
in functionality correspond to increasing the complexity of the cross-section.

Figure 1.2: Buckling modes, instability types.

It is well-known that the modulus of elasticity of aluminium is one third of


steel which makes aluminium more susceptible to instability. Two types of insta-
bility are usually considered: overall instability and cross-sectional instability (see
Figure 1.2). The stability problem of the entire structure (translation and rotation)
is called overall instability, e.g. flexural, torsional or flexural-torsional buckling.
Cross-sectional instability concerns stability of the cross-section, like local buckling
and distortional buckling. In classification of cross-sectional instability, local buck-
ling implies changes in geometry with only rotation occurring at the fold lines of
the section. For distortional buckling rotation and also translation take place at the
fold lines of the section with changes in the cross-sectional geometry, Yu and Schafer
[2006]. Previous research has shown that cross-sectional instability has a consider-
able influence on the stability of the whole structural element (Mennink [2002]).
When using aluminium elements such as beams and columns, the design re-
sistance should be determined. Local instability effects often determine the struc-
tural resistance because optimization of the material (weight) results in the applica-
tion of thin-walled sections with complex cross-sectional shapes. The complexity of
thin-walled cross-sectional shapes makes the prediction or determination of cross-
sectional instability one of the decisive design issues for aluminium extrusions.

2
1.2 Problem statement and research goals
For building structures, analytical design rules are used to calculate the structural
resistance. As mentioned before, cross-sectional instability influences the stability
of the whole structure. There are several codes on aluminium structures which
deal with the aspect of local buckling. The commonly used approach in Europe
is given in Eurocode 9 CEN [2007], described by Mazzolani [1985]. The cross-
section is usually seen as a number of plates connected by nodes. Design standards
consider the cross-sectional instability as the buckling of individual plate elements
which compose the cross-section and do not consider the interaction of the cross-
sectional plates. Unfortunately, such an approach is only sufficient for traditional
cross-sectional shapes, like rectangular or so-called I-sections.
Hence, the current design rules used by engineers are limited with respect to
cross-sectional instability: only a limited range of cross-sections is covered and only
with a limited accuracy. Extensive research into cross-sectional instability of alu-
minium structural elements concerning variety of shapes is required.
Due to the lack of knowledge in designing aluminium elements, the following
research goals are stated:

• To study cross-sectional instability effects (local and distortional buckling) for


aluminium structural elements with various cross-sectional shapes;

• To extend the current analytical formulas for calculating the ultimate load of
aluminium sections regarding local buckling to a practical design recommen-
dation predicting the local and, moreover, distortional buckling resistance of
aluminium profiles.

Thus, the main objective of the current research is to result in an improved pre-
diction model, which is not only applicable to traditional cross-sectional shapes (as
covered by the design rules of the current codes) but could be also applied to more
complex cross-sectional shapes. Development of the prediction model is based on
the existing model of Mennink [2002], see section 2.8. In the current research Men-
nink’s model is extended by taking distortional buckling mode into account.

1.3 Research approach and thesis outline


In general, the objectives of the current research can be achieved by using experi-
mental data and numerical methods, such as the finite strip method (FSM) and the
finite element method (FEM).
A scheme of the research approach is shown in Figure 1.3. The contents of the
thesis correspond to the presented research approach. First two chapters with in-
troduction and literature survey are given. These are followed by the description
and results of the research steps indicated in the presented layout. Brief commen-
taries for the scheme are given below; respective chapter numbers are included in
the figure.

3
Figure 1.3: Research approach scheme.

• Based on the results of finite strip analyses, various cross-sectional shapes are
selected for an experimental program. The results of the finite strip analyses
are also compared with those of the Euler buckling analyses according to the
finite element program DIANA (Hendriks and Wolters [2007]). The outcome
of the selection procedure according to the finite strip analyses results is pre-
sented in Chapter 3.

• The experimental investigation includes 85 compression tests for aluminium


columns with L-shaped, Z-shaped and C-shaped sections by varying the di-
mensions of the cross-section and also the length of the specimen. The exper-
iments are expected to show a wide range of instability modes, such as local
buckling, distortional buckling, torsional and flexural buckling. Complete de-
scription of the experimental program is given in Chapter 3.

• The aim of the experiments is mainly to use the data as input for the finite ele-
ment (FE) model and also for validation of the FE-model by comparing the re-
sults of the finite element (FE) calculations with the test data. Therefore, at the
beginning, the FE-model is built by varying, for example, loading conditions,
mesh fineness and load step size. Then, the model is adjusted by comparison
with the test results. When a suitable accuracy and validity of the numeri-
cal model is achieved, the FE-model is assumed to be valid to investigate the
buckling resistance of various cross-sections (see Chapter 4).

• The purpose of Chapter 5 is to validate Mennink’s prediction model regard-


ing cross-sectional shapes included in the current investigation. Description
of Mennink’s model for local buckling prediction is introduced and possible
extension of the model for the distortional buckling phenomenon is discussed.

4
• A parameter study is executed in Chapter 6, using the validated FE-model.
The purpose of the parameter study is to estimate the influence of various pa-
rameters and properties (e.g. cross-sectional geometry parameters, mechani-
cal properties) on distortional buckling resistance. Hence, the ultimate buck-
ling resistance of C-shaped sections with various dimensions and geometry is
determined. Based on results of the parameter study and Mennink’s model
validation, an improved prediction model for distortional buckling behaviour
of aluminium C-shaped members is developed.

The thesis finalizes with a last chapter (Chapter 7), which contains conclusions
and recommendations for further research.

5
Chapter 2
State of the art: cross-sectional instability

Abstract
In this chapter the stability theory for columns and plates is briefly described based
on previous research of Timoshenko and Gere [1961], Walker [1975], Galambos [1998]
and Yu [2000]. The concept of effective width is one of the common design approaches.
The description of aluminium properties and characteristics is given. The definition of
buckling modes is introduced in the context of aluminium plate buckling. Thus, the
current chapter provides an overview of the existing knowledge on the cross-sectional
instability problem of aluminium structural members.

2.1 Column Buckling

T he buckling theory of a column loaded by a compressive force is based on the


work of Euler that is described by many authors, e.g. Timoshenko and Gere
[1961]. An understanding of the strength of centrally loaded columns is essential
to the development of design criteria for compression members in general. Consid-
ering the Euler column as a mathematically straight, pin-ended, perfectly centrally
loaded column, the buckling load or critical load or bifurcation load is equal to:
π 2 EI
Ncr;E = (2.1)
L2
where EI is the elastic stiffness and L is the axial length. The Euler load Ncr;E is the
reference value to which the strength of actual columns is usually compared. If end
conditions differ from perfectly frictionless pins, the critical load is expressed by:
π 2 EI
Ncr;EK = (2.2)
(KL)2
where KL is an effective length defining the portion of the deflected shape between
points of zero bending moment. In other words, KL is the length of an equivalent
pin-ended column buckling at the same load as the end-restrained column. The
general approach for column buckling is to determine the failure or design load as
a function of the critical and squash loads.

2.2 Buckling Theory of Plates


2.2.1 Elastic critical buckling
This section regards buckling of rectangular thin plates subjected to uniform com-
pression. Figure 2.1 shows a possible buckling shape in the case of uniform com-

7
pression. An important contribution to the plate-buckling theory has been made
by von Kármán et al. [1932] and Winter [1947]. The plate-buckling phenomenon is
summarized in many books, see for example: Timoshenko and Gere [1961], Walker
[1975], Kirby and Nethercot [1979], Murray [1984], Narayanan [1987], Rhodes and
Harvey [1971], Galambos [1998], and Yu [2000]. Earlier investigations established
that plates exhibit a stable equilibrium with a substantial amount of post-buckling
strength which is not the case for a uniformly compressed column.

Figure 2.1: Rectangular plate loaded by uniform pressure.

The theoretical critical load for a plate is not necessarily a satisfactory basis for
design, since the ultimate strength can be much greater than the critical buckling
load. For example, a plate loaded in uniaxial compression, with both longitudinal
edges supported, provides postbuckling support which means the plate will un-
dergo stress redistribution as well as develop transverse tensile membrane stresses
after buckling. Initial imperfections in such a plate may cause bending to begin
below the buckling load, yet unlike an initially imperfect column, the plate may
sustain loads greater than the theoretical buckling load.
The rectangular plate of Figure 2.1 allows various boundary and loading condi-
tions. The buckling problem of uniformly compressed plate with no lateral loading
was also formulated by Saint-Venant [1883]. This problem was solved first by Bryan
[1890]. All four plate-edges are supported, i.e. in-plane translations are restricted
whereas rotations are allowed, while one pair of opposite edges is subjected to a
uniform compressive stress σ . The elastic critical stress has been defined as σcr :
· µ ¶ ¸2
Dπ 2 b n2 ³ a ´
σcr = m + (2.3)
tb2 a m b

where
Et3
D= (2.4)
12 (1 − ν 2 )
The critical value of σcr , i.e. the smallest value, can be obtained by taking n equal
to 1. The physical meaning of this is that a plate buckles in such a way that there
can be several half-waves in the direction of compression but only one half-wave

8
in the perpendicular direction. m is the number of half-waves into which the plate
buckles.
For the case of a simply supported rectangular plate under uniform compression
the buckling coefficient kcr equals:

m2 ϕ2
kcr = + +2 (2.5)
ϕ2 m2
By introducing the plate-slenderness parameter (ϕ = a/b) in equation 2.5, equa-
tion 2.3 gives the critical stress as:

π 2 Et2
σcr = kcr (2.6)
12(1 − ν 2 )b2

Equation 2.6 is similar to Equation C.2 except that the plate width b is squared
rather than column length L and the Poisson’s ration ν is included in the flexural
rigidity D for plates. The plate buckling coefficient kcr allows for different boundary
conditions.

Figure 2.2: Buckling coefficient kcr of a simply supported plate.

Figure 2.2 presents kcr , for a simply supported plate, as a function of the plate
slenderness ϕ. In structural engineering long plates have a relatively large a/b ratio.
Such plates can be used to describe the behaviour of plates in cross-sections, for
example the webs of square hollow sections. In those cases, kcr reduces to the well-
known value of 4.0.
Various boundary conditions have been studied by different authors. Figure 2.3
by Gerard and Becker [1957] summarizes five common cases of supports, where
c stands for clamped, ss for simply supported, and free for an unsupported edge.
Values of the buckling coefficient for different boundary conditions are given in Fig-
ure 2.1. The buckling coefficient, kcr , for a long rectangular plate simply supported
along three sides, with one unloaded free edge equals: kcr = 0.425. However, when
the restraining effect of the web is considered, kcr may be taken as 0.5 for the de-
sign of the unstiffened compression flange. The supporting effect of edges, which
defines kcr , is therefore highly important.

9
Figure 2.3: Buckling coefficients kcr of Table 2.1: Values of kcr for determining
uniformly compressed rectangular plates the critical buckling stress.
with various boundary conditions.

2.2.2 Postbuckling Strength


Unlike one-dimensional structural members such as columns, stiffened compres-
sion elements will not collapse when the buckling strength is reached. In the plate,
the stress distribution is uniform prior to its buckling, as is shown in Figure 2.4(a).
After buckling, a portion of the prebuckling load of the center transfers to the edge
zone of the plate. As a result, a non-uniform stress distribution is developed, as
shown in Figure 2.4(b). The redistribution of stress continues until the stress at the
edge reaches the yield point fy and the plate begins to fail (Figure 2.4(c)). The post
buckling behaviour of a plate can be analyzed by using large deflection theory.

Figure 2.4: Stages of stress distribution in simply supported compressed plates, according
to Yu [2000].

2.2.3 Effective Width Approach


The concept of effective width was introduced by von Kármán et al. in 1932. In
this approach, instead of considering the non-uniform distribution over the entire
width of the plate b, it is assumed that the total load is carried by a fictitious effective

10
width bef f , subject to a uniformly distributed stress equal to the edge stress σmax ,
as shown in Figure 2.5.

Figure 2.5: Definition of effective width.

The width b is selected so that the area under the curve of the actual non-uniform
stress distribution is equal to the sum of the two parts of the equivalent rectangular
shaded area with a total width bef f and an intensity of stress equal to the edge stress
σmax , that is:
Zb
σ dx = beff σmax (2.7)
0

It may also be considered that the effective width represents a particular width of
the plate that just buckles when the compressive stress reaches the yield point of
material. This results in the Von Kármán formula:
s
E
beff = Ct (2.8)
fy

where C equals for µ = 0.3:


π
C=p = 1.9 (2.9)
3(1 − µ2 )

Based on his extensive investigation on light-gauge cold-formed steel sections, Win-


ter [1947] indicated that the equation is equally applicable to the element in which
the maximum stress σmax is below the yield point fy . This results in the Winter
formula: r
E
beff = Ct (2.10)
σmax

11
In addition, results of tests previously conducted by Sechler [1968] and Winter led
to an improvement of the term C:
" µ ¶r #
t E
C = 1.9 1 − 0.475 (2.11)
b σmax

which led consequently to the modified Winter formula Winter [1947] and Winter
[1948] for plates simply supported along both longitudinal edges:
r " µ ¶r #
E t E
beff = 1.9t 1 − 0.475 (2.12)
σmax b σmax

It should be noted that the effective width not only depends on the edge stress but
also on the b/t ratio.

2.2.4 Effective Thickness Approach


Design rules on local buckling for steel are based on the effective width approach.
In some recently adopted national standards the effective thickness is used since
it leads to simple calculations. In Eurocode 9 CEN [2007], the effective thickness
is used for calculation of resistance. However, in many cases an effective width
concept is used for calculation of stiffness (see Hoglund [1972]). A description of
the effective thickness approach can be found in section 2.4.

2.3 Stability: influence of material and imperfections


2.3.1 Material characteristics
The structural analysis for aluminium is based on a material with a generalized in-
elastic behaviour. Several authors have formulated various models for inelastic be-
haviour of material, which are described in Kutanova and Soetens [2006]. The most
common way is to separate the stress-strain relationship diagram into three regions.
The first region corresponds to elastic behaviour, the second region to inelastic be-
haviour and the third region identifies strain-hardening behaviour. With respect to
these three regions, models proposed by Baehre [1966] and Mazzolani [1972] define
the limits and non-dimensional relations for each region. The stress-strain curve
of aluminium alloys is often described as a ”round-house” type DeMartino et al.
[1990], which can be modeled accurately using the Ramberg-Osgood expression.
A generalized stress (σ) - strain (ε) law ε = f (σ) has been proposed by Ramberg
and Osgood [1943] for aluminium alloys:
µ ¶n
σ σ
ε= + 0.002 , (2.13)
E f0.2

where E = Young’s modulus at the origin, f0.2 = 0.2% proof stress, ε = variable
strain, σ = variable stress.

12
The exponent n of the Ramberg-Osgood law is a characteristic of the strain-
hardening rate of the inelastic portion of the σ − ε diagram and can be expressed
as:
ln 2
n= ³ ´ (2.14)
ln ff0.2
0.1

where f0.1 = 0.1% proof stress , which has to be determined by experiment.

Figure 2.6: Stress-strain diagram based on the Ramberg-Osgood law.

The higher the value of the exponent, n, is, the sharper the knee of the stress-
strain curve, and the slower the slope of the strain-hardening portion of the curve.
The exponent n is a function of f0.2 and f0.1 , and hence this form of the law is in
terms of parameters that can all be determined experimentally from a tensile test.
The Ramberg-Osgood law is now widely used because its predicted behaviour is
very close to the actual behaviour of aluminium alloys, in particular for the first
part of the stress-strain diagram.
Sometimes it is not possible to get the value of f0.1 and specifications usually do
not provide this value with the other mechanical properties. Thus, there are other
methods to determine the value of the exponent n. Two proposals of this type are
explained below.
The Steinhardt’s proposal [1971] is very simple and concise: E and f0.2 values
are assumed to be the minimum required by specifications, and it is also assumed
that:
10n = f0.2 (2.15)

This simple expression is based on experimental results that showed a connec-


tion between the 0.2% tensile proof stress and the exponent of the Ramberg-Osgood
law, n. Non-heat-treated alloys have an exponent, n, in the range of 10-20, and heat-
treated alloys have an exponent, n, in the range of 20-40. If the analysis concerns the
range of elastic deformations or so-called elastic range (CEN [2007]), Steinhardt’s
proposal provides a sufficient prediction of the value of n.

13
If the analysis concerns the range of plastic deformations or so-called plastic
range, Mazzolani’s expression can be used where three values of the material spec-
ifications are taken into account [1974]. Starting from the minimum values required
by specification for mechanical properties: f0.2 - elastic limit, σmax - ultimate strength,
εmax - elongation at rupture, the following approximate expression for the exponent
ń is proposed:
ln 2
ń = (2.16)
ln(1 + kχ)
where
σmax − f0.2 σmax
χ= [N/mm2 ] (2.17)
10εmax f0.2
and k is a dimensional constant. This expression has been verified using the statis-
tical results of tests carried out at Liège University by Bernard et al. [1973]. On this
basis it has been assumed that k = 0.028 mm2 N−1 .

2.3.2 Geometrical Imperfections


Imperfections affect the load-bearing capacity of the members and cannot be ig-
nored (Mazzolani [1995]). Unavoidable imperfections are produced during the fab-
rication process of structural members, e.g. initial out-of-straightness, thickness de-
viation, initial curvature, and load eccentricities. In this section aluminium proper-
ties and initial geometrical imperfections are discussed.
By means of geometrical imperfections the differences between the nominal and
the actual shape of the structural element, or so-called initial out-of-straightness is
considered. Extensive studies for initial out-of-planeness definition have been car-
ried out for steel columns: Beer and Schultz [1970], Fukumoto et al. and Essa and
Kennedy [1993]. The most common approaches to define plate deflections is to mea-
sure the largest out-of-plane deflection of the plate from initial straightness, or to
describe the whole deflection pattern in sinusoidal shape by the use of Fourier anal-
yses. The post welding shape of welded steel plating in ships has been determined
by Antoniou et al. [1984], Antoniou [1980], Carlsen and Czujko [1978] and Kmiecik
et al. [1995]. Carlsen and Czujko [1978] found the following expression for the max-
imum out-of-plane deflection of a plate:

wmax = 0.016b − 0.36t (2.18)

Corresponding results found by Antoniou [1980] can be expressed as:

wmax = 0.014b − 0.32t (2.19)

Similar studies are known in aluminium, e.g. Clarke [1987] found that the initial
imperfection amplitude varies from 0.0066b to 0.021b. It was concluded that slender
plates show larger initial imperfection amplitudes than stocky plates.
Due to the production process of aluminium plates, variation in thickness might
occur. The plates will have different values of the thickness in different parts of

14
the plates, and the thickness will differ from nominal thickness. According to Maz-
zolani [1995] the different national specifications for extruded profiles allow a de-
viation from nominal thickness equal to 5 percent. This value can reach 10 percent
in the case of thin profiles with a thickness less than 5 mm. From the results of im-
perfections measurements, it was concluded that out-of-straightness of aluminium
extruded bars is usually less than in steel. The amplitude of the geometrical imper-
fections for extruded bars is limited by the value of about L/2000, for welded bars
this limit increases to L/1300.

2.4 Local Buckling


It was mentioned in Chapter 1 that applied thin-walled aluminium structural ele-
ments (beams and columns) often consist of complex cross-sectional shapes that are
composed of connected plate elements. When a plate is subjected to compression,
bending or shear, or a combination of loads, the plate may buckle locally before the
whole structural member becomes unstable.

Local buckling is characterized by distortion of the cross-section with only rotation


at the fold lines of the section. Local buckling is usually prevalent and occurs
with repeated short buckling waves of individual plate elements.

2.4.1 Slenderness parameter


The prediction of the ultimate resistance of the structural member due to the oc-
currence of local buckling is rather complex even for simple symmetrical cross-
sectional shapes. One of the design methods for buckling analysis makes use of
a non-dimensional strength design curve for a column, which is based on a non-
dimensional slenderness.
The buckling load depends on the slenderness of the cross-section element, which
is normally determined by the ratio of the width divided by the thickness (β = b/t).
In many cases the more general parameter for slenderness, λ is used
r
f0.2
λ= , (2.20)
σcr
where f0.2 is the 0.2% proof stress and σcr is the elastic buckling stress for a perfect
flat
p plate without initial buckles or residual stresses. λ is proportional to b/t and
f0.2 /E and depends on the loading conditions and the boundary conditions, e.g.
the connection to other cross-sectional elements.
For a plate simply supported along the edges, the critical buckling stress is ex-
pressed by Equation 2.6. For an axially loaded plate with simply supported edges,
kcr = 4 (see Figure 2.1). The slenderness parameter can now be determined by
inserting expression 2.6 into 2.20.

15
s r
f0.2 b f0.2
λ= π2 E t2
= 0.526 , for kcr = 4 and ν = 0.3. (2.21)
kcr 12(1−ν 2 ) b2
t E

A principal relationship between axial resistance N (N = σA, where A is the


cross-sectional area ) and the slenderness parameter λ for a slender plate is given
in Figure 2.7. Failure of the plate elements is represented by the axial squash load
N0.2 = f0.2 A , critical axial load Ncr = σcr A and ultimate load Nu = σmax A . It
can be shown that the load exceeds the buckling load, which is called post-buckling
strength. The resistance in the post-buckling stage is important for slender plates.

Figure 2.7: Relationship between slenderness parameter λ, buckling load Ncr and ultimate
load Nu .

2.4.2 Classification of cross-sections


The edge restraining conditions of the plate elements of the cross-section influence
the occurrence of local buckling (see Figure 2.3 and Table 2.1). From this point of
view, two groups of plate elements are recognized: flat internal and flat outstanding
elements. Cross-sectional shapes composed by different plate elements (e.g. a rect-
angular hollow section composed by only internal flat elements) are investigated
by many authors. An extensive experimental program for aluminium square hol-
low sections, rectangular hollow sections, channels and angles has been developed
to calibrate a prediction approach for local buckling ( Mazzolani et al. [1996], Maz-
zolani et al. [1997] and Mazzolani et al. [2001]). The aim of the experiments was
to provide Eurocode 9 CEN [2007] with the necessary data for classification of alu-
minium members.
Design methods control the capability of members to reach a given limit state.
Referring to a relationship between a relative force F/Fmax and a relative displace-
ment u/umax in Figure 2.8, the above classes correspond to the following behaviour
according to Eurocode 9 (CEN [2007]):

16
Figure 2.8: Classification of cross-sections (from CEN [2007]).

1. Class 1 ductile cross-sections can develop the ultimate resistance without the
occurrence of buckling, reaching the ultimate value of deformation with - de-
pending on the type of alloy (collapse limit state) - considerable plastic strains;

2. Class 2 compact cross-sections are capable of developing ultimate plastic resis-


tance but have limited deformation capacity, which is prevented by plastic
instability phenomena (plastic limit state α1 );

3. Class 3 semi-compact cross-sections are able to reach the proof strength f0.2 of
the material, but the development of the important plastic deformations is
prevented. Only small plastic deformations are allowed, sometimes giving
rise to a brittle behavior ( elastic limit state α2 );

4. Class 4 slender cross-sections, whose strength is governed by the occurrence of


local buckling phenomena, are not able to reach the proof strength f0.2 . No
plastic deformation is allowed within the section (elastic buckling limit state
α3 ).

Therefore, cross-sections of Class 4 need to be checked for local buckling. A


slender cross-section must be designed according to resistance formulae calculated
with reference to the effective thickness of the element. The effective thickness is
determined by the reduction factor for local buckling.
Three different types of elements are recognized in a cross-section: flat outstand,
flat internal and curved internal elements. The basic parameter for the classification
criterion is the width-to-thickness ratio (b/t) of each element. Experiments have
been executed to identify the influence of the width-to-thickness ratios on the local
buckling phenomenon (Mazzolani et al. [1996]). On the basis of tests, the classi-
fication of cross-sectional internal and outstand elements linked to the values of

17
the slenderness parameter has been introduced in Eurocode 9 CEN [2007]. Clas-
sification in Eurocode 9 does not account for the interaction between flat internal
elements and flat outstand elements. Thus, another classification criterion has been
developed by Faella et al. [2000], taking into account the interaction between the
slenderness parameters of the plate elements. Updating the classification method
provided by Eurocode 9 was advised.

2.5 Distortional buckling


The distortional type buckling mode was first detected by Lundquist and Stow-
ell [1942] and later also by Gallaher and Boughan [1971] in their investigation on sta-
bility of panels with Z-stiffeners. Van der Maas [1954] studied the buckling behavior
of hat-section columns and noticed another kind of local buckling phenomenon in-
volving flange-stiffener rotations. Later, Sharp [1966] called the attention to ”flange-
stiffener” instability and developed a simple formula for the buckling load of lipped
channel columns failing in this mode. The influence of edge and intermediate stiff-
eners on the buckling behaviour of lipped channel columns and beams has been
studied further by Desmond, Pekoz and Winter [1977], [1981] at Cornell University.
A significant amount of experimental and analytical investigations for ”flange-
stiffener” buckling has been carried out at the University of Sydney by Hancock
et al. [1990] and Hancock [1997]. Finally, distortional buckling received its current
name. As a result of this research, an analytical prediction model for the distortional
buckling stress of thin-walled cold-formed carbon steel sections has been developed
by Lau and Hancock [1986]. This analytical method has been implemented in the
Australian/New Zealand codes [2005].
A clear definition for distortional buckling was given in the work of Yu and
Schafer [2003]:

Distortional buckling involves changes in geometry when both rotation and trans-
lation occur at the flange/lip fold lines of the section. The wavelength of dis-
tortional buckling is in between that of local buckling and torsional-flexural
buckling, while torsional-flexural buckling is characterized by one half-wave
over the length of the structural member.

In general, for distortional buckling a reduction is applied in addition to local buck-


ling. Existing design rules do not distinguish between local and distortional buck-
ling. One of the reasons of the lack of knowledge of distortional buckling, is that it
is hard to perform experiments for distortional buckling.
Distortional buckling tests on cold-formed steel beams with C and Z sections
have been executed by Yu and Schafer [2006]. The results of the distortional buck-
ling tests have been compared with the results of local buckling tests carried out
by Yu and Schafer [2003]. In both tests steel beams with identical geometry and
material have been used and identical test set-up, except the restraint conditions for
the compression flange. In the distortional buckling test the compression flange was

18
not restrained by attachment to paneling, while for local buckling test the compres-
sion flange was restrained by a special fastener configuration to be able to prevent
the distortional buckling mode. It was discovered, that the elastic stiffness for local
buckling and distortional buckling tests was the same, but failure in case of dis-
tortional buckling was initiated before failure according to local buckling occurred.
The results of the local and distortional buckling tests have been compared with
design specifications and with a direct strength method developed by Schafer and
Peköz [1998]. It was concluded that the so-called direct strength method provides
an adequate prediction of the distortional buckling capacity. A description of this
method can be found in section 2.7.

2.6 Numerical Investigations


2.6.1 Finite Element Method
Various numerical methods can be used to analyse the buckling behaviour of alu-
minium thin-walled members. The finite element method (FEM) is one of the most
popular, as it successfully handles almost any buckling problem. In the current
investigation, the finite element program DIANA is used. General aspects are de-
scribed here, detailed information can be found in the DIANA manual, Hendriks
and Wolters [2007].
Finite element calculations of the thesis are performed by three analyses: linear
elastic, stability and non-linear analysis. Linear elastic analysis assumes the first
order analysis with elastic material characteristics. In most large deformation prob-
lems, linear elastic assumptions are not satisfactory. Therefore, material and geo-
metrical nonlinear analysis is considered for the best approximate solutions. How-
ever, the result of the linear elastic analysis is an essential input for the non-linear
analysis. The stability analysis according to the finite element code (Euler Eigen-
value analysis) results in several buckling modes. Each mode represents a value of
the buckling load and a buckling deformation pattern. The deformation pattern of
the first Euler buckling mode is applied as imperfection pattern for the geometri-
cal non-linear analysis. The Newton-Raphson iterative method is used to solve the
nonlinear problem. For material hardening behaviour, the von Mises yield criterion
and a work-hardening stress-strain relation is applied.

2.6.2 Finite Strip Method


FSM uses a discretization of a cross-section into a series of longitudinal elements,
so-called strips. Based on these strips elastic and geometric stiffness matrices can
be formulated. Hancock [1978] presented the formulation of a semi-analytical fi-
nite strip method (FSM) to perform buckling analyses of thin-walled members. This
paper introduced first the concept of the ”signature curve”: a curve representing
the relation of the buckling stress with the member length, which exhibited a local
minimum, identified with local buckling, and a descending branch, identified with

19
lateral-torsional buckling. Based on numerical study of columns with edge stiff-
ened cross-sections, Hancock [1985] has shown later that the ”signature curve” has
two local minima, identified with local buckling for short columns and distortional
buckling for intermediate columns. Furthermore, is has been shown that distor-
tional buckling may be critical.

Figure 2.9: Buckling modes of lipped channel in compression (from Hancock [2008]).

A result of the finite strip buckling analysis of the lipped channel section, sub-
jected to a uniform compression, is shown in Figure 2.9 as the buckling load versus
the half-wavelength. Point A corresponds to the local buckling mode and the buck-
ling shape includes deformation of the web without translation at the flange/lip
junction. Point B corresponds to the distortional buckling mode and the buckling
shape includes deformation of the web and translation at the flange/lip junction.
Points C, D and E correspond to torsional-flexural and flexural buckling.

The constrained finite strip method (cFSM) is a new extension to the FSM. The
constrained method is developed by Adany and Schafer [2006] and is able to pro-
vide both modal decomposition and modal identification. It means that local, dis-
tortional and global deformation modes can be recognized and decomposed, and
participation of each buckling mode can be defined. The conventional finite strip
method combined with the constrained finite strip method is implemented in an
open source program CUFSM Schafer [2006].

20
2.6.3 Generalized Beam Theory
Distortional buckling of thin-walled members can be analyzed using a generalized
beam theory. The theory of beams for thin-walled members developed by Vlasov
[1940] has been extended to a first order Generalized Beam Theory (GBT) by Schardt
[1989]. GBT accounts for cross-sectional distortion and flexural/flexural-torsional
modes. Davies[1998] extended the first order GBT to second order analysis. It al-
lows local and distortional buckling modes to be considered. A substantial number
of recent publications devoted to the development of GBT was produced by Camo-
tim et al. [2004] at the Technical University of Lisbon. For example, Dinis et al. [2006]
derived a GBT formulation for the buckling analysis of thin-walled members with
arbitrary ”branched” open cross-sections by studying the local, distortional and
global buckling behaviour of I-section beams with unequal stiffened flanges. The
code GBTUL, Bebiano et al. [2008], has been recently developed as a freely available
software program to perform buckling analyses of open-section thin-walled mem-
bers with arbitrary end support and (non-uniform) loading conditions.

2.7 Direct Strength Method


Experimental research has been executed by Kwon and Hancock [1992] to evalu-
ate the ultimate strength of cold-formed steel lipped channel columns, with and
without an intermediate web stiffener, failing in local and distortional modes. The
distortional buckling mode was always the critical one. However, interaction be-
tween local and distortional buckling modes was observed. Based on experimen-
tal results of Lau and Hancock [1986] for plain lipped channel columns and Kwon
and Hancock [1992] for plain and web-stiffened lipped channel columns, Hancock
et at. [1994] showed that the existing design specifications do not provide a suffi-
ciently accurate prediction concerning distortional buckling. Collecting all the data,
Lau and Hancock [1986] proposed a design approach using the Johnson parabola to
determine the critical stress fn :

For λd < 1.414 µ ¶


λ2
fn = 1 − d fy (2.22)
4
For λd ≥ 1.414 µ ¶
1
fn = fy (2.23)
λ2d
where the non-dimensional slenderness is given by:
r
Ny
λd = (2.24)
Nod
The critical stress fn can be multiplied by the cross-sectional area to get the de-
sign column strength Nc . The load Ny = Afy is the squash load of the section and
the load Nod is the elastic distortional buckling load.

21
An alternative formulation for the distortional buckling strength based on the
Winter equation for the postbuckling behaviour of plates was given by Kwon and
Hancock [1992]:

For λd ≤ 0.561
Ncd = Ny (2.25)
For λd > 0.561 " µ ¶0.6 # µ ¶0.6
Nod Nod
Ncd = 1 − 0.25 Ny (2.26)
Ny Ny
where λd is given by Equation 2.24.
Based on the distortional design curve developed by Hancock et al., Schafer
and Peköz [1998] proposed the newly developed Direct Strength Method (DSM).
DSM was implemented and is fully described in design standards for cold-formed
steel: American Iron and Steel Institute (AISI) [2006] and Standards of Australia and
Standards of New Zealand (SA-SNZ) [2005]. The DSM assumes a buckling load or
moment for the whole section either as local or distortional buckling. If the mode
is distortional, then Equations 2.24, 2.25 and 2.26 apply. For local buckling strength
prediction, Schafer and Peköz proposed the following equations:

For λl ≤ 0.776
Ncl = Nce (2.27)
For λl > 0.776 " µ ¶0.4 # µ ¶0.4
Nol Nol
Ncl = 1 − 0.15 Nce (2.28)
Nce Nce
where the non-dimensional slenderness is given by:
r
Nce
λl = (2.29)
Nol
Nce is the column strength in flexural or flexural-torsional mode and Nol is the local
buckling load.
Yang and Hancock [2004] proposed the direct strength formulas to determine the
nominal axial strength of lipped channel sections undergoing interaction between
local and distortional buckling. The formulas are similar to represented above equa-
tions but column strength in flexural or flexural-torsional mode is substituted with
the distortional buckling strength. Thus, interaction between local and distortional
buckling is described by:

For λl1 ≤ 0.776


Ncl = Ncd (2.30)
For λl1 > 0.776 " µ ¶0.4 # µ ¶0.4
Nol Nol
Ncl = 1 − 0.15 Ncd (2.31)
Ncd Ncd

22
where the non-dimensional slenderness is given by:
r
Ncd
λl1 = (2.32)
Nol
Ncd is the distortional buckling strength computed by equations 2.25 and 2.26.

2.8 Mennink’s prediction model


The model of Mennink [2002] has been developed on the base of experimental and
numerical data for aluminium rectangular hollow sections, square hollow sections
and I-sections. These sections have in common that they are doubly symmetrical
and consist of only two sets of identical plates (webs and flanges). The observed
behaviour showed a remarkable coincidence with plate buckling specifics, thus ob-
taining an almost analytical background.
Mennink’s model is based on the actual buckling behaviour. It combines an accu-
rate description of both elastic and inelastic buckling and the elastic post-buckling
strength, and provides a safe approach with respect to the inelastic post-buckling
resistance. Key aspect of the model is the application and determination of the ac-
tual critical stresses due to initial buckling (numerically determined) and secondary
buckling (analytically determined). Contrary, current design rules underestimate
initial buckling as calculations are based on the critical stresses of individual plates.
In case of inelastic buckling, the specimen fails at the inelastic critical stress. Men-
nink’s model is able to determine this load based on the elastic critical stress. It
should be noted that Mennik’s prediction model can be used only with respect to
initial instability of the cross-section and not the plate elements. The occurrence of
the post-buckling resistance is limited to the elastic range. Though this underesti-
mates the failure strength, it is not extended as inelasticity may cause a sudden and
unpredictable occurrence of secondary buckling modes that could lead to unsafe
results.
The prediction model was verified with experimental results on rectangular hol-
low sections, U-sections and some complex sections. It was concluded, that the
prediction model could accurately describe the local buckling behaviour of profiles
with actual simple cross-sections. The occurrence of overall and distortional buck-
ling was not accounted for. Only those specimens were considered of which the
failure pattern could be associated with local buckling.

23
2.9 Literature Study Evaluation
The well-known aspects of stability problems were discussed at the beginning of
this chapter, considering overall buckling of centrally loaded columns and plates
under uniform compression. This theory is the basis for more complex instability
effects, such as local and distortional buckling, as well as the interaction of buckling
modes.
In daily practice, aluminium extruded profiles with thin wall thicknesses are
usually applied. Because of this and the high ratio between strength and modulus
of elasticity, aluminium structures are sensitive to local instability. Local buckling
has an appreciable effect on the stability behaviour of the whole structure. Con-
sidering the local buckling effect, the following parameters are of importance: the
slenderness ratio, loading and boundary conditions, material characteristics.
Overall buckling, which is flexural and flexural-torsional buckling, is well de-
scribed in the current design rules for aluminium. In case of local buckling or more
complicated distortional buckling, the analytical rules are limited for specific cross-
sections. Studying local instabilities of the structural element, the cross-section is
built up as a number of individual plates and plate buckling phenomena are con-
sidered. In general, this leads to conservative results.
Mennink’s prediction model gives more accurate results for local buckling pre-
diction of simple cross-sections. For more complex cross-sectional shapes the model
has to be checked. Cross-sectional instability of aluminium structural members de-
mands to be investigated for distortional buckling phenomenon.

24
Chapter 3
Experimental Investigation

Abstract
To study the actual cross-sectional instability of aluminium structural members it is
essential to have experimental data. Therefore, an experimental program was defined
for aluminium elements with profiles well-selected to study the local and distortional
buckling effects of aluminium complex cross-sectional shapes. Tests were executed for
extruded and cold-formed specimens at the Pieter Van Musschenbroek laboratory at
Eindhoven University of Technology. The current chapter contains the description of
test set-up, actual dimensions and imperfection measurements, determination of mate-
rial characteristics, and the results of compression tests. The experimental results are
used for the validation of a finite element model and development of a practical analytical
design approach in the following chapters.

3.1 Program Description

C
ross-sectional shapes were chosen for the experimental program to provide data
for local and distortional buckling resistance of complex aluminium sections.
The selection procedure was based on using the finite strip method according to the
CUFSM program (Schafer [2006]). Detailed description of the selection procedure
can be found in Kutanova et al. [2007]. In total, eight shapes were selected (see
Figure 3.1). The selection of the cross-sectional shapes resulted from two research
concepts described below.

Figure 3.1: Examples of cross-sections selected for investigation

For the first research concept (Figure 3.2) the investigation starts with a simple
shape (angle-section) and goes forward by adding plate elements to this section.
In other words, by adding complexity to the original shape the effects of changing
the cross-sectional geometry are examined. Studying the contribution of the vari-
ous plate elements to the load-bearing capacity of the structural element is one of

25
the ideas for investigation of the cross-sectional behaviour of complex shapes. Re-
garding cross-sectional instability two modes are considered: local and distortional
buckling modes. Specimens with represented shapes were produced of aluminium
alloy EN AW - 6060 T66.

Figure 3.2: Series of cross-sections with gradually increased geometrical complexity.

The second research concept is focused on the interaction of local and distor-
tional buckling modes. Two different types of Z-sections are considered (see Fig-
ure 3.3). Specimens of the first shape are aluminium extrusions made of EN AW-
6060 T66 aluminium alloy. This shape is expected to represent the pure local buck-
ling mode. To study the occurrence of distortional buckling mode, Z-shaped sec-
tions with additional flange stiffeners have been selected. That kind of shape is not
available in stock in industry, therefore Z-shaped profiles with different dimensions
and thicknesses have been cold-formed. It was decided to manufacture specimens
of EN AW 6082-T6 aluminium alloy plus identical specimens made of EN AW 5083-
H111 aluminium alloy to be able to investigate the influence of the material on the
stability behaviour.

Figure 3.3: Selected shapes to study the material influence and distortional buckling

A complete description of the experimental program is reported in Kutanova


et al. [2008a]. A total of 85 compression tests are performed to observe the influ-
ence of the cross-sectional instability on the strength and behaviour of aluminium
specimens with L-, Z- and C-shaped profiles. All considered specimens are divided
into two experimental subprograms based on the production method. As described
earlier, shape numbers one to seven of Figure 3.1 were produced out of aluminium
extrusions, thus the specimens of this group of cross-sections are included into the
”extruded” subprogram. Specimens with Z-shaped profile with additional flange
stiffeners (number 8) were cold-formed and belong to the the ”cold-formed” sub-

26
program. Tables 3.1 - 3.3 provide the specification of profiles selected for the exper-
imental program.
Specification of all cross-sectional sides can be read while looking to the cross-
section from top left to bottom right, the value of thickness and the length at the end,
see Figure 3.1. Numbers in the specification correspond to (d) x (a) x (h) x (b) x (c) x
(t)-(L), where (d) is the top flange stiffener, (a) is the top flange, (h) is the web, (b) is
the bottom flange, (c) is the bottom flange stiffener, (t) is the thickness and (L) is the
length. Table 3.1 gives the dimensions of extruded cross-sections according to the
appropriate number of the shape in Figure 3.1. For the specimens of groups 3C80E7
and 3C80E8 in Table 3.1, the same type of cross-section is used but the lengths of the
aluminum specimens are different.

Table 3.1: Specification for ”Extruded” subprogram, 6060-T66 alloy.

Specimen Specimen Shape Specification h a b c t L


notation quantity No. [mm] [mm] [mm] [mm] [mm] [mm]
2Z40E1- 5 1 20x40x20x2-250 40 20 20 2 250
3Z35E2- 5 1 20x35x30x3-350 35 20 30 3 350
3L80E3- 5 2 80x80x3-700 80 80 3 700
3C80E4- 5 3 80x80x20x3-700 80 80 20 3 700
3C80E5- 5 4 40x80x80x3-700 80 40 80 3 700
3C80E6- 5 5 40x80x80x20x3-500 80 40 80 20 3 500
3C80E7- 5 6 50x80x80x10x3-500 80 50 80 10 3 500
3C80E8- 5 6 50x80x80x10x3-1300 80 50 80 10 3 1300
3C80E9- 5 7 40x80x80x40x3-500 80 40 80 40 3 500

The Z-shaped specimens in Tables 3.2 and 3.3 were produced by folding alu-
minium plates. It should be noted that when using the folding process, it is im-
possible to obtain sharp corners. The radiuses in cross-sectional corners were ap-
proximately 2-2.5 times the plate thickness. Two types of alloys of aluminium plates
were used: 6082-T6 and 5083-H111. For each type of alloy four different profiles
were manufactured. For each group of profiles five identical specimens have been
prepared.

Table 3.2: Specification for ”Cold-formed” subprogram, 6082-T6 alloy.

27
Table 3.3: Specification for ”Cold-formed” subprogram, 5083-H111 alloy.

Notation for all the specimens is in line with the specification. The first num-
ber corresponds to the value of thickness. The second letter indicates the shape of
the profile: ”Z” for Z-shaped section, ”L” for angle section, ”C” for C-shaped pro-
file. The following number is the cross-sectional web length (or value (h) in the
tables). The next letter means the production method: ”E” - extruded, ”F” - folded.
Last number is an ordinal number for each batch of profiles. Ordinal numbers are
in accordance with the experimental report Kutanova et al. [2008a]. Cold-formed
specimens are identical, but made of different aluminium alloys. Therefore, in the
specification of cold-formed specimens the last letter ”A” in the profile number in-
dicates that the specimen was manufactured out of 6082-T6 aluminium alloy. The
final letter ”B” in the profile number indicates that the specimen was produced out
of 5083-H111 aluminium alloy. An example of the notation is shown in Figure 3.4.

Figure 3.4: Specimen notation.

3.2 Measuring Dimensions


The actual dimensions of all experimental specimens are measured. The dimensions
are measured using an electronic sliding caliper. For dimensions with a length less
than 160 mm, a caliper with an accuracy of 0.01 mm is used; for dimensions with a
length greater than 160 mm but less than 450 mm a caliper with an accuracy of 0.1
mm is used. For long profiles with a length more than 500 mm a ruler of 1 meter
length has been used with an accuracy of 0.5 mm.
It is not easy to measure the radiuses at the angles of the specimen, which were
formed by the folding process. To measure radiuses at the angles of the profile,
gages of different radiuses have been used trying to find the best fit with the angles.
The best fit defines the radius at the angle. More accurate measurements are carried
out for two profiles with the most round angles. Two thin samples have been cut
out of two different profiles with the most round angles (see Figure 3.5). Such thin
samples have been zoomed up with a Nikon profile projector (Model V-16D), which

28
allowed to define precisely the radius at the angles. It was observed that radiuses be-
tween flanges and flange stiffeners are equal to each other and higher than radiuses
between web and flanges. That is why there are two values of radiuses given for
each specimen: r1 corresponds to the radius of curvature between web and flange,
r2 corresponds to the radius of curvature between flange and flange stiffener.

Figure 3.5: Z-shaped profile cut to measure radiuses.

Measured dimensions for extruded and cold-formed profiles are presented in


Appendix A. Measured dimensions are used further as an input for the finite ele-
ment model. The FE-model and the prediction model require the cross-section to
be substituted by a number of plates connected at nodes. Nodes are placed at the
intersection of the heart-lines of the cross-section. Therefore, the plate thicknesses
are identical to the measured ones, but the plate widths are taken as heart-to-heart
values (see Figure 3.6). The following notations are used for heart-to-heart values:
”h1 ” - is the web height, ”a1 ” and ”b1 ” - are the dimensions of flanges, ”d1 ” and
”c1 ” - are the dimensions of flange stiffeners. The axial length of the specimen is
measured for each side and averaged to one value L.

Figure 3.6: Measured heart-to-heart values of cross-sectional dimensions.

3.3 Measuring Imperfections


3.3.1 Introduction
This section describes the way of measuring the initial out-of-planeness of alu-
minium profiles selected for the experimental program. The aim of the imperfec-
tions measurements is to determine a value e, which represents the magnitude of the
initial imperfections. This value can be applied in the subsequent finite element (FE)
analysis as an imperfection amplitude of the deformation pattern, which is identi-
cal to that of the first Euler buckling mode. Based on the results of measurements,
the real imperfection pattern can be defined as well and implemented into the FE

29
analysis. Hence, the imperfection measurements for all tested specimens have to
be obtained by determining the deformation pattern, the magnitude and the ampli-
tude of the initial imperfections. All results and observations experienced during
the imperfection measurements can be found in Kutanova et al. [2008a].

3.3.2 Measuring concept


It is important to take into account that initial out-of-planeness is a combination of
local and overall imperfections. It is assumed that the amplitude of local imperfec-
tions occurs at the middle of the internal element (web buckling). The amplitude
of overall imperfections occurs at the unsupported edge of the outstanding element
(flange buckling). Distortional buckling causes initial bending or twisting of the out-
standing members (flanges and lips), that is why it is assumed that the amplitude
of imperfections due to distortional buckling is equal to the amplitude of overall
buckling. Current research and experimental program is focused on local and dis-
tortional buckling. Therefore, local and overall imperfections are taken into account.
The method of calculation of the local and overall imperfections amplitudes is dis-
cussed below.
An illustration of imperfection measurements for the web plate of the specimen
is represented in Figure 3.7. It can be noticed that a substantial number of mea-
surements have been done. For the sides subjected to local or distortional buckling,
i = 1 .. m points are measured in lateral (x−) direction and j = 1 .. n in axial (y−)
direction. Approximately the values of 72 points are found to represent the imper-
fection pattern.

Figure 3.7: Schematization of the measuring process.

The starting position of the measurements is indicated on the right drawing of


the figure. At the corner of the specimen’s side the digital gauge is set as zero, which
means that the value of initial deviations or imperfections are assumed to be equal
to zero at this point. In other words, the starting point is a reference point at location
x = y = z = 0. However, it should be noted that this particular starting point has
no reference regarding the whole specimen and only indicates the ”zero” point for
one plate element. From the starting point the specimen is shifted in both the axial
and lateral directions of the plate. This procedure is repeated for all sides of the
specimen.

30
3.3.3 Test set-up
A computerized bench is used to accurately measure the flatness of each side of the
specimen (see Figure 3.8).

Figure 3.8: Imperfection test set-up.

It is very important to place the specimen in one plane. It has been noticed before
that measurement results are influenced by the reference plane that is determined
by the specimen’s placement. Cold-formed Z-shaped profiles have rounded corners.
Due to the corners, it is difficult to put cold-formed specimens on the bench in such a
way that they are stable and flat. It was decided not to measure the area close to the
corners because in this case the imperfection measurement results might contain
a substantial influence of the corners. Figure 3.9 shows an example of specimen
placement on the bench and the way of supporting the specimen.

Figure 3.9: Imperfection measurements for web and flange.

Specimens with lengths of 1200 and 1300 mm did not fit to the test set-up. For
these specimens the measurements start from the middle to the left or to the right
direction depending on the placement of the specimen.

31
Figure 3.10: Test set-up for long profile.

3.3.4 Test results processing


Fitting the results of measurements: When a set of observations is taken, the data
may include errors. The Least Square Method can be used to adjust the modeling
function to the measured data.

Figure 3.11: Fitting plane for measured points, using the Least Square method.

In the Least Square Theory the equation for the plane is: ź = ax́ + bý + c, where
the dependent ź−value is a function of the independent x́− and ý− values. Values a
and b are coefficients corresponding to each x́−, ý− values, and c is a constant value.
Knowing the expression of the fitting plane (ź = ax́ + bý + c) we can get the values
of the ź in a vector (ź1 , ź2 , ... źn ) and then calculate the distance from each measured
point (z1 , z2 , ... , zn ) to that plane resulting in ”linearized” values of imperfections.
The maximally registered values, referred to as emax and emin , provide an indication
of the magnitude for the maximum and minimum values of imperfections.

32
Method for calculation of the amplitude of imperfections: One of the proposed
procedures for calculation of the amplitude of imperfections is described in Men-
nink [2002] and Mennink and Schillemans [2002]. Similar calculations are applied
for the considered specimens. As it was mentioned before, the amplitude of local
and overall imperfections are calculated (see section 3.3.2).
At first, the calculation of the overall imperfection amplitude is described. The
amplitude of overall imperfection occurs at the unsupported edge of the outstand-
ing element (flange buckling). For buckling of the outstand element, calculations
are made for i = 1 and i = m (longitudinal edges in Figure 3.12). It means that
two values of amplitudes are found for each longitudinal edge and the maximum
of these two values is the amplitude of the overall imperfection.

Figure 3.12: Schematization of the measured points.

There are two cases in the calculation of imperfections: (1) zi;1 < zi,n , when the
value z of the first point is smaller than the value z of the last measured point in
longitudinal direction; (2) zi;1 > zi,n , when the value z of the first point is higher
than the value z of the last measured point in longitudinal direction.
If zi;1 < zi,n , then the graph of Figure 3.13 is relevant:

Figure 3.13: Imperfection amplitude calculation (case 1).


In the case when the value z of the first point is lower than the value z of the
last measured point in longitudinal direction, we draw a line from the first point
zi;1 to the last measured point zi;n . The distances between each measured point and
projection to that line are calculated. Calculations are performed with the formula
given below. The value of the overall imperfection amplitude is indicated with a
thick line and it can be found by:
µ µ¯ ¯¶¶
¯ zi;n − zi;1 ¯
¯
eov = max max ¯zi;j − zi;1 − (yi;j − yi;1 )¯¯ (3.1)
i=1;m j=1..n yi;n − yi;1

33
If zi;1 > zi,n , then the graph of Figure 3.14 is relevant:

Figure 3.14: Imperfection amplitude calculation (case 2).

In the case when the value z of the first point is higher than the value z of the
last measured point in longitudinal direction, we draw a line from the last measured
point zi;n to the first measured point zi;1 and through point zi;n parallel to the y-axis,
until the intersection with the axis z (line l). It can be seen from the drawing that
a triangle is formed. The distances between each measured point and projection to
the drawn line are calculated. Calculations are performed with the formula given
below. µ µ¯ ¯¶¶
¯ zi;1 − zi;n ¯
¯
eov = max max ¯zi;j − zi;n − (yi;n − yi;j )¯¯ (3.2)
i=1;m j=1..n yi;n − yi;1
To find a value of amplitude of imperfections due to the buckling of an internal
element (local buckling), the resulting values of imperfection (z values) in transverse
direction are considered. Two cases are considered: 1. z1;j < zm,j , when the value
z of the first point is lower than the value z of the last measured point in transverse
direction; 2. z1;j > zm,j , when the value z of the first point is higher than the value
z of the last measured point in transverse direction. For j = 1 .. n there are number
of m imperfection points in x direction. The values of the local imperfection am-
plitude are calculated with the same idea as was represented before for the overall
imperfection amplitude. The resulting formulas lead to:

1. if z1;j < zm,j , then


µ µ¯ ¯¶¶
¯ zm;j − z1;j ¯
eloc = max max ¯¯zi;j − zi;1 − (xi;j − x1;j )¯¯ (3.3)
j=1..n i=1..m xm;j − xm;1

2. if z1;j > zm,j , then


µ µ¯ ¯¶¶
¯ z1;j − zm;j ¯
¯
eloc = max max ¯zi;j − zm;j − (xm;j − xi;j )¯¯ (3.4)
j=1..n i=1..m xm;j − x1;j

3.3.5 Imperfections results


Tables A.3 and A.4 of Appendix A contain all calculated values of the imperfection
amplitude for each side of the tested specimens. The amplitudes of local imperfec-
tions are calculated as the value of (eloc ), the amplitudes of overall imperfections

34
are represented by the value (eov ). The maximum of these two values is defined for
each side of the specimen. The average value for all sides of the specimen is given in
the last column of the tables. Note, that for some plate components of cold-formed
specimens, rounded corners exerted undesirable influence on the imperfection mea-
surements results. Hence, extreme values appeared (eloc , eov ≥ 1 mm), which are
not taken into account when the average value is calculated. The average value
will be applied to the FE-analysis as a magnitude of the imperfection amplitude. It
should be noted that there were no imperfections measurements performed for two
long profiles: 3C80E8-4 and 3C80E8-5. Thus, knowing that specimens are almost
identical for the same batch, measurements for three specimens have been executed
and the results can be applied for the other two specimens as well.
It should be noted that for several specimens the imperfection measurements
have been done twice to check the repeatability of testing. It was found that the im-
perfection measurement results differ within only 2%. Table A.3 exhibits that initial
imperfections for extruded specimens are less than e < 0.39 mm, which is about
10 ∼ 15 % of the plate thickness. Table A.4 contains the measured values of imper-
fections for cold-formed specimens that are, in general, higher than imperfections
for extruded specimens. Extreme values of the calculated imperfection amplitude
are neglected. When the width of the plate member is short and the influence of the
rounded corners is high, it is not possible to perform several measurements on the
side, thus the local imperfections amplitude is equal to zero. In average, imperfec-
tions for cold-formed specimens are less than e < 0.94 mm.
According to imperfection measurement results, the imperfection pattern can be
defined (see Figure 3.15). The shape of the real imperfection pattern will be imple-
mented into the FE-model. Alternatively, the elastic buckling mode will be used as
imperfection pattern.

Figure 3.15: Initial deflections of the surface.

35
3.4 Material characteristics
3.4.1 Introduction
The tensile test is the most widely used test to determine the mechanical proper-
ties of materials. In this test, a piece of material - a tensile coupon - is pulled until
it fractures. During the test, the specimen’s elongation and applied load is mea-
sured. Strain and stress are calculated from these values, and are used to form a
stress-strain curve. From this curve, the elastic modulus and yield strength are de-
termined. The highest load in the tensile test gives the tensile or ultimate strength.
Thus, from one test a large amount of information can be obtained about the mate-
rial properties. Hence, this section gives a description and results of the tensile test
program executed at Eindhoven University of Technology (Kutanova et al. [2008a]).

3.4.2 Tensile coupon


Extruded sections of Table 3.1 were prepared out of extruded Z-section, L-section
and square hollow section (SHS) aluminium beams. Two Z-shaped beams of differ-
ent wall thicknesses, one L-shaped beam and six SHS beams have been used. Three
tensile tests are executed for each beam, thus 18 tensile coupons are manufactured
using the spark erosion process.
The tensile tests are carried out following the code NEN-EN 10002-1 (NNI [2001]).
The dimensions of a typical tensile coupon are shown in Figure 3.16. The width of
the parallel length is 20 mm. The parallel length is 100 mm, while the gauge length
is 80 mm. The total length of the tensile specimen is 300 mm.

Figure 3.16: Tensile coupon, dimensions in [mm].

Four aluminium plates with different wall thicknesses and different aluminium
alloys have been used for manufacturing of the Z-shaped cold-formed profiles (spec-
imens of Tables 3.2 and 3.3). Plates of EN AW-6082 T6 and EN AW-5083 H111 types
of alloys with two different thicknesses (2 mm and 2.5 mm) were used. All spec-
imens were prepared by folding the aluminium plate in one direction (called the
”bend”-direction). The fabrication process of aluminium (bending) might generate
changes in the metal structure and correspondingly changes of material properties
(Altenpohl [2002]). Therefore, it is reasonable to produce tensile coupons in ”bend”
direction and also in ”non-bend” direction. As can be seen in Figure 3.17, two tensile
coupons are cut in the longitudinal direction (”non-bend”) and another two in the

36
transverse (”bend”) direction. Thus, for each aluminium plate, four tensile tests are
executed. Note, that tensile coupons are produced out of aluminium plates and not
out of resulted specimens. Hence, the actual material properties after the bending
process are not measured.

Figure 3.17: Tensile coupons for cold-formed program. Dimensions in [mm].

3.4.3 Test set-up


A 250 kN test bench is used, see Figure 3.18. The load is applied by displacement
control. The test specimen is clamped at its edges by so-called wedge grips. A
displacement meter with an initial length L0 = 50 mm or L0 = 80 mm is also used
for determining the actual elasto-plastic behavior.

Figure 3.18: Test set-up for tensile test in 250 kN test bench.

Before the specimen is placed to the test rig, the width and the thickness of the
coupon are measured using a micrometer. Two strain gauges are applied at the
center of the opposite sides of the specimen. The measuring length of strain gauges
is 20 mm. It should be noted that strain gauges were not applied for all tested
tensile coupons. The load is applied at a speed of 0.3 mm/min until a stress of
approximately 120∼160 N/mm2 is reached. The test coupon is unloaded until the
stress of approximately 20∼40 N/mm2 is reached and then reloaded. The unloading
and reloading cycles aid to achieve an accurate elastic part of the material curve.

37
3.4.4 Tensile test results
This section provides an explanation of the tensile test data treatment. The tests
result in the applied force (F ), the measured strains (ε) of the two strain gauges
and a displacement (u) measured with the displacement meter. These results are
transformed to the stress-strain (σ − ε) relation: F is divided by the measured area
(thickness * width), u is divided by the measurement length (L0 = 50 mm or L0 = 80
mm). Figure 3.19 shows an example of the obtained stress-strain diagram, which
includes one curve according to the displacement meter (LVDT) and the results of
two strain gauges.

Figure 3.19: Tensile test results: stress-strain relation.

The stiffness-strain diagram is presented for all tensile coupons. Figure 3.20
shows the tangential stiffness ET of the stress-strain diagram based on the displace-
ment measurements with the displacement meter LVDT and the measurements of
two strain gauges. For load step i, ET is determined according to:
σ(i+2) − σ(i−2)
ET ;i = (3.5)
ε(i+2) − ε(i−2)
The average results of the reloading branch of the strain gauges are considered
to represent a linear behaviour that is well suited to determine the elastic stiffness
E of the material. In case when the strain gauges were not applied, the material pa-
rameters are defined from the measurements of the displacement meter. The value
of E is determined using the following formula:
σII − σI
E= (3.6)
εII − εI
Positions I and II are indicated in the graph (Figure 3.20). Depending on the spec-
imen’s dimensions and alloy these positions correspond to stresses of respectively
40 and 110 N/mm2 .

38
Figure 3.20: Tensile test results: stiffness-strain relation.

Knowing the magnitude of the modulus of elasticity, the values of f0.1 and f0.2
are determined according to Figure 2.6 and the following equations:
σi σi
f0.1 = σi when + 0.001 = εi ; f0.2 = σi when + 0.002 = εi . (3.7)
E E
Finally, the ultimate stress of the stress-strain curve (fu ) is determined (similarly
to Figure 2.6).
The results of the tensile tests are presented in Tables A.5-A.11 of Appendix A.
The measured and calculated data for material characteristics are given. For each
tensile specimen the modulus of elasticity E, the 0.1% proof stress f0.1 and 0.2%
proof stress f0.2 , and the ultimate limit stress fu are provided. Note, that only the
essential results are included. All results of the tested tensile coupons can be found
in Kutanova et al. [2008a].

3.4.5 Summary
The obtained characteristics are greater than material data given in Eurocode 9 (CEN
[2007]), because material properties in Eurocode 9 are characteristic values. How-
ever, the differences between the strength properties of the three aluminium alloys
(6060-T66, 6082-T6 and 5083-H111) are in accordance with Eurocode 9. The result-
ing modulus of elasticity E varies correspondingly to alloy: E = 66900 N/mm2
for 6060-T66, E = 71325 N/mm2 for 6082-T6, E = 71490 N/mm2 for 5083-H111.
The difference between test results for tensile coupons produced in rolling and non-
rolling directions is negligible for the 5083-H111 aluminium alloy and less than 2.3%
for the 6082-T6 aluminium alloy. This difference is neglected in the rest of this re-
search.

39
3.5 Test set-up for compression test
Due to the proposed experimental program (see section 3.1), two test set-ups are
used in the current investigation. The first test set-up for the Eindhoven experimen-
tal program is presented in Figure 3.21(a). The specimens are placed, freely, within
fixed supports. Thus, the specimens resemble clamped columns that are loaded in
compression by a uniform axial edge displacement. The tests are executed on a
250-kN bench. The bench is set to measure the axial shortening of the specimens.
For the specimens longer than 700 mm, a frame is built (see Figure 3.21(b)). Spec-
imens are placed horizontally on the bench, which is operated by displacement con-
trol of the hydraulic cylinders, imitating the first set-up. Support conditions are
performed as similar as possible to the first test set-up.
Two perfectly flat support plates are fixed to the test rig. The specimens are
placed between these plates, which means that the specimen is loaded through con-
tact between the specimen and the support plates. Therefore, it is important to
ensure that the edges of the specimen are flat and perpendicular to the specimen
longitudinal axis.

(a) Short speci- (b) Long specimens (more than 700 mm)
mens
Figure 3.21: Compression test set-up for specimens.

Displacements are measured using four digital gages (Mitutoyo IDC-112B). In


some experiments, strain gauges are applied at mid-length of all side walls of the
cross-section.
During experiments a gap between the specimen and support plates is observed
for many of the tested specimens (see Figure 3.22). It means that cross-sectional
side walls of the specimens are not flat enough and not parallel to each other. As
a result, the load is not applied simultaneously for all cross-sectional sides of the
specimen. To avoid the gap between the specimen and supports it is decided to
apply a layer of material that is softer than aluminium. Different types of material
(soft aluminium, copper, lead, Teflon) are tried in a way to compensate the gap and
to improve test conditions. In most of the cases, double layers of aluminum plates

40
(softer aluminium compared to the material of the tested specimens) are applied
in-between the specimen and the support plates, in order to obtain loading of all
cross-sectional sides right from the beginning of the test.

Figure 3.22: Gap between the specimen and support.

3.6 Compression test results


3.6.1 Agreement with FSM-prediction
Detailed observations of the compression tests and test results including load versus
axial shortening diagrams can be found in Kutanova et al. [2008a]. The results for
test 2Z200F10A are considered here as an example. First of all, this Z-shaped profile
has been selected to represent interaction between local and distortional buckling
modes. The selection analysis has been performed using the finite strip method as
it was indicated in section 3.1. Figure 3.23 represents the resulting output CUFSM-
screen (Schafer [2006]) based on the finite strip analysis for section 2Z200F10A.

Figure 3.23: CUFSM result for the selected Z-shaped profile.

41
It can be seen that the first critical point for this cross-sectional shape is related
to the initiation of the local buckling mode, while the second critical point defines
the interaction of local, distortional and overall buckling modes.
A similar behaviour has been observed during the compression test. The results
of the compression test are presented in a load-displacement curve, which relates
the measured reaction force (N ) in [kN] to the applied axial displacement (u) in mm.
As can be seen from compression test results for considered specimen 2Z200F10A,
local buckling initiated first at the load of about 30 kN (see Figure 3.24). Reduction
in stiffness can be noticed after the critical point is reached. The left photograph
accompanied with the diagram was taken right after the first critical load corre-
sponding to local buckling. Buckling waves of the specimen’s web are visible on
the picture. Later the interaction of different buckling modes occurred. The right
photograph was taken after the second critical load corresponding to interaction of
local, distortional and overall buckling.

Figure 3.24: Experimental result for selected Z-shaped profile.

3.6.2 Extruded experimental subprogram

A summary of test results for the extruded experimental subprogram is shown be-
low. Figure 3.25 contains the resulting deformed specimens for each batch of pro-
files. Compression tests for extruded specimens with various cross-sectional shapes
resulted in different buckling modes and interaction of buckling modes. Table 5.1
contains the resulting values of the critical loads due to local and distortional buck-
ling, the ultimate loads for the extruded specimens and their resulting buckling
modes. It can be seen that local, distortional, flexural and torsional buckling oc-
curred during compression tests. Corresponding cross-sectional shape is indicated,
nominal cross-sectional dimensions can be found in Table 3.1.

42
Figure 3.25: Compression test results: extruded specimens. From left to right:
2Z40E1, 3Z35E2, 3L80E3, 3C80E4, 3C80E5, 3C80E6, 3C80E7, 3C80E8, 3C80E9.

Table 3.4: Compression test results for extruded subprogram.

Specimen Shape Ncr; l Ncr; d Nu Buckling


No. [kN] [kN] [kN] mode
2Z40E1 1 - - 28 LF
3Z35E2 1 - - 46.2 F
3L80E3 2 18.3 - 41.9 T
3C80E4 3 45.1 - 54.6 LF
3C80E5 4 55.5 - 80.9 L
3C80E6 5 - - 111.4 LF
3C80E7 6 - - 99.9 LD
3C80E8 6 - - 72.9 DF
3C80E9 7 - - 132.3 L

L=local buckling/ D=Distortional buckling / F=Flexural buckling / T=Torsional buckling

Observations during compression tests have shown that some of the compressed
specimens failed at the supports. This can be explained by concentration of forces
close to the edges of the specimen or by the occurrence of a gap. For some of the
specimens visual gaps between the specimen edge and the support were observed
(see Figure 3.22).

43
3.6.3 Cold-formed experimental subprogram
The obtained values of the critical loads due to local and distortional buckling, ul-
timate loads for the cold-formed experimental subprogram are given in Table 5.2.
Nominal dimensions for specimens can be found in Tables 3.2–3.3. However, it
should be noted that all tested cold-formed specimens have rounded angles with
a radius of about 5.5-7.2 mm between the web and flanges and a radius of 6-7.8
mm between flanges and lips (see section 3.2). Figure 3.26 contains the resulting
deformed specimens for each batch of the cold-formed profiles.

Figure 3.26: Compression test results: deformed specimens. From left to right: 2Z200F10A,
2Z200F10B, 2Z100F11A, 2Z100F11B, 2Z50F12A, 2Z50F12B, 2.5Z50F13A, 2.5Z50F13B.

Table 3.5: Compression test results for cold-formed subprogram.


Specimen Mat Ncr; l Ncr; d Nu Md
[kN] [kN] [kN]
2Z200F10A 6082 30 57 100.9 LD
2Z200F10B 5083 30 57 73.1 LD
2Z100F11A 6082 75 - 107.7 L
2Z100F11B 5083 - - 70.6 L
2Z50F12A 6082 - - 111.3 L
2Z50F12B 5083 - - 69 L
2.5Z50F13A 6082 - - 57.2 DF
2.5Z50F13B 5083 - - 49.2 DF

Mat=Alloy type/ Md=Mode/ L=local buckling/ D=Distortional buckling / F=Flexural buckling

3.6.4 Results of the geometry influence


Figure 3.27 shows the average results of the compression tests for the selected L-
shaped and C-shaped profiles. It can be noticed that adding flanges of different
dimensions to the original L-shape results in a load capacity increase, except for the

44
specimen number 5. Cross-sectional area of specimen No. 4 is equal to the cross-
sectional area of specimen No.5. However, specimen No. 5 has a lower ultimate
load, which can be explained by the diminished value of one of the flanges. Flange
of a small value is similar to a stiffener that can initiate the distortional buckling
behavior. Therefore, the load-bearing capacity of specimen No. 5 is limited due to
the occurrence of distortional buckling.

Figure 3.27: Compression test results: ultimate values for the specimens with gradually in-
creased complexity of the cross-sectional geometry (values in kN).

The investigation for extruded L-shaped and C-shaped profiles aimed to study
the influence of a gradual increase of the complexity of geometry on the buckling
behaviour. It was shown that additional elements of the cross-sectional geometry
have an effect on the structural capacity and behaviour. However, it is important to
pay attention to the dimensions of additional plate elements and also to the type of
elements (internals or outstands).

45
3.6.5 Local vs distortional buckling
The results of the experimental program show that local and distortional buck-
ling behaviour of different cross-sections occurred during experiments. Figure 3.28
shows the observed buckling behaviour in a local and a distortional mode for two
specimens with different lengths but identical cross-sectional shape. In case of dis-
tortional buckling, the ultimate strength is lower when compared to local buckling.
It was also observed that for distortional buckling, the deformations exhibited long
wavelengths in the flange and/or flange stiffener, while local buckling occurred
with shorter wavelengths in the web.

Figure 3.28: Comparison of local and distortional buckling tests for the same profile.

3.6.6 Test repeatability


For each type of profile five identical specimens have been produced to check the
repeatability of tests. Compression test results for five specimens of one batch of the
cold-formed profiles is given below as an example. Figure 3.29 represents the load-
displacement curves for these five specimens. Similar structural behaviour (stiffness
and ultimate loads) is present for the five identical specimens. Values of the ultimate
load for one batch of profiles obtained during experiments can be found in Table 3.6.
The difference between the values of the ultimate load is almost negligible.

Table 3.6: Repeatability of tests.

Specimen Nu [kN] Buckling mode


2Z200F10B-1 72.8 Local+Distortional
2Z200F10B-2 73.3 Local+Distortional
2Z200F10B-3 72.8 Local+Distortional
2Z200F10B-4 73.1 Local+Distortional
2Z200F10B-5 73.0 Local+Distortional

46
Figure 3.29: Load-displacement curves for the batch of specimens 2Z200F10B.

3.6.7 Results of the material study


The influence of the material was also examined (Figure 3.30). Six Z-shaped speci-
mens with three different web heights and two different materials are considered. It
can be seen that specimens made of 5083-H111 aluminium alloy resulted in a lower
structural capacity compared to the identical specimens made of 6082-T6 aluminium
alloy. This fact can be explained by the difference in material characteristics that was
observed in the tensile tests. Thus, experiments show a material strength influence
on the structural behaviour, while the influence of the shape of the material stress-
strain curve (steepness) is not clearly distinguished.

Figure 3.30: Compression test results: material influence.

47
3.7 Conclusions and summary of observations
In general, it can be concluded that the results of the compression tests satisfy the
expectations. Based on observations and results of the experimental program, the
following conclusions can be drawn:

• The resulting buckling modes of various cross-sections selected for the exper-
imental research agree well with the expected buckling behaviour according
to the finite strip method, executed using the CUFSM-program.

• Observations during compression tests show that there were many compressed
specimens that failed at the supports. This can be explained by concentration
of forces close to the edges of the specimen due to the occurrence of a gap. For
some specimens visual gaps between the specimen edge and the support were
observed.

• It was shown that in a batch of identical specimens identical behaviour oc-


curs, if support conditions are the same. Values of the ultimate load for the
same type of sections mainly differ within 4%. Thus, it can be concluded that
experiments are well reproducible.

• It was shown to be a good choice to select Z-shaped profiles with additional


flange stiffeners to observe local and distortional buckling.

• The investigation of extruded C-shaped profiles aimed to study the influence


of a gradual increase of the complexity of geometry on the buckling behaviour.
It was shown that additional elements of the cross-sectional geometry affected
the structural capacity and behaviour. However, it is important to pay atten-
tion to the dimensions of additional plate elements and also to the type of
additional elements (internals or outstands).

• The influence of the material was also examined. It was observed that material
strength influences the structural behaviour. However, the influence of the
material non-linear behaviour was not clearly distinguished.

• Tests were carried out to provide data for local and distortional buckling phe-
nomena of aluminium structural elements with various cross-sectional shapes.
The experimental data will contribute to the validation of a finite element
model discussed in the following Chapter 4.

48
Chapter 4
Finite Element Model Validation

Abstract

The aim of the current chapter is to present the validation of a finite element (FE) model.
The FE-model validation is performed on the results of the experimental program de-
scribed in Chapter 3. Thus, the FE-model is built based on the measured geometry, ini-
tial imperfections and material characteristics of extruded and cold-formed specimens.
The resulting buckling modes, values of the ultimate load, stiffness and post-buckling
behaviour of extruded and cold-formed specimens according to numerical analyses and
tests are compared. The insight of the current chapter is used in further development
and validation of the prediction model for local and distortional buckling of aluminium
members with various cross-sectional shapes.

4.1 Scope of the chapter

A
n experimental investigation has been executed on the selected cross-sectional
shapes at Eindhoven University of Technology (Chapter 3). Development of a
numerical model is necessary to be able to perform an extensive parameter study
for arbitrary shapes. The advantages of the numerical model over the costly tests
are obvious. However, the numerical model should provide results with the same
validity as tests.
The finite element method (FEM) is used to simulate the compression tests for
extruded and cold-formed specimens. The details and characteristics of the FEM-
analyses are given in this chapter. The set-up of the FE-model contains applied ge-
ometry, imperfections and material characteristics. A sensitivity analysis is carried
out to investigate the influence of the FE-model input parameters on the final result.
Validation of the FE-model is realized by comparison of the numerical results with
the experiments.
This chapter is a summary of the finite element calculation results which have
been described in the reports: Kutanova et al. [2008b] and [2008c]. The chapter re-
sults in a validated finite element model, which is used for a further parameter study
of aluminium structural elements with various cross-sectional shapes.

4.2 Numerical program


The numerical program reproduces nine compression tests on specimens with L-, Z-
and C-shaped extruded profiles and eight compression tests on specimens with Z-

49
shaped cold-formed profiles. Specifications of corresponding cross-sectional shapes
are given in Chapter 3 in Tables 3.1-3.3. Measured dimensions and imperfections are
implemented for the FE-simulations to repeat numerically the tests with maximum
accuracy.
The dimensions of the FE simulations are defined based on actually measured
characteristics (see section 3.2). The plate thicknesses are identical to the experimen-
tal values, but the plate widths are taken as heart-to-heart values (see Figure 3.6).
For cold-formed specimens radiuses between the plate elements of the section cor-
respond to the measured radiuses. Specimen dimensions used in the FE-analyses
are given in Appendix B according to Tables B.1 and B.2.

4.2.1 Applied imperfections


Imperfection measurements have been done for all tested specimens. Basic purpose
of the imperfection measurements was determination of the deformation pattern
and the magnitude of the amplitude of the initial imperfections. These data can be
applied in the finite element (FE) analyses.
The measured deformation pattern is implemented for one specimen (3C80E7)
only to perform an imperfection sensitivity analysis. For all other specimens the
real imperfection pattern is not applied. Instead, it is assumed that the imperfection
pattern is identical to that of the first Euler buckling mode. The accuracy of this
approach is verified in section 4.5. Thus, in all numerical calculations the deforma-
tion pattern of the first Euler buckling mode is used and the calculated value of the
imperfection amplitude based on imperfection measurements is applied.
At the imperfection measurements two types of imperfections were considered:
internal and overall, corresponding to local and distortional buckling (see section 3.3).
Based on results of imperfection measurements, the values of the imperfection am-
plitude were calculated. The calculation procedure is described in detail in sec-
tion 3.3.4. The resulting values of imperfection amplitude for each side of the profile
are given in Appendix A in Tables A.3 and A.4. The average value is applied in the
finite element analysis. Tables B.3 and B.4 give the magnitude of the imperfection
amplitude for considered specimens.

4.2.2 Material input


The material input for the FE-analyses is described using a Von Mises yield crite-
rion and a work-hardening stress-strain relation as recommended in Hendriks and
Wolters [2007]). This approach requires the modulus of elasticity E, a proof stress
(f0.2 ) and a hardening diagram for the plastic strains and stresses (σpl and εpl ).
The tensile tests result in relations between the load and elongation of the tested
specimen (see section 3.4.4). These relations are transformed into stress-strain curves.
An example of the resulting stress-strain diagram for tensile coupon 6P2II1 of 6082-
T6 aluminium alloy is represented in Figure 4.1. The values of true stresses and true
strains were calculated and their relation is drawn in the graph. The hardening di-

50
agram is described as a function of the modulus of elasticity E and the elastic and
plastic strains (εel and εpl ):
σ = Eεel + εpl .
For the FE-model a hardening diagram for the plastic true strains and true stresses
is applied: σ(true) = εpl (true) . It is common practice to represent the results of a
tensile test as an equation. Therefore, a curve for the plastic strains based on the
Ramberg-Osgood relationship (see Eq. 2.13 and Eq. 2.14) is shown in the diagram:
µ ¶n
σpl ln 2
εpl = 0.002 , n= ³ ´
f0.2 ln ff0.2
0.1

Figure 4.1: Material stress-strain curves.

4.3 FE-model
4.3.1 Element types
The test specimens are simulated using so-called CQ40S eight-node curved shell el-
ements. The curved shell elements in DIANA (Hendriks and Wolters [2007]) are
based on iso-parametric solid approach by including the following shell theory hy-
pothesis: normals remain straight, but not necessarily normal to the reference sur-
face; the normal stress in the normal direction is equal to zero. Each node has
three translations as well as two rotations (the rotation in normal direction is not
included); a total of 40 degrees of freedom exist.
A quadratic interpolation and Gauss integration scheme is applied for each ele-
ment. Seven integration points have been applied over the plate thickness, in each

51
of the 2*2 in-plane shell integration points. Note that the calculation time increases
linearly with the number of thickness integration points. However, the accuracy
of the described stress-pattern improves significantly when a number of seven is
taken instead of default number of three. Existing work by Mennink [1999] (with 3,
7 and 9 thickness integration points) shows that the number of seven is sufficiently
accurate.

4.3.2 Mesh
The model and mesh of the specimen are generated using the actual dimensions of
the test specimens. Each side of the specimen is modeled as a rectangular plate.
Each plate is divided into 32 elements over the specimen’s height and 8 over its
width.

Figure 4.2: Mesh applied for the FE-model.

It was defined that this mesh density is sufficiently accurate for the description
of the elastic range, ultimate load and the post-buckling strength. In section 4.5 the
influence of the mesh density will be considered.

4.3.3 Supports, loading and modelled gap


The model is supported at both the ”bottom” and the ”top” edges of the specimen.
All translations and rotations at the bottom and the top edges are restricted. Thus,
the column is modeled to resemble a two-sided fixed column. The applied load is
performed as a uniform axial displacement. During the experiment, the support
conditions allowed the compressive forces to be transmitted from the test set-up to
the specimens, but transmission of tensile forces were not allowed.
The compression tests for extruded specimens were not perfect and loading dis-
tribution for all the sides of the cross-section was not achieved. The edges of the

52
specimens were not exactly flat and parallel to each other, which was discussed
in Chapter 3. As a result, a gap between the test set-up and specimens edges has
been observed (as shown in Figure 3.22). Therefore, the gap is included into the FE-
model. It should be noted that cold-formed specimens have been more accurately
prepared for the tests than extruded specimens. Hence, the gap is not introduced
for the cold-formed specimens.
The size of the gap was not measured during experiments. However, based on
visual observations it is assumed that the value of the gap was approximately 1 mm,
but the gap size is different for different specimens.
The gap is modeled with spring elements. Spring elements have been added on
the top and the bottom edges of the specimen. SP2TR spring elements have been
used. This SP2TR element is a two-node translation spring. Spring elements are
connected to the CQ40S curved shell elements. Figure 4.3 shows the properties of
the spring elements: high stiffness in compression (k = 7000000 [N/mm]) and al-
most zero stiffness in tension (k = 0.07 [N/mm]). The nonlinear elasticity model for
spring elements applied on three sides of the profile specifies a high stiffness in com-
pression (see left figure 4.3(a) ). The nonlinear elasticity model for spring elements
applied for the side with a gap gradually specifies a low stiffness in compression
until the value of the gap is reached and high stiffness when the gap is closed (see
right figure 4.3(b) ).

(a) Without a gap (b) With a gap

Figure 4.3: Material models for applied springs.

The gap is gradually increased for the mesh elements of the side non-parallel to
the support (see Figure 4.4). It means that only the last node has a gap equal to the
size of the gap and the first mesh node of the side has a gap equal to zero. Hence,
the gap is imitating the real situation observed during experiments. For each node
of the side with a gap, stiffness properties shown of Figure 4.3(b) are applied, where
the value of the gap is varied from zero to the maximum size of the gap. The load is
applied on the spring elements of the top of the profile as a uniform displacement,
similar to tests.

53
Figure 4.4: Gap during the test and modelled gap.

4.3.4 Analyses types


Numerical calculations are executed using the general-purpose DIANA program.
Simulation of compression tests is performed by a geometrical and physical non-
linear analysis, where physical non-linearity means a nonlinear material model. The
nonlinear analysis uses the output of the Euler analysis: the deformation pattern of
the first Euler buckling mode is used as an imperfection pattern. A Regular Newton-
Raphson integration scheme has been chosen: the stiffness matrix is evaluated every
iteration. More information about the Regular Newton-Raphson method is found
in the DIANA manual (see Hendriks and Wolters [2007]).
The load is applied as a uniform axial shortening of the specimen. Therefore,
correct specification of the load step size is highly important to get good accuracy.
The selected load step size of 0.01 mm is small enough to provide necessary accuracy
and to result in a smooth load-displacement diagram.

4.4 Results
4.4.1 Extruded subprogram
This section contains the results of the numerical simulations for the compression
tests of the extruded experimental subprogram. The deformed specimens according
to experiments were shown in Figure 3.25 in Chapter 3. The FE-model with mea-
sured data input and a modelled gap has been applied (see section 4.3.1). The result-
ing deformed specimens according to the FE-analyses can be seen in Figure 4.5. The
designation for extruded specimens with various shapes and obtained failure modes
are indicated in Figure 4.5 as well. Note, that L=local, D=distortional, F=flexural,
and T=torsional buckling modes. Sizes of the gap applied for extruded specimens
are indicated in figure caption.
For all extruded specimens a gap between the specimen and the support has
been modelled using spring elements. Due to the reason that the size of the gap was
not measured, the gap size has been determined by comparison with tests trying to

54
Figure 4.5: FEM-results: extruded specimens. From left to right: 2Z40E1 (gap 0.36 mm),
3Z35E2 (gap 0.1 mm), 3L80E3 (gap 1.12 mm), 3C80E4 (gap 0.36 mm), 3C80E5 (gap 0.48 mm),
3C80E6 (gap 1.68 mm), 3C80E7 (gap 0.2 mm), 3C80E8 (gap 0.1 mm), 3C80E9 (gap 0.1 mm).

Figure 4.6: Comparison of FEM results with and without gap modelling (3C80E7).

find the best fit. According to FEM the size of the gap varied from 0.1 mm to 1.68
mm. An example of the comparison of results for the FE-model with and without
gap modelling is shown in Figure 4.6 for the specimen 3C80E7. This diagram rep-
resents experimental and FEM curves, which describe a relationship between the
reaction force (N ) in kN and the applied axial shortening (u) in mm.

55
From the results it is clear that the deviation in stiffness between the FE-model
and test is caused by the occurrence of a gap, in this case the gap value is 0.2 mm.
This statement is also supported by the graph in Figure 4.7(a) where experimental
and numerical curves are compared to the tensile test material curve. It can be seen
that appearance of a gap results in a lower stiffness. The shape of the experimental
and numerical N − u curves compare very well when the gap is included into the
FE-model. The ultimate loads show excellent agreement and the stiffness does as
well.

(a) Load-displacement curve. (b) Load-deflection curve.

Figure 4.7: Comparison of experimental and numerical curves (specimen 3C80E7).

The experiment failed due to local buckling. The experimental curve shows elas-
tic behaviour up to a peak load where failure occurred. The load-displacement curve
does not represent any buckling limits, because local buckling occurs in the inelas-
tic range. Buckling happens in the inelastic range if the critical Euler buckling load
is greater than the proportional limit of the material (σcr > fp ). It is considered
that proportional limit of the material indicates inelasticity. This is illustrated in the
load-deflection diagram of Figure 4.7(b) . The load-deflection (N − w) diagram cor-
responds to the relation between the amplitude of the out-of-plane deformations of
the FE-analysis w and the axial load N . Note that w represents the absolute value
of the maximum deflection of one side of the specimen. While w has not been mea-
sured in the experiments, only the FE results are presented. The diagram also pro-
vides (as straight horizontal lines) the value of the ultimate load measured during
the experiment (Nu ), the proportional limit of material according to tensile test and
the initial Euler buckling load (Ncr ) obtained from Euler analysis results. The N − w
curve shows a short perturbation at the beginning (N ≈ 18 [kN]) due to the gap
and further gradual increase of the out-of-plane deformations up to the load, which
relates to the proportional material limit (Np = 90 [kN]). Inelasticity results in a
substantial reduction of stiffness. It is known that inelastic buckling exhibits sud-
den failure (see Mennink [1999]). Thus, the peak load is reached at (Nu = 100 [kN]),
where inelastic local buckling occurs. After the peak load the load-bearing capacity
decreases.

56
Figure 4.8: Comparison of tangential stiffness of the experiment (LVDT’s and strain gauges),
tensile test and numerical analysis (3C80E7).

Figure 4.8 illustrates the tangential stiffnesses (ET ) of the tensile test, the com-
pression test (LVDT’s, strain gauges) and FE-analysis. It can be seen that there is
no agreement in represented curves. The stiffness of the compression test is much
lower compared to the tensile test. One of the main reasons of such a difference is
the occurrence of a gap. Numerical stiffness curve is close to the compression test
results according to the LVDT’s. It should be noted that the material characteris-
tics of tensile test were applied for the FE-model. A gap between one side and the
support has been introduced for the model, which caused the reduction in stiffness.
Identical to specimen 3C80E7, comparison analyses were performed for all con-
sidered extruded specimens. Detailed validation analyses of the FE-model for ex-
truded specimens can be found in Kutanova et al. [2008b]. In total nine different
profiles are evaluated, important characteristic values of both the experiments and
FE-analyses are given in Table 4.1. The table presents the following values of both
the experiments and FE-analyses: the compression stiffness (Ec ), the critical and
failure loads (Ncr and Nu ) and their corresponding modes, as well as the plastic
load carrying capacity of the specimen (N0.2 = f0.2 A). The most right column of
the table compares the critical and ultimate loads according to test and FE-model.
If critical local buckling or distortional buckling happened in the elastic range and,
therefore, can be defined from experimental load-displacement graph, the value of
Ncr;F E /Ncr;exp is given in the table.

57
Table 4.1: Comparison of experimental and FE-results for extruded profiles.

Experiment Finite Element Model FEM/Exp


Buckling Failure Euler Failure Cr. Fail.
Spec. Ec Md Ncr Md Nu Ec Md Ncr Md Nu N0.2
2Z40E1 50.3 - - LF 28 55.1 L 78.4 LF 28 32.1 - 1.00
3Z35E2 55.5 - - F 46.2 58.6 L 135 F 46.6 50.1 - 1.01
3L80E3 51.7 L 18.3 T 41.9 53.9 L 20.2 T 45.5 98.6 1.1 1.08
3C80E4 53.5 L 45.1 LF 54.6 56.1 L 46.1 LF 56.5 109.3 1.02 1.03
3C80E5 55.8 L 55.5 L 80.1 55.8 L 54.6 L 81.1 120.7 0.98 1.01
3C80E6 53.1 - - LF 111.5 54.3 L 187.5 L 111.6 136 - 1.00
3C80E7 53.2 - - LD 99.9 58.5 LD 124.9 LD 99.9 130.7 - 1.00
3C80E8 54.7 - - DF 72.9 58.4 D 87.1 DF 72.9 134.8 - 1.00
3C80E9 54.2 - - L 132.3 57.0 L 188.7 L 132.3 142.3 - 1.00
Md=Mode/ L=local buckling/ D=Distortional buckling / F=Flexural buckling / T=torsional buckling /
FT=flexural-torsional buckling

It is concluded that the finite element model provides an accurate result. The
experimental values of the ultimate load could be approached within less than 1%
accuracy.

4.4.2 Cold-formed subprogram


This section contains the results of the numerical simulations for compression tests
of the cold-formed experimental subprogram. The resulting deformed specimens
according to the experiments were shown in Figure 3.26 in Chapter 3. The FE-
model with measured input data without inclusion of the gap has been used (see
section 4.3.1). The resulting deformed specimens according to the finite element
analyses can be seen in Figure 4.9.

Figure 4.9: FEM-results: extruded specimens. From left to right: 2Z200F10A, 2Z200F10B,
2Z100F11A, 2Z100F11B, 2Z50F12A, 2Z50F12B, 2.5Z50F13A, 2.5Z50F13B.

58
It should be mentioned that for all cold-formed specimens, the first Euler mode is
used as an imperfection pattern and the calculated value of the measured imperfec-
tion amplitude is applied to the model. The designation for cold-formed specimens
with various shapes and obtained failure modes are indicated in the figure as well.
Note, that L=local, D=distortional, and F=flexural buckling modes.
Validating the FE-model based on the results of the cold-formed experimental
subprogram has been executed similar to the extruded subprogram. Detailed val-
idation analysis of the FE-model for cold-formed specimens can be found in Ku-
tanova et al. [2008c]. An example of the comparison of the structural behaviour ac-
cording to experimental and numerical results for the specimen 2Z100F11A is given
in Figure 4.10.

Figure 4.10: Comparison of test and FEM results for the specimen 2Z100F11A.

It can be seen that the FE-model gives a conservative prediction for the ultimate
load-bearing capacity. However, the difference in the ultimate load is not extreme
(< 5%) and can be explained by the applied imperfections. Firstly, first Euler mode
is used as the deformation pattern and not the real imperfection pattern, which leads
to more conservative results. Secondly, the measured value of the imperfections
amplitude might be overestimated, because measurement inaccuracies could have
occurred due to the influence of the rounded corners of the specimens. Hence, the
applied value of imperfection’s amplitude might be higher than in reality, resulting
in a lower ultimate value for the FE-model. This fact explains a difference between
the ultimate loads according to experimental and FE-model results. Application
of the lower value of imperfection’s amplitude will result in higher value of the
ultimate load.

59
It was discovered that the difference in stiffness between the FE-model results
and experimental results can be explained by the possible inequality in support con-
ditions. According to experimental research of B.Young [2003], additional fixed-end
bearings have to be applied at the column edges during the test, especially in case of
distortional buckling test. The fixed-end bearings are able to restrain axis rotations
as well as twist rotations and warping. In the current investigation no additional
constraints have been applied for the distortional buckling test. Thus, inaccuracies
might occur in the test due to support conditions.
Seven different cold-formed profiles have been evaluated. Results are combined
in one table to be able to draw conclusions. The resulting characteristic values of
both the experiments and FE-analyses are shown in Table 4.2.

Table 4.2: Comparison of experimental and FE-results for cold-formed profiles.


Experiment Finite Element Model FEM/Exp
Buckling Failure Euler Failure Cr. Fail.
Spec. Ec Md Ncr Md Nu Ec Md Ncr Md Nu N0.2
2Z200F10A 55 L 30 LD 101 54 L 32 LD 106 199 0.99 1.05
2Z200F10B 50 L 30 LD 73 54 L 30 LD 69 115 1.02 0.95
2Z100F11A 61 L 75 L 108 64 L 75 L 101 147 1.00 0.94
2Z100F11B 60 - - L 71 69 L 71 L 68 83 - 0.95
2Z50F12A 55 - - L 111 57 L 110 L 99 134 - 0.90
2Z50F12B 61 - - L 69 65 L 91 L 65 68 - 0.94
2.5Z50F13A 52 - - DF 71 60 D 136 DF 72 97 - 1.00
2.5Z50F13B 51 - - DF 49 61 D 123 DF 46 54 - 0.94

Md=Mode/ L=local buckling/ D=Distortional buckling / F=flexural buckling

The table presents the following values of both the experiments and FE-analyses:
the compression stiffness (Ec ), the critical and failure loads (Ncr and Nu ) and their
corresponding modes, as well as the plastic load carrying capacity of the specimen
(N0.2 = f0.2 A). The most right column of the table compares the critical and ultimate
loads according to test and FE-model. If critical local buckling or distortional buck-
ling happened in the elastic range and, therefore, can be defined from experimental
load-displacement graph, the value of Ncr;F E /Ncr;exp is given in the table.
The FE-model is considered to provide an accurate result and deviations from
experiments are explained by uncertainties in imperfection and support conditions
as explained before.

4.5 Influence of the modelling parameters


4.5.1 General
Hereafter the influence of several parameters on the results of the finite element
calculations are studied. This sensitivity study is performed for two extruded C-
shaped specimens (3C80E6 and 3C80E7) and one cold-formed Z-shaped specimen
(2Z200F10A). The experimental results for these three specimens are presented in

60
Chapter 3. The FE-model described in section 4.3.1 is varied applying different pa-
rameters. The following aspects are considered:

• Mesh density

• Material anisotropy

• Material model with enhanced corner properties

• Imperfections

4.5.2 Mesh density


Three meshes are used to check the influence of the mesh density on the results of
the FE-calculations for specimen 3C80E6.
A - 832 elements (medium)

B - 280 elements (wide-meshed)

C - 1536 elements (fine)


Medium meshing (case A) is considered in the general FE-model (see section 4.3.1).
In case of medium mesh refinement the regarded C-shaped specimen is consisted
of eight elements in the web, eight elements in each flanges and one element in the
stiffener. The number of elements in longitudinal direction is 32. In case of wide
or fine mesh the number of elements is relatively decreased or increased. Different
meshes are shown in Figure 4.11.

Figure 4.11: Mesh density.

The resulting load-displacement curves for the three different meshes and the
experimental curve are represented in Figure 4.12. It can be seen that there is a sig-
nificant difference in structural behaviour for the wide-meshed model comparing to

61
the medium and fine meshes. On comparison with the experiment it is concluded
that the wide-meshed model is not accurate enough. Regarding the denser meshes
the results are accurate. Varying mesh density, the load-carrying capacity was af-
fected only a little; the accuracy refers mainly to the displacements. Variation from
analysis A: for B - 0.06%, for C - 0.02%. It can be summarized that mesh A is suffi-
ciently accurate for the prediction of the structural behaviour till the failure occurs.

Figure 4.12: Load-displacement diagram varying mesh density.

4.5.3 Material anisotropy


In section 4.2 it was mentioned that the Von Mises yield criterion with isotropic
hardening is used. Hopperstad [1993] discovered that aluminium alloys demon-
strate anisotropic yield. Hence, the Von Mises yield criterion might provide inac-
curate results. Other work by Moen [1999] considered anisotropy of the material,
where also parameters defining the anisotropic surface were given. In his research,
Moen has shown that anisotropy substantially affected the post-buckling capacity,
but insignificantly affected the ultimate load-bearing capacity. According to exper-
imental results of Kutanova et al. [2008a], there was no substantial difference be-
tween material characteristics measured in two directions. The same conclusion has
been drawn by Frey and Mazzolani [1977] for extruded aluminium profiles. Ac-
cording to this work, the alterations of the value of the 0.2% proof stress in differ-
ent directions were low and did not significantly affect the load-bearing capacity of
compressed members. Thus, it can be concluded that the influence of the material
anisotropy is not significant for the current research.

62
The Bauschinger effect is not considered in the current investigation. The de-
scription of the Bauschinger effect can be found in Mazzolani [1995]. During man-
ufacturing, the extruded material has been stretched to form the required profile.
Due to the Bauschinger effect, the compressive strength is lower than the tensile
strength. The influence of Bauschinger effect has been checked for extruded speci-
mens and reported in Kutanova et al. [2008a]. Compression tests for a stub column
have been performed to determine the material characteristics in compression. Ma-
terial characteristics based on tensile and compression tests were evaluated. There
was no substantial difference between the results of tensile and compression tests.
Therefore, it has been concluded that there is no influence of the Bauschinger effect.
In section 3.4.2, characteristics are given for cold-formed tensile test coupons,
which were cut in rolling and perpendicular to rolling directions. It was decided
to check how changes in the material properties affected the load carrying capacity
and also to see if there is an influence of the fabrication process. Therefore, material
curves for 6082-T6 aluminium alloy obtained for tensile coupons in rolling and non-
rolling directions (6P2II and 6P2L in Table A.8) are applied in the FE-model. The
Ramberg-Osgood approach is also considered. The resulting material characteristics
are presented in Figure 4.13.

Figure 4.13: Tensile test results for two coupons in rolling and ⊥ to rolling directions.

The obtained load-displacement diagrams are presented in Figure 4.14. There


are four graphs for the corresponding material characteristics of Figure 4.13. The in-
fluence of the tensile test results for coupons in rolling and perpendicular to rolling
directions can be observed and also results of the application of the Ramberg-Osgood
approach based on the measured material characteristics.

63
It can be seen that there is no significant difference in structural behaviour when
slightly different material characteristics in two directions are compared. The ob-
tained variation in the values of the ultimate load applying material properties in
rolling and perpendicular to rolling directions is 2%. Application of the Ramberg-
Osgood relationship based on the measured values results in practically the same
behaviour as when the actual material curves are introduced. It can be concluded
that the Ramberg-Osgood approach provides sufficiently accurate results.

Figure 4.14: Load-displacement diagram varying material characteristics.

4.5.4 Material model with enhanced corner properties


Due to the fact that all cold-formed profiles have rounded corners, it was proposed
to apply different material characteristics for the corners. It should be noted that ma-
terial characteristics at the rounded corners were not measured for the experimental
program. However, for the sensitivity study it is assumed that due to cold-forming,
the properties at the corners might be 10% higher than the measured stress-strain re-
lation. Thus, the basic material model was modified to include the enhanced corner
properties, which were applied strictly to the corner geometry of the section.
Comparison of the ”general” FE-model (uniform material properties), the FE-
model with application of the enhanced material properties for rounded corners and
experimental results for aluminium specimen 2Z200F10A are shown in Figure 4.15.
It can be concluded that there is no influence on the results of the ultimate load
bearing capacity, when slightly enhanced material characteristics are applied for the
rounded corners.

64
Figure 4.15: Comparison FE-model and test results, applying enhanced material properties
for the rounded corners.

4.5.5 Imperfection sensitivity


The sensitivity of the FE-analyses as a result of the imperfection input parameters is
discussed below. The following imperfection parameters are varied:

• imperfection pattern: a) real (measured); b) first Euler buckling mode;


c) Euler buckling shape different than the first mode.

• value of the amplitude of imperfections.

In section 4.2 it was mentioned that for the numerical calculations the first Euler
buckling mode is used as the deformation pattern and not the real imperfection pat-
tern. Based on imperfection measurement results the realistic imperfection pattern
has been defined for some of the specimens (Figure 3.15). Therefore, it is decided
to compare the results of application of the realistic deformation pattern with the
results of using the first Euler buckling pattern.
First, a short overview of the Euler buckling analysis for the investigated speci-
men is considered. Figure 4.16 illustrates the FE-results of the Euler buckling anal-
ysis for the specimen 3C80E7. Four buckling modes are represented. It is important
to pay attention to the buckling shapes and Euler critical loads for each of the rep-
resented modes, because these results of Euler buckling analysis are used for the
imperfection sensitivity study. Applying different Euler buckling modes as an im-
perfection pattern for non-linear analysis, the calculated value of the imperfections
amplitude is used (see Table B.3).

65
Figure 4.16: Euler buckling analysis results for specimen 3C80E7.

The sensitivity of the results to the value of imperfection amplitude is investi-


gated using different magnitudes of imperfections while the imperfection pattern is
the first Euler mode. The calculated value of the imperfection amplitude is applied.
Due to the fact that the sign of the imperfection amplitude is not known, positive
and negative values are compared. Hence, the calculated values according to mea-
surements are:

A e0 = 0.15 [mm]

B e0 = −0.15 [mm]

Arbitrary scattered values of the imperfection amplitude are also applied to the
FE analyses to better understand the influence of the magnitude of the imperfection
amplitude on the ultimate load. However, these arbitrary values relate to the plate
width value. The selected values are:

C (pl.width/10000=80/10000): e0 = 0.008 [mm]

D (pl.width/1000=80/1000): e0 = 0.08 [mm]

E (pl.width/100=80/100): e0 = 0.8 [mm]

Table 4.3: Sensitivity analysis results for the specimen 3C80E7.

Imperfection Pattern Value of imperf. e0 [mm] Fu [kN]


1st Euler mode 0.15 99.9
1st Euler mode -0.15 99.7
1st Euler mode 0.008 100.0
1st Euler mode 0.08 100.0
1st Euler mode 0.8 97.4
2nd Euler mode 0.15 99.9
3rd Euler mode 0.15 100.2
4th Euler mode 0.15 100.3
Real pattern 101.4

66
It is known that applying a higher value of the imperfection amplitude results in
a lower ultimate load. It can be noticed from Table 4.3 that all the results correspond
to this statement. Varying the sign of the magnitude of imperfections amplitude
does not give a substantial effect on the results of the FE-analysis.
Based on the results represented in Table 4.3, it can be noticed that the intro-
duction of the real measured imperfection pattern provides the highest value of the
ultimate load. Hence, it can be concluded that the application of the first Euler mode
as an imperfection pattern is safe.

4.6 Conclusions
A numerical model is developed and validated for cross-sectional instability of alu-
minium specimens using the results of the experiments as presented in Chapter 3.
The following is concluded:

• Compression tests for extruded and cold-formed specimens can be accurately


simulated using the FE-model.

• For extruded specimens, the FE-model with a modelled gap between the spec-
imen and supports using spring elements, provides a sufficiently accurate re-
sult compared to the experiment.

• For the cold-formed specimens, especially in case of distortional buckling, ad-


ditional end restraints have to be used in a test set-up to enable a proper simu-
lation. This conclusion allows to assume that the FE-model is accurate enough
and deviations from experiments can be explained by uncertainties in imper-
fections and support conditions.

• The FE-model is validated based on results of the experimental program. The


resulting deformed specimens according to the FE-model and tests are compa-
rable. The experimental structural behaviour (ultimate load, stiffness) could
be approached with good accuracy for the extruded sections and with accept-
able accuracy for the cold-formed specimens.

• For the FE-model, the first Euler buckling mode is used as an imperfection
pattern and the amplitude of the applied imperfection mode is based on the
measured average value of initial out-of-planeness. The average value of the
imperfection amplitude might be greater than in reality due to measuring inac-
curacies. This explains the difference between the experimental and FE-results
with respect to ultimate load. Application of a lower value of the imperfection
amplitude will result in a higher value of the ultimate load.

• It was shown that application of the first Euler buckling mode as an imperfec-
tion pattern is safe.

• Anisotropy can be ignored for compressed members.

67
• Enhanced corner material properties of approximately 10% greater than the
measured tensile test characteristics do not have influence on the ultimate
strength.

• A significant amount of information has been gained during the numerical


study. The finite element model is a useful tool for prediction of the structural
behaviour of uniformly compressed aluminium members with various cross-
sectional shapes (L-, C- and Z-shapes).

• Results of this chapter contribute to the further investigation of the distor-


tional buckling phenomenon and to the development of a prediction model
for distortional buckling.

68
Chapter 5
Assessment of the existing model with respect
to distortional buckling

Abstract

In the current chapter, a prediction model for local buckling earlier developed by Men-
nink [2002] is applied for the investigated specimens. First, Mennink’s model for lo-
cal buckling prediction is described with some additional improvements. Inclusions for
Mennink’s model to adjust the model for distortional buckling are proposed. Hence, eval-
uation of the existing prediction model for local buckling is presented and, furthermore,
an extended model for distortional buckling is proposed.

5.1 Scope of the chapter

T his research is focused on the distortional buckling phenomenon, therefore


specimens subjected to distortional buckling mode are highlighted. Mennink’s
model has been developed for local buckling prediction only. The prediction model
is illustrated on the results of one Z-shaped specimen as an example. It is assumed
that the model can be extended for the prediction of distortional buckling behaviour.
Existing knowledge for distortional buckling in literature together with experimen-
tal and numerical observations of the current research allow to suggest proposals
for Mennink’s model to be extended for the distortional buckling case.
The extended model is applied for the cross-sectional shapes of the tested spec-
imens to predict the structural resistance due to local and distortional buckling.
Nominal values of dimensions, imperfections and material characteristics are con-
sidered for specimens with L-shaped, Z-shaped and C-shaped profiles. The struc-
tural behaviour of the regarded specimens with nominal characteristics is numeri-
cally calculated using the validated finite element model (see Chapter 4). To eval-
uate the accuracy of the prediction model, the results of the extended model are
compared to the FEM results. This chapter results in conclusions, which are an im-
portant basis for the development of a prediction model for distortional buckling.

5.2 Specimen example for model evaluation


In the experiments, the Z-shaped specimen 2Z200F10A failed due to local-distortional
buckling interaction (see Figure 3.24). Hence, this cross-sectional shape is an appro-
priate example for illustrating Mennink’s model and its possible extension. A speci-

69
men, identical to specimen 2Z200F10A, but with nominal heart-to-heart dimensions
is considered (Figure 5.1).

Figure 5.1: Z-shaped specimen dimensions.

Finite strip method (FSM) results for the considered specimen correspond to the
results of specimen 2Z200F10A that were represented in Figure 3.23. According to
CUFSM, two critical points are defined with the following characteristics:
2
Local buckling: Lcr; l = 160[mm]; σcr; l = 36[N/mm ]; Fcr; l = 26[kN].
2
Distortional buckling: Lcr; d = 700[mm]; σcr; d = 74[N/mm ]; Fcr; d = 54[kN].

FSM indicates that local buckling is initiated first and distortional buckling might
occur for higher stresses. However, the length of the considered specimen is L = 700
[mm], which is equal to the critical length for distortional buckling Lcr; d . Thus,
the length of the specimen might be insufficiently long for distortional buckling
development.
The finite element nonlinear analysis normally results in the load-displacement
(F − u) relationship which is transferred into the stress-strain (σ − ε) relationship.
It is common practice to present the relations between these parameters in a non-
dimensional form. Therefore, the initial buckling of the cross-section (critical stress
for the local buckling mode and modulus of elasticity) is used to convert parameters
in dimensionless values (Eq. 5.1).

σ F Eε Eu
σ∗ = = ε∗ = = , (5.1)
σcr; l A σcr; l σcr; l σcr; l L

where the modulus of elasticity E = 70000[N/mm2 ] and σcr; l is the value of the
critical stress for local buckling.
To get a clear understanding of interaction between plate elements, the FE-model
result for the Z-shaped specimen is shown in Figure 5.2. There are three curves: one
corresponds to the flange/lip assembly, one to the web and one to the whole sec-
tion at half the specimen length. It should be noted that the two flange/lip groups
behave identically, only one result is presented.

70
Figure 5.2: Buckling behaviour of cross-sectional plates and critical limits.

Figure 5.3: Cross-sectional deformation at different load limits.

It can be seen that at the point of initial local buckling (ε∗ = 1), the web and the
whole section starts to buckle (A). As a result, the post-buckling stiffnesses of the
web and the section reduce. No stiffness reduction can be noticed for the flange/lip
assembly until the proportional limit εp . The flange/lip assembly behaves elastically

71
until the proportional limit of the material is reached. After the proportional limit
the stiffness of the outstanding elements reduces and the post-buckling strength due
to local buckling of the web decreases. At the value of the ultimate load ε∗ = 5.5, the
dimensionless stress-strain curves shows the occurrence of mode jumping (point B
in Figure 5.2). Mode jumping can be better demonstrated by the resulting deforma-
tion patterns. Figure 5.3 contains three deformation patterns at the corresponding
loading steps: initial local buckling load, ultimate load and post-failure load. Pure
local buckling happens at the critical load (A), while the presence of the distortional
mode can be noticed at the ultimate load (B). The post-failure deformation shape
indicates that the specimen failed due to local-distortional mode interaction (C).

5.3 Mennink’s model for local buckling


5.3.1 General
Hereafter the prediction model of Mennink is discussed [2002]. The input data for
the model are: Lcr; l - the critical length for local buckling initiation, σcr; l - the criti-
cal stress for local buckling initiation, cross-sectional dimensions and material prop-
erties. Values of the critical length Lcr; l and critical stresses σcr; l can be obtained
using Eigenvalue analyses by analytical or numerical (finite element, finite strip)
methods. For reasons of simplicity, the model is divided into two parts: elastic
buckling prediction with post-buckling strength and inelastic buckling prediction.
It is assumed that the proportional limit of the material, represented by εp and fp ,
distinguishes the elastic range from the inelastic part.
The concept of Mennink’s model is based on the interaction of cross-sectional
plate elements. A cross-section is divided into: plate elements that buckle (group I)
and plates that provide support (group II). Figure 5.2 illustrates these two groups
of plates: group I is the web, group II is the outstanding flange/lip assembly. Plate
elements with the critical buckling stresses lower than the critical buckling stress of
the whole cross-section belong to plate group I. It is assumed that after the propor-
tional limit is reached immediate failure occurs (see Figure 5.2). Concerning plate
group II, two situations are distinguished: plate group II behaves elastically until
the proportional limit of the material or buckling of the plate group II occurs. Buck-
ling of the plate group II is called secondary buckling. The post-buckling strength
after secondary buckling is limited to the proportional limit. Therefore, initial and
secondary buckling modes have to be defined for the cross-section.

5.3.2 Initial and secondary buckling determination


Traditionally the initial critical buckling stress is determined for each plate element
analytically, based on buckling of a rectangular plate uniformly compressed in one
direction. In section 2.2.1 the general theoretical expression for elastic critical stress
is given:

72
µ ¶2
Eπ 2 t
σcr = kcr , (2.6)
12(1 − ν 2 ) b
where kcr is the buckling coefficient, which depends on support conditions and the
length of the corresponding plate element.
Eigenvalue analysis results in the values of the critical stress (σcr ) and the critical
half-wavelength (Lcr ) indicating the initiation of buckling for a whole cross-section.
Thus, considering the cross-section itself and not only one plate element, theoretical
expression (Eq. 2.6) can be rewritten by substituting the length of the plate element
L by a critical length Lcr of the cross-section. In this case, the buckling coefficient kcr
depends on the rotational stiffness transferred from the connected plates (support
conditions) and the critical length of the cross-section (Lcr ). Thus, such an approach
takes plate interaction into account.

Figure 5.4: Initial and secondary buckling stress determination for local buckling (Mennink
[2002]).

Using the results of Figure 2.3 by Gerard and Becker [1957], Mennink specified
the area and limits of the buckling coefficient values for plates with various support
conditions (see Figure 5.4). Initial buckling behavior of plate elements of the cross-
section will result in one buckle over the critical length Lcr . Second buckling mode

73
is the upper bound for the area where the buckling coefficient value is expected to
be. Therefore, kcr for internal plates is positioned in the area between curves A and
C, while kcr for outstanding plates is positioned in between curves D and E.
Thus, it is possible to determine the buckling coefficient kcr;i and the critical
stress σcr;i for each plate element i in the cross-section. However, it is obvious that
all supported plates buckle at the initial critical stress of the cross-section σcr;sec .
Thus, all plate elements whose individual critical stresses σcr;i are less than the
critical stress of the cross-section σcr;sec belong to supported plate group I, corre-
spondingly, supporting plate group II consist of all remaining plates. The secondary
buckling stress is the lowest critical stress of plate group II.
In case of extreme values of kcr the current approach may provide an insufficient
result. Therefore, a safe upper bound for this approach is applied comparing with
the case A of Figure 5.4 that coincides with kcr = 7.0 at ϕ > 5.0.

5.3.3 Elastic local buckling prediction: (σcr < fp )


The cross-section will show post-buckling resistance if the elastic critical stress is less
than the proportional limit. To calculate the post-buckling resistance the following
steps should be taken:

1. Schematize the cross-section to nodes and plates.


Each individual plate i has a width (bi ) and a thickness (ti ). Plates are con-
nected at nodes. Nodes are positioned at the intersection of the heart-lines of
the individual plates.

2. Execute Eigenvalue analysis.


Using an analytical approach or, more appropriate for arbitrary cross-sections,
using finite strip or finite element programs, the values of the critical length
Lcr , the critical stress σcr and, respectively, the critical strain εcr = σEcr are
obtained.

3. Determine the elastic critical stress of each plate (σcr;i ).


π2 E
Determine: D = 12(1−ν 2 )
Lcr
Plate slenderness: ϕi = bi

1
For internals: kcr;i = ϕ2i
+ ϕ2i + 2 but : kcr;i ≤ 7
1
For outstands: kcr;i = 0.456 + ϕ2i
³ ´2
Critical stress: ti
σcr;i = kcr;i · D · bi

4. Determine plate groups

74
• Plate group I (pg1) - Supported plates
The first plate group consists of all plates that would buckle before the
critical buckling stress of the cross-section is reached (σcr;i ≤ σcr ).
• Plate group II (pg2) - Supporting plates
The second plate group consists of all plates that are not within plate
group I. The secondary buckling stress is equal to the lowest critical stress
of any plate j within plate group II:
σcr;2
σcr;2 = min(σcr;j ); εcr;2 =
E

• Determine the cross-sectional areas


For each plate i of plate group I and plate j of plate group II, the cross-
sectional area is determined from their plate width and thickness. Subse-
quently, this results in the following representations of the cross-sectional
areas: X
Plate group I Apg1 = b i ti
i
X
Plate group II Apg2 = bj tj
j
Cross-section A = Apg1 + Apg2

5. Resistance plate group I (σpm;pg1 )


σpm;pg1 (ε) = Eε for ε < εcr
If plate group I consists of at least one internal plate:
2
σpm;pg1 (ε) = 0.67Eεcr + 0.35Eε − 0.015E εεcr for εcr < ε < εp
If plate group I consist solely of outstands:
2
σpm;pg1 (ε) = 0.49Eεcr + 0.52Eε − 0.01E εεcr for εcr < ε < εp

6. Resistance plate group II (σpm;pg2 )


The plates of plate group II behave elastically up to either the proportional
limit is reached, or secondary buckling occurs:
σpm;pg2 (ε) = Eε for ε < min(εp ; εcr;2 )
If secondary buckling occurs in the elastic range an additional post-buckling
resistance is available:
ε2
σpm;pg2 (ε) = 0.49Eεcr;2 + 0.52Eε − 0.01E εcr;2 for εcr;2 < ε < εp

7. Predict the axial resistance (Nu;pm )

Nu;pm = σpm;pg1 (εp )Apg1 + σpm;pg2 (εp )Apg2

75
5.3.4 Inelastic buckling prediction: (fp ≤ σcr )
If the elastic critical stress is higher than the proportional limit of the material, buck-
ling in the inelastic range is considered. The structural behaviour is similar to the
material behaviour until the inelastic critical strength (σcr;T ) is reached, after which
a sudden failure occurs. The inelastic critical strength (σcr;T ) can be predicted by
using buckling curves presented in design codes or using equations for the inelastic
buckling coefficient χT , specified below:

σcr.T = χT (σcr )f0.2

The failure load for inelastic buckling results in:

Nu;pm = χT (σcr )f0.2 · A


The following section provides values of the inelastic buckling coefficient for three
aluminium alloys and a general procedure for inelastic coefficient calculation.

5.3.5 Inelastic buckling coefficient


The current research, experiments and numerical calculations are executed for ex-
truded specimens of 6060–T66 aluminium alloy and for cold-formed specimens of
6082–T6 and 5083–H111 aluminium alloys. The same alloys are considered later for
the parameter study and the prediction model development. It is assumed however,
that no test data are available for the considered alloys. Thus, aluminium alloys are
selected using material characteristics according to Eurocode 9 (CEN [2007]). The
value of the Ramberg-Osgood exponent is calculated according to Steinhardt’s pro-
posal n = f0.2 /10 (Steinhardt [1971]), which showed to provide a sufficient descrip-
tion for the elastic range of the material curve (Eq. 2.15). Material properties for the
regarded alloys are summarized in Table 5.1.

Table 5.1: Material characteristics used for parameter study.

Alloy E [N/mm2 ] f0.2 [N/mm2 ] fu [N/mm2 ] n


6060-T66 70000 160 215 16
6082-T6 70000 250 290 25
5083-H111 70000 110 270 11

Buckling curves for the three considered alloys giving the buckling coefficient
χT can be found in section C.2 of Appendix C. Mennink’s linearization for buck-
ling behaviour of these three alloys is described in section C.3, which results in the
following expressions:

6060-T66 χT = 1.08 − 0.3λ̄


6082-T6 χT = 1.06 − 0.22λ̄
5083-H111 χT = 1.10 − 0.4λ̄

76
Improved approach

According to theoretical derivation of section C.1 in Appendix C based on Shan-


ley’s approach [1947], the inelastic buckling coefficient would be:
r
1 1 f0.2
χT = 0.002En n−1 λ̄2
, where λ̄ = (5.2)
1+ f χT σcr
0.2

Figure 5.5 shows the proposed approximation for description of the inelastic part
of the considered buckling curves. It can be seen that it is more accurate to describe
the material curve with two approximated lines (dotted lines) instead of one line
as is stated in Mennink’s model (see section C.3). In Mennink’s inelastic buckling
approximation only the part λ̄ < 1 or σcr ≥ f0.2 is considered (compare with Fig-
ure C.2). It is possible that a situation occurs, where fp ≤ σcr < f0.2 . Therefore, the
part of the buckling curve from fp to f0.2 can be included in the prediction of the
non-linear behaviour.

Figure 5.5: Proposed inelastic buckling approximation.

Three important points can be assigned: σcr;T = f0.2 (χT = 1), σcr = f0.2 (λ̄ = 1)
and σcr = fp (Figure 5.5).
If σcr ≥ f0.2 , then Mennink’s approximation is used equal to Equation C.15:
s
χ∗T − 1 χ∗T − 1 f0.2
χT = q λ̄ − q · + 1, (5.3)
1− f0.2
1− f0.2 f0.2+0.002En
f0.2+0.002En f0.2+0.002En

77
where χ∗T is a single solution for two graphs:

1
f (χT ) = , where 0 < χT ≤ 1 (5.4)
χT
µ ¶
0.002En
and f (χT ) = χn−1
T +1 (5.5)
f0.2

When fp ≤ σcr < f0.2 , the mathematical expression for an approximated line can
be found from two points, where σcr = f0.2 (λ̄ = 1) and σcr = fp .
For the first point where σcr = f0.2 the value of the χT coordinate can be found
graphically according to Equations 5.4 and 5.5.
The second point coordinates are known:
s
f0.2 fp
λ̄ = and χT = . (5.6)
fp f0.2

Knowing the expression for a line between two points, the approximation for
non-linear material curve when fp ≤ σcr < f0.2 can be written as:

fp − χ∗T f0.2 fp − χ∗T f0.2


χT = ³q ´ · λ̄ − ³q ´, (5.7)
f0.2 f0.2
f0.2 fp − 1 f0.2 fp − 1

where χ∗T is a solution according to Equations 5.4 and 5.5.


Therefore, the proposed approximation results in the following expressions for
the materials considered:

6060-T66 χT = 1.08 − 0.3λ̄ for σcr ≥ f0.2


χT = 1.35 − 0.58λ̄ for fp ≤ σcr < f0.2
6082-T6 χT = 1.06 − 0.22λ̄ for σcr ≥ f0.2
χT = 1.4 − 0.56λ̄ for fp ≤ σcr < f0.2
5083-H111 χT = 1.10 − 0.4λ̄ for σcr ≥ f0.2
χT = 1.20 − 0.5λ̄ for fp ≤ σcr < f0.2

Summary and calculation procedure for χT

It can be concluded that Mennink’s model is improved by the inclusion of the


proposed method to define the inelastic buckling coefficient. The final expressions
for the general calculation procedure of the inelastic buckling coefficient χT are sum-
marized below.

If fp ≤ σcr < f0.2 , then Eq. 5.7 is applicable


If σcr ≥ f0.2 , then Eq. 5.3 is applicable

χ∗T can be determined graphically from Equations 5.4 and 5.5.

78
5.4 Model extension for distortional buckling

5.4.1 Description of the behaviour

Mennink’s model covers the prediction of local buckling behaviour. Thus, model
extension for distortional buckling is needed. In this section, distortional buckling,
local-distortional and distortional-local interactions are discussed, based on experi-
mental observations, results of the numerical calculations and literature research.
Compression tests for Z-shaped specimen 2Z200F10A have shown that initially
local buckling happened, then mode jumping to distortional buckling mode has
been detected by changes in the deformed shape of the specimen (Figure 3.24).
Post-buckling strength for distortional buckling has been observed during the ex-
periment. Hence, Mennink’s prediction model should include distortional post-
buckling strength.
FEM results did not indicate the distortional buckling initiation for the same
Z-shaped profile with nominal characteristics (see Figure 5.2). Unfortunately, the
example of section 5.2 does not provide a good representation of the distortional
buckling presence. As already discussed, the total length of the specimen is not
enough to allow distortional buckling to develop freely. Therefore, a parameter
study to investigate the distortional buckling behaviour is necessary.
Based on the evaluation of the buckling behaviour of the tested Z-shaped spec-
imen, it is assumed that secondary buckling of the cross-section is associated with
distortional buckling initiation. Following the theory of section 5.3.1, local-distortional
interaction can be described as initial local buckling of plate group I and secondary
distortional buckling of plate group II. Similarly, for distortion-local interaction the
initial buckling of plate group I relates to the distortional mode, while secondary
buckling of plate group II corresponds to the local buckling mode. It is assumed that
secondary buckling can be predicted by the finite strip method. If secondary buck-
ling is not defined by the FSM, a method for the secondary buckling determination
proposed by Mennink can be applied (section 5.3.2). This method is slightly modi-
fied for the distortional buckling case, which is discussed further in section 5.4.2.
From the research of Dinis and Camotim [2004], it is known that the distor-
tional post-buckling resistance is much lower than the local post-buckling resis-
tance. Kwon and Hancock [1992] state that in case of local-distortional buckling
interaction, the conventional methods for local buckling calculation can be used,
but the maximum stress for distortional buckling serves as the limiting stress rather
than the proportional limit stress. Therefore, if distortional buckling occurs after lo-
cal buckling, the post-buckling strength due to local buckling can be limited by the
distortional buckling critical stress. Similarly, in case of distortional-local interaction
the post-buckling strength due to distortional buckling can be limited by the local
buckling critical stress.

79
5.4.2 Post-buckling resistance
The general ideas of Mennink’s model for local buckling have been described in
section 5.3.1. An illustration of the local buckling behaviour of the cross-section is
given in Figure 5.6. It can be seen that the section behaves elastically until the first
critical point is reached (εcr1 ). Local buckling of the plate group I occurs, resulting
in stiffness reduction. The post-buckling strength of plate group I is limited with the
proportional limit of the material (εp ): line ”LL” in the figure. Plate group II behaves
elastically until the second critical point is reached (εcr2 ), where local buckling of the
plate group II happens. The post-buckling strength of the plate group II is limited
with the material proportional limit, where failure occurs.

Figure 5.6: Illustration of the local and distortional buckling behaviour of the cross-section,
according to the model

The general ideas for distortional buckling behaviour are illustrated in Figure 5.6
together with local buckling behaviour. It is assumed that distortional buckling be-
haviour is similar to local buckling behaviour described above. However, the post-
buckling strength for distortional buckling is lower than in case of local buckling,
which can be noticed from the figure. The post-buckling strength for distortional
buckling of plate group I is influenced by the secondary buckling initiation of plate
group II: line ”DD” in the figure. When the secondary buckling of plate group II

80
is reached, no additional support is provided for plate group I. As a safe approach,
it is assumed that there is no post-buckling strength for plate group I after the sec-
ondary distortional buckling. Therefore, the post-buckling strength for plate group
I is limited with the secondary distortional buckling stress of plate group II. Local-
distortional and distortional-local interactions are also represented in the figure, ac-
cording to research of Kwon and Hancock [1992]. If initial buckling corresponds to
local buckling and secondary buckling to distortional buckling, the post-buckling
strength for local buckling of plate group I is limited with the secondary distor-
tional buckling stress of plate group II (line ”LD”). Vice versa for distortional-local
interaction, the post-buckling strength for distortional buckling of plate group I is
limited with the secondary local buckling stress of plate group II (line ”DL”).

5.4.3 Model proposals


To extend Mennink’s model for distortional buckling the following proposals are
formulated:

1. For local-distortional interaction, distortional buckling can be associated with


secondary buckling. Similarly for distortional-local interaction, local buckling
can be associated with secondary buckling.

2. In case of local-distortional interaction in the elastic range (fp > σcr; d > σcr; l ),
the elastic post-buckling strength for local buckling should be limited with the
distortional buckling strength. Similarly, in case of distortional-local interac-
tion in the elastic range (fp > σcr; l > σcr; d ), the post-buckling strength for
distortional buckling should be limited with the local buckling strength.

3. If the distortional buckling strength is in the inelastic range (fp ≤ σcr; d ), the
conventional method is applied limiting the post-buckling strength to the pro-
portional material strength (see 5.3.3).

5.5 Application of the extended prediction model


Finite element analyses are executed for all experimental profiles, described in sec-
tion 3.1. As it was mentioned before, nominal values for dimensions are applied.
Material characteristics according to Eurocode 9 and Steinhardt’s approach are con-
sidered. Material properties for the aluminium alloys considered can be found in
Table 5.1. Imperfections with the shape of the first Euler buckling mode are in-
troduced into the FE-model. It is decided to select the value of the imperfection
amplitude as 10% of the plate thickness. In Chapter 4 for the finite element model
validation, fixed supports have been applied. The same approach is used in the
current section: the boundary conditions for all specimens are fixed.
Local and distortional buckling for all profiles considered are predicted accord-
ing to Mennink’s model with the inclusion of an improved approach for the inelastic
buckling coefficient calculation (section 5.3.5). Inclusions for distortional buckling

81
(section 5.4.3) are not evaluated because the specimens considered do not repre-
sent local-distortional or distortional-local interactions . The results of Mennink’s
model calculations and the numerical calculations are shown in Tables 5.2–5.4. Fi-
nite strip analysis results according to the CUFSM-program, FEM results and pre-
diction model output are presented.
It is important to mention that the finite strip analysis is able to define not only
the local buckling critical values, but also the distortional buckling initiation. There-
fore, two sets of CUFSM results are given in the tables: Lcr; l and σcr; l correspond
to the critical length and critical stress due to local buckling, Lcr; d and σcr; d cor-
respond to the critical length and critical stress due to distortional buckling. The
specimen’s total length determines whether the specimen is subjected to local or
distortional buckling mode. The critical length for distortional buckling is usually
longer than the critical length for local buckling. Thus, specimens that (according
to CUFSM) are subjected to both buckling modes, but with longer total length, are
considered to exhibit distortional buckling. These specimens are marked with color
and they are calculated applying CUFSM critical values for distortional buckling.
For many of the specimens, distortional buckling did not occur, thus no data for
distortional buckling is given in the tables.
Mennink’s model can be divided into two parts: elastic buckling and inelastic
buckling. If the value of the critical stress for local or distortional buckling is less
than the material proportional limit, the procedure for elastic buckling calculation
with post-buckling strength is used. Otherwise, it is considered that buckling hap-
pens in the inelastic range and then the procedure for inelastic buckling calculation
is used. To be able to evaluate both cases, it is mentioned in the tables if the critical
stresses according to the Eigenvalue analysis are within the elastic range or not. In
the field of ”El.” (Elastic) the word ”Yes” means that the elastic buckling calculation
with post-buckling strength according to Mennink’s model has been applied; the
word ”No” means that no elastic buckling occurred and the procedure for inelastic
buckling has been used.
The last number in the specimen’s designation is the corresponding cross-sectional
shape number of Figure 5.7.

82
Table 5.2: Comparison of the FE-results with Mennink’s prediction model results. ”Ex-
truded” profiles, 6060-T66 aluminium alloy.

Specimen Length CUFSM FEM Pr.Model Pr/FE


L Lcr;l σcr;l Lcr;d σcr;d Nu Nu;pr Mode El.
[mm] [mm] [kN] [mm] [kN] [kN] [kN] [-]
2Z40E1(1) 250 50 505 - - 24.2 22.2 L No 0.92
3Z35E2(1) 350 60 642 - - 34.1 F No -
3L80E3(2) 700 700 40 - - 46.3 - T Yes -
3C80E4(3) 700 160 87 - - 54.8 49.8 L Yes 0.91
3C80E5(4) 700 160 88 - - 69.5 55.9 L Yes 0.80
3C80E6(5) 500 90 310 - - 94.5 87.5 L No 0.92
3C80E7(6) 500 110 212 350 138 83.8 83.0 L No 0.99
3C80E8(6) 1300 110 212 350 138 76.0 77 D No 1.01
3C80E9(7) 500 100 299 - - 103.3 95.4 L No 0.92
L=local buckling/D=Distortional buckling/F=flexural buckling/T=torsional buckling/El.=Elastic

Figure 5.7: Considered cross-sectional shapes.

Table 5.3: Comparison of the FE-results with Mennink’s prediction model results. ”Cold-
formed” profiles, 6082-T6 aluminium alloy.
CUFSM FEM Pr.Model Pr/FE
Specimen L Lcr;l σcr;l Lcr;d σcr;d Nu Nu;pr Mode El.
[mm] [mm] [kN] [mm] [kN] [kN] [kN] [-]
2Z200F10A(8) 700 150 37 700 74 102 89 L Yes 0.87
2Z100F11A(8) 400 80 140 600 188 110 90 L Yes 0.82
2Z50F12A(8) 450 60 300 93 91 L No 0.98
2.5Z50F13A(8) 660 40 909 220 568 71 74 D No 1.04
L=local buckling/D=Distortional buckling/El.=Elastic

Table 5.4: Comparison of the FE-results with Mennink’s prediction model results. ”Cold-
formed profiles”, 5083-H111 aluminium alloy.
CUFSM FEM Pr.Model Pr/FE
Specimen L Lcr;l σcr;l Lcr;d σcr;d Nu Nu;pr Mode El.
[mm] [mm] [kN] [mm] [kN] [kN] [kN] [-]
2Z200F10B(8) 700 150 37 700 74 51 35 L Yes 0.69
2Z100F11B(8) 400 80 140 600 188 54 43 L No 0.80
2Z50F12B(8) 450 60 300 - - 45 40 L No 0.88
2.5Z50F13B(8) 660 40 909 220 568 31 33 D No 1.06
L=local buckling/D=Distortional buckling/El.=Elastic

83
The most right column of the tables compares the ultimate loads according to the
FE-model and the prediction model. The following observations can be highlighted:

• Mennink’s model for cross-sectional instability is not applied for specimens


3Z35E2 and 3L80E3 (see Table 3.1), because these specimens failed due to over-
all buckling. Specimen 3Z35E2 is a non-symmetrical profile and flexural buck-
ling has been observed. According to the FEM-results, specimen 3L80E3 failed
due to torsional buckling.

• For the specimens representing pure local buckling, occurring in the elastic
range, the prediction according to the model is too conservative (see speci-
mens 3C80E4, 3C80E5, 2Z200F10A, 2Z200F11A and 2Z200F10B).

• For the specimens representing pure local buckling, occurring in the inelastic
range, the prediction model provides a good result.

• For the specimens representing pure distortional buckling in the inelastic range,
the prediction is good (e.g. specimens 2.5Z50F13A and 2.5Z50F13B).

• Specimens 2Z200F10A, 2Z200F11A and 2Z100F10B, 2Z100F11B are subjected


to local buckling followed by distortional buckling. These specimens are not
calculated for distortional buckling, because the total length of the specimen
is not long enough to allow distortional buckling to develop. An example for
this situation was presented in section 5.2. Hence, this type of specimens is
calculated for local buckling only, which gives a very conservative result.

84
5.6 Summary and conclusions
The current chapter is summarized with the following conclusions:

• An additional approach for the inelastic buckling coefficient calculation is in-


cluded into Mennink’s model, which results in an improved prediction model.

• Inclusion of distortional buckling effects into Mennink’s model are proposed.


According to experimental observations, it is assumed that in case of local-
distortional interaction, distortional buckling can be introduced into the model
as secondary buckling. The critical value for distortional buckling can be de-
fined by the finite strip method. If the critical values for distortional buckling
initiation are not determined by the FSM, a theoretical approach is proposed.

• Mennink’s model for local buckling is evaluated considering various cross-


sectional shapes tested in the experimental program. According to the com-
parison analysis between the FE-model and Mennink’s model for the speci-
mens with various shapes, it can be concluded that the prediction model pro-
vides a good result for the inelastic local buckling calculation. However, in
case of an elastic local buckling calculation with post-buckling strength, the
result is much too conservative.

• The calculation procedure for inelastic local buckling prediction, based on


Shanley’s [1947] approach, is applied for inelastic distortional buckling predic-
tion. For the specimens representing pure distortional buckling in the inelastic
range, the prediction is very good.

• An improved model with additional inclusions for distortional buckling is


proposed. Based on experimental research and literature observation, it is
assumed that the prediction model of Mennink is adaptable for distortional
buckling. A parameter study for distortional buckling is needed, which is the
subject of Chapter 6.

• It can be noticed that among the specimens considered there is no situation


that distortional buckling happened in the elastic range. Thus, the prediction
model for elastic distortional buckling prediction is not evaluated. An exten-
sive parameter study is needed to evaluate distortional buckling behaviour
and to develop a prediction model for elastic distortional buckling.

85
Chapter 6
Prediction model for distortional buckling

Abstract

As a conclusion of Chapter 2, the current design rules for aluminium do not provide
a general approach for the prediction of distortional buckling. Therefore, buckling be-
haviour of aluminium structural members with complex cross-sectional shapes might
be inaccurately predicted by aluminium design standards. This chapter reviews the ac-
tual distortional buckling behaviour of a C-shaped cross-section following discussions of
Chapter 5. Special attention is given to elastic distortional buckling and post-buckling
distortional strength. A parameter study for C-shaped sections results in the description
of the general behaviour of plates that are composing the cross-section subjected to dis-
tortional buckling. A prediction model for distortional buckling behaviour of C-shaped
sections is developed.

6.1 Scope of the chapter

T his chapter deals with the results of a parameter study on distortional and
local-distortional (distortional-local) buckling effects of C-shaped specimens
with various slenderness parameters. The purpose of the parameter study is to in-
vestigate the interaction between cross-sectional plate elements and to develop a
prediction model for distortional buckling. To achieve the objectives of the research,
the following steps are taken:

• A C-section subjected to pure distortional buckling is selected for the parame-


ter study. A method to exclude the occurrence of overall buckling is proposed
for the FE-model. This FE-model is used for the parameter study.

• The selected C-shaped specimen is a reference shape in the parameter study.


Varying cross-sectional slenderness parameters, eleven C-shaped specimens
are chosen for the current investigation. Finite element analyses are executed
for the selected specimens. Similar to Mennink’s model described in sec-
tion 5.3.3, a distinction is made between distortional buckling occurring in
the elastic and inelastic ranges. To study elastic distortional buckling, elastic
buckling is considered applying ideal linear material characteristics according
to Hooke’s law (σ = Eε).

• Based on results of elastic buckling analyses a prediction model for elastic dis-
tortional buckling of C-shaped specimens is developed. Local-distortional and

87
distortional-local buckling interaction is also discussed. This newly developed
prediction model is applied for the C-shaped specimens considered of three
aluminium alloys (6082–T6, 6060–T66, 5083–H111). The prediction model re-
sults are compared with the results of inelastic buckling analyses.

6.2 Selection of the cross-sectional shape for distortional


buckling
The appropriate specimen selection for distortional buckling investigation is achieved
using the CUFSM program based on the finite strip method (Schafer [2006]). The
selected cross-sectional shape should be able to provide the development of distor-
tional buckling avoiding the occurrence of local and overall buckling.
Distortional buckling is known as a flange-stiffener phenomenon (see Hancock
[1978]). Therefore, it is proposed to select a simple symmetrical shape with flange
stiffeners. The commonly used C-shaped profile is selected. Figure 6.1 shows the
selected profile. To allow dominance of distortional buckling and not local buckling,
the length of the flange stiffener is taken small. To exclude any occurrence of local
buckling, the thickness of the web is twice as thick as the thickness of flanges and
lips.

Figure 6.1: Cross-sectional shape selected for parameter study.

Specimen dimensions are given in Table 6.1 in accordance with Figure 6.1. Spec-
imen designation 1(2)C5 is in line with the specification. The first number corre-
sponds to the value of thickness for flanges and lips, while the second number in
brackets relates to the value of thickness for the web. The letter indicates the shape
of the profile: ”C” for C-shaped profile. The last number is the cross-sectional stiff-
ener length (or value (d) and (c) in the table).

Table 6.1: Specimen specification for parameter study.

Specimen h a c th ta tc
1(2)C5 100 50 5 2 1 1

The signature curve and deformed shape for the considered profile as a result of
the finite strip analysis are given in Figure 6.2.

88
Figure 6.2: CUFSM results for selected C-profile with defined buckling shapes for local and
distortional modes.

Figure 6.3: CUFSM results for selected C-profile with defined buckling curves and critical
points for local and distortional modes.
It can be noticed from the CUFSM results that local buckling might occur and
not only distortional buckling. Figure 6.2 shows the occurrence of the local buck-
ling mode and shows the resulting local buckling shapes. Critical points for local
buckling are not defined by the CUFSM, however signature curves are visible. Fig-
ure 6.3 specifies one distortional signature curve, one curve for local/distortional

89
interaction when local buckling governs and one signature curve for local buckling.
It means that distortional buckling initiates first and local buckling governs later.
This observation is very important for the prediction of cross-sectional instability
behaviour.
Finite strip analyses results have shown that applying additional constraints at
the flange/web junction for C-shaped sections ”aids” to exclude the occurrence of
flexural buckling. Two springs for prevention of displacements in two axial direc-
tions are applied for one web/flange juncture, while one spring to prevent displace-
ments in horizontal direction is applied for another web/flange junction (see Fig-
ure 6.4).

Figure 6.4: Applied springs for C-shaped profile.

Figure 6.5: CUFSM result for C-shaped profile with additional springs.

90
The CUFSM signature curve and the deformed shape for the considered profile
with additional springs as a result of the finite strip analysis are given in Figure 6.5.
Critical points for local buckling are added similar to Figure 6.2. Thus, distortional
buckling is initiated first, while local/distortional interaction and local buckling pro-
ceeds later.
Comparing finite strip results for the specimen with springs (Figure 6.2) with
the results of specimen without springs (Figure 6.5), it can be noticed that the critical
stress for the specimen with springs are slightly higher. This can be explained by the
fact that in the CUFSM program springs are applied ”virtually” for the whole length
of the specimen. Different spring applications for the FE-model will be discussed in
the next section 6.3.
Based on the CUFSM result of Figure 6.2 it can be concluded that specimen
1(2)C5 of Table 6.1 is subjected to pure distortional buckling and local buckling for
higher stresses. However, it is better to ensure that overall buckling is also restricted.
For that purpose, additional springs of high stiffness are applied at the web/flanges
junctions. Figure 6.5 shows the results of spring application. It can be seen that the
occurrence of overall buckling is avoided. Thus, specimen 1(2)C5 of Table 6.1 with
additional springs is an appropriate choice for the parameter study. However, one
should be careful to ensure that additional springs do not influence the distortional
buckling behaviour.

6.3 FEM set-up for the parameter study


A finite element analysis is executed for the selected specimen 1(2)C5 of Table 6.1.
The set-up for the finite element model is similar to the one described in section 4.3.1.
Finite strip analysis input includes only the cross-section geometry, varying dif-
ferent lengths and determining the critical length for buckling initiation. For the
finite element model the total length of the specimen has to be introduced. The
specimen length has to be long enough to allow initiation of distortional buckling.
It is a common approach to base the choice of the specimen length on the critical
buckling length. In case of local buckling the total length of the specimen is usually
selected five times longer than the critical buckling length. However, overall buck-
ling might initiate for the longer lengths. In case of distortional buckling according
to the discussion on specimen length selection in Kutanova et al. [2009], the total
length of the specimen is taken three times longer than the critical length to allow
initiation of distortional buckling. This approach is used for all parameter study
FEM calculations.
The amplitude and the shape of the initial geometrical imperfections play a sub-
stantial role in the resulting buckling and post-buckling behaviour. Imperfections
in the shape of the first Euler buckling mode are introduced into the FE-model. For
the parameter study the influence of imperfections is reduced, using very small im-
perfections of 1/1000 of the plate thickness. Minimizing the influence of imperfec-
tions helps to recognize the pure distortional buckling behaviour. The application

91
of very small imperfections is sufficient, because the size of imperfections is small
for aluminium extrusions, which has been shown in the experimental program. The
sign of imperfections refers to the inward (positive sign) or outward (negative sign)
deformed shape. Dinis et al. [2006] have shown that columns with negative imper-
fections sign (outward buckling) result in a slightly lower post-buckling strength
than columns with positive sign (inward buckling). In the current investigation, the
opposite behaviour has been observed (see Kutanova et al. [2009]). However, the
influence of the imperfections sign is found less than 4% and can be neglected. For
the evaluation of the parameter study results a unified approach is used: outward
deformations are considered.
For the considered C-shaped specimen, the following mesh refinement provides
a sufficiently accurate result: 20 elements in the web, 10 elements in each flange and
one or two elements in each stiffener depending on the size of the stiffener. The size
of the element in longitudinal direction is maximum 33 mm. All rotations and all
translations are restricted for the specimen edges.
Three different ways of spring application were discussed in Kutanova et al.
[2009]. Spring elements were applied along the total length of the specimen, at
the middle of the specimen and at two points on a distance of the critical buck-
ling length from both edges. It was important to assure that the applied springs
do not influence the distortional buckling behaviour, but do restrict overall buck-
ling to develop. Based on the results of the different way to apply springs, it was
concluded that highly stiff springs can be applied for two points on a distance of
the critical buckling length from both edges. Due to the fact that the total length
is three times longer than the critical length, springs are applied at 1/3 and 2/3 of
the specimen length. Three translational spring elements SP1TR can be applied for
the FE-model at the web/flange junctions, following the CUFSM calculation as in
Figure 6.4. The stiffness of the spring elements is high and is equal to k = 7000000
[N/mm]. Schematization of the springs applied can be seen in Figure 6.6.

Figure 6.6: Spring application at two points on a critical length distance from both edges.

92
6.4 Parameter study definition
The parameter study is executed for the selected C-shaped profile showing the dis-
tortional buckling mode (see Figure 6.5). To be able to develop a theoretical ap-
proach for distortional buckling behaviour, a range of C-shaped profiles is studied.
Dimensions of the selected profile are varied according to the slenderness of the
cross-section.
The cross-sectional slenderness λ is commonly defined as:
r
f0.2
λ= .
σcr
It is proposed to vary the cross-sectional slenderness value within a range of
[0.5;2.2]. To choose sufficient specimens, values of the cross-sectional slenderness
based on the 0.2% proof stress of 6082-T6 aluminium alloy are considered. Subse-
quently, the cross-sectional slenderness is increased and decreased to the desired
values by reducing or increasing the plate thickness. The length of the stiffener is
also varied: increasing the stiffener length results in the development and, more-
over, in the prevalence of local buckling. Therefore, the stiffener length is varied
within a particular range to study distortional-local and local-distortional interac-
tions.
The selection of specimens for the parameter study is performed with the finite
strip method. Table 6.2 gives the dimensions in accordance with Figure 6.4 for the se-
lected C-shaped specimens. Table 6.3 contains the resulting critical lengths, stresses
and slenderness values λ for all considered specimens. Similar to Figure 6.5, critical
points are manually defined for the local buckling mode. For some specimens two
critical points for local buckling can be distinguished. Thus, critical values for the lo-
cal buckling mode (Lcr;l1 , σcr;l1 ; Lcr;l2 , σcr;l2 ) and critical values for the distortional
buckling mode (Lcr;d , σcr;d ) are included in the table.
It can be noticed that the total length of the specimens is selected three times
longer than the critical buckling lengths as mentioned in section 6.3. Specimen num-
ber 9 is marked with color for recognizing the reference specimen 1(2)C5, which was
selected in the previous section.

93
Table 6.2: Parameter study C–shaped specimens specification.

No. Specimen h a c th ta tc L
[mm] [mm] [mm] [mm] [mm] [mm] [mm]
1 5(6)C5 100 50 5 6 5 5 420
2 4.5(5)C5 100 50 5 5 4.5 4.5 450
3 3.5(4)C5 100 50 5 4 3.5 3.5 480
4 2.5(3)C5 100 50 5 3 2.5 2.5 510
5 1(2)C10 100 50 10 2 1 1 1050
6 2(2.5)C5 100 50 5 2.5 2 2 540
7 1(2)C7.5 100 50 7.5 2 1 1 870
8 1.5(2)C5 100 50 5 2 1.5 1.5 600
9 1(2)C5 100 50 5 2 1 1 630
10 0.75(2)C5 100 50 5 2 0.75 0.75 720
11 0.75(1.5)C5 100 50 5 1.5 0.75 0.75 750

Table 6.3: CUFSM results for C–shaped specimens of the parameter study (6082-T6 alu-
minium alloy proof stress is considered f0.2 = 250 [N/mm2 ]).
No. Specimen CUFSM
Lcr;l1 σcr;l1 Lcr;l2 σcr;l2 Lcr;d σcr;d λd
[mm] [N/mm2 ] [mm] [N/mm2 ] [mm] [N/mm2 ] -
1 5(6)C5 - - - - 140 516 0.7
2 4.5(5)C5 - - - - 150 399 0.8
3 3.5(4)C5 - - - - 160 265 1.0
4 2.5(3)C5 70 270 - - 170 160 1.2
5 1(2)C10 90 106 40 140 350 136 1.5
6 2(2.5)C5 90 166 - - 180 119 1.5
7 1(2)C7.5 100 105 40 139 290 102 1.6
8 1.5(2)C5 100 109 200 84 1.7
9 1(2)C5 100 97 40 133 210 70 1.9
10 0.75(2)C5 40 78 - - 240 57 2.1
11 0.75(1.5)C5 100 58 - - 250 49 2.2

6.5 Parameter study results: elastic buckling


In this section the elastic distortional buckling behaviour of uniformly compressed
C-shaped specimens selected for the parameter study is considered. Specimen di-
mensions are given in Table 6.2. FE-analyses are executed applying spring elements
on a distance of the initial critical buckling length from both edges, using elastic ma-
terial characteristics and using negligible imperfections (see section 6.3). Numerical
results for all selected C-shaped profiles are presented in detail in Kutanova et al.
[2009]. Important aspects of the elastic distortional study and interaction between
distortional and local buckling are introduced below.

6.5.1 Pure distortional buckling


FE-analyses results for the first specimen of Table 6.2 are discussed here. the CUFSM
output screen is represented to observe the mode decomposition (Figure 6.7). Pure

94
distortional buckling can be observed. The CUFSM critical values for distortional
buckling are given in Table 6.3.

Figure 6.7: CUFSM results for specimen 5(6)C5.

The FEM stability (Euler) analysis results in several buckling modes. The re-
sults of the first Euler buckling mode are comparable to the results of the finite strip
analysis. Differences between FSM and FEM critical values for initial buckling are
negligible (Table D.1 in Appendix D). Figure 6.8 shows the resulting buckling shape
of the first Euler mode according to the FEM. It can be noticed that the first Euler
mode represents a pure distortional buckling mode, the same as it was predicted by
the CUFSM calculation. Three waves of the flange/stiffener group can be observed.
The shape of the initial buckling is outward.
The FEM non-linear analysis results in deformed shapes at corresponding load
steps. The first Euler buckling mode corresponds to distortional buckling. The de-
formation patterns at the critical and post-failure loads are shown in Figure 6.9:
specimen failure is caused by pure distortional buckling.

Figure 6.8: Initial deformed shape according to Euler analysis (mode 1).

95
Figure 6.9: FEM deformed shapes for specimen 5(6)C5 according to non-linear analysis.

These FE-analysis results can be represented in load-displacement and load-


deflection diagrams (Figure 6.10). Critical Euler load for distortional buckling ac-
cording to the FSM results is included in the load-deflection graph.
The FE-model results for the web (internal element), flange+lip (outstanding
groups) and the whole specimen 5(6)C5 are introduced in Figure 6.11. The results
are represented in a non-dimensional diagram: stresses are divided by the critical
stress due to distortional buckling and strains are divided by the critical strains (see
results representation in section 5.2).

Figure 6.10: Load-displacement (left) and load-deflection (right) plots for specimen 5(6)C5.

96
Figure 6.11: FE-results for plate elements of specimen 5(6)C5.

Figure 6.12: Tangential stiffness for plate elements of specimen 5(6)C5.

According to Figure 6.11, the flange/lip assembly buckles first. Following the
concept of Mennink’s model described in section 5.3.1, plate group I consists of plate
elements most susceptible to buckling. Thus, in case of distortional buckling of this
C-shaped specimen, the flange/lip groups correspond to plates of group I. Respec-

97
tively, the web of the C-shaped specimen relates to the plate group II. The flange/lip
groups start to buckle at the initial critical stress for distortional buckling, while the
web of plate group II provides a support. A subsequent load increase may result in
secondary buckling of the web, which can be defined using the approach presented
in section 5.4.2.
Figure 6.12 introduces the relationship between the tangential stiffness E ∗ and
the axial strain ε∗ for plate elements and the whole C-section. Post-buckling stiffness
values for the whole section and flange/lip group are included in this diagram.
The initial buckling develops when the critical point (ε∗ = 1) is reached, which
results in a gradual stiffness decrease of the web and the section. For the flange/lip
assemblies, the initial buckling causes a sudden drop in stiffness.
Load redistribution over the cross-section occurs, which can be noticed by the
stiffness increase of the web (see Figure 6.12). It means that at the initial buckling,
the flange/lip assemblies partly unload unto the web. It should be also noted that
the average axial stresses of the cross-section are taken locally in the middle of the
specimen’s length. However, the average axial strains are global, because they are
obtained from the axial shortening of the specimen.

6.5.2 Secondary buckling and distortional-local interaction


FE-analyses results for specimen No. 4 of Table 6.2 are discussed here. The CUFSM
output screen is represented to observe the mode decomposition (Figure 6.13). It can
be seen that distortional buckling is governing first, while for higher stresses local
buckling might occur. CUFSM critical values for local and distortional buckling are
given in Table 6.3.

Figure 6.13: CUFSM results for specimen 2.5(3)C5.

98
The resulting deformed specimens according to the non-linear FE-analysis are
shown in Figure 6.14. Different deformation shapes can be recognized depending
on the prevalent buckling mode. The specimen is subjected to distortional buck-
ling first, while after the second critical point, local buckling starts to govern. The
specimen failed due to distortional-local buckling interaction.
Results of the FE-analysis are plotted in Figure 6.15. The first critical Euler load
corresponding to distortional buckling is included in the load-deflection graph. The
critical local buckling load according to the CUFSM calculation is also added in the
load-deflection diagram. It can be seen that critical local buckling load according
to CUFSM coincides with the FEM load level where mode jumping occurs. Hence,
a perfect prediction of the secondary buckling initiation can be noticed. The load-
deflection diagram gives a good illustration of the secondary buckling initiation.

Figure 6.14: FEM deformed shapes for specimen 2.5(3)C5 according to non-linear analysis.

Figure 6.15: Load-displacement (left) and load-deflection (right) plots for 2.5(3)C5.

99
Figure 6.16: FE-results for plate elements of specimen 2.5(3)C5.

Figure 6.17: Tangential stiffness for plate elements of specimen 2.5(3)C5.

The FE-model results for the web, outstanding group (flange+lip) and the whole
section 2.5(3)C5 are introduced in Figure 6.16. Figure 6.16 shows that initial buck-
ling occurs at the critical stress σcr; d due to instability of the outstanding element.

100
Secondary buckling or buckling of the web is well-defined from the CUFSM results
(see Figure 6.13). For larger deformations mode jumping can be observed. It means
that there is mode jumping from distortional buckling into local buckling mode.
Plate buckling behaviour changes as well after the mode jumping: the web is more
susceptible to buckling, which indicates that local buckling starts to govern.
Figure 6.17 introduces the relationship between the tangential stiffness E ∗ and
the axial strain ε∗ for the plate elements and the whole C-section. At the beginning,
distortional buckling behaviour can be recognized similar to the specimen 5(6)C5
described in section 6.5.1. For higher deformations the curves intersect at one point
and mode jumping can be observed. There is a clear representation of mode jump-
ing from distortional to local buckling.

6.5.3 Secondary buckling and local-distortional interaction


The FE-analyses results for specimen No.5 of Table 6.2 are discussed here. The
CUFSM output screen is represented to observe the mode decomposition (Figure 6.18).
It can be seen that local buckling is initiated first, whereas for the higher stresses dis-
tortional and secondary local buckling occur. It is assumed that secondary buckling
of the cross-section corresponds to distortional buckling of the flange/lip group.
The manually indicated second critical point for local buckling is not considered,
because it is almost equal to the distortional buckling critical stress and represents a
local-distortional interaction.

Figure 6.18: CUFSM results for specimen 1(2)C10.

The resulting deformed specimens according to the FE-analysis are shown in


Figure 6.19. The specimen is subjected to local buckling initially, whereas after the

101
second critical point, distortional buckling starts to govern. The value belonging to
secondary buckling is identical to the CUFSM critical value for distortional buck-
ling. The specimen fails due to local-distortional interaction.

Figure 6.19: FEM deformed shapes for specimen 1(2)C10 according to non-linear analysis.

Results of the FE-analysis are plotted in Figure 6.20. The critical loads corre-
sponding to local buckling and distortional buckling, defined by the CUFSM, are
included in the load-deflection graph. Similar to the results of the previously de-
scribed specimen No. 4, a perfect prediction of the secondary buckling initiation
can be observed.

Figure 6.20: Load-displacement (left) and load-deflection (right) plots for specimen 1(2)C10.

102
Figure 6.21: FE-results for plate elements of specimen 1(2)C10.

Figure 6.22: Tangential stiffness for plate elements of specimen 1(2)C10.

The FE-model results for the web, outstanding group (flange+lip) and the whole
section 1(2)C10 are introduced in Figure 6.21. Figure 6.22 shows the relationship
between the tangential stiffness E ∗ and the axial strain ε∗ for the plate elements and

103
the whole C-section. It can be concluded that at first buckling of the web initiates,
which refers to local buckling. When the critical point for secondary buckling is
reached, mode jumping occurs and distortional buckling is dominant. It is a clear
graphic representation of mode jumping from local buckling to distortional buck-
ling and local-distortional interaction.

6.5.4 Elastic buckling of the reference C-shape


The FE-analyses results for the reference specimen No. 9 of Table 6.2 are discussed
here. The CUFSM output screen was represented before in Figure 6.5. Accord-
ing to the CUFSM calculations, distortional buckling is initiated first, whereas for
higher stresses two critical points for local buckling are present. It is assumed that
secondary buckling of the cross-section is the local buckling of the web, which cor-
responds to the most critical point for local buckling. The manually defined second
critical point for local buckling is considered as well to check if it has any physical
meaning.
The resulting deformed specimens according to the FEM non-linear analysis are
shown in Figure 6.23. The specimen is subjected to pure distortional buckling first
and later local buckling occurs, which can be noticed by the deformed shapes. The
specimen failed due to distortional-local buckling.

Figure 6.23: FEM deformed shapes for specimen 1(2)C5 according to non-linear analysis.

Results of the FE-analysis are plotted in Figure 6.24. Based on the CUFSM re-
sults, three critical points have been defined. The critical values for initial distor-
tional buckling, local buckling and manually defined secondary local buckling are
included in the load-deflection graph. According to the load-deflection diagram, the
second critical value for local buckling does not indicate any effect on the buckling
behaviour.

104
Figure 6.24: Load-displacement (left) and load-deflection (right) plots for specimen 1(2)C5.

The FE-model results for the web, outstanding group (flange+lip) and the whole
section 1(2)C5 are represented in Figure 6.25. Figure 6.25 shows that initial buck-
ling occurs at the critical stress σcr; d due to instability of the outstanding element.
Secondary buckling or buckling of the web is well-defined from the CUFSM results
(see Figure 6.5). Again, no particular influence of the second critical value for local
buckling can be recognized. Therefore, it is concluded that only two CUFSM results
should be considered: initial buckling and manually indicated secondary buckling
(which is the most critical after initial buckling).

Figure 6.25: FE-results for plate elements of specimen 1(2)C5.

105
Figure 6.26: Tangential stiffness for plate elements of specimen 1(2)C5.

Figure 6.26 shows the relationship between the tangential stiffness E ∗ and the
axial strain ε∗ for the plate elements and the whole C-section. Initial buckling devel-
ops when the critical point (ε∗ = 1) is reached, which results in a gradual decrease
of stiffness for the web and section and a sudden drop for flange/stiffener assem-
blies. A subsequent load increase results in secondary buckling of the web. Mode
jumping from distortional mode into local mode is not observed. Specimen failure
is attributed to distortional-local interaction.

6.5.5 Evaluation of elastic buckling results


Based on the finite element analyses results for 11 selected specimens (Kutanova
et al. [2009]), the following conclusions can be drawn:

• All regarded specimens failed due to distortional, local-distortional and distortional-


local buckling. No occurrence of overall buckling has been observed.

• In case of distortional buckling, the outstanding group (flange+lip) buckles


first. The internal element (web) provides initial support for the outstanding
elements. In case of local buckling, the internal element (web) buckles first,
while the outstanding components provide initial support.

• Pure distortional buckling occurred for the following specimens: No. 1, No. 2,
and No. 3 (see Tables 6.2–6.3). Finite element analyses results of these speci-
mens can be used for the development of the prediction model for elastic dis-
tortional buckling behaviour. No secondary buckling can be defined for these

106
specimens from the results of the finite strip analysis. Thus, it is assumed that
initial deformations of the flanges and lips cause initiation of secondary buck-
ling of the web. Critical stresses for secondary buckling can be found using a
theoretical approach of the buckling coefficient determination kcr for the sup-
porting plate group as proposed in section 5.4.2. Figure 6.27 illustrates the
structural behaviour of plate elements. First, initial buckling of the outstand-
ing elements (pgI) occurs, the critical stress σcr coincides with the bifurcation
load according to the finite strip analysis or Euler buckling analysis. Due to
the fact that imperfections are very small, it is assumed that the whole cross-
section behaves elastically up to the initial critical stress. Secondary buckling
of the supporting internal plate group (pgII) occurs at the value of σcr;2 , which
is defined using the theoretical approach given in section 5.4.2.

• Distortional buckling with defined secondary buckling (CUFSM) occurred for


the following specimens: No. 6, No. 8, and No. 9. For these specimens no
mode jumping was observed. The finite element analyses results of the listed
specimens can be used for the development of the prediction model for the
elastic distortional buckling behaviour. Secondary buckling of the web is pre-
dicted by FSM. A perfect agreement has been found between FSM and FEM re-
sults for secondary buckling of the web. Thus, the CUFSM results can be used
for the initial buckling prediction and also for the secondary buckling predic-
tion. Figure 6.28 illustrates the structural behaviour of the cross-sectional plate
elements, similar to Figure 6.27. However, in this case secondary buckling is
determined using the CUFSM.

Figure 6.27: Initial buckling (outstanding plates group I) and secondary buckling (in-
ternal plates group II). Secondary buckling is defined theoretically.

107
Figure 6.28: Initial buckling (outstanding plates group I) and secondary buckling (in-
ternal plates group II). Secondary buckling is determined by CUFSM.

Figure 6.29: Post-buckling behaviour. Distortional-local interaction.

• Pure distortional buckling with mode jumping at larger deformations occurred


for specimen number 4. Local buckling or secondary buckling of the web is
well-predicted by FSM. Finite element analyses results for this specimen can

108
be also included for the development of the prediction model. However, only
results prior to mode jumping should be considered.

• Distortional buckling with subsequent local buckling occurred for specimens


No. 7, No. 10 and No. 11. These specimens failed due to distortional-
local and local-distortional interaction. Secondary buckling is well-detected
by the CUFSM. The CUFSM indicates the governing of local buckling. The
distortional-local interaction behaviour has shown that initial distortional buck-
ling of the outstanding elements should be limited with the critical stress for
the secondary buckling. When the web starts to buckle, no additional support
is provided to flange/lip assemblies (see Figure 6.29).

• Local buckling followed by jumping into the distortional buckling mode oc-
curred for specimen number 5. Secondary buckling is perfectly predicted by
CUFSM. Figure 6.30 presents the resulting structural behaviour of this speci-
men. It is a clear representation of local-distortional interaction.

Figure 6.30: Local-distortional interaction.

6.6 Development of a prediction model for distortional


buckling of C-shaped specimens
The parameter study results for C-shaped specimens are summarized in this sec-
tion. First, initial buckling of the flange/stiffener assembly is considered and then,
secondary buckling of the web.

109
6.6.1 Initial buckling
The FE results for outstanding plate elements (flange/stiffener assembly) of C-shaped
sections subjected to pure distortional buckling are combined in Figure 6.31.
The plotted curves are fitted with three straight dashed lines. These lines indicate
that initial buckling develops at ε∗ = εεcr = 1, resulting in a stiffness decrease until
the value of the average post-buckling stiffness E ∗ . The equations for the average
straight line are stated in Table 6.4.

Figure 6.31: Investigation of the scatter in distortional buckling behaviour of the outstanding
plate elements.

Table 6.4: Equations for the average post-buckling stiffness of outstanding components.


Eave = 0.36 − 0.05ε∗

R ∗
σave = Eave dε∗ = 0.36ε∗ − 0.025(ε∗ )2 + const
ε σave
If ε∗ = εcr

= 1, then σave = σcr =1 =⇒

σave = 0.66 + 0.36ε∗ − 0.025(ε∗ )2 =⇒

2
σave = 0.66Eεcr + 0.36Eε − 0.025 εεcr

110
6.6.2 Secondary buckling
The FE results for internal plate elements (web) of C-shaped sections subjected to
pure distortional buckling are combined in Figure 6.32. The diagram corresponds
to the behaviour of supporting webs (pgII):
ε σ
ε∗∗ = ; σ ∗∗ =
εcr;2 σcr;2

Critical stresses for secondary buckling σcr;2 are defined numerically (FSM). From
the results of Mennink [2002], it is known that the supporting plates (pgII) for rect-
angular hollow sections and I-sections behave like imperfect plates. Therefore, the
curve of an individual ss-free plate (a simply supported plate whose unloaded sides
are free to wave), ss-straight plate (a simply supported plate whose unloaded edges
are straight) and imperfect ss-free plate with imperfections value equal to 1/10 of
plate thickness are included in Figure 6.32. It can be seen that the ss-free curve
presents the average approximation for the C-shaped sections considered. The sup-
porting plate element (pgII) resembles an imperfect ss-free plate and can, thus, be
described by the ss-free plate behaviour.

Figure 6.32: Investigation of the scatter in distortional buckling behaviour of the internal
plate elements.

111
The equations for the average approximation, described by the ss-free individual
plate, are given in Mennink [2002] and presented in Table 6.5.

Table 6.5: Equations for the average post-buckling stiffness of internal components.

∗∗
Eave = 0.42 − 0.03ε∗∗
∗∗
R ∗∗
σave = Eave dε∗∗ = 0.42ε∗∗ − 0.015(ε∗∗ )2 + const
ε σave
If ε∗∗ = εcr;2
∗∗
= 1, then σave = σcr;2 =1 =⇒
∗∗ ∗∗ ∗∗ 2
σave = 0.6 + 0.42ε − 0.015(ε )

2
ε
σave = 0.6Eεcr;2 + 0.42Eε − 0.015E εcr;2

The equations of Tables 6.4–6.5 can be used for the prediction of distortional
buckling behaviour of C-shaped aluminium specimens. In sections 6.7 and 6.8 the
calculation procedure for distortional buckling prediction will be dealt with.

6.7 Prediction model for distortional buckling behaviour


of C-shaped aluminium structural elements
The input data for distortional buckling prediction are: Lcr; d - the critical length for
distortional buckling initiation, σcr; d - the critical stress for distortional buckling ini-
tiation, the cross-sectional dimensions and the material properties. If local buckling
takes place for a profile considered, the input data will include the critical length
and stress for local buckling development.
The following steps should be taken:

1. Determine the eigenvalue by either using analytical solutions or, more ap-
propriate for arbitrary cross-sections, finite-element or finite-strip programs.
This results in the determination of the critical length Lcr; d and the critical
stress σcr; d of the cross-section. The resulting buckling shape indicates flange-
stiffener or distortional buckling. If the eigenvalue analysis results in local
buckling, Mennink model can be applied (see section 5.3.3) according to the
buckling behaviour illustrated in Figure 5.6.

2. Determine secondary critical buckling stress by FSM. Local buckling proceed-


ing after distortional buckling implies secondary buckling of the internal ele-
ments of the cross-section. The most critical local buckling stress defined on
the CUFSM curve is related to the secondary buckling stress. Therefore, two
plate groups are determined:

112
• The first plate group (pg1) consists of all outstanding plate elements (flange-
stiffener assembly) that buckle at the initial critical stress for distortional
buckling σcr; d .
• The second plate group (pg2) consists of all internal plates (web) that
buckle at the secondary buckling critical stress for the local mode σcr; l .

3. If secondary buckling is not determined by FSM, Mennink’s approach accord-


ing to the model in section 5.3.3 can be used.

4. Determine the cross-sectional areas


For each plate i of plate group I and plate j of plate group II, the cross-sectional
area is determined from their plate width and thickness. Subsequently, this
results in the following representations
X of the cross-sectional areas:
Plate group I Apg1 = bi ti
Xi
Plate group II Apg2 = bj tj
j
Cross-section A = Apg1 + Apg2

5. ELASTIC BUCKLING with post-buckling strength


It is assumed that the proportional limit of the material, represented by εp
and fp , roughly defines the difference between the elastic and inelastic buck-
ling range. The cross-section will show post-buckling resistance if the elastic
critical stress is less than the proportional limit (σcr < fp ). If local buckling
or secondary buckling takes place, the cross-section will show post-buckling
resistance until the critical stress for local buckling (σcr < σcr; l ). The post-
buckling resistance for distortional buckling is determined from:
Plate group I post-buckling resistance (σpm;pg1 ):

σpm;pg1 (ε) = Eε for ε < εcr; d


ε2
σpm;pg1 (ε) = 0.66Eεcr; d +0.36Eε−0.025E εcr; d
for εcr; d ≤ ε ≤ min(εcr; l ; εp )
ε2cr; l
σpm;pg1 (ε) = 0.66Eεcr; d + 0.36Eεcr; l − 0.025E εcr; d
for ε ∈ [εcr; l ; εp ]

Plate group II post-buckling resistance (σpm;pg2 ):

The plates of plate group II behave elastically up to either the proportional


limit is reached, or secondary buckling occurs:
σpm;pg2 (ε) = Eε for ε < min(εp ; εcr;2 )
If secondary or local buckling occurs in the elastic range an additional post-
buckling resistance is available:
ε2
σpm;pg2 (ε) = 0.6Eεcr;2 + 0.42Eε − 0.015E εcr;2 for εcr;2 ≤ ε ≤ εp

Predict the axial resistance (Nu;pm ) as a function of ε

Nu;pm (ε) = σpm;pg1 (ε)Apg1 + σpm;pg2 (ε)Apg2

113
6. INELASTIC BUCKLING
For distortional buckling prediction in the inelastic range, Shanley’s [1947] ap-
proach as for local buckling is used. The inelastic buckling calculation proce-
dure is described in section 5.3.4.

6.8 Inelastic distortional buckling


The influence of the material inelasticity on the distortional buckling behaviour of C-
sections is investigated in this section. Three aluminium alloys which are commonly
used in structural applications are considered in the parameter study: 5083–H111,
6060–T66 and 6082–T6. Material characteristics for these aluminium alloys were
given in Table 5.1. Numerical calculations are executed applying three different
material characteristics for the eleven specimens of Table 6.2.
Tables D.2–D.4 in Appendix D contain the FE-results for all specimens apply-
ing material properties of the three aluminium alloys. Figures D.1– D.6 show the
resulting shapes of the compressed specimens at the value of the ultimate load. Ta-
bles E.1–E.3 in Appendix E contain the results of the prediction model applying the
characteristics of the three aluminium alloys.
Comparisons between the prediction model and the FEM results for all speci-
mens applying 6082-T6, 6060-T66 and 5083-H111 material characteristics are pre-
sented in Tables 6.6–6.8. Similar to section 5.5, in the field of ”El.” (Elastic) the word
”Yes” means that the elastic buckling calculation with post-buckling strength ac-
cording to the newly developed model has been applied; the word ”No” means that
Shanley’s approach for the inelastic buckling calculation has been used.
The non-linear calculation results for all specimens using 6082-T6, 6060-T66 and
5083-H111 alloy material properties are presented in Figures 6.33–6.35. This stress-
strain plot includes the calculated inelastic critical stresses (dots) that correspond
to the word ”No” in the table and the results of the elastic prediction model for
distortional buckling (squares) that correspond to the word ”Yes” in the table.

114
Table 6.6: FEM results for C-profiles selected for the parameter study (6082-T6).

No. Specimen Length Slend. FEM Pr. model Pr/FEM


L λd Nu Nu;pr El.
[mm] [kN] [kN] [-]
1 5(6)C5 420 0.7 291 261 No 0.90
2 4.5(5)C5 450 0.8 246 220 No 0.90
3 3.5(4)C5 480 1.0 178 166 No 0.94
4 2.5(3)C5 510 1.2 100 101 Yes 1.02
5 1(2)C10 1050 1.4 40 38 Yes 0.94
6 2(2.5)C5 540 1.4 69 73 Yes 1.06
7 1(2)C7.5 870 1.6 36 39 Yes 1.09
8 1.5(2)C5 600 1.7 44 43 Yes 0.98
9 1(2)C5 630 1.8 34 35 Yes 1.02
10 0.75(2)C5 720 2.0 30 29 Yes 0.95
11 0.75(1.5)C5 750 2.2 21 20 Yes 0.96

Figure 6.33: Comparison of the inelastic distortional buckling (6082–T6 alloy) with the elastic
prediction model at εp (squares) and inelastic prediction model σcr;T (dots).

115
Table 6.7: FEM results for C-specimens (6060-T66).

No. Specimen Length Slend. FEM Pr. model Pr/FEM


L λd Nu Nu;pr El.
[mm] [kN] [kN] [-]
1 5(6)C5 420 0.6 193 167 No 0.87
2 4.5(5)C5 450 0.6 166 141 No 0.85
3 3.5(4)C5 480 0.8 124 105 No 0.85
4 2.5(3)C5 510 1.0 78 71 No 0.91
5 1(2)C10 1050 1.1 34 33 No 0.96
6 2(2.5)C5 540 1.1 54 51 No 0.93
7 1(2)C7.5 870 1.3 33 31 No 0.96
8 1.5(2)C5 600 1.4 34 34 Yes 1.00
9 1(2)C5 630 1.5 27 28 Yes 1.02
10 0.75(2)C5 720 1.6 24 24 Yes 1.01
11 0.75(1.5)C5 750 1.8 15 15 Yes 1.05

Figure 6.34: Comparison of the inelastic distortional buckling (6060–T66 alloy) with the
elastic prediction model at εp (squares) and inelastic prediction model σcr;T (dots).

116
Table 6.8: FEM results for C-shaped specimens (5083-H111).

No. Specimen Length Slend. FEM Pr. model Pr/FEM


L λd Nu Nu;pr El.
[mm] [kN] [kN] [-]
1 5(6)C5 420 0.5 141 116 No 0.82
2 4.5(5)C5 450 0.5 121 97 No 0.81
3 3.5(4)C5 480 0.6 91 73 No 0.80
4 2.5(3)C5 510 0.8 59 49 No 0.82
5 1(2)C10 1050 0.9 27 24 No 0.91
6 2(2.5)C5 540 0.9 43 37 No 0.86
7 1(2)C7.5 870 1.0 26 24 No 0.92
8 1.5(2)C5 600 1.1 27 25 No 0.92
9 1(2)C5 630 1.2 22 20 No 0.87
10 0.75(2)C5 720 1.3 18 16 No 0.87
11 0.75(1.5)C5 750 1.5 12 12 Yes 0.99

Figure 6.35: Comparison of the inelastic distortional buckling (5083–H111 alloy) with the
elastic prediction model at εp (squares) and inelastic prediction model σcr;T (dots).
According to results of the comparison analysis for the three aluminium alloys,
it can be concluded that the prediction model provides sufficiently accurate results.
In case of elastic distortional buckling calculation with post-buckling strength, the
deviation of the model prediction is less than 7%. In case of inelastic distortional
buckling prediction, using Shanley’s approach [1947], the result is less accurate but
conservative.

117
6.9 Chapter conclusions
In this chapter a prediction model for distortional buckling of aluminium C-shaped
specimens has been developed:

• A parameter study for C-sections with various slenderness parameters and


different materials has been performed. The eleven selected specimens failed
due to distortional or local-distortional (distortional-local) buckling. In case of
distortional buckling, the outstanding plate groups (flange+lip) buckle first,
while the internal element (web) provides initial support to the outstanding
elements. In case of local buckling, the internal element (web) buckles first,
while the outstanding components provide initial support.

• For the distortional-local buckling interaction, local buckling of the web is a


secondary buckling for the cross-section. Similarly, for the local-distortional
buckling interaction, distortional buckling of the flange/lip assemblies is a
secondary buckling for the cross-section. Secondary buckling can be predicted
by FSM. A perfect agreement has been found between FSM and FEM results
of secondary buckling prediction. Thus, CUFSM results can be used for initial
buckling prediction and also for secondary buckling prediction. If secondary
buckling is not predicted by FSM, Mennink’s approach can be used.

• The distortional-local buckling interaction has shown that initial distortional


buckling of the outstanding plate elements should be limited by the critical
stress of secondary buckling. Similarly, for local-distortional interaction, initial
local buckling of the internal element should be limited by the critical stress of
secondary buckling of outstanding plate elements.

• The prediction model for distortional buckling has been applied for C-shaped
specimens of three aluminium alloys. Comparing the numerical results with
the results of the newly developed prediction model, it can be concluded that
the prediction model provides sufficiently accurate results.

118
Chapter 7
Conclusions and recommendations

Abstract

This thesis provides insight into distortional buckling behaviour of aluminium structural
elements with complex cross-sectional shapes. The research includes a large amount of
experimental and numerical data. Compression tests on aluminium members with L-
shaped, Z-shaped and C-shaped profiles have been executed to observe the local, distor-
tional and local-distortional buckling effects. The results of tests have been used for the
finite element model validation. A parameter study for distortional buckling has been
performed applying the validated FE-model. Based on the parameter study results and
assessment of Mennink’s prediction model for local buckling, a prediction model for dis-
tortional buckling of C-shaped aluminium members has been developed.

7.1 Conclusions
7.1.1 Distortional buckling in the current design rules
Cross-sectional instability often determines the structural resistance of aluminium
members with slender cross-sectional shapes. Local and distortional buckling phe-
nomena correspond to cross-sectional instability. The cross-sectional complexity
makes the prediction of local and distortional buckling behaviour one of the most
important design aspects. Current design rules for aluminium provide design rules
for cross-sectional instability with limited accuracy. Therefore, an extensive study
into local buckling of aluminium extrusions has been carried out by Mennink [2002],
resulting in an accurate prediction model for local buckling resistance. However, a
prediction model for distortional buckling of aluminium structural members is still
not available.

7.1.2 Experimental research


Executing compression tests is an appropriate way to investigate the local and dis-
tortional buckling behaviour of thin-walled aluminium elements with complex cross-
sectional shapes. Cross-sectional complexity can be investigated studying simple
cross-sections, the contribution of the various plate elements to the load-bearing ca-
pacity of the structural member. To study distortional buckling, it is sufficient to
consider cross-sectional shapes with flange stiffeners, e.g. Z-shaped or C-shaped
specimens.

119
For the compression tests, it is necessary to provide fixed support conditions.
In this case an accurate way of loading can be achieved and overall buckling is
prevented. Cross-sectional side walls of the specimen have to be flat and parallel to
each other. Despite the effort taken in specimen preparation, the necessary accuracy
was not achieved. Many compressed specimens failed at the supports and gaps
between the specimen edges and supports were observed. As a result of this gap, the
load is not uniformly distributed over all cross-sectional sides of the specimen. To
avoid the influence of the gap, double layers of aluminum plates (softer aluminium
compared to the material of the tested specimens) can be applied in-between the
specimen and the supports, in order to obtain homogeneous loading of all cross-
sectional sides.

7.1.3 Numerical research


Specimen selection for the current investigation is performed with the finite strip
method. CUFSM is a user-friendly program that allows to vary easily the dimen-
sions of the profile and to predict the occurrence of local and/or distortional buck-
ling modes. The results of the finite strip analysis are comparable to the results of
the first Euler buckling mode according to the finite element method.
The finite element model is validated based on experimental results. Deviations
from experiments can be explained by uncertainties in imperfections and support
conditions. To be able to compare the results of tests and FEM calculations a gap
between the specimen and supports is modelled. The FE-models with added gap
provide an excellent agreement with tests. The resulting deformed specimens ac-
cording to the FE-models and tests are comparable. The experimental structural
behaviour (ultimate load, stiffness) could be approximated with good accuracy. Im-
perfection sensitivity analysis shows that application of the first Euler mode as an
imperfection pattern is safe. However, introduction of the measured imperfection
pattern for the FE-model results in better simulation of the actual cross-sectional
instability behaviour, but is rather laborious.

7.1.4 Prediction model for distortional buckling


Mennink’s model for local buckling prediction is validated in the current investiga-
tion. It is assumed that the model can be extended for the prediction of distortional
buckling. Following the concept of Mennink’s model, a cross-section is divided into
plate elements that buckle (supported plates) and plates that provide support (sup-
porting plates). Plate elements with critical buckling stresses lower than the critical
buckling stress of the whole cross-section belong to the group of supported plates.
The second group of supporting plates consists of all remaining plate elements of
the cross-section. It is assumed that the proportional limit of the material separates
the elastic buckling range from the inelastic one. Buckling in the elastic range re-
sults in substantial post-buckling resistance. Buckling in the inelastic range leads to
immediate failure. Initial buckling of the cross-section occurs due to buckling of the

120
supported plates. The supporting plates behave either elastically until the propor-
tional limit of the material or they buckle. Buckling of the supporting plates is called
secondary buckling. An analytical approach is proposed for the secondary buckling
determination. The post-buckling strength after secondary buckling is limited by
the proportional limit. Thus, initial and secondary buckling modes are important
parameters for the prediction of cross-sectional instability.
Distortional buckling is known as a flange-stiffener phenomenon. Therefore, dis-
tortional buckling is studied on C-shaped specimens. In case of distortional buck-
ling, flange/lip assemblies or outstanding plate elements buckle first and belong to
the supported group. The internal element provides initial support for outstanding
elements. Local buckling of the internal plate is considered as secondary buckling
of the cross-section. The secondary buckling of the internal plate can be predicted
by the finite strip method. A perfect agreement is found between FSM and FEM
results for secondary buckling of the supporting plate. Thus, the CUFSM results
can be used for initial buckling prediction and also for secondary buckling predic-
tion. If secondary buckling is not detected by the FSM, secondary buckling of the
supporting internal plates can be sufficiently defined using a theoretical approach.
The initial post-buckling resistance of the supported group is limited by the critical
stress of secondary buckling. The secondary post-buckling resistance of the sup-
porting group is limited to the proportional limit of the material.
A prediction model for distortional buckling is developed. This model is based
on the actual distortional buckling behaviour of C-shaped specimens. It gives an
accurate prediction of the elastic distortional buckling and the elastic post-buckling
strength. The model also provides a safe approach for the inelastic buckling. The
newly developed model is applied to C-shaped specimens of three aluminium al-
loys. Comparing the numerical results with the results of the prediction model, it
can be concluded that the prediction model provides sufficiently accurate results.
In case of elastic distortional buckling with post-buckling strength, the deviation of
the model prediction in terms of ultimate buckling strength is less than 7%.

121
7.2 Recommendations
In the current thesis, cross-sectional instability of aluminium complex cross-sectional
shapes is studied on relatively simple cross-section. It is assumed that the concept
of local and distortional buckling prediction developed on a base of structural be-
haviour of common cross-sectional shapes is appropriate for complex shapes. How-
ever, future research for cross-sectional instability of aluminium extrusions should
focus on complex cross-sectional shapes. Prediction models for complex shapes will
contribute to the development of a general design approach.
Numerical tools are efficient for the description of the cross-sectional instability.
Numerically determined critical values are more accurate than the theoretical ones.
For mode identification, it is recommended to use the finite strip method rather than
the finite element method. The finite strip method defines mode decomposition,
which helps to recognize the occurrence and the prevalence of buckling modes.
Tests carried out in the current research are not satisfying in spite of the efforts
made. Initial imperfections of aluminium specimens could not be measured with
great accuracy. When an accurately measured real imperfection pattern of the spec-
imen is required, it is recommended to use a different test set-up, probably with a
more advanced way of scanning the complete surface of the specimen. For com-
pression tests, the length of the specimen was restricted due to the test set-up. In
further research, longer lengths should be considered, which allow to observe dis-
tortional buckling and combinations of local and distortional buckling. For distor-
tional buckling tests, it is recommended to apply additional fixed-end bearings at
the specimen’s edges. The fixed-end bearings are able to restrain axis rotations as
well as twist rotations and warping.
The current prediction model for distortional buckling of C-shaped specimens is
an extension of the existing model for local buckling prediction. The developed
prediction model has no limitations for any further extensions to more complex
cross-sectional shapes. Local-distortional and distortional-local buckling interaction
is described by the model, but have not been extensively investigated. It is recom-
mended to study the local-distortional (distortional-local) buckling interaction on
specific C-shaped profiles that have almost identical critical stresses for local and
distortional buckling.

122
Bibliography

S. Adany and B.W. Schafer. Buckling mode decompostition of single-branched open cross-
section members via finite strip method: derivation. Thin-Walled Structures, 44(5):563–584,
2006.

D. Altenpohl. Aluminium viewed from within. TU/e internal report, (02.14), 2002.

American Iron and Steel Institute (AISI). Direct Strength Method (DSM) Design Guide. CF06-
1, 2006.

A.C. Antoniou. On the maximum deflection of plating in newly built ships. Journal of Ship
Research, 24(1):31–39, 1980.

A.C. Antoniou, M. Lavidas, and G. Karvounis. On the shape of post-welding deformations


of plate pantes in newly built ships. Journal of Ship Research, 28(1):1–10, 1984.

R. Baehre. Comparison between structural behaviour of elastoplastic materials. Tekn. Dr. Arne
Johnson Ingenjorsbyra, 16, 1966.

R. Bebiano, P. Pina, N. Sivestre, and D. Camotim. GBTUL - a GBT-based code for thin-
walled member analysis. Proceedings of Fifth International Conference on Thin-Walled Struc-
tures (ICTWS 2008-Brisbane, 18-20/6), 2008.

H. Beer and G. Schultz. Theoretical basis for the European column curves. Construction
Metallique, (3), 1970.

A. Bernard, F. Frey, J. Janss, and Ch. Massonnet. Research on the buckling behaviour of
aluminium columns. IABSE Mem., 33-1:1–33, 1973.

G.H. Bryan. On the stability of a plane plate under thrust in its own plane. In Proc. of the
London Mathematical Society, volume 22, 1890.

D. Camotim, N. Silvestre, R. Gonçalves, and P.B. Dinis. GBT Analysis of Thin-Walled Mem-
bers: New Formulations and Applications. Thin-Walled Structures: Recent Advances and
Future Trends in Thin-Walled Structures Technology, pages 137–168, 2004.

C.A. Carlsen and J. Czujko. The specification of post-welding distortion tolerances for stiff-
ened plates in compression. The Structural Engineer, 56A(5):133–141, 1978.

123
CEN. EN1999-1-1, Eurocode 9: Design of aluminium structures - Part 1-1: General Structural
Rules. CEN, Brussels, 2007.

J.D. Clarke. Buckling of aluminium alloy stiffened plate ship structure. Proceedings of Interna-
tional Conference on Steel and Aluminium Structures, Cardiff, 1987.

J.M. Davies. Generalised Beam Theory (GBT) for Coupled Instability problems, part iv. Cou-
pled Instabilities in Metal Structures, 1998.

A. DeMartino, R. Landolfo, and F.M. Mazzolani. The use of the ramberg-osgood law for
materials of round-house type. Mater. Struct., 23:59–67, 1990.

T.P. Desmond. The behaviour and strength of thin-walled compression members with longi-
tudinal stiffeners. Report No. 369, 1977.

T.P. Desmond, T. Peköz, and G. Winter. Intermediate stiffeners for thin-walled members.
Journal of Structural Engineering (ASCE), 107(4):627–648, 1981.

P.B. Dinis and D. Camotim. Local-plate and distortional post-buckling behavior of cold-
formed steel columns: elastic and elastic-plastic fem analysis. Proceedings of SSRC annual
stability conference, pages 475–498, 2004.

P.B. Dinis, D. Camotim, and N. Silvestre. GBT formulation to analyse the first-order and
buckling behaviour of thin-walled members with arbitrary ’branched’ open cross-sections.
Thin-Walled Structures, 44(1):20–38, 2006.

H.S. Essa and D.L.J. Kennedy. Distortional buckling of steel beams. Structural Engineering
report No.185, 1993.

C. Faella, F.M. Mazzolani, V. Piluso, and G. Rizzano. Local buckling of aluminium members:
testing and classification. J. Struct. Eng., 126:353–360, 2000.

F. Frey and F.M. Mazzolani. Buckling behaviour of aluminium-alloy extruded members. In-
ternational Colloquium on Stability of Steel Structures, pages 85–94, 1977.

Y. Fukumoto, D.A. Nethercot, and T.V. Galambos. Experimental data for the buckling of steel
structures - NDSS Stability of metal structures. Proceedings of the 3rd International Coloquium
SSRC, pages 609–630.

T. V. Galambos. Guide of Stability design criteria for metal structures. John Wiley & Sons, USA,
5th edition, 1998.

G.L. Gallaher and R.B. Boughan. A method of calculating the compressive strength of Z-
stiffened panels that develop local instability. NACA Report TN-1482, Langley Memorial
Aeronautical Laboratory, (47), 1971.

G. Gerard and H. Becker. Handbook of Structural Stability. Part 1 - Buckling of Flat Plates. Tech-
nical note 3781, NACA, USA, 1957.

G.J. Hancock. Design of Cold-Formed Steel Structures to AS/NZS 4600:2005. Australian Steel
Institute, 4th edition edition, 2008.

G.J. Hancock. Local, distortional, and lateral bucklig of i-beams. Journal of the Structural
Division (ASCE), 104(11):1787–1798, 1978.

124
G.J. Hancock. Distortional buckling of steel storage rack columns. Journal of Structural Engi-
neering, 111(12):2770–2783, 1985.

G.J. Hancock. Design for distortional buckling of flexural members. Thin-Walled Structures,
27(1):3–12, 1997.

G.J. Hancock, A.J. Davis, P.W. Key, S.C.W. Lau, and K.J.R. Rasmussen. Recent developmens
in the buckling and nonlinear analysis of thin-walled structural members. Thin-Walled
Structures, 9(1-4):309–338, 1990.

G.J. Hancock, Y.B. Kwon, and E.S. Bernard. Strength design curves for thin-walled sections
undergoing distortional buckling. Journal of Construction Steel Research, 31(2-3):169–186,
1994.

A.N. Hendriks and J.B.M. Wolters. DIANA Finite Element Analysis. TNO DIANA BV, Delft,
The Netherlands, 9.2 edition, 2007.

T. Hoglund. Design of thin plate I girders in shear and bending with special reference to web buckling.
Royal Inst of Technology, Dept. of Building Statics and Structural Engineering, Stockholm,
1972.

O.S. Hopperstad. Modelling of cyclic plasticity with application to steel and aluminium struc-
tures. Department of Structural Engineering, 1993.

T. von Kármán, E.E. Sechler, and L.H. Donnell. The Strength of Thin Plates in Compression,
volume 54. Transactions ASME, 1932.

P.A. Kirby and D.A. Nethercot. Design for Structural Stability. Granada Publishing, UK, 1979.

M. Kmiecik, T. Jastrzebski, and J. Kuzniar. Statistics of ship plating distortions. Marine Struc-
tures, (8):119–132, 1995.

N. Kutanova and F. Soetens. Literature review on aluminium extrusions with complex cross-
sectional shapes. NIMR internal report, (P06.1.060), 2006.

N. Kutanova, F. Soetens, and H.H. Snijder. Selection procedure for cross-sectional shapes
appropriate to further experimental program. TU/e internal report, (0-2007.2), 2007.

N. Kutanova, F. Soetens, H.H. Snijder, and J. Mennink. Experimental investigation on struc-


tural behaviour of aluminium members with various cross-sectional shapes under com-
pression. TU/e internal report, (0-2008.7), 2008a.

N. Kutanova, F. Soetens, H.H. Snijder, and J. Mennink. Validating the finite element model.
extruded profiles. TU/e internal report, (0-2008.14), 2008b.

N. Kutanova, F. Soetens, H.H. Snijder, and J. Mennink. Validating the finite element model.
cold-formed profiles. TU/e internal report, (0-2008.24), 2008c.

N. Kutanova, F. Soetens, H.H. Snijder, and J. Mennink. Parameter study. TU/e internal report,
(0-2009.4), 2009.

Y.B. Kwon and G.J. Hancock. Tests of cold-formed channels with local and distortional buck-
ling. Journal of Structural Engineering (ASCE), 118(7):1786–1803, 1992.

125
S.C.W. Lau and G.J. Hancock. Distortional buckling formulas for channel columns. Journal of
Structural Engineering, 113(5):1063–1078, 1986.

E.E. Lundquist and E.Z. Stowell. Restraint provided a flat rectangular plate by a sturdy
stiffeer along an edge of the plate. NACA Report 735, Langley Memorial Aeronautical Lab-
oratory, 1942.

C.J. van der Maas. Charts for the calclation of the critical compressive stress for local instabil-
ity of columns with hat sections. Journal of the Aeronautical Sciences, 21(6):399–403, 1954.

F.M. Mazzolani. Aluminium Structural Design (CISM Course 443). Wien: Springer-Verlag, 2002.

F.M. Mazzolani. Characterisation of the law and buckling of aluminium columns. Constr.
Metall., 3, 1972.

F.M. Mazzolani. Proposal to classify the aluminium alloy on the basis of the mechanical
behavior. ECCS Committee 16, (Doc. 16-74-2), 1974.

F.M. Mazzolani. Aluminium alloy structures. Chapman & Hall, London, UK, 2nd edition, 1985.

F.M. Mazzolani. Aluminium Alloy Structures (second edition). E & FN SPON, an imprint of
Chapman & Hall, London, London SE1 8HN, UK, 1995. ISBN 0-419-17770-1.

F.M. Mazzolani, C. Faella, V. Piluso, and G. Rizzano. Experimental analysis of aluminium al-
loy SHS-members subjected to local buckling under uniform compression. 5th International
colloquim on structural stability, SSRC, Brazilian Session, pages 281–291, 1996.

F.M. Mazzolani, C. Faella, V. Piluso, and G. Rizzano. Local buckling of aluminium alloy RHS-
members: Experimental analysis. XVI Congresso C.T.A., Italian Conference on Steel Construc-
tion, 1997.

F.M. Mazzolani, V. Piluso, and G. Rizzano. Experimental analysis of aluminium alloy chan-
nels subjected to local buckling under uniform compression. Congresso C.T.A., Italian Con-
ference on Steel Construction, 2001.

J. Mennink. Preliminary FEM-simulations on aluminium stub columns. TU/e internal report,


(BCO-Report 99.09), 1999.

J. Mennink. Cross-sectional stability of aluminium extrusions. PhD Thesis, Eindhoven University


of Technology, The Netherlands, 2002.

J. Mennink and J. Schillemans. Experimental research on the stability behaviour of axially


compressed aluminium extrusions with complex cross-sectional shapes. TU/e internal re-
port, (02.14), 2002.

L.A. Moen. Rotational capacity of aluminium alloy beams. Department of Structural Engineer-
ing, (N-7034), 1999.

N.W. Murray. Introduction to the theory of thin-walled structures. Oxford University Press, UK,
1984.

R. Narayanan. Aluminium Structures: Advances, Design & Connections. Elsevier Applied Sci-
ence, UK, 1987.

126
NNI. Metalic materials - tensile testing - part 1: Methods of test (at ambient temperature).
(NEN 10002-1), 2001.

W. Ramberg and W.R. Osgood. Description of stress-strain curves by three parameters. NACA
Technical Note, 902, 1943.

J. Rhodes and J.M. Harvey. Effects of eccentricity of load or compression on the buckling and
postbuckling behaviour of flat plates. Int. J. Mech. Sci., 13:867–879, 1971.

A.J.C.B. De Saint-Venant. Théorie de l’elasticité des corps solides. Clebsch, 1883.

B.W. Schafer. CUFSM 3.11 Elastic buckling prediction. 2006.

B.W. Schafer and T. Peköz. Direct strength prediction of cold-formed steel members using
numberical elastic buckling solutions. 14th Int. Specialty Conf. on Cold-Formed Steel Struc-
tures, 1998.

R. Schardt. Verallgemeinerte Technische Biegetherie (Generalised Beam Theory). Springer Verlag,


Berlin, Heidelberg, 1989.

E.E. Sechler. Elasticity in engineering. Dover, 1968.

F.R. Shanley. Inelastic column theory. Journal of Aeronautical Science, 14(5), 1947.

M.L. Sharp. Longitudinal stiffeners for compression members. Journal of the Structural Division
(ASCE), 92(5):187–211, 1966.

Standards of Australia and Standards of New Zealand (SA-SNZ). Australian/New Zealand


Standard on Cold-Formed Steel Structures. AS/NZS 4600 (second edition), 2005.

G. Steinhardt. Aluminium in engineered construction. Aluminium, (47), 1971.

S.P. Timoshenko and M.J. Gere. Theory of elastic stability 2nd Ed. McGraw-Hill, USA, 1961.

V.Z. Vlasov. Thin-walled elastic bars. Moscow, 1940.

A.C. Walker. Design and analysis of cold-formed sections. International Textbook Company, UK,
1975.

G. Winter. Strength of Thin Steel Compression flanges. Cornel University Engineering Experi-
ment Station, 1947.

G. Winter. Performance of Thin Steel Compression Flanges, preliminary publication, 3rd Congress.
The Association for Bridge and Structural Engineering, 1948.

D. Yang and G.J. Hancock. Compression tests of high strength steel channel columns with
interaction between local and distortional buckling. Journal of Structural Engineering, 130
(12):1954–1963, 2004.

B. Young and K.J.R. Rasmussen. Measurement techniques in the testing of thin-walled struc-
tural members. Society for Experimental Mechanics, 43(1):32–38, 2003.

C. Yu and B.W. Schafer. Local bucklin tests on cold-formed steel beams. J. Struct. Eng., 129
(12):1596–1606, 2003.

127
C. Yu and B.W. Schafer. Distortional buckling tests on cold-formed steel beams. J. Struct. Eng.,
132(4):515–528, 2006.

W.W. Yu. Cold-formed steel design. John Wiley & Sons, USA, 3rd edition, 2000.

128
Appendix A
Experimental results

A.1 Measured dimensions

Table A.1: Measured dimensions ”Extruded” subprogram.

129
Table A.2: Measured dimensions, ”Cold-formed” subprogram

130
A.2 Measured imperfections

Table A.3: Measured imperfections ”Extruded” subprogram.

131
Table A.4: Measured imperfections, ”Cold-formed” subprogram

132
A.3 Measured material characteristics
Aluminium alloy EN AW-6060 T66, extruded profiles

Table A.5: Z-section, 2 mm thickness

Specimen E [N/mm2 ] f0.1 [N/mm2 ] f0.2 [N/mm2 ] fu [N/mm2 ]


2Z 66700 192.52 195.75 212.41

Table A.6: Z-section, 3 mm thickness

Specimen E [[N/mm2 ] f0.1 [N/mm2 ] f0.2 [N/mm2 ] fu [N/mm2 ]


3Z 66100 211.64 213.86 235.88

Table A.7: Square hollow section, 3 mm thickness

Specimen E [[N/mm2 ] f0.1 [N/mm2 ] f0.2 [N/mm2 ] fu [N/mm2 ]


B1 66800 207.29 210.84 236.53
B2 66600 208.84 212.00 237.33
B3 66800 202.81 206.29 233.64
B4 66960 207.45 210.24 236.32
B5 66672 200.34 203.66 230.98
B6 66882 208.14 211.08 237.00

133
Aluminium alloy EN AW-6082 T6, cold-formed profiles

Table A.8: Aluminium plate 2 mm thickness

Specimen E [N/mm2 ] f0.1 [N/mm2 ] f0.2 [N/mm2 ] fu [N/mm2 ]


6P2II 70700 275.56 286.91 335.54
6P2L 71900 291.21 302.08 347.51

Table A.9: Aluminium plate 2.5 mm thickness

Specimen E [N/mm2 ] f0.1 [N/mm2 ] f0.2 [N/mm2 ] fu [N/mm2 ]


6P2.5II 70500 257.36 267.14 320.99
6P2.5L 72200 261.11 270.73 326.19

Aluminium alloy EN AW-5083 H111, cold-formed profiles

Table A.10: Aluminium plate 2 mm thickness

Specimen E [N/mm2 ] f0.1 [N/mm2 ] f0.2 [N/mm2 ] fu [N/mm2 ]


5P2II 71700 159.60 161.48 316.10
5P2L 71600 155.62 155.38 301.70

Table A.11: Aluminium plate 2.5 mm thickness

Specimen E [N/mm2 ] f0.1 [N/mm2 ] f0.2 [N/mm2 ] fu [N/mm2 ]


5P2.5II 71375 158.57 160.19 314.50
5P2.5L 71300 157.22 157.38 303.66

134
Appendix B
Numerical analysis

B.1 Applied dimensions

Table B.1: Dimensions used in FE simulations (in mm) for extruded specimens.

Table B.2: Dimensions used in FE simulations (in mm) for cold-formed specimens.

135
B.2 Applied imperfections

Table B.3: Imperfections used in FE simulations (in mm) for extruded specimens.

Table B.4: Imperfections used in FE simulations (in mm) for cold-formed specimens.

136
Appendix C
Inelasticity coefficient derivation

C.1 Theory
The inelastic critical stress according to initial buckling (σcr;T ) can be determined from the
elastic critical stress (σcr ) and the inelasticity coefficient χT . Theoretical derivation of the
inelastic coefficient χT is given below.
By differentiation of the Ramberg-Osgood relation with respect to σ the tangential stiffness
ET of the material characteristic can be determined:
µ ¶n−1
∂ε 1 0.002n σ
= + (C.1)
∂σ E f0.2 f0.2
Inversion of this equation results:
∂σ 1
ET = = ³ ´n−1 (C.2)
∂ε 1 0.002n σ
E
+ f0.2 f0.2

According to Mennink’s model, Shanley’s approach is used (see Shanley [1947]). Shanley’s
approach assumes that inelastic buckling load (or stress) is obtained by replacing the elastic
modulus of elasticity with its tangential equivalent:

ET (σcr;T )
σcr;T = σcr (C.3)
E
The inelastic buckling coefficient χT can be introduced as a function of inelastic critical
stress and the 0.2% proof stress f0.2 :

σcr;T
σcr;T = χT (σcr )f0.2 or χT (σcr ) = (C.4)
f0.2
Instead of using the elastic critical stress, the relative plate slenderness λ̄ is proposed:
r
f0.2 f0.2
λ̄ = or σcr = 2 (C.5)
σcr λ̄
This results for elastic (σcr;T = σcr ) and, respectively, inelastic material:

1
χ= (C.6)
λ̄2
ET 1
χT = (C.7)
E λ̄2
Substitution of ET from eq. C.2 and σ = σcr;T yields to the following expression for inelastic
buckling coefficient determination:

1 1
χT = 0.002En n−1
(C.8)
1+ f0.2
χT λ̄2

137
C.2 Buckling curves for three aluminium alloys
For three considered materials, the relationship between buckling coefficient χT and plate
slenderness λ̄ can be exhibited in Figure C.1.

Figure C.1: Buckling curves for three aluminium alloys.

Material behaves elastically when the tangential stiffness ET = E. Thus, a part of the
buckling curve for higher range of the plate slenderness indicates elastic behaviour. Vertical
lines for three materials, where ET = E, distinguishes the elastic range from the inelastic
one. Hence, for the inelastic part yields: ET < E. For determination of the inelastic buckling
coefficient it is considered that the critical stress is laying in the inelastic range. Inelastic
range of the material can be characterized by the proportional limit fp and 0.2% proof stress
f0.2 . Therefore, the value of the critical stress belongs to one of the following intervals:
fp ≤ σcr < f0.2 or fp < f0.2 ≤ σcr . Taking into account expressions for the χ and λ̄ values,
different cases are distinguished:
r
1 ET 1 1 f0.2
χT = = n−1 2 , where λ̄ =
λ̄ E 1 + 0.002En
f
0.2
χT λ̄ σcr

1. σcr = fp The proportional limit is reached. At this point ET = E =⇒


s
f0.2 1 fp
λ= , χT = 2 = . (C.9)
fp λ̄ f0.2

2. fp < σcr < f0.2 Inelastic behaviour;

138
3. σcr = f0.2 The 0.2% proof stress is reached. Therefore, λ̄ = 1 =⇒
1
χT = 0.002En n−1
;
1+ f0.2
χT

χT f0.2 + 0.002Enχn T = f0.2 ;


µ ¶
0.002En
χnT + χT = 1;
f0.2
µ ¶
0.002En 1
χn−1
T +1= . (C.10)
f0.2 χT
4. σcr > f0.2 Strain-hardening behavior.
5. σcr;T = f0.2 Inelastic buckling stress is equal to the 0.2% proof stress.

σcr;T = χT (σcr )f0.2 =⇒ χT = 1 =⇒


1 1
χT = 0.002En n−1
= 1;
1+ f0.2
χT λ̄2
1 f0.2
=1
f0.2 + 0.002En λ̄2
s
f0.2
λ̄ = . (C.11)
f0.2 + 0.002En
6. σcr;T > f0.2 Plastic range, which is not considered.

139
C.3 Mennink’s approximation
Mennink proposed to apply a piecewise idealization of the material curve (see Mennink
[2002]). It is the simplest way to model the curves with two lines representing an
elastic-hardening diagram. The elastic part of the diagram is the Euler curve χT = λ̄12 , which
is shown in Figure C.2. Line approximation is drawn for the strain-hardening part of the
diagram for the smaller values of the plate slenderness. The intersection of the approximated
line with Euler curve defines the conventional value fp0 of the elastic limit of proportionality.

Figure C.2: Piecewise idealization of the curves.

According to the presented approximation, the non-linear material curve can be described
linearly. Mathematical expression for this approximated line can be found from two points,
where σcr;T = f0.2 (χT = 1) and σcr = f0.2 . These points were just mentioned above, see
Equations C.10 and C.11. The first point coordinates are known:
s
f0.2
χT = 1 and λ̄ = . (C.12)
f0.2 + 0.002En

For the second point where σcr = f0.2 the value of the χT coordinate can be found
graphically according to Equation C.10:

λ̄ = 1 and χ∗T is a single solution for two graphs:

140
1
f (χT ) = , where 0 < χT ≤ 1 (C.13)
χT
µ ¶
0.002En
and f (χT ) = χn−1
T +1 (C.14)
f0.2

Thus, the approximation for non-linear material curve can be expressed by:
s
χ∗T − 1 χ∗T − 1 f0.2
χT = q λ̄ − q · + 1, (C.15)
1− f0.2
1− f0.2 f 0.2+0.002En
f0.2+0.002En f0.2+0.002En

where χ∗T is a solution according to Equations C.13 and C.14.


Therefore, Mennink’s approximation results in the following expressions for considered
materials:

6060-T66 χT = 1.08 − 0.3λ̄


6082-T6 χT = 1.06 − 0.22λ̄
5083-H111 χT = 1.10 − 0.4λ̄

141
Appendix D
Parameter Study Results

Table D.1: Comparison of the FSM and Eigenvalue analysis results for C-shaped specimens
of the parameter study.

No. Specimen CUFSM FEM


Lcr;l σcr;l Lcr;d σcr;d Lcr;d σcr;d
[mm] [N/mm2 ] [mm] [N/mm2 ]] [mm] [N/mm2 ]
1 5(6)C5 - - 140 516 140 513
2 4.5(5)C5 - - 150 399 150 399
3 3.5(4)C5 - - 160 265 160 268
4 2.5(3)C5 70 270 170 160 170 166
5 1(2)C10 90 106 350 136 90 104
6 2(2.5)C5 90 166 180 119 180 124
7 1(2)C7.5 100 105 40 102 40 101
8 1.5(2)C5 100 109 200 84 200 87
9 1(2)C5 100 97 210 70 210 74
10 0.75(2)C5 40 78 240 57 240 60
11 0.75(1.5)C5 100 58 250 49 250 51

Table D.2: FEM results for C-shaped specimens of the parameter study (6082-T6).

No. Specimen Length CUFSM FEM


L Lcr;l σcr;l Lcr;d σcr;d Nu Md
[mm] [mm] [N/mm2 ] [mm] [N/mm2 ] [kN] [-]
1 5(6)C5 420 - - 140 516 291 DB
2 4.5(5)C5 450 - - 150 399 246 DB
3 3.5(4)C5 480 - - 160 265 178 DB
4 2.5(3)C5 510 70 270 170 160 100 DB
5 1(2)C10 1050 90 106 350 135 40 LB/DB
6 2(2.5)C5 540 90 166 180 119 69 DB/LB
7 1(2)C7.5 870 100 105 290 102 36 DB/LB
8 1.5(2)C5 600 100 109 200 84 44 DB/LB
9 1(2)C5 630 100 97 210 70 34 DB/LB
10 0.75(2)C5 720 40 78 240 57 30 DB/LB
11 0.75(1.5)C5 750 100 58 250 49 21 DB/LB

143
Table D.3: FEM results for C-shaped specimens of the parameter study (6060-T66).

No. Specimen Length CUFSM FEM


L Lcr;l σcr;l Lcr;d σcr;d Nu Md
[mm] [mm] [N/mm2 ] [mm] [N/mm2 ] [kN] [-]
1 5(6)C5 420 - - 140 516 193 DB
2 4.5(5)C5 450 - - 150 399 166 DB
3 3.5(4)C5 480 - - 160 265 124 DB
4 2.5(3)C5 510 70 270 170 160 78 DB
5 1(2)C10 1050 90 106 350 135 34 LB/DB
6 2(2.5)C5 540 90 166 180 119 54 DB
7 1(2)C7.5 870 100 105 290 102 33 DB/LB
8 1.5(2)C5 600 100 109 200 84 34 DB
9 1(2)C5 630 100 97 210 70 27 DB/LB
10 0.75(2)C5 720 40 78 240 57 24 DB/LB
11 0.75(1.5)C5 750 100 58 250 49 15 DB/LB

Table D.4: FEM results for C-shaped specimens of the parameter study (5083-H111).

No. Specimen Length CUFSM FEM


L Lcr;l σcr;l Lcr;d σcr;d Nu Md
[mm] [mm] [N/mm2 ] [mm] [N/mm2 ] [kN] [-]
1 5(6)C5 420 - - 140 516 141 DB
2 4.5(5)C5 450 - - 150 399 121 DB
3 3.5(4)C5 480 - - 160 265 91 DB
4 2.5(3)C5 510 70 270 170 160 59 DB
5 1(2)C10 1050 90 106 350 135 27 LB/DB
6 2(2.5)C5 540 90 166 180 119 43 DB
7 1(2)C7.5 870 100 105 290 102 26 DB/LB
8 1.5(2)C5 600 100 109 200 84 27 DB
9 1(2)C5 630 100 97 210 70 22 DB
10 0.75(2)C5 720 40 78 240 57 18 DB
11 0.75(1.5)C5 750 100 58 250 49 12 DB

144
Figure D.1: Compressed specimens 6082 aluminium alloy.

Figure D.2: Compressed specimens 6082 aluminium alloy.

145
Figure D.3: Compressed specimens 6060 aluminium alloy.

Figure D.4: Compressed specimens 6060 aluminium alloy.

146
Figure D.5: Compressed specimens 5083 aluminium alloy.

Figure D.6: Compressed specimens 5083 aluminium alloy.

147
Appendix E
Prediction model results

Table E.1: Prediction model results for C-profiles of the parameter study (6082-T6).

No. Specimen Material Input Model


fp E ApgI ApgII σcr;l σcr;d Nu;pr
[N/mm2 ] [N/mm2 ] [mm2 ] [mm2 ] [N/mm2 ] [N/mm2 ] [kN]
1 5(6)C5 185 70000 550 600 - 516 261
2 4.5(5)C5 185 70000 495 500 - 399 220
3 3.5(4)C5 185 70000 385 400 - 265 166
4 2.5(3)C5 185 70000 275 300 270 160 101
5 1(2)C10 185 70000 120 200 106 135 38
6 2(2.5)C5 185 70000 220 250 166 119 73
7 1(2)C7.5 185 70000 115 200 105 102 39
8 1.5(2)C5 185 70000 165 200 109 84 43
9 1(2)C5 185 70000 110 200 97 70 35
10 0.75(2)C5 185 70000 82.5 200 78 57 29
11 0.75(1.5)C5 185 70000 82.5 150 58 49 20

Table E.2: Prediction model results for C-profiles of the parameter study (6060-T66).

No. Specimen Material Input Model


fp E ApgI ApgII σcr;l σcr;d Nu;pr
[N/mm2 ] [N/mm2 ] [mm2 ] [mm2 ] [N/mm2 ] [N/mm2 ] [kN]
1 5(6)C5 98 70000 550 600 - 516 167
2 4.5(5)C5 98 70000 495 500 - 399 141
3 3.5(4)C5 98 70000 385 400 - 265 105
4 2.5(3)C5 98 70000 275 300 270 160 71
5 1(2)C10 98 70000 120 200 106 135 33
6 2(2.5)C5 98 70000 220 250 166 119 51
7 1(2)C7.5 98 70000 115 200 105 102 31
8 1.5(2)C5 98 70000 165 200 109 84 34
9 1(2)C5 98 70000 110 200 97 70 28
10 0.75(2)C5 98 70000 82.5 200 78 57 24
11 0.75(1.5)C5 98 70000 82.5 150 58 49 15

149
Table E.3: Prediction model results for C-profiles of the parameter study (5083-H111).

No. Specimen Material Input Model


fp E ApgI ApgII σcr;l σcr;d Nu;pr
[N/mm2 ] [N/mm2 ] [mm2 ] [mm2 ] [N/mm2 ] [N/mm2 ] [kN]
1 5(6)C5 54 70000 550 600 - 516 116
2 4.5(5)C5 54 70000 495 500 - 399 97
3 3.5(4)C5 54 70000 385 400 - 265 73
4 2.5(3)C5 54 70000 275 300 270 160 49
5 1(2)C10 54 70000 120 200 106 135 24
6 2(2.5)C5 54 70000 220 250 166 119 37
7 1(2)C7.5 54 70000 115 200 105 102 24
8 1.5(2)C5 54 70000 165 200 109 84 25
9 1(2)C5 54 70000 110 200 97 70 20
10 0.75(2)C5 54 70000 82.5 200 78 57 16
11 0.75(1.5)C5 54 70000 82.5 150 58 49 12

150
Samenvatting

Aluminium extrusieprofielen zijn van belang voor verschillende industriële sectoren met
constructieve toepassingen zoals de bouw- en de transportsector. Met het extrusieproces
kunnen structurele elementen relatief gemakkelijk geoptimaliseerd worden om te voldoen
aan de ontwerpeisen. Optimalisatie van de doorsnedevorm van aluminium elementen
resulteert vaak in het gebruik van dunwandige doorsneden , waardoor de complexiteit van
de doorsnede toeneemt. Als gevolg van die dunwandige elementen heeft doorsnede
instabiliteit - met name plooi en ”distortional buckling” -een aanzienlijk effect op de
constructieve gedrag.
Bij de classificatie van doorsnede instabiliteit impliceert plooi veranderingen in geometrie
zonder zijdelingse verplaatsing of verdraaiing, terwijl voor distortional buckling zijdelingse
verplaatsing en verdraaiing plaatsvinden met veranderingen in de doorsnede geometrie. De
huidige rekenregels gebruikt door ingenieurs zijn beperkt tot plooi van simpele en
symmetrische doorsneden. Daar komt bij dat deze ontwerpregels niet voorzien in een
nauwkeurige beschrijving van distortional buckling gedrag en niet gebruikt kunnen worden
voor meer complexe doorsnedevormen. Uitgebreid onderzoek naar doorsnede instabiliteit
van aluminium structurele elementen betreffende distortional buckling is vereist, hetgeen
het belangrijkste onderwerp van dit proefschrift is.
Een experimenteel programma is uitgevoerd om de uiterste draagkracht van aluminium
structurele elementen te voorspellen in geval van doorsnede instabiliteit. Dit
proevenprogramma bestaat uit axiale drukproeven op geëxtrudeerde en koudgevormde Z-,
Hoek- en C-profielen. Uitgebreide metingen zijn uitgevoerd op de initiële imperfecties van
geextrudeerde en koudgevormde proefstukken. De invloed van een geleidelijke toename
van de complexiteit van de geometrie en materiële invloed op het knik gedrag zijn
onderzocht. Het experimentele programma heeft geresulteerd in een set van gegevens over
doorsnede instabiliteit van aluminium structurele elementen met verschillende
doorsnedevormen, met inbegrip van plooi en distortional buckilng, alsook de
interactievormen. Deze gegevens zijn gebruikt voor de validatie van het numerieke model.
Een eindige-elementenmodel is ontwikkeld en gevalideerd op basis van de resultaten van
het experimentele programma. Daartoe zijn de experimenten gesimuleerd met toepassing
van de gemeten geometrie, imperfecties en materiaaleigenschappen. De vergelijking tussen
het eindige-elementenmodel en de testresultaten geeft de nauwkeurigheid van de
numerieke voorspelling aan. Het is aangetoond dat het eindige-elementenmodel een nuttig
middel is voor de voorspelling van het stabiliteitsgedrag van gelijkmatig gedrukte
aluminium proefstukken met diverse doorsnedevormen. Het gevalideerde

151
eindige-elementenmodel is gebruikt voor een nauwkeurig onderzoek van het
daadwerkelijke distortional buckling gedrag en de plooi-distortional (distortional-plooi)
interactie.
Een aanzienlijk aantal van analyses zijn uitgevoerd om het distortional buckling fenomeen
met behulp van het gevalideerde eindige-elementenmodel te onderzoeken. Aan de hand
van de numerieke resultaten van het daadwerkelijke distortional buckling gedrag van de
C-vormige profielen is een voorspellingsmodel opgesteld voor de berekening van de
uiterste draagkracht van C-vormige drukprofielen die zijn onderworpen aan doorsnede
instabiliteit. Daarmee zijn het bestaande model voor plooi voorspelling , het nieuw
ontwikkelde model voor distortional buckling voorspelling en de stabiliteitsinteractie in
staat een nauwkeurige beschrijving te geven van de doorsnede instabiliteit van aluminium
C-profielen. Daarnaast zijn deze modellen een belangrijke stap in het onderzoek van de
doorsnede instabiliteit van complexe aluminium doorsnedevormen.
Curriculum Vitae

Personal data
Name Natalia Kutanova
Nationality Russian
Date of birth May 16th, 1981
Place of birth Kavalerovo, Primorsk region, Russia
Present address Valkenboslaan 127, 2563CJ The Hague, The Netherlands
E-mail natalia.kutanova@gmail.com

Education
1998-2004 Saint-Petersburg State University
Faculty of Applied Mathematics and Mechanical Science
Masters degree in Mechanical Science
Thesis ”Delayed fracture of elasto-viscous substances under
the influence of UV-radiation and elevated temperatures”

1996-1998 Apatity High School # 2, Murmansk region, Russia

PhD research
2004-2009 Eindhoven University of Technology
Faculty of Architecture, Building and Planning
Department of Structural Design

You might also like