Download as pdf or txt
Download as pdf or txt
You are on page 1of 1373

STABILITY AND DUCTILITY OF STEEL STRUCTURES 2019

PROCEEDINGS OF THE INTERNATIONAL COLLOQUIA ON STABILITY AND DUCTILITY


OF STEEL STRUCTURES, PRAGUE, CZECH REPUBLIC, SEPTEMBER 11-13, 2019

Stability and Ductility of Steel


Structures 2019

Editors
František Wald & Michal Jandera
Czech Technical University in Prague, Czech Republic
CRC Press/Balkema is an imprint of the Taylor & Francis Group, an informa business

© 2019 Czech Technical University in Prague, Czech Republic

Typeset by Integra Software Services Pvt. Ltd., Pondicherry, India

All rights reserved. No part of this publication or the information contained herein may be
reproduced, stored in a retrieval system, or transmitted in any form or by any means,
electronic, mechanical, by photocopying, recording or otherwise, without written prior
permission from the publisher.

Although all care is taken to ensure integrity and the quality of this publication and the
information herein, no responsibility is assumed by the publishers nor the author for any
damage to the property or persons as a result of operation or use of this publication and/or
the information contained herein.

Published by: CRC Press/Balkema


Schipholweg 107C, 2316XC Leiden, The Netherlands
e-mail: Pub.NL@taylorandfrancis.com
www.crcpress.com – www.taylorandfrancis.com

ISBN: 978-0-367-33503-8 (Hbk)


ISBN: 978-0-429-32024-8 (eBook)
DOI: https://doi.org/10.1201/9780429320248
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Table of contents

Foreword xv
Committees xvii

Keynote lectures
Structural steel design by advanced analysis with strain limits 3
L. Gardner, A. Fieber & L. Macorini
Cold-formed high strength steel RHS under combined bending and web crippling 16
H.T. Li & B. Young
Stability design of steel structures: From members to plates and shells 28
L.S. da Silva, T. Tankova & J.P. Martins
Advancements in the stability design of steel frames considering general nonprismatic
members and general bracing conditions 42
D.W. White, R. Slein & O. Toğay
Design by advanced analysis–2016 AISC specification 55
R.D. Ziemian & Y. Wang

Regular papers
On the modal buckling of longitudinally stiffened plates 71
S. Adany & M.Z. Haffar
Strength characterisation of a CFS section with initial geometric imperfections 80
H.S.S. Ahmed & S. Ghosh
Shear behaviour of sandwich panel fasteners in fire 88
T. Arha, K. Cábová, N. Lišková & F. Wald
Bracing details for trapezoidal steel box girders 96
S.V. Armijos-Moya, Y. Wang, T. Helwig, M. Engelhardt, E. Williamson & P. Clayton
Behaviour of slender plates in case of fire of different stainless steel grades 106
F. Arrais, N. Lopes, P. Vila Real & C. Couto
Numerical modelling of cold-formed steel members at elevated temperatures 115
F. Arrais, N. Lopes, P.Vila Real & M. Jandera
Experimental study on the general behaviour of stainless steel frames 124
I. Arrayago, E. Real, E. Mirambell & I. González de León
Experimental investigation of flexural buckling of sandwich panels with steel facings 133
I. Balázs & J. Melcher
Nonlinear behavior and instability of deployable arches 139
A.B.G. Barcellos, M.V.B. Santana & P.B. Gonçalves
Numerical advanced analysis of steel-concrete composite beams and columns under fire 147
R.C. Barros, R.A.M. Silveira, P.A.S. Rocha, D. Pires & Í.J.M. Lemes

v
Buckling of spatial laced columns composed of built-up cold-formed channel members 155
C.C.D.O. Bastos & E.M. Batista
Local-distortional buckling interaction of cold-formed steel columns design approach 164
E.M. Batista, G.Y. Matsubara & J.M.S. Franco
Solutions to simplified von Karman plate equations 173
J. Becque
Simplified method for lateral torsional buckling of beams with lateral restraints 181
A. Beyer & A. Bureau
Buckling resistance of mono-symmetric I-/H-section members in biaxial bending, axial
compression, and torsion 189
A.-L. Bours, R. Winkler & M. Knobloch
Cyclic plastic behavior of steel material under uniaxial load paths 197
V. Budaházy & L. Dunai
Simulation based imperfections and their effects on stability resistance 205
V. Budaházy, D. Kollár & L.G. Vigh
Development of an innovative multi-performance system for LWS structures 213
A. Campiche
Seismic design criteria for CFS steel-sheathed shear walls 221
A. Campiche, S. Shakeel, L. Fiorino & R. Landolfo
Behaviour of a concrete slab in compression in composite steel-concrete frame joints 229
P. Červenka & J. Dolejš
On the development of IoT platforms for the detection of damage in steel railway bridges 235
R. Chacón, A. Rodriguez, P. Sierra, X. Martínez & S. Oller
Experimental and numerical studies on shear behaviour of stainless steel plate girders with
inclined stiffeners 244
X.W. Chen, H.X. Yuan, X.X. Du & E. Real
Ultimate strength analysis of steel-concrete cross-sections at elevated temperatures 253
C.G. Chiorean, M. Selariu & S.M. Buru
Seismic design of two-storey X-bracings 262
S. Costanzo, M. D’Aniello, G. Di Lorenzo, A. De Martino & R. Landolfo
Experimental tests on bolted end-plate connections using thermal insulation layer attached
to steel structures 269
M. Couchaux, A. Alhasawi & A. Ben Larbi
Assessment of Eurocode fire design rules for structural members made of high strength steels 277
C. Couto & P. Vila Real
Numerical investigation on thin-walled CFS columns in fire 286
H.D. Craveiro, J. Henriques, A. Santiago, L. Laím & J. Henriques
FEM analysis of the buckling behavior of thin-walled CFS columns. Part I—channel (C) and
Double Channel (I) cross-sections 295
H.D. Craveiro, L.S. da Silva & J.P. Martins & J. Henriques
FEM analysis of the buckling behaviour of thin-walled CFS columns.
Part II—monosymmetric (R) and double symmetric built-up box cross-sections 303
H.D. Craveiro, J. Henriques, L.S. Silva & J.P. Martins

vi
Comparison of two different innovative solutions for IPE beam to CHS column connections 312
R. Das, A. Kanyilmaz & H. Degee
Performances of moment resisting frames with slender steel and composite sections in low
and moderate seismic areas 321
H. Degée, Y. Duchêne & B. Hoffmeister
Laser technology for innovative connections in steel construction: An overview of the project
LASTEICON 329
H. Degée, A. Kanyilmaz, C. Castiglioni, L. Calado, M. Couchaux & B. Hoffmeister &
F. Morelli
Analytical assessment of CFS wall-panels sheathed with MgO board 337
A. Dewangan, G. Bhatt & C. Sonkar
Direct Strength Method (DSM) design of simply supported short-to-intermediate hot-rolled
steel equal-leg angle columns 345
P.B. Dinis & D. Camotim
Stability of ring stiffened steel liners under external pressure—comparison of the existing
design concept with 3D-FEM analysis 354
A. Ecker & H. Unterweger
Numerical investigation of steel built-up columns composed of track and channel
cold-formed sections 363
M.A. El Aghoury, E.A. Amoush, A.M. El Hady & S.M. Ibrahim
Fatigue failure of skew beam grid steel bridges—causes and assessment 371
M.A. El Aghoury, I.M. El Aghoury & Amr M. El Hady
Lateral torsional buckling of hybrid steel–glass beams 379
M. Eliasova & I. Pravdova
A yielding criterion for seismic gusset plates in tension 388
M.D. Elliott & L.H. Teh
Experimental calibration of centrally loaded built-up battened compression members 394
G.M. El-Mahdy
Stainless steel fillet weld tests 402
N. Feber, M. Jandera, L. Forejtova & L. Kolarik
Experimental investigation of compressed stainless steel angle columns 409
A. Filipović, J. Dobrić, Z. Marković, M. Spremić, N. Fric & N. Baddoo
Accurate and efficient account of geometrical imperfections in Koiter analysis of elastic
solid-like shells 417
G. Garcea, F.S. Liguori, L. Leonetti, D. Magisano & A. Madeo
Ductility of different types of shear connectors—experimental and numerical analysis 427
N. Gluhović, M. Spremić, B. Milosavljević, Z. Marković & J. Dobrić
GBT-based semi-analytical solutions for the elastic/plastic stability analysis of stainless steel
thin-walled columns exposed to fire 434
R. Gonçalves, R. Marçalo Neves & D. Camotim
A comparative analysis on the stability and ultimate strength of steel plated girders with
planar and corrugated webs 443
A. González, L. Vallelado & M.A. Serna

vii
Experimental testing of plastic buckling of moderately thick circular rings under uniform
lateral loading 452
F. Guarracino
Ultimate bending resistance of pipes: Testing arrangements and design approaches.
A multi-year perspective 460
F. Guarracino
On the GNI analysis of simple thin-walled beams with using linear buckling mode as
geometric imperfection 468
M.Z. Haffar, M.H. Taher & S. Ádány
U-shaped steel plate dissipative connection for concentrically braced frames 476
J. Henriques, L. Calado, C.A. Castiglioni & H. Degée
Critical buckling load on transversally and longitudinally stiffened steel plate girders
subjected to patch loading 483
J. Herrera & R. Chacón
Buckling behavior and strength of corroded steel shapes under axial compression 491
K. Hisazumi & R. Kanno
Modal analysis of thin-walled members with transverse plate elements using the constrained
finite element method 499
T. Hoang & S. Ádány
Experimental verification of shear connection of thin-walled steel built-up members 508
M. Horáček, J. Melcher & O. Ceh
Buckling analysis of circular arches with trapezoidal corrugated web 515
J.R. Ibañez, R. Díez, C. López & M.A. Serna
Effect of stiffener position on buckling behavior of H-shaped steel beam with upper flange
restraint 523
N. Igawa & K. Ikarashi
Local buckling strength of vertical haunch H-shaped beam under shear bending 531
W. Ishida & K. Ikarashi
Imperfection sensitivity of corrugated web girders subjected to lateral-torsional buckling 539
B. Jáger, M. Kachichian, L. Dunai & C. Égető
Enhanced buckling capacity of axially compressed stiffened plates taking into account the
shear-lag effect 548
A. Jäger-Cañás
Axial buckling behavior of welded ring-stiffened shells 556
A. Jäger-Cañás, Z. Li & H. Pasternak
Ductility assessment of structural steel and composite joints 564
J.-P. Jaspart, A. Corman & J.-F. Demonceau
Study on the influence of reduced beam sections on the seismic behaviour of a moment
resisting frame 570
A. Jiménez, E. Mirambell & E. Real
Stability of double-symmetric sections subjected to axial force, bending moments and torsion 578
F. Jörg & U. Kuhlmann

viii
Distortional buckling of stiffeners in stainless steel profiled sheeting 587
J. Jůza & M. Jandera
Appropriate spring stiffness models for the end support of bolted single steel angle members
in compression 596
M. Kettler, H. Unterweger & T. Harringer
Global buckling strength of built-up cold-formed steel column under compression 605
T. Kobashi & N. Shimizu
Coupled buckling of steel LC-beams under bending 614
Z. Kołakowski, T. Kubiak & M. Kamocka
Crashworthiness performance of tubular energy absorbing structures with triggers 622
M. Kotełko, M. Ferdynus & K. Okoń
Failure plastic mechanisms in TWCFS columns under eccentric compression 629
M. Kotełko, V. Ungureanu & D. Dubina
Experimental study of beam-to-column connection with bolted joints 638
Y. Koyama, A. Sato, H. Idota, Y. Sato, S. Yagi, S. Takaki & M. Kamada
Material strength statistics and reliability aspects for the reassessment of end-of-service-life
steel bridges 646
R. Kroyer & A. Taras
The influence of stiffeners width on buckling modes of steel LC-beams subjected to bending 655
T. Kubiak, Z. Kolakowski & F. Kazmierczyk
Design of slender compressed plates in structural steel joints by component based finite
element method 664
M. Kuříková, F. Wald & J. Kabeláč
Experimental verification of lateral-torsional buckling of steel I-beams with tapered flanges 673
J. Kuś
Composite floor system with cold-formed trussed beams and pre-fabricated concrete slab 682
L.A.A. de S. Leal, E. de M. Batista
Experimental investigation of stability behavior of members supported by sandwich panels at
elevated temperature 691
A. Lendvai & A.L. Joó
Ultimate shear resistance of cylindrically curved steel panels 699
F. Ljubinković, J.P. Martins, H. Gervásio & L.S. Silva
Nonlinear finite element analysis of delta hollow flange girders subjected to patch loading 708
N. Loaiza, C. Graciano & E. Casanova
Critical loads of semi-rigid columns subjected to non-linear temperature distributions 717
T. Ma & L. Xu
The stability of semi-braced steel frames containing members with stepped segments 727
T. Ma & L. Xu
A quasi-static nonlinear analysis for assessing the fire resistance of steel 3D frames exploiting
time-dependent yield surfaces 735
D. Magisano, F. Liguori, L. Leonetti & G. Garcea
Fire design of class 4 tapered steel beams with the general method—a proposal 744
É. Maia, C. Couto, P. Vila Real & N. Lopes

ix
Progressive collapse assessment of storage racks due to localized failures. Explicit
consideration of dynamic effects 753
I. Marginean, F. Dinu & D. Dubina
Elastic buckling strength of lipped channel section beam restrained on upper flange subjected
to bending 761
H. Masuda & K. Ikarashi
Experimental study on SCFs of empty SHS-SHS T-joints subjected to static out-of-plane
bending 770
F.N. Matti & F.R. Mashiri
Experimental study of cold-formed high strength steel circular hollow sections 778
X. Meng & L. Gardner
Towards automated identification of structural steel components from 3D-point clouds to
subsequent GMNA-stability-analysis 787
C. Merkl & A. Taras
Static effects of modular structures made of containers 795
O. Miller, V. Křivý, D. Mikolášek, P. Pařenica & R. Cuřín
Calibration of European web-crippling equations for cross-sections with one web 804
T. Misiek & A. Belica
Explanatory notes to buckling design of longitudinally welded aluminium compression
members 813
T. Misiek, B. Norlin & T. Höglund
Buckling of circular hollow section stainless steel columns in fire 821
A. Mohammed & S. Afshan
The capacity of bolted cold-formed steel connections in bending 830
S.M. Mojtabaei, J. Becque & I. Hajirasouliha
Study on the deformation and rotation capacity of HSS hollow sections 839
A. Mueller & A. Taras
An analytical solution for the compressed simply-supported plate with initial geometric
imperfections 847
M. Nedelcu
Experimental study on the shear connections of composite girders with trapezoidally
corrugated web 854
G. Németh, B. Jáger, N. Kovács & B. Kövesdi
Numerical study of end-plate steel connections with two and four bolts-per-row 863
D.L. Nunes & A. Ciutina
Experimental study on square steel tubular columns under compressive force with biaxial
bending moment 872
T. Onogi & A. Sato
Considering realistic weld imperfections in load bearing capacity calculations of ring-
stiffened shells using the analytical numerical hybrid model 882
H. Pasternak, Z. Li, C. Stapelfeld & B. Launert, A. Jäger-Cañás
Seismic response of steel dual eccentrically braced frames with equal-strength joints 890
Č. Penelov & N. Rangelov

x
Tests and design of built-up section columns 898
D.K. Phan & K.J.R. Rasmussen
Buckling and strength of prestressed steel stayed columns 906
R. Pichal & J. Machacek
Numerical simulation and analysis of axially restrained stainless steel beams in fire 911
A. Pournaghshband, S. Afshan & M. Theofanous
Experimental and numerical investigations of unstiffened steel girders with non-rectangular
panels subjected to bending and shear 921
V. Pourostad & U. Kuhlmann
Studying bolt force distribution in ultra-large capacity end-plate connections 929
A.A. Ramzi, I.M. El Aghoury, S.M. Ibrahim & A.I. El-Serwi
Sensitivity of the stiffness reduction model used to analyze the ultimate load condition of
steel frames 937
B. Rosson, T. Villalon-Camacho, H. Gurneian & R. Ziemian
Statistical evaluation of the bearing capacity of short polygonal columns 946
G. Sabau, E. Koltsakis, O. Lagerqvist & P. Manoleas
Direct Strength Method (DSM) design of end-bolted cold-formed steel columns failing in
distortional modes 954
W.S. Santos, A. Landesmann & D. Camotim
Design limitations for the steel beam-column to ensure full plastic moment 963
A. Sato, M. Aoyama, K. Inden, T. Ono & K. Mitsui
Experimental study on LTB behaviour and residual stresses of welded I-section members 972
L. Schaper, R. Winkler, F. Jörg, U. Kuhlmann & M. Knobloch
Behavior of column base plates under bi-axial bending moment 981
L. Seco, M. Couchaux, M. Hjiaj & L.C. Neves
Quantifying the seismic ductility of lightweight steel lateral force resisting systems through
procedures of FEMA P695 989
S. Shakeel
Numerical modelling of a two storey LWS building braced with gypsum-based panels 997
S. Shakeel, A. Campiche & R. Landolfo
Elastic buckling strength of H-shaped beams subjected to end moment and uniformly
distributed load 1005
D. Shinohara & K. Ikarashi
On the incorporation of cross-section restraints in Generalised Beam Theory (GBT) 1015
T.G. da Silva, C. Basaglia & D. Camotim
Stability interaction effects in 3D steel frames—a case study 1025
H.H. Snijder, L.H.J.D. Titulaer, P.A. Teeuwen & H. Hofmeyer
Experimental investigation on the instability phenomenon in stainless steel connections—plate
curling 1034
K. Sobrinho, A. Tenchini, M. Cordeiro, P. Vellasco, L. Lima & J. Henriques
Experimental and analytical study of Cold-Formed Steel (CFS) single-stud walls sheathed
with FCB, CSB and MgO under compression 1042
C. Sonkar, A. Kr. Mittal, S. Kr. Bhattacharyya, S. Kumar & A. Dewangan

xi
Lateral-torsional buckling of stainless steel beams with slender cross section 1051
M. Šorf & M. Jandera
Improve load capacity calculations by considering realistic imperfections induced by welding
for plates and shells 1059
C. Stapelfeld, B. Launert, H. Pasternak, N. Doynov & V.G. Michailov
Plastic collapse loads of rectangular plate assemblies with constant and linear load
distribution 1068
S. Stehr & N. Stranghöner
Stability of axially compressed cylindrical shells made of stainless steel for different
imperfection patterns 1077
N. Stranghöner & E. Azizi
Proposal for improving the consistency between Eurocode 3-1-8 and Eurocode 8-1 1086
A. Stratan & D. Dubina
Assumption of imperfections for the LTB-design of members based on EN 1993-1-1 1095
R. Stroetmann & S. Fominow
Welds on high-strength steels—influence of the welding process and the number of layers 1103
R. Stroetmann & T. Kästner
Validation of the Overall Stability Design Methods (OSDM) for tapered members 1111
J. Szalai, F. Papp & G. Hajdú
Stability design of cable-stayed columns 1120
T. Tankova, L.S. da Silva & J.P. Martins
Influence of geometrical imperfection of rib stiffeners on beam-to-column joint behaviour 1128
R. Tartaglia, M. D’Aniello, G.M.Di Lorenzo & R. Landolfo
The fire behaviour of extended stiffened joints designed for seismic actions 1136
R. Tartaglia, M. Zimbru, A. Linguiti, M. D’Aniello, R. Landolfo & F. Wald
Buckling length assessment with finite element approach 1145
T. Tiainen, K. Mela & M. Heinisuo
Experimental and numerical analysis of the local and interactive buckling behaviour of
hollow sections 1151
A. Toffolon, A. Müller, A. Taras & I. Niko
Proposal of a design curve for the overall resistance of cold-formed rectangular and square
hollow sections 1159
A. Toffolon & A. Taras
Behavior of extended end-plate connections under cyclic alternate loading 1167
I.C. Tomăscu, R.M. Bâlc
Thermorheological testing and modelling of seismic bearing elastomers 1176
C. Treib, M.A. Kraus & A. Taras
Built-up cold-formed steel beams with web openings 1185
V. Ungureanu, I. Both, C. Neagu, M. Burca, D. Dubina & A.A. Cristian
Numerical investigation of built-up cold-formed steel beams with corrugated web 1193
V. Ungureanu, I. Lukačević, I. Both, M. Burca & D. Dubina
Study on the out-of-plane stability of steel portal frames 1201
M. Vassilev & N. Rangelov

xii
Warping transfer superelement model for bolted end-plate connections subject to 3D loads 1210
B. Vaszilievits-Sömjén, J. Szalai & R.M. Movahedi
Tests of gusset plate connection under compression 1218
J. Vesecký, K. Cábová & M. Jandera
Numerical modelling of gusset plate connections under eccentric compression 1227
J. Vesecký, M. Jandera & K. Cábová
Buckling of columns during welding 1236
M. Vild & M. Bajer
Design of gusset plate connection with single-sided splice member by component based finite
element method 1243
M. Vild, J. Kabeláč, M. Kuříková & F. Wald
Beam-to-column joints for slim-floor systems in seismic zones: Numerical investigations and
experimental program 1251
C. Vulcu, R. Don & A. Ciutina
A reexamination of high strength steel Q690 plasticity model 1260
Y. Wang, Y. Wang, G. Li & Y. Lyu
Analysis of mechanical properties of cold formed high strength steel in weld area 1269
M. Werunský & J. Dolejš
Degradation processes in normalized mild- and low-alloy steel building structures in service 1275
W. Wichtowski & J. Hołowaty
Effect of the steel grade on equivalent geometric imperfections for lateral torsional buckling 1283
R. Winkler & M. Knobloch
Influence of collision damage on load-carrying capacity of steel girder 1292
E. Yamaguchi, Y. Tanaka & T. Amamoto
Modelling of one-sided unstiffened beam-to-column joint 1300
J. Zamorowski & G. Gremza
Modelling of roof bracings of single-storey industrial buildings 1309
J. Zamorowski & G. Gremza
Buckling assessment of cylindrical steel tanks with top stiffening ring under wind loading 1317
Ö. Zeybek & C. Topkaya
Slim-floor beam bending moment resistance considering partial shear connection 1326
Q. Zhang & M. Schäfer
Stainless steel SHS and RHS beam-columns 1334
B. Židlický & M. Jandera
Calibration of parameters of combined hardening model using tensile tests 1342
C.I. Zub, A. Stratan & D. Dubina

Author index 1351

xiii
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Foreword

The series of International Colloquia on Stability and Ductility of Steel Structures have been
supported by the Structural Stability Research Council (SSRC) for more than forty years and
its objective is to present the progress in theoretical, numerical and experimental research in
the field of stability and ductility of steel and steel-concrete composite structures. Special
emphasis is laid on new concepts and procedures concerning the analysis and design of steel
structures and on the background, development and application of rules and recommenda-
tions either appearing in recently published Codes or Specifications or about to be included in
their upcoming versions. This International Colloquium series started in 1972 in Paris and its
subsequent editions took place in different cities with the last five being held in: Timisoara,
Romania (1999), Budapest, Hungary (2002), Lisbon, Portugal (2006), Rio de Janeiro, Brazil
(2010) and Timisoara, Romania (2016).
The 2019 edition of SDSS is organized by the Czech Technical University in Prague. The
university held the second edition of the Eurosteel conference in 1999 and the first three edi-
tions of Applications of Structural Fire Engineering (ASFE) conference (2009, 2011 and
2013).

František Wald

Michal Jandera

xv
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Committees

ORGANIZING COMMITTEE

K. Cábová
J. Dolejš
M. Eliášová
M. Jandera
M. Kuříková
J. Macháček
P. Ryjáček
Z. Sokol
M. Šorf
F. Wald
B. Židlický

SCIENTIFIC COMMITTEE

Chairman: František Wald


Scientific Secretary: Michal Jandera

G.A. Altay (Turkey)


I. Balaz (Slovakia)
C. Baniatopoulos (Greece)
A. Bureau (France)
E.M. Batista (Brazil)
R. Beale (UK)
R. Bjorhovde (USA)
M.A. Bradford (Australia)
B. Brune (Germany)
L. Calado (Portugal)
D. Camotim (Portugal)
R. Chacon (Spain)
S.L. Chan (Hong Kong, China)
T.M. Chan (Hong Kong, China)
R. Casciaro (Italy)
K.F. Chung (Hong Kong, China)
C. Chiorean (Romania)
M. D’Aniello (Italy)
H. Degée (Belgium)
J.-F. Demonceau (Belgium)
F. Dinu (Romania)
J. Dobric (Serbia)
D. Dubina (Romania)
L. Dunai (Hungary)
S. Easterling (USA)
A. Elghazouli (UK)
M. Fontana (Switzerland)
D. Frangopol (USA)

xvii
L. Gardner (UK)
M. Garlock (USA)
G. Garcea (Italy)
P. Gonçalves (Brazil)
F. Guarracino (Italy)
J. Hajjar (USA)
G.J. Hancock (Australia)
M. Hjiaj (France)
B. Izzuddin (UK)
M. Knobloch (Germany)
R. Landolfo (Italy)
N. Lopes (Portugal)
R. Leon (USA)
J.R. Liew (Singapore)
J. Loughlan (UK)
J. Macháček (Czech Republic)
M. Mahendran (Australia)
F. Mazzolani (Italy)
E. Mirambell (Spain)
D. Nethercot (UK)
J. Packer (Canada)
J. Paik (South Korea)
S. Pajunen (Finland)
H. Pasternak (Germany)
N. Rangelov (Bulgaria)
K.J.R. Rasmussen (Australia)
E. Real (Spain)
B. Rossi (Belgium)
F. Roure (Spain)
A. Sato (Japan)
R. Sause (USA)
B.W. Schafer (USA)
L.S. Silva (Portugal)
N. Silvestre (Portugal)
H. Snijder (Netherlands)
R. Stroetmann (Germany)
A. Taras (Germany)
J.G. Teng (Hong Kong, China)
V. Ungureanu (Romania)
H. Unterweger (Austria)
B. Uy (Australia)
I. Vayas (Greece)
P. Vellasco (Brazil)
P. Vila Real (Portugal)
A. Wada (Japan)
Y.B. Yang (Taiwan, China)
N. Yardimci (Turkey)
B. Young (Hong Kong, China)

xviii
Keynote lectures
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Structural steel design by advanced analysis with strain limits

L. Gardner, A. Fieber & L. Macorini


Imperial College London, London, UK

ABSTRACT: The design of steel structures traditionally involves two steps: first, a structural
analysis is performed to determine the internal forces and moments; then, design checks are carried
out to verify the stability of individual structural members. In design by advanced analysis, both
material and geometric nonlinearities are captured during the analysis, hence eliminating the need
for subsequent member checks. However, steel frames are typically analysed using beam elements,
which cannot capture local buckling. Hence, steel design specifications use the concept of cross-
section classification to limit the strength and deformation capacity of a cross-section. A more
sophisticated approach is set out herein, whereby strain limits are employed to mimic local buck-
ling. This allows cross-sections of all classes to be analysed in a consistent advanced analysis
framework. The approach has been applied in this paper to individual members and indeterminate
systems and shown to be more consistent and accurate than current steel design specifications.

1 INTRODUCTION

The structural analysis of steel frames is typically performed using beam finite elements. These
elements are able to accurately capture the overall elastic-plastic load-deformation response of
structures composed of compact cross-sections. However, conventional beam elements do not
account for local cross-section deformations (i.e. local instabilities in either the elastic or inelastic
range). To overcome this limitation, steel design specifications, such as EN 1993-1-1 (2005), use the
concept of cross-section classification, whereby class-dependent restrictions on the cross-section
resistance and permissible analysis type (i.e. elastic or plastic) are specified. As a result, plastic ana-
lysis and design methods, which allow for the beneficial effect of force and moment redistribution,
are restricted to structures composed of compact Class 1 cross-sections. Structures composed of
more slender cross-sections (i.e. Class 2 to 4) must be analysed elastically and step-wise resistance
functions are employed to limit the capacity of the cross-sections and members.
Recently, a novel design approach based on geometrically and materially nonlinear struc-
tural analysis, also referred to as advanced analysis, using beam finite elements in conjunction
with strain limits, has been proposed (Gardner et al., 2019). The benefits of design by
advanced analysis are widely recognised (Liew et al. 2000; Chen 2008; Kim & Chen 1999; Tra-
hair & Chan 2003; Buonopane & Schafer 2006; Rasmussen et al. 2016; Surovek 2012). Com-
pared to the traditional approach to structural design, whereby the structural analysis is
followed by individual member and cross-section checks, in advanced analysis global sway
(P–Δ) and member (P–δ) instabilities are captured directly and the need for subsequent
member checks is eliminated. In the proposed approach, the strain limits are determined from
the continuous strength method (CSM) (Gardner 2008) and are used to mimic the effects of
local buckling. Hence, a consistent advanced analysis framework using beam finite elements
with strain limits can be employed to analyse and design structures comprising cross-sections
of any class, with failure defined as the lower of (1) the peak load factor reached during the
analysis or (2) the load factor at which the strain limit is first reached.
In this paper, the proposed method of design by advanced analysis is briefly outlined, including
a description of the quad-linear material model for hot-rolled steel (Yun & Gardner 2017) and the
continuous strength method (CSM) strain limits. Then, the most recent developments of the

3
proposed design approach are presented, including its application to beam-columns consisting of
hot-rolled steel I-shaped and SHS/RHS sections, continuous beams and a planar frame example.

2 TRADITIONAL STEEL DESIGN

In traditional steel design, the treatment of material nonlinearity, namely the spread of plasti-
city and the redistribution of forces in a structure, is tied to the classification of the cross-
sections. Plastic design is allowed when the cross-sections have sufficient rotation capacity to
enable plastic hinges to develop and to be sustained. Plastic design can account for the redis-
tribution of forces as plasticity spreads and can thus offer more accurate capacity predictions
when it can be used. Typically an elastic-perfectly plastic material model is assumed (i.e. strain
hardening is ignored). Alternatively, elastic analysis can be used for all cross-section classes.
The classification process is based on the width-to-thickness ratio of the critical isolated plate
element within the cross-section, and currently does not account for any flange-web interaction.
To facilitate the design process, EN 1993-1-1 (2005) defines four discrete cross-section classes, each
representing an idealised cross-sectional response with a corresponding moment and rotation cap-
acity. Class 1 sections can reach their full plastic moment capacity and have sufficient plastic hinge
ductility to be used in plastic design. An idealized moment-curvature relationship with unrestricted
rotation capacity is generally assumed in plastic design methods, while the beneficial effects of
strain hardening are typically ignored. Class 2 cross-sections can also attain their full plastic
moment capacity, yet are insufficiently ductile for plastic design purposes and consequently elastic
analysis methods have to be used. Local buckling prevents Class 3 sections from developing their
full plastic moment capacity and hence their resistance is limited to their elastic moment capacity.
The design of Class 4 sections reflects explicit allowance for the occurrence of local buckling below
the first yield resistance and generally involves somewhat lengthy calculations using effective sec-
tion properties.
After determining the member forces and moments from structural analysis, a series of
design checks are performed to ensure strength and stability requirements are satisfied.
Strength checks are applied to the most heavily stressed cross-section, with the bending resist-
ance limited to the plastic, elastic or effective section moment capacity (Mpl, Mel and Meff
respectively) and the compression resistance limited to the plastic or effective squash load (Npl
and Neff respectively), depending on the cross-section class. However, experiments (Lay 1964;
Gioncu & Petcu 1997) have shown that the maximum in-plane bending resistance of
a member subjected to a moment gradient can exceed that under uniform bending; a moment
gradient is of course present in most practical applications. While EN 1993-1-1 (2005) utilises
equivalent moment factors to account for the influence of the bending moment distribution
along the member length on member stability, it does not consider the beneficial effects of
local moment gradients on local stability.
The latest editions of many structural design codes permit the use of geometrically and materially
nonlinear analysis, also known as advanced analysis, for the design of steel structures consisting of
compact cross-sections. Examples include Section 5 of EN 1993-1-1 (2005), Appendix D of AS
4100 (1998) and Appendix 1 of AISC 360 (2016). Typically, any limit state that is not included in
the structural analysis must be accounted for using appropriate design checks. For example,
a linear (first order) analysis does not capture member buckling and thus a corresponding member
buckling check is required. Advanced analysis methods reduce the number of required design
checks by incorporating various limit states into the analysis itself.

3 DESIGN BY ADVANCED ANALYSIS WITH STRAIN LIMITS

3.1 Introduction
Recently, a novel approach for structural steel design based on geometrically and materially
nonlinear analysis including imperfections, also referred to as GMNIA or advanced analysis,

4
with strain limits has been proposed (Gardner et al., 2019). In the proposed approach, the
continuous strength method (CSM) (Gardner 2008) strain limits are used to mimic local buck-
ling in beam finite elements. In conjunction with a standardised material model for hot-rolled
steel (Yun & Gardner 2017), it is hence possible to employ the same advanced analysis frame-
work using beam elements for structures comprising cross-sections of all four classes since the
strain limits control the degree of plastic redistribution in a rational manner. Failure is defined
as the lower of (1) the peak load factor reached during the analysis or (2) the point at which
the CSM strain limit is reached. Details of the proposed method of design by advanced ana-
lysis are outlined in the following subsections.

3.2 Material modelling


The stress-strain relationship for structural carbon steels is often idealised by an elastic-perfectly
plastic model, though this model fails to capture the characteristic strain hardening behaviour of
carbon steels. Nevertheless, this simplified model generally forms the basis of the current design
provisions in EN 1993-1-1 (2005). For advanced design methods, such as the continuous strength
method (CSM) (Gardner 2008), an accurate representation of the stress-strain response of the
material is important, particularly for stocky cross-sections that may benefit from strain hardening.
A quad-linear material model, as illustrated in Figure 1 and described by Equation (1), has been
proposed (Yun & Gardner 2017) to represent accurately the yield plateau and strain hardening
behaviour of hot-rolled structural carbon steels. The material model has been calibrated against
a large database of tensile coupon test results and depends only on three commonly available
parameters: the Young’s modulus E, the yield stress fy and the ultimate stress fu. Two material coef-
ficients (C1 and C2) are used in Equation (1): C1 defines a ‘cut-off’ strain to avoid over-predictions
of material strength and is also included in the CSM base curve, as described in Section 3.3. and C2
is employed in Equation (2) to determine the strain hardening slope Esh. These two coefficients
may be determined from Equations (3) and (4) respectively, which are expressed in terms of the
strain hardening strain εsh and the ultimate strain εu.
8
> E" for "  "y
>
< fy for "y 5 "  "sh
f ð"Þ ¼ f þ E ð"  " Þ for "sh 5 "  C1 "u ð1Þ
>
> y sh sh
: f f
fC1 "u þ "uu CC11 ""uu ð"  C1 "u Þ for C1 "u 5 "  "u

Figure 1. Quad-linear material model for hot-rolled structural carbon steel (Yun & Gardner 2017).

5
Figure 2. Comparison of quad-linear material model with an experimental stress-strain curve for grade
S355 steel tested by Chan and Gardner (Chan & Gardner 2008).

fu  fy
Esh ¼ ð2Þ
C2 "u  "sh
"sh þ 0:25ð"u  "sh Þ
C1 ¼ ð3Þ
"u
"sh þ 0:4ð"u  "sh Þ
C2 ¼ ð4Þ
"u

The values of εsh and εu may be predicted from Equations (5) and (6), respectively.

fy
"sh ¼ 0:1  0:055; but 0:015  "sh  0:03 ð5Þ
fu
 
fy
"u ¼ 0:6 1  ; but "u  0:06 ð6Þ
fu

A typical comparison between a measured stress-strain curve (Chan & Gardner 2008) and the
quad-linear material model is shown in Figure 2. Further comparisons are presented and discussed
by Yun and Gardner (2017). The model has recently been incorporated into the CSM for the
design of hot-rolled steel cross-sections (Yun et al. 2018a; Yun et al. 2018b), showing improved
accuracy over the EN 1993-1-1 design provisions.

3.3 Continuous strength method (CSM) and strain limits


The continuous strength method (CSM) is a deformation based design approach that relates the
cross-section deformation capacity to the cross-section slenderness (Gardner 2008). The CSM con-
sists of two key components: (1) a base curve that defines the peak compressive strain εcsm that
a cross-section can endure and (2) a material stress-strain model, which has already been introduced
in Section 3.2. Combined, the CSM strain limit and the adopted material model define the stress
distribution at failure, which in turn, when integrated over the depth of the cross-section, defines
the CSM cross-section capacity. The CSM has been validated extensively and shown to be more
accurate than traditional design methods for the structural design of stainless steel (Gardner 2008;

6
Figure 3. Continuous strength method (CSM) base curve.

Zhao et al. 2016; Afshan & Gardner 2013), aluminium alloy (Ashraf & Young 2011; Su et al. 2016)
and hot-rolled steel (Yun et al. 2018b; Yun et al. 2018a; Liew & Gardner 2015).
The CSM base curve, shown in Figure 3, provides a continuous relationship between the
local slenderness of a cross-section l p and its deformation capacity, which is defined in nor-
malised form as εcsm/εy, where εcsm is the maximum compressive strain a cross-section can
sustain and εy is the yield strain. The base curve is split in two parts: Equation (7) applies to
non-slender sections with l p ≤ 0.68, which are referred to as Class 1 to 3 cross-sections in
EN 1993-1-1 (2005), and Equation (8) applies to slender cross-sections with l p > 0.68, which
are currently referred to as Class 4 cross-sections. The cross-section slenderness is defined by
Equation (9) and discussed further below.
 
"csm 0:25 "csm C1 "u
¼ 3:6 ; but  min ; for lp  0:68 ð7Þ
"y lp "y "y
!
"csm 0:222 1
¼ 1  1:05 for lp 40:68 ð8Þ
"y lp lp
1:05

Two upper limits are defined for the cross-section deformation capacity εcsm/εy in Equa-
tion (7). The first limit of Ω defines the permissible level of plastic deformation and may
be defined on a project-by-project basis. For example, a high value of say Ω =30 may be
specified where extensive plasticity is tolerable at ultimate limit state and a suitably duc-
tile steel is being used. Conversely, a value of Ω =1 would be used when material yielding
is not desirable, for example for serviceability limit checks. A value of Ω =15 is generally
recommended to prevent excessive deformations and to ensure that the material ductility
requirements from EN 1993-1-1 (2005) are satisfied. The second limit of C1εu/εy prevents
over-predictions of material strength when the simplified CSM resistance functions are
used (Gardner et al. 2017; Yun et al. 2018a).
The accuracy of the method is dependent on how accurately the cross-section slenderness
can be determined. The cross-section slenderness is a dimensionless parameter that quantifies
susceptibility to local instability and is defined by Equation (9), where fy is the yield stress and
σcr is the elastic critical buckling stress.

7
sffiffiffiffiffiffi
fy
lp ¼ ð9Þ
cr

Various methods are available to calculate the elastic critical buckling stress of a cross-
section. Following Eurocode 3 (EN 1993-1-1 2005; EN 1993-1-5 2006), the buckling
stress of the full cross-section may be taken as the buckling stress of the most slender
plate element within the cross-section, though this approach conservatively neglects any
element interaction. Full element interaction can be accounted for using numerical
approaches (e.g. the finite strip method as implemented in CUFSM (Li & Schafer 2010)).
Alternatively, approximate expressions calibrated from numerical studies (Seif & Schafer
2010) may be used to determine the elastic buckling stress of the full cross-section. While
the expressions presented in (Seif & Schafer 2010) account for element interaction, they
are only applicable to members subjected to pure compression and pure major/minor axis
bending. To overcome this shortcoming, Gardner et al. (2019) recently developed predict-
ive expressions for the elastic critical buckling stress of standard steel profiles subjected
to compression, bending and combined loading, which have been used throughout the
present study.

3.4 Application of CSM strain limits to advanced analysis


Including the CSM strain limits in advanced analysis enables the CSM cross-section capacity
to be computed directly since numerical integration is performed at each load increment of the
analysis. In the calculation of the CSM cross-section resistance, the strain limit εcsm is applied
to the peak compressive strain in the cross-section (i.e. the strain at the extreme fibre). The
cross-section can withstand the required strain demand if the design strain εEd is less than the
CSM strain limit εcsm, as given by Equation (10).

"Ed
 1:0 ð10Þ
"csm

Beam finite elements typically output strain values at the centreline of the wall thickness (i.e. at
a distance of half the plate thickness away from the extreme fibre). It was found that the differ-
ence in capacity predictions is negligible when applying the strain limit to the compatible design
strain at either the extreme fibre εEd or the centreline εEd,cl; this is because the slightly lower
design strain at the centreline of the wall thickness is offset by the slightly lower strain limit
εcsm,cl at this location. The application of the strain limit is shown schematically in Figure 4.
The CSM base curve has been calibrated based on experimental results on cross-sections
under uniform compression (i.e. stub columns) and uniform bending (i.e. the central region of
laterally restrained beams under four-point bending). Thus, the influence of residual stresses
and local geometric imperfections on the cross-section strength are directly accounted for. To
account for the effects of initial imperfections (i.e. member out-of-straightness and residual
stresses) on the member capacity, equivalent geometric imperfections are employed in the pro-
posed approach (Lindner et al. 2018).
In cases other than uniform compression and uniform bending, the local stability of the
cross-section is enhanced by the presence of a strain gradient (i.e. the local moment gradient
along the member length). To account for the beneficial effects of local moment gradients, the
CSM strain limit can be applied to strains that are averaged over a characteristic length along
the member, rather than simply (conservatively) to the peak strain. The length over which the
strains are averaged in the proposed approach is the local buckling half-wavelength of the full
cross-section Lb,cs, since local buckling is the limit state controlling the resistance of the cross-
section in both the elastic and inelastic regimes. The local buckling half-wavelength is calcu-
lated from the predictive expressions presented in (Fieber et al., submitted), which requires

8
Figure 4. Application of strain limits to compatible design strains at either extreme fibre or centreline of
wall thickness for (a) I-sections and (b) SHS/RHS subjected to major axis bending.

minimal additional calculation effort when the full cross-section local buckling stress is calcu-
lated from the expressions presented in (Gardner et al., 2019).
The strain averaging approach is an extension to the cross-section strain check given in
Equation (10). Instead of considering the peak compressive strain along the member, the
strain contributions εi of all n elements located completely within the critical local buckling
half-wavelength Lb,cs are considered. Assuming equal length elements, the strain averaging
approach is expressed in Equation (11), where εEd,Lb is the averaged design strain. The
weighted average strain must be determined in cases where the lengths of elements located
within Lb,cs are not equal. The strain averaging approach can account for the beneficial effects
of local moment gradients and reduces the sensitivity of strength predictions to the mesh dens-
ity employed in the finite element model.

"Ed;Lb 1Xn
 1:0 where "Ed;Lb ¼ "i and n  1 ð11Þ
"csm n 1

The application of the strain averaging approach is shown schematically in Figure 5 for the case
where n = 2. Assuming that first-order beam finite elements with a single integration point
located at the centre of the element are used, the continuous strain distribution is approximated
by the beam FE model in step-wise increments as shown in Figure 5. The considered member is
able to withstand the applied design loads if the averaged strain εEd,Lb = ε1 + ε2 is less than the
CSM strain limit εcsm.

4 APPLICATION OF PROPOSED APPROACH

The accuracy of the proposed method of design by advanced analysis with strain limits is
assessed in this section for isolated beam-columns, continuous beams and frames. The cap-
acity predictions obtained from beam FE advanced analyses with strain limits are compared
against those obtained from the benchmark shell FE models and conventional EN 1993-1-1
design. In the comparisons made herein, partial factors are set to unity. In practical design,

9
Figure 5. Schematic representation of the strain averaging approach for the case of n = 2. Note that the
strain contribution ε3 is not included in the averaging approach since element 3 is not located completely
within the local buckling half-wavelength Lb,cs.

the structural capacity obtained through the proposed design approach would be divided by
the partial safety factor for the relevant limits state, typically γM1

4.1 Members subjected to combined compression and bending


In this subsection, the proposed method of design by advanced analysis is applied to
a series of pin-ended beam-columns of varying member slenderness (l y = 0.5, 1.0 or 1.5).
The complete loading range from pure compression to pure major axis bending was
covered by varying the ratio of applied compression to bending. The capacity predictions
of the proposed method of design by advanced analysis were compared against the
results of shell FE modelling and two alternative EN 1993-1-1 (2005) design approaches:
(i) a member check using the interaction factor kyy from Annex B of EN 1993-1-1 and
(ii) a linear interaction cross-section check applied to a second-order elastic analysis with
equivalent imperfections (i.e. geometrically nonlinear analysis with imperfections, denoted
as GNIA in Figure 6). Note that the CSM strain limits were determined based on the
first-order stress distribution in the critical cross-section and that the ultimate capacities
from the proposed method were taken as the lesser of the peak load factor, indicated by
empty circles in Figure 6, and the point at which the CSM strain limit was reached, indi-
cated by the filled circles.
The normalised moment-axial force interaction diagram for a series of RHS 120х60х5
beam-columns are shown in Figure 6. Overall, the proposed method of design by advanced
analysis is able to accurately reflect the shell FE model behaviour, though in compression
dominated cases, the assumed imperfection magnitude may result in slightly conservative cap-
acity predictions.
The beam-columns considered in Figure 6 were subjected to a uniform first-order moment
gradient. Members under moment gradients are considered in Figure 7. As the level of first-
order moment gradient increases, the EN 1993-1-1 member check becomes increasingly con-
servative since the effects of local moment gradients and strain hardening are not considered.
This may be seen in the normalised moment-axial force interaction diagram for a series of
HEB 100 beam-columns in Figure 7. For clarity, only the results for l y = 0.5 and 1.5 are
shown. It can be seen that in compression dominated cases, failure is defined by the peak load
factored reached during the analysis, while in bending dominated cases the strain limits

10
Figure 6. Normalised ultimate capacity of RHS 120 х 60 х 5 pin-ended beam-columns subjected to com-
bined compression and major axis bending.

govern the member capacity. While the Eurocode 3 member checks become increasingly con-
servative in bending dominated cases, the proposed method of design by advanced analysis is
able to accurately reflect the shell FE model behaviour. Furthermore, it can be seen that when
the upper bound strain limit is relaxed to Ω = 30, the proposed method accurately captures
the strain hardening behaviour of the shell FE model. Benefits of up to 25% may be achieved
compared to EN 1993-1-1 capacity predictions.

4.2 Continuous beams


The ultimate collapse load of a continuous beam may significantly exceed the load at
which the first cross-section reaches its capacity, as shown in Figure 8, in which αel is the
load factor at which the first cross-section reaches its elastic moment and αult is the ultim-
ate load factor. The system capacity depends not only on the local cross-section strength,
but is also influenced by the beneficial effects of local moment gradients and moment
redistribution. Traditional design according to EN 1993-1-1 (2005) only permits full plas-
tic moment redistribution in Class 1 cross-sections. This results in a discrete jump in cap-
acity from plastic analysis for Class 1 cross-sections to elastic analysis for Class 2 to 4
cross-sections, which is clearly a simplistic representation of reality. As a result, the Euro-
code 3 capacity predictions tend to be overly conservative for Class 2 and 3 cross-
sections. The significant capacity reserves due to strain hardening in stocky cross-sections
with lp < 0.4 are also not captured in Eurocode 3, as shown in Figure 8.
Using a consistent materially nonlinear advanced analysis for all cross-section classes pro-
vides a more rational exploitation of moment redistribution, as well as strain hardening in
stocky cross-sections. Figure 8 shows that the CSM strain limits are able to predict the appro-
priate levels of redistribution for each cross-section. Particularly in the typical hot-rolled slen-
derness range from l p = 0.35 to 0.55 there is an excellent match to the shell FE results. Below
a slenderness of lp = 0.32, the CSM limit of Ω = 15 prevents excessive deformations. However,
by increasing the strain limit to Ω = 30, the beam element models allow for additional redistri-
bution and accurately follow the shell model trends again.

11
Figure 7. Normalised ultimate capacity of HEB 100 pin-ended beam-columns subjected to combined
compression and major axis bending.

Figure 8. Normalised ultimate capacity of three span continuous beams with varying local cross-section
slenderness.

12
4.3 Frames
In this section, the proposed method of design by advanced analysis is applied to an asymmet-
ric low-rise frame. The frame geometry and considered load case is shown in Figure 9. All
members were made from grade S355 steel and the beam to column connections were mod-
elled as fully fixed. The load was applied proportionally and the normalised load-
displacement responses of the frame obtained using different analysis methods are shown in
Figure 10. Failure in the shell FE model was governed by inelastic local buckling at the top of
the column C4 and side-sway, at a maximum load factor of α = 1.02. The deformed shape of
the critical region in the shell FE model at the peak load factor is shown in Figure 10.
Following EN 1993-1-1 (2005), the columns are Class 3 cross-sections and the beams are
Class 1 cross-sections. Hence, elastic analysis must be used. The critical elastic buckling load
factor of the frame at the peak load factor of the shell FE model is αcr = 3.10; hence, the frame
is sensitive to second-order sway effects. Various different design approaches are permissible
according to EN 1993-1-1, namely: (1) a second-order elastic analysis with equivalent imper-
fections (GNIA) with cross-section checks, (2) a second-order elastic analysis (GNA) with
member and cross-section checks using the non-sway effective length, (3) a linear elastic ana-
lysis (LEA) with amplified sway moments using kamp followed by member checks using non-
sway effective lengths (with the non-sway effective length factor denoted keff), or (4) a linear
elastic analysis (LEA) with member checks using the sway effective lengths (with the sway
effective length factor denoted keff,sway) to account for second-order effects. All four Eurocode
approaches correctly predict failure in column C4. Approaches (1) to (3) predict a similar load
carrying capacity, with a maximum load factor of around α = 0.84, where the non-sway effect-
ive length of column C4 is taken as keff = 0.745. Accounting for second-order effects in a LEA
through the sway effective length, taken as keff,sway = 2.1 for column C4, results in an overly
conservative capacity prediction of α = 0.68.
In the proposed method of design by advanced analysis, the strain limit for column C4 is
εcsm = 1.67εy. Thus, in contrast to Eurocode 3, some plastification is allowed and failure is pre-
dicted at a load factor of α = 0.96, just 6% short of the shell FE model and 12% above the best
Eurocode 3 prediction. Furthermore, the strain limits accurately mimic local buckling in the
beam FE model, with the first strain limit reached at the top of column C4. Note that if no

Figure 9. Asymmetric low-rise frame example: geomtry and applied loading.

13
Figure 10. Normalised load-deformation response of asymmetric low-rise frame example.

strain limits were applied to the beam FE advanced analysis, an unconservative peak load
factor of α = 1.05 would be reached.

5 CONCLUSIONS

The most recent developments of a new method of design by advanced analysis with strain
limits are presented. The design method is based on a geometrically and materially nonlinear
analysis with imperfections (i.e. advanced analysis) and employs the continuous strength
method strain limits to mimic local buckling in beam finite elements. The proposed method is
able to capture the beneficial effects of local moment gradients and predict realistic levels of
force and moment redistribution. Application of the method is demonstrated for a series of
beam-columns, continuous beams and a frame. It is shown that the proposed method is con-
sistently more accurate compared to the current Eurocode 3 design approach and that
advanced analysis with strain limits is a viable design tool for structures composed of cross-
sections of all classes.

REFERENCES

Afshan, S. & Gardner, L., 2013. The continuous strength method for structural stainless steel design.
Thin-Walled Structures, 68, 42–49.
AS 360–16, 2016. Specification for Structural Steel Buildings, American Institute of Steel Construction.
AS 4100, 1998. Australian Standard. Steel Structures, Sydney: Standards Australia.
Ashraf, M. & Young, B., 2011. Design formulations for non-welded and welded aluminium columns
using Continuous Strength Method. Engineering Structures, 33, 3197–3207.
Buonopane, S.G. & Schafer, B.W., 2006. Reliability of steel frames designed with advanced analysis.
Journal of Structural Engineering, ASCE, 132(2), 267–276.

14
Chan, T.M. & Gardner, L., 2008. Bending strength of hot-rolled elliptical hollow sections. Journal of
Constructional Steel Research, 64, 971–986.
Chen, W.F., 2008. Advanced analysis for structural steel building design. Frontiers of Architecture and
Civil Engineering in China, 2(3), 189–196.
EN 1993-1-1, 2005. Eurocode 3: Design of steel structures - Part 1-1: General rules and rules for buildings,
Brussels: European Committee for Standardization.
EN 1993-1-5, 2006. Eurocode 3: Design of steel structures – Part 1-5: Plated structural elements, Brussels:
European Committee for Standardization.
Fieber, A., Gardner, L. & Macorini, L., submitted for review. Formulae for determining elastic local
buckling half-wavelengths of structural steel cross-sections. Journal of Constructional Steel Research.
Gardner, L., Yun, X., Macorini, L. & Kucukler, M., 2017. Hot-rolled steel and steel-concrete composite
design incorporating strain hardening. Structures, 9, 21–28.
Gardner, L., Yun, X., Fieber, A. & Macorini, L., 2019. Steel design by advanced analysis: material mod-
elling and strain limits. Engineering. 5(2), 243–249.
Gardner, L., 2008. The continuous strength method. Proceedings of the Institution of Civil Engineers -
Structures and Buildings, 161(3), 127–133.
Gardner, L., Fieber, A. & Macorini, L., 2019. Formulae for calculating elastic local buckling stresses of
full structural cross-sections. Structures. 17, 2–20.
Gioncu, V. & Petcu, D., 1997. Available rotation capacity of wide-flange beams and beam-columns Part
2. Experimental and numerical tests. Journal of Constructional Steel Research, 43, 219–244.
Kim, S.E. & Chen, W.F., 1999. Design guide for steel frames using advanced analysis program. Engineer-
ing Structures, 21, 352–364.
Lay, M.G., 1964. The experimental bases of plastic design. WRC, Bulletin No. 99, Publication No. 258.
Li, Z. & Schafer, B.W., 2010. Buckling analysis of cold-formed steel members with general boundary
conditions using CUFSM: Conventional and constrained finite strip methods. In Twentieth Inter-
national Speciality Conference on Cold-Formed Steel Structures. Saint Louis, Missouri, USA.
Liew, A. & Gardner, L., 2015. Ultimate capacity of structural steel cross-sections under compression,
bending and combined loading. Structures, 1, 2–11.
Liew, J.Y.R., Chen, W.F. & Chen, H., 2000. Advanced inelastic analysis of frame structures. Journal of
Constructional Steel Research, 55, 245–265.
Lindner, J., Kuhlmann, U. & Jörg, F., 2018. Initial bow imperfections e0 for the verification of flexural
buckling according to Eurocode 3 Part 1-1 - additional considerations. Steel Construction, 11(1), 30–
41.
Rasmussen, K.J.R., Zhang, H., Cardoso, F. & Liu, W., 2016. The direct design method for cold–formed
steel structural frames. In Eighth International Conference on Steel and Aluminium Structures.
Hong Kong: Hong Kong, 1–14.
Seif, M. & Schafer, B.W., 2010. Local buckling of structural steel shapes. Journal of Constructional Steel
Research, 66(10), 1232–1247.
Su, M.N., Young, B. & Gardner, L., 2016. The continuous strength method for the design of aluminium
alloy structural elements. Engineering Structures, 122, 338–348.
Surovek, A.E., 2012. Advanced analysis in steel frame design. Guidelines for direct second-order inelastic
analysis. A. E. Surovek, ed., Reston, Virginia: ASCE.
Trahair, N. & Chan, S.L., 2003. Out-of-plane advanced analysis of steel structures. Engineering Struc-
tures, 25, 1627–1637.
Yun, X. & Gardner, L., 2017. Stress-strain curves for hot-rolled steels. Journal of Constructional Steel
Research, 133, 36–46.
Yun, X., Gardner, L. & Boissonnade, N., 2018a. The continuous strength method for the design of
hot-rolled steel cross-sections. Engineering Structures, 157, 179–191.
Yun, X., Gardner, L. & Boissonnade, N., 2018b. Ultimate capacity of I-sections under combined load-
ing – Part 2: Parametric studies and CSM design. Journal of Constructional Steel Research, 148, 265–
274.
Zhao, O., Gardner, L. & Young, B., 2016. Behaviour and design of stainless steel SHS and RHS
beam-columns. Thin-Walled Structures, 106, 330–345.

15
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Cold-formed high strength steel RHS under combined bending


and web crippling

Hai-Ting Li
School of Civil and Environmental Engineering, Nanyang Technological University, Singapore
Department of Civil Engineering, The University of Hong Kong, Hong Kong, China

Ben Young
Department of Civil and Environmental Engineering, The Hong Kong Polytechnic University,
Hong Kong, China
Department of Civil Engineering, The University of Hong Kong, Hong Kong, China

ABSTRACT: This paper presents experimental and numerical investigations of cold-formed


high strength steel (CFHSS) rectangular hollow sections (RHS) under combined bending and
web crippling. In the experimental investigation, 5 pure bending tests and 28 combined bend-
ing and web crippling tests were conducted on RHS with measured 0.2% proof stress in the
flat portion of the section ranged from 679 to 971 MPa. The combined bending and web crip-
pling tests were performed using the Interior-One-Flange loading condition that specified in
the North American Specification for cold-formed steel structures. The specimens were tested
at various lengths to examine the interaction relationship between bending moment and con-
centrated interior bearing load. Finite element (FE) models were developed and validated
against the test results for members under combined bending and web crippling as well as
pure bending. Upon validation of the FE models, a parametric study was performed using the
validated models to generate further numerical data over a wide range of web slenderness
ratio, bearing length to web thickness ratio and bearing length to web flat portion ratio. The
ultimate strengths obtained from experimental and numerical investigations were compared
with nominal strengths calculated using the European Code. It is shown that the codified
bending and web crippling interaction formula can be used for the CFHSS RHS members,
while more accurate predictions can be achieved by using the recently proposed web crippling
design rules.

1 INTRODUCTION

Cold-formed steel (CFS) rectangular hollow sections (RHS) are widely used in various engin-
eering applications due to their structural efficiency and aesthetic appearance. The CFS rect-
angular (includes square in the context of this paper) hollow sections are often applied in
construction without transverse stiffeners; therefore, the webs of the unstiffened CFS RHS
may cripple when subjected to high local intensity of bearing forces or reactions. Moreover,
the web crippling capacity of CFS RHS under local transverse forces may reduce noticeably
due to the presence of bending moment, especially in the vicinity of the loading point within
a span or at the interior support of a continuous beam. Therefore, combined bending and
bearing check is crucial in designing CFS tubular structural members.
The North American Specification AISI-S100-16 (AISI 2016), Australian/New Zealand
Standard AS/NZS 4600 (AS/NZS 2005) and Eurocode EN 1993-1-3 (CEN 2006) provide
design provisions for CFS members under combined bending and web crippling; the codified
combined bending and web crippling design rules are generally empirical in nature. This is due
to the fact that theoretical analysis of web crippling is rather complicated (Yu & LaBoube

16
2010) and the pure web crippling rules in the aforementioned codes of practice are empirically
derived based upon experimental investigations conducted by researchers from the 1940s
onwards. Therefore, the codified combined bending and web crippling provisions are only
applicable for certain materials and cross-section profiles. High strength steel tubular mem-
bers are increasingly attractive in construction owing to the structural and architectural
advantages (Zhao et al. 2014, Ma et al. 2017). The applicability of the existing combined bend-
ing and web crippling provisions to cold-formed high strength steel (CFHSS) structural mem-
bers shall be investigated.
In this study, a test program on CFHSS RHS under combined bending and web crippling is
presented. A total of 33 tests, including 5 pure bending tests as well as 28 combined bending
and web crippling tests, were undertaken on RHS that had nominal yield strengths of 700 and
900 MPa. Four-point bending tests were conducted to obtain the moment capacities of the
CFHSS RHS under constant bending moment. The combined bending and web crippling
tests were carried out under the Interior-One-Flange (IOF) load case as per the AISI (2016);
various specimen lengths were tested to study the interaction of bending moment and localized
interior bearing load. The pure web crippling test results of IOF load case reported by the
authors in Li & Young (2017a) are also used in this paper to non-dimensionalize the combined
bending and web crippling test results. In addition, finite element (FE) models were developed
and validated against the test results; upon validation, a parametric study was undertaken
using the validated models to generate further numerical data. The results gained from the
experimental and numerical investigations were compared against the nominal resistances as
per the EN 1993-1-3 (CEN 2006) to assess the suitability of the codified combined bending
and web crippling provisions to CFHSS RHS.

2 EXPERIMENTAL INVESTIGATION

A test program was undertaken to study the CFHSS RHS under bending and web crippling.
Three types of experiments were carried out: pure bending tests; pure web crippling tests; com-
bined bending and web crippling tests. The RHS specimens of the abovementioned three types of
tests were from the same batch of tubes, the material properties of which have been previously
reported by the authors in Li & Young (2017a, 2018a). The measured static 0.2% proof stresses
of the RHS gained from the tensile flat coupon tests varied between 679 and 971 MPa. Table 1
tabulated the mechanical properties gained from longitudinal tensile flat and corner coupon tests,
namely, the Young’s moduli (E), static 0.2% proof stresses (σ0.2) and static tensile strengths (σu).

2.1 Pure bending tests


The CFHSS RHS specimens of pure bending series had measured web heights H ranging
between 50.1 to 160.1 mm, flange widths B ranging from 50.0 to 160.2 mm, thicknesses
t ranging between 3.896 to 3.971 mm, and inside corner radii r ranging between 4.6 to 6.8 mm.

Table 1. Mechanical properties obtained from tensile flat and corner coupon tests.
Flat coupon^ Corner coupon†

Section (H×B×t) E (GPa) σ0.2 (MPa) σu (MPa) E (GPa) σ0.2 (MPa) σu (MPa)

H80 × 80 × 4 210.9 725 834 214.2 877 945


H160 × 160 × 4 212.4 751 829 216.2 899 992
H50 × 100 × 4 211.3 679 820 207.2 860 932
V100 × 100 × 4 205.3 971 1079 219.6 1073 1175

Note: ^Conducted by Li & Young (2017a); †Conducted by Li & Young (2018a).

17
Table 2. Measured dimensions and test/FE results of pure bending test specimens (Li & Young
(2019b)).
Specimen H (mm) B (mm) t (mm) r (mm) L (mm) MExp (kN·m) MExp/MFEA

B-H80 × 80 × 4 80.1 80.1 3.896 4.7 1390 29.0 1.04


B-H160 × 160 × 4 160.1 160.2 3.971 5.1 2188 87.8 0.98
B-H50 × 100 × 4 50.1 100.3 3.931 4.6 1389 18.0 1.07
B-H100 × 50 × 4 100.4 50.0 3.930 4.6 1389 31.6 1.08
B-V100 × 100 × 4 99.9 100.2 3.946 6.8 1592 51.7 1.00
Mean 1.04
COV 0.043

Figure 1. Pure bending test setup.

The section H50 × 100 × 4 was bent about both major and minor axes. The specimen labels, as
shown in Table 2, illustrated the test type, nominal yield strength and cross-sectional dimen-
sions. For instance, the labels B-H100 × 50 × 4 and B-V100 × 100 × 4 defined the following spe-
cimens: the first letter B indicated they were pure bending specimens; the second letter
indicated the material of the RHS, where H and V indicated the nominal yield strengths (0.2%
proof stresses) of 700 and 900 MPa, respectively; the following symbols were the nominal
cross-sectional dimensions, arranged as H × B × t, in millimetres, where the H, B and t are the
height, width and thickness of the RHS, respectively.
Four-point bending tests were undertaken by Li & Young (2019b) to obtain the moment cap-
acities of the CFHSS RHS under constant bending moment. The four-point bending test setup is
shown in Figure 1. Simply supported boundary conditions have been achieved by half-round and
roller supports. The lengths of moment spans were 500 mm, except for the specimen B-H160 ×
160 × 4, of which the moment span is 600 mm. Calibrated linear variable displacement trans-
ducers (LVDTs) were set at the bottom flange of the two load points and the specimen midspan.
The curvatures of the RHS specimens were obtained from these LVDT readings. Compressive
loads were imposed on the RHS specimens via displacement control through a servo-controlled
hydraulic actuator; the applied load rate was 0.5 mm/min for all the tests. The moment
capacities MExp gained from the pure bending experiments are reported in Table 2.

2.2 Summary of pure web crippling tests


Pure web crippling tests of CFHSS RHS were conducted and have been previously reported
by the authors (Li & Young, 2017a). The pure web crippling tests were performed under the

18
Figure 2. Pure web crippling test setup.

four codified web crippling load cases in the CFS specifications, namely, the AISI (2016) and
AS/NZS (2005); the pure web crippling specimen lengths were employed as per the AISI
(2016). The loading or reaction forces were imposed to the RHS via bearing plates. Figure 2
shows the setup of a pure web crippling test under the Interior-One-Flange (IOF) load case.
The web crippling capacities per web (PExp) of the IOF specimens of sections H80 × 80 × 4,
H160 × 160 × 4, H50 × 100 × 4, H100 × 50 × 4 and V100 × 100 × 4 were employed in this study
to non-dimensionalize the combined bending and web crippling test results, which will be
described in Section 4 of this paper. Details of the pure web crippling tests are available in Li
& Young (2017a).

2.3 Combined bending and web crippling tests


The measured dimensions of the combined bending and web crippling specimens are reported
by Li & Young (2019b) and shown in Table 3. The test specimens have measured H ranging
from 50.0 to 160.1 mm, B ranging from 50.0 to 160.3 mm, t ranging from 3.902 to 3.979 mm,
and r varying between 4.6 to 6.8 mm. The h/t ratios ranging between 8.3 to 35.8. In general,
the lengths L of the combined loading series specimens equalled to 2a+90 mm, where a is the
distance (in millimetres) from the support point to the midspan as shown in Figure 3. For
specimens of the C-H160 × 160 × 4N150 series (i.e. bearing length N = 150 mm), the L were
designed to be 2a+150 mm to avoid any possible failure at the end of the specimens. The dis-
tance a was varied to study the interaction relationship between bending moment and local-
ized interior bearing load; this method has been previously employed by many researchers
(e.g. Zhao & Hancock 1991, 1992, Young & Hancock 2002, Zhou & Young 2007). The dis-
tance a = kMExp/PExp, in which, MExp is the moment capacity from the four-point bending
test; PExp is the experimental web crippling capacity per web under the IOF load case of the
same RHS; and k is the interaction factor. The k values were selected to allow the interaction
of bending moment and localized bearing load over a wide range; the selected k values varied
between 0.7 to 1.8 in this test program. Test specimens with lower k values resulted in lower
ratios of moment-to-localized bearing force, whilst specimens with higher k values led to
higher ratios of moment-to-localized bearing force.
The specimens, as shown in Table 3, were labelled so that the test type, nominal yield
strength, cross-sectional dimensions, bearing length and interaction factor can be identified.
For example, the label C-H80×80×4N50-k1.0-R defined the following specimen: the first
letter C indicated it was a combined loading specimen; the second letter H showed the nominal

19
Table 3. Measured dimensions and test/FE results of combined bending and web crippling test speci-
mens (Li & Young (2019b)).
Specimen H (mm) B (mm) t (mm) r (mm) L (mm) PC,Exp (kN) PC,Exp/PC,FEA

C-H80×80×4N90-k1.0 80.1 80.1 3.922 4.7 572 111.4 1.00


C-H80×80×4N90-k1.3 80.1 80.1 3.925 4.7 717 90.9 1.02
C-H80×80×4N90-k1.8 80.2 80.1 3.919 4.7 957 66.1 1.01
C-H80×80×4N50-k0.7 80.1 80.1 3.949 4.7 488 99.4 1.00
C-H80×80×4N50-k1.0 80.1 80.1 3.901 4.7 658 81.5 1.02
C-H80×80×4N50-k1.0-R 80.1 80.1 3.942 4.7 658 81.1 1.00
C-H80×80×4N50-k1.5 80.1 80.1 3.902 4.7 943 60.3 1.02
C-H160×160×4N150-k0.8 160.1 160.2 3.953 5.1 1053 144.9 0.95
C-H160×160×4N150-k1.1 160.1 160.1 3.974 5.1 1392 119.0 0.93
C-H160×160×4N150-k1.6 160.0 160.2 3.970 5.1 1957 89.7 0.93
C-H160×160×4N90-k0.7 160.0 160.3 3.979 5.1 1015 121.6 0.98
C-H160×160×4N90-k1.0 160.0 160.1 3.956 5.1 1411 97.7 0.95
C-H160×160×4N90-k1.5 160.0 160.2 3.979 5.1 2072 74.4 0.94
C-H50×100×4N50-k0.9 50.0 100.4 3.944 4.6 395 96.3 1.10
C-H50×100×4N50-k1.2 50.1 100.3 3.958 4.6 496 78.6 1.11
C-H50×100×4N50-k1.7 50.1 100.3 3.932 4.6 665 57.8 1.09
C-H100×50×4N50-k0.7 100.3 50.0 3.948 4.6 539 95.2 1.01
C-H100×50×4N50-k0.7-R 100.4 50.2 3.963 4.6 539 93.7 0.98
C-H100×50×4N50-k1.0 100.3 50.0 3.952 4.6 732 77.9 0.99
C-H100×50×4N50-k1.5 100.3 50.1 3.926 4.6 1053 57.8 0.99
C-H100×50×4N30-k0.7 100.3 50.1 3.967 4.6 617 77.4 1.02
C-H100×50×4N30-k0.7-R 100.3 50.1 3.940 4.6 617 75.3 1.00
C-H100×50×4N30-k1.0 100.3 50.1 3.926 4.6 843 62.9 1.02
C-H100×50×4N30-k1.0-R 100.3 50.1 3.937 4.6 843 61.3 0.98
C-H100×50×4N30-k1.5 100.3 50.0 3.933 4.6 1220 46.8 0.99
C-V100×100×4N50-k0.7 100.4 100.0 3.957 6.8 717 95.5 0.94
C-V100×100×4N50-k1.0 100.2 99.9 3.907 6.8 986 77.2 0.94
C-V100×100×4N50-k1.5 100.1 100.1 3.931 6.8 1434 59.0 0.95
1.00
0.048

Figure 3. Specimen length design for combined bending and web crippling tests.

20
Figure 4. Combined bending and web crippling test setup.

0.2% proof stress was 700 MPa; the following symbols showed the nominal H×B×t were
80×80×4 in millimetres; the N50 indicated the bearing length N=50 mm; the k1.0 indicated
the interaction factor was 1.0; the R at the end (if any) indicated it was a repeated test.
Regarding the combined bending and web crippling tests, various specimen lengths were
employed using the IOF web crippling load case as per the AISI (2016). The combined bending
and web crippling test setup is shown in Figure 4. A bearing plate (BP) was employed to trans-
fer the bearing force to the RHS via a half round at the midspan. Two steel plates, supported
by two rollers, were applied at the specimen ends to allow symmetric boundary conditions; the
width of these two plates were 90 mm except for the specimens of C-H160×160×4N150 series,
where 150 mm width plates were adopted. Similar to the IOF test setup previously detailed by
the authors in Li & Young (2017a), steel stiffening plates were employed at the specimen ends.
Displacement control mode was selected and the tests were conducted at a load rate of 0.5 mm/
min; this is the same as the pure bending tests and pure web crippling tests. The ultimate loads
per web (PC,Exp) of the combined bending and web crippling specimens are reported in Table 3.
The experimentally gained PC,Exp was utilized to obtain the ultimate moment of the RHS speci-
men by using MC,Exp = a PC,Exp. Out-of-plane deformation was not observed during testing for
all the 28 combined bending and web crippling specimens.

3 NUMERICAL MODELLING

In parallel to the experimental program, a numerical study (Li & Young 2019b) was con-
ducted by using ABAQUS (2012). Finite element (FE) models were built to replicate the tests.
Upon validation, a parametric study was undertaken to gain further numerical data on
CFHSS RHS structural members under combined bending and web crippling. The numerical
models were established based upon measured specimen geometries. The S4R element in the
ABAQUS (2012), which has been successfully used in similar previous FE simulations by the
authors (Li & Young 2017b, 2018a, b, 2019a, b), was employed herein to model the CFHSS
RHS members. The applied meshes in the RHS flat regions varied between 4×4 to 12×12 mm,
which depended on the RHS sizes; finer meshes at the RHS corners were adopted to accur-
ately represent the corner regions. The nonlinearity of the CFHSS materials was incorporated
based upon measured engineering stress-strain data gained from the flat and corner coupon
tests; the engineering stress-strain data have been converted to gain the true stress and true
plastic strain relationships prior to being put into the FE models. In this study, the corner
properties were applied to the RHS corners with a 2t extension to the adjacent flat portions.

21
The boundary conditions were replicated according to the experiments. With regard to the
pure bending tests, the loading points and supports were modelled using reference points,
which were coupled to the contact surfaces between the specimens and load transfer plates.
The half-round support was modelled by restraining the corresponding reference point against
all degrees of freedom (DOF) except for rotation about the bending axis, whereas the roller
support was simulated by allowing an extra DOF of longitudinal movement. The loads were
imposed by applying axial displacements to the reference points that modelled the loading
points. The FE modelling of the pure web crippling tests has been previously reported by the
authors in Li & Young (2018a); regarding the combined bending and web crippling modelling,
the same technique as employed by Li & Young (2018a) for the IOF load case was applied
herein. The concentrated bearing forces were transferred to the RHS specimens by bearing
plates (BP). The BP were simulated by discrete rigid 3D solid elements. The surface interactions
of the BP and the RHS were defined using contact pairs. The loads were imposed by applying
displacements to the BP, which is the same as the experiments using displacement control.
The developed numerical models were validated against the experiments. For the CFHSS
RHS under pure bending, the moment capacities obtained experimentally (MExp) and numer-
ically (MFEA) were compared and shown in Table 2; the mean ratio of the MExp/MFEA
equalled to 1.04 and the corresponding coefficient of variation (COV) was 0.043. The failure
modes and moment-curvature curves derived from the finite element analyses (FEA) were
also validated against their experimental counterparts. The numerical validation for IOF spe-
cimens undergoing pure web crippling was performed and has been reported by Li & Young
(2018a). Regarding the CFHSS RHS members under combined bending and web crippling,
the validation is carried out herein; the ultimate loads per web (PC,FEA) predicted by the FEA
were compared to their experimental counterparts PC,Exp. The mean ratio of the PC,Exp/PC,FEA
was 1.00 and the COV was 0.048, as tabulated in Table 3. Typical numerical failure mode,
load-web deformation curves and load-midspan deflection curves derived from FEA were
compared to their corresponding experimental ones, as illustrated in Figures 5–7. It is shown

Figure 5. Experimental and numerical failure modes for combined bending and web crippling specimen
C-H160×160×4N90-k1.5. (a) Experimental failure mode. (b) Numerical failure mode.

22
Figure 6. Experimental and numerical load-web vertical deformation curves for combined bending and
bearing specimens C-H80 × 80 × 4N90-k1.0, C-H80 × 80 × 4N90-k1.3 and C-H80 × 80 × 4N90-k1.8.

Figure 7. Experimental and numerical load-midspan deflection curves for specimens C-H80 × 80 ×
4N90-k1.0, C-H80 × 80 × 4N90-k1.3 and C-H80 × 80 × 4N90-k1.8.

that the developed FE model successfully replicate the combined bending and web crippling
tests, and therefore is deemed suitable to be used for parametric study.
After validation of the FE models, a parametric study was undertaken to gain further numer-
ical data. Similar as the test program, the nominal yield strengths of the RHS were 700 MPa for
the H series and 900 MPa for the V series herein; the measured properties gained from material
tests of the sections H160×160×4 and V100×100×4 were applied for the H and V series, respect-
ively. An extensive range of 18 RHS was studied. These RHS had H ranging between 150 to
400 mm and t ranging between 2 to 8 mm; the h/t ratios ranging from 13.8 to 106.0. The bearing
lengths N varying between 75 to 220 mm, and were chosen to be either N = B or N = 0.5B. The
N/t ratios ranging between 9.4 to 110.0 and the N/h ratios varying between 0.3 and 1.4. In total,
18 parametric specimens were modelled under pure bending and 188 results were generated for
specimens under combined bending and web crippling.

4 COMPARISON OF TEST AND NUMERICAL RESULTS WITH CURRENT


DESIGN PROVISIONS

In order to investigate the CFHSS RHS members under combined bending and web crippling,
the pure bending and pure web crippling endpoints in the interaction curves are needed. The
moment capacities Mu gained from the pure bending test and FE programs (Li & Young
2019b) were compared with nominal moment resistances obtained as per the Eurocode EN

23
Figure 8. Comparison of pure bending test and FE results with nominal resistances predicted by EC3.

1993-1-1 (CEN 2014) (MEC3), as shown in Figure 8. It is shown that the MEC3 provided quite
conservative predictions; the Mu/MEC3 had a minimum value of 1.01 and the mean ratio of
the Mu/MEC3 was 1.14 with the corresponding COV of 0.087.
The web crippling capacities per web Pu gained from the experimental (Li & Young 2017a)
and numerical (Li & Young 2018a) studies have been compared to the codified nominal web
crippling resistances, as shown in Figure 9. It has been demonstrated that the CEN (2006) pre-
dictions were overly conservative; therefore, improved design rules have been proposed for
CFHSS RHS undergoing pure web crippling, as detailed in Li & Young (2018a). The Pu of
the IOF specimens, which are required in this study to non-dimensionalize the combined
bending and web crippling capacities, were compared against the nominal resistances per web
based upon the CEN (2006) (PEC3) and Li & Young (2018a) (PL&Y) provisions. It is shown
that the PL&Y, which has been proposed based upon modification of the AISI (2016) design
rules, provided much-improved predictions than the current Eurocode predictions. The mean

Figure 9. Comparison of pure web crippling test and FE results with nominal resistances predicted by
EC3 and Li & Young (2018a).

24
Figure 10. Comparison of combined bending and web crippling test/FE capacities with nominal resist-
ances (non-dimensionalized with respect to PEC3 and MEC3).

ratio of the Pu/PL&Y was 1.07 with the corresponding COV of 0.099, while the Pu/PEC3 had
a mean ratio of 1.76 with the corresponding COV of 0.098.
The obtained combined bending and web crippling results (Li & Young 2019b) were com-
pared to the nominal resistances calculated using the CEN (2006) and the bending and web
crippling interaction equation in the CEN (2006) is illustrated in Eq. (1).
   
PC MC
þ  1:25 ð1Þ
PEC3 MEC3

in which, PC is the maximum concentrated interior bearing load per web in the presence of
bending moment; MC is the maximum bending moment of the RHS. The PC and MC were
non-dimensionalized with respect to the nominal web crippling resistances per web PEC3 and
the nominal moment resistances MEC3. The comparisons with the CEN (2006) interaction
curve were depicted in Figure 10, where great conservatism of the EC3 predictions is observed.
This over conservatism is mainly due to the over-pessimistic predictions of the PEC3
endpoints.
In this study, the PC were also non-dimensionalized with respect to the PL&Y; the compari-
sons with the CEN (2006) interaction curves by using PL&Y as pure web crippling endpoints in
the interaction equation are shown in Figure 11, in which the vertical axis is still the
ratio MC/MEC3 as its counterpart in Figure 10, but the horizontal axis is changed into PC/PL&Y.
It is demonstrated that the CEN (2006) provided much-improved predictions when using the
web crippling strengths predicted by Li & Young (2018a) as the web crippling endpoints.
Hence, improved design rules for CFHSS RHS under combined bending and web crippling can
be sought through the adoption of the PL&Y as the web crippling endpoints in the interaction
curve.
Furthermore, the experimentally and numerically derived PC and MC were also non-
dimensionalized with respect to the corresponding Pu and Mu that obtained from the pure
web crippling specimens and pure bending specimens. This is to appraise the appropriateness
of the CEN (2006) interaction coefficients for CFHSS RHS. The interaction equations in the
CEN (2006) can be therefore expressed as Eq. (2).
   
PC MC
þ  1:25 ð2Þ
Pu Mu

25
Figure 11. Comparison of combined bending and web crippling test/FE capacities with nominal resist-
ances (non-dimensionalized with respect to PL&Y and MEC3).

Figure 12. Comparison of combined bending and web crippling test/FE capacities with EC3 interaction
curve.

The PC/Pu and MC/Mu are plotted in Figure 12 with the codified interaction curve. As illus-
trated by Figure 12, the interaction curves in the CEN (2006) were conservative, although
may slightly overestimate the strengths of a few specimens with high moment-to-concentrated
load ratios. Overall, the codified interaction equations are deemed appropriate for the CFHSS
RHS members under combined bending and web crippling.

5 CONCLUSIONS

Design of cold-formed high strength steel (CFHSS) rectangular hollow sections (RHS) under
combined bending and web crippling has been appraised. A test program was undertaken on
cold-formed RHS of high strength steel with measured 0.2% proof stresses ranging from 679
to 971 MPa. The combined bending and web crippling tests were undertaken using the IOF
web crippling load case. The specimens were tested at different lengths and various bending
moment-to-concentrated bearing load ratios were achieved. Finite element models were built
and validated with the experiments; a parametric study was performed thereafter and 188
parametric results were generated for specimens under combined bending and web crippling.

26
The experimentally and numerically obtained results were compared against the nominal
resistances predicted by the CEN (2006). It has been demonstrated that the combined bending
and web crippling resistances calculated from the CEN (2006) were overly conservative. It has
been illustrated that the codified interaction equations can be used for the CFHSS RHS mem-
bers under combined bending and web crippling, while improved predictions can be achieved
through the adoption of the PL&Y as the pure web crippling endpoints in the interaction curve.

ACKNOWLEDGEMENTS

The authors are grateful to Rautaruukki Corporation for providing the test specimens. The
research work described in this paper was supported by a grant from the Research Grants
Council of the Hong Kong Special Administrative Region, China (Project No. 17209614).

REFERENCES

ABAQUS. (2012). Abaqus/Standard user’s manual volumes I-III and Abaqus CAE manual. Version
6.12. Hibbitt, Karlsson & Sorensen, Inc., Pawtucket, USA.
AISI (American Iron and Steel Institute). 2016. North American specification for the design of
cold-formed steel structural members. AISI-S100-16, Washington, DC, USA.
AS/NZS (Australian/New Zealand Standard). 2005. Cold-formed steel structures. AS/NZS 4600, Sydney,
Australia.
CEN (European Committee for Standardization). 2006. Eurocode 3: Design of steel structures - Part 1-3:
General rules - Supplementary rules for cold-formed members and sheeting. EN 1993-1-3, Brussels,
Belgium.
CEN (European Committee for Standardization). 2014. Eurocode 3: Design of steel structures - Part 1-1:
General rules and rules for buildings. EN 1993-1-1:2005+A1:2014, Brussels, Belgium.
Li, H.T. & Young, B. 2017a. Tests of cold-formed high strength steel tubular sections undergoing web
crippling. Engineering Structures 141, 571–583.
Li, H.T. & Young, B. 2017b. Cold-formed ferritic stainless steel tubular structural members subjected to
concentrated bearing loads. Engineering Structures 145, 392–405.
Li, H.T. & Young, B. 2018a. Design of cold-formed high strength steel tubular sections undergoing web
crippling. Thin-Walled Structures 133, 192–205.
Li, H.T. & Young, B. 2018b. Web crippling of cold-formed ferritic stainless steel square and rectangular
hollow sections. Engineering Structures 176, 968–980.
Li, H.T. & Young, B. 2019a. Behaviour of cold-formed high strength steel RHS under localised bearing
forces. Engineering Structures 183, 1049–1058.
Li, H.T. & Young, B. 2019b. Cold-formed high-strength steel tubular structural members under com-
bined bending and bearing. Journal of Structural Engineering, ASCE. (In press)
Ma, J.L. Chan, T.M. & Young, B. 2017. Tests on high-strength steel hollow sections: a review. Proceed-
ings of the Institution of Civil Engineers170(SB9): 621–630.
Young, B. & Hancock, G.J. 2002. Tests of channels subjected to combined bending and web crippling.
Journal of Structural Engineering 128(3): 300–308.
Yu, W.W. & LaBoube, R.A. 2010. Cold-formed steel design. Fourth Edition, New York: John Wiley &
Sons.
Zhao, X.L. & Hancock, G.J. 1991. T-joints in rectangular hollow sections subject to combined actions.
Journal of Structural Engineering 117(8): 2258–2277.
Zhao, X.L. & Hancock, G.J. 1992. Square and rectangular hollow sections subject to combined actions.
Journal of Structural Engineering 118(3): 648–667.
Zhao, X.L. Heidarpour, A. & Gardner, L. 2014. Recent developments in high-strength and stainless steel
tubular members and connections. Steel Construction 7(2): 65–72.
Zhou, F. & Young, B. 2007. Experimental investigation of cold-formed high-strength stainless steel tubu-
lar members subjected to combined bending and web crippling. Journal of Structural Engineering 133
(7): 1027–1034.

27
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Stability design of steel structures: From members to plates


and shells

L. Simões da Silva, Trayana Tankova, & João Pedro Martins


ISISE, Department of Civil Engineering, University of Coimbra, Portugal

ABSTRACT: This paper presents an overview of the current design procedures for the
buckling design of members, plates and curved panels. It highlights recent developments in
these fields, namely methodologies for the evaluation of the buckling resistance of generic
beam-columns with variable cross section, loading and boundary conditions, the evaluation of
the reliability of the Winter curve in slender webs and the development of design guidance for
curved panels. Finally, the incorporation of some of these developments in the current revi-
sion of Eurocode 3 is discussed.

1 INTRODUCTION

Nowadays, structural engineering is a highly standardized activity whereby current design


procedures evolved over many years based on accumulated experience and a continuous effort
to reflect a sound mechanical basis. Additionally, design standards in Europe are bound to
product and execution standards which are aimed to guarantee that the adopted assumptions
concerning material properties and the fabrication processes reflect reality.
Historically, verification procedures were aimed at simple design processes, thus providing
expressions suitable for hand calculations, very often empirical, calibrated to experimental tests
whenever possible. More recently, there is a trend to transform the simpler expressions into
design procedures that contain the mechanical background of the phenomenon under consider-
ation, complemented by extensive numerical validation that in some cases leads to accurate
design expressions that are statistically calibrated (Tankova et al, 2014). This is only made pos-
sible due to the tools available to the engineers – computers and sophisticated software.
These advances have reached design codes as well. Both Eurocode 3 (2005) and AISC
(2016) allow the design of steel structures using numerical analyses with different level of
sophistication. There are three basic levels (Simões da Silva et al, 2016): i) global analysis
which accounts for all imperfections and all 2nd order effects followed by cross-section verifi-
cations; ii) global analysis that partially accounts for the imperfections (only global) and 2nd
order effects and followed by member and cross-section verification buckling length equal to
the member geometrical length or the simplest version iii) basic cases where individual equiva-
lent member checks are performed using buckling lengths corresponding to the global buck-
ling mode of the structure/member and cross-section verifications.
Both Eurocode 3 (2005) and AISC (2016) present fairly similar recommendations for the
global analysis of the structure. Both design codes recommend the use of second order ana-
lysis for the determination of the design forces. In both cases, it is allowed to use fictitious
horizontal forces in order to account for the global sway imperfections of the structure.
In case of Eurocode 3, the designer has the choice of including member imperfections in
the second-order analysis and further verifying only the cross-section resistance or neglecting
the P-δ effects and accounting for them by further verification according to its specifications in
section 6.3 for member stability using the geometrical member length (L) or the non-sway buck-
ling length. In case of AISC (2016), second order analysis is always followed by a member veri-
fication; however, the member length is taken equal to the unbraced member length KL = L.

28
Even though it is conceptually simple, second order analysis with global and local imperfec-
tions is not yet the preferred method and it is yet to become the daily basis of design in
a design office. This is due to the requirement of more in deep knowledge about member
imperfections but also due to the software availability to apply the imperfections directly.
Hence, the most common approach used by the engineers today is the second order global
analysis with global imperfections and then followed by the member checks using as buckling
length the member length.
The member design rules in the modern codes also follow very similar concepts – for col-
umns and beams the buckling curve approach, whereas beam-columns are designed using
interaction equations which combine the utilization due to axial force and bending moments
through calibrated factors.
The verification format for the flexural buckling of columns is directly derived from the dif-
ferential equations for a column with an initial imperfection. It is achieved by applying a first
yield criterion at mid-span for a simply-supported column (i.e., equating the normal stress at
the most compressed fiber to fy):

N N^v0
þ ¼ fy ð1Þ
A 1  N=Ncr;i

which is transformed to

χ
χþη 2
¼ 1:0 ð1Þ
1  χλ

where λ is the normalized slenderness, χ is the buckling reduction factor and η is the general-
ized imperfection factor (accounting for out-of-straightness and residual stresses), with
W denoting the elastic section modulus relative to the buckling axis. Eq.(1) is the buckling
curve equation where the only unknown is the amplitude of the imperfection. This advantage
of the Ayrton-Perry representation of the problem was used in the calibration of the European
design rules.
The buckling curves were established in the 1970’s, and their development was based on an
extensive experimental programme carried out by the European Convention for Constructional
Steelwork (ECCS) in several European countries Sfintesco (1970); on theoretical developments
Beer & Schulz (1970) by thorough analysis of the experimental programme; and on reliability
assessment by a Monte Carlo simulation Strating & Vos (1973). Finally, Maquoi & Rondal
(1978) derived the analytical Ayrton-Perry format of the design verification and the curves
were put into equation. Presently, the code proposes 5 buckling curves, a to d were originally
calibrated on the basis of the ECCS experimental programme. The most relaxed curve a0, was
added later to account for the more favourable properties of high strength steels (HSS).
Similarly, the AISC design specification is based on SSRC experimental programme which
initially specified 3 curves which were subsequently reduced to a single curve, SSRC curve 2P
(Ziemian, 2010). Yet, this curve is divided into three regions – the inelastic and purely elastic
which are calculated using three different expressions unlike the single expression of Euro-
code 3.
The stability of unrestrained beams can be verified according to two methods in Eurocode 3.
The first approach, the so-called General case, assumes that columns and beam behave alike,
and so the compression flange is considered as an equivalent column. This results in identical
imperfection factors as for flexural buckling of columns. The method, however, adopts differ-
ent splits according to the cross-section geometry, accounting for their different torsional
rigidity. Furthermore, according to the European specifications for stability design given in
ECCS No. 119, these curves were meant to be used with deep and slender sections which are
outside of the scope of rolled sections. The method is of general application but too conserva-
tive when applied to members with variable bending moment diagram (Rebelo et al. (2009),
Taras (2010)).

29
As an alternative for hot-rolled and equivalent welded sections, Eurocode 3 provides
another set of buckling curves, the Special case. The method was calibrated on the basis of
extensive numerical studies in the Research Project Lateral-torsional buckling of steel and
composite beams (1993) and Salzgeber (2000). Additional calibration was carried out by
Grainer & Kaim (2001) on the basis of experimental results by Janss & Maquoi. In order to
justify, the plateau at 0.4, experimental results by Byfield & Nethercot (1998) were used in the
assessment according to Annex Z of ENV 1993-1-1:1992. However, the method was shown to
be unconservative when compared to numerical results Snijder & Hoenderkamp (2007),
Rebelo et al. (2009).
The European interaction formula for verification of members subject to bending and com-
pression currently uses two sets of interaction coefficients: the “French-Belgian” team which
was responsible for the interaction coefficients in Method 1 and the “Austrian-German” team
responsible for the development of the coefficients associated with Method 2. The main differ-
ence between the two methods is the way of considering the various effects which affect the
beam-column behaviour. The interaction coefficients associated with Method 1 were devel-
oped aiming to distinguish each structural effect in the interaction coefficient (plasticity,
equivalent moment factors, lateral-torsional buckling), therefore laying the ground for any
further modifications, if necessary, and to directly identify the impact of each physical phe-
nomenon. Method 2, aimed at easier practical implementation, combines the non-linearities
into global interaction factors kii and kij calibrated on the basis of a wide-range of numerical
simulations.
At cross-section level, in Eurocode 3, the ability of the cross-section to either develop its full
plastic capacity, behave elastically or being unable reach the elastic limit due to local instabil-
ity is distinguished by the cross-section classification. Depending on the cross-section class,
the member stability design rules are applied with plastic, elastic or effective properties.
In the specification of the effective properties for slender sections, it is considered that
a redistribution of compressive stresses takes place due to reduction of stresses in the middle-
buckled region which leads to increase of stresses near the plate edges (Beg et al., 2010). The
non-linear stress distribution is simplified by either an effective width method (based on an
appropriate cross-section reduction in central buckled part of the plate) or a reduced stress
method (by calculating a reduced average stress).
The concept of effective width was initially developed by von Karman (1932) by assuming
that the effective cross-section works under the yield stress (fy). However, in this work, von
Kaman did not consider the presence of initial imperfections and residual stresses. Later,
Winter (1947, 1948) performed experimental studies on long plates stiffened along both longi-
tudinal edges such as web of channels and I-sections. More than 100 tests under compression
of thin walled cold-formed C-channels and I-sections with different b/t ratios and yield stress
were performed. This experimental programme was the basis of the buckling curve considered
in Eurocode 3-1-5 for the estimation of the effective width, it is often referred as the Winter
curve. Since it was developed for plates that were supported at the edges (channels and I-sec-
tions) by normally thicker flanges, the level of rotational restraint provided is somehow
incorporated in the design expression.
This paper addresses the design procedures and their reliability for the stability design of
members, plates and curved panels and discusses current developments and their incorpor-
ation in the current revision of Eurocode 3.

2 BEAM-COLUMN DESIGN

Even though the development of stability design rules currently used and presented in this sec-
tion were developed using analytical developments, experimental tests and statistical assess-
ments, stability design continues to attract the attention of researchers. Nowadays, the efforts
are focused in the incorporation of the enhanced material properties, new cross-section geom-
etries and the possibility of computer aided calculations into the design process. In the

30
following, a brief overview of the recent developments in the scope of the European design
rules for beams, columns and beam-columns is summarized.
There are several works which build upon the Ayrton-Perry format for flexural buckling of
columns, i.e. extending it to beam-columns, non-uniform members and members subject to
arbitrary loading. It was succeeded by combining the first and second order effects corres-
ponding to the case in consideration and calibration of overstrength factors which account for
the variation of the most unfavourable cross-section depending on the loading. This concept
was used to extend design verifications to flexural-torsional and lateral-torsional buckling
once the correct analytical behaviour was considered by Taras (2010). In Naumes (2009), the
equilibrium equation for the flexural buckling of tapered members was also established; in
this derivation, the shape of the initial imperfection was considered eigenmode conform. It
was shown that the Ayrton-Perry design format can be adopted for the design of non-uniform
members. However, the proposed expressions are not applicable for practical design verifica-
tion due to lack of recommendations for the determination of the design location. Further-
more, Taras (2010) offers the same type of model for flexural buckling of beam-columns
which was combined with calibrated factors accounting for the plasticity effects at low slen-
derness. Marques (2012) provided design equations for flexural buckling of web-tapered col-
umns and lateral-torsional buckling of web-tapered beams in the same Ayrton-Perry format
which account in a systematic way for the position of the critical for verification design
location.
Simplifications of the current rules also exist, an approximate verification format for beam-
columns is proposed by Hoglund (2014) where the stress utilizations due to axial force and
bending moments are specified as power functions. The method claims to provide better repre-
sentation of the transition between Class 2 and Class 3 sections, which however are already
implemented in the final draft of prEN 1993-1-1 (CEN/TC250, 2018) following the recom-
mendations of the European project SEMICOMP+ (Greiner et al., 2011). It is also incorpor-
ated in Eurocode 9 for verification of aluminium beam-columns.
Recent research by Szalai (2017) shows the extension of the Ayrton-Perry equation to pris-
matic simply supported members subject to arbitrary loading. The author does not provide
calibration of the corresponding imperfection factors but shows that it is theoretically possible
to achieve this format for various buckling modes. Based on this development, Szalai & Papp
(2017) built their proposal for reformulation of the General method, by putting it into the
derived Ayrton-Perry proposal for prismatic simply supported members subject to arbitrary
loading, which is its major flaw, being unable to capture the specific aspects of non-uniform
members.
Even though the Ayrton-Perry based developments have the advantage of considering the
first and second order effect in a consistent manner, in most of the cases the necessity of cali-
bration additional factors was adopted, due to variation of the critical location along the
member. This is a major drawback when non-conventional cases need to be considered.
Driven by these limitations, there are a few developments supporting the design by use of
numerical analyses. Their proposals were mainly focused on the definition of the equivalent
geometrical imperfection to be considered in the design Chladný & Stujberová (2013a,b),
Aguero et al. (2015a,b), Papp (2016) and a mixture between LBA conform imperfection and
a reduction factor calculation Badari & Papp (2015).
Recently, a new method, the so-called general formulation (GF) was proposed by (Tankova
et al. (2018)). The proposed general formulation for the stability design of steel members com-
prises a linear interaction equation that combines the first-order normal stresses due to
applied forces and the normal stresses due to second order forces. This linear interaction
between first and second order stresses is consistent with the Eurocode 3 procedures where the
reduction factor χ was also derived on the basis of a Ayrton-Perry design philosophy and
assuming that the shape of the initial imperfection follows the same shape as the buckling
mode.
The proposed general formulation also adopts the buckling mode as the shape of the initial
imperfection and the amplitude previously calibrated for the standard prismatic simply-
supported columns and beams in Eurocode 3, while the amplitude for the standard prismatic

31
simply-supported beam-column buckling out-of-plane is obtained from recent work by Tan-
kova (2018a) Since the terms concerning the stress utilization due to second order forces are
amplified by the imperfection factors from Eurocode 3 they are consistent with the rules for
prismatic members. Finally, the verification is implemented as a sequence of cross-section veri-
fications along the member length. The fact that the stress utilizations due to first and second
order forces are added at each location allows to avoid the calibration of additional factors
for each specific case, in contrast with the case of web tapered members (Marques et al., 2012,
Marques et al., 2013). where specific generalized imperfection factors had to be adjusted. The
developed interaction equation needs to be applied for all potential failure modes.
The utilization ratio of the generic single member may be expressed by equating the total
longitudinal stress, due to first and second order forces, to the yield stress, fy:

NðxÞ My ðxÞ Mz ðxÞ MyII ðxÞ M II ðxÞ MwII ðxÞ


þ þ þ þ z þ  1:0 ð2Þ
AðxÞ fy Wy ðxÞ fy Wz ðxÞ fy Wy ðxÞ fy Wz ðxÞ fy Ww ðxÞ fy

where A(x) is the cross-section area, Wy (x) and Wz (x) are the section moduli relative to the y-
and z axes, respectively, and Ww (x) = Iw(x)/wmax(x) is the warping modulus at location
x along the member, with wmax(x) = hb/4. It is noted that for section classes 1 and 2 the plastic
section moduli should be used while for class 3 sections the semi-compact approach of Annex
B of prEN 1993-1-1 may be used. Then, as long as the second order contributions can be
determined, the buckling resistance may be verified for an appropriate number of locations
along the member.
The verification of a single member with variable geometry, boundary conditions, subject
to arbitrary loading, is done by verifying Eq. (2) at a sufficient number of locations along the
member. At each position, the respective values of the first order axial force, N(x), bending
moments My(x), Mz(x), second order contributions obtained from the relevant buckling mode
and cross-section properties are to be used. Figure 1 illustrates the procedure for a tapered
single member:

Figure 1. Schematic illustration of the verification procedure.

32
This verification shall be performed for the global buckling modes which may affect the
studied member, for instance in and out-of-plane buckling. For instance, the verification of
a beam-column shall be done by applying Eq. (3) for in-plane buckling and out-of-plane buck-
ling. In addition, the requirement to check the cross-section resistance at the extremities of the
member is automatically included, as the member is checked for a sufficient number of cross-
sections, including the end-sections, as explained above. Table 1 illustrates the calculation of
the second-order forces in Eq. (3)

3 LOCAL BUCKLING

3.1 General remarks


The stochastic nature of the governing parameters in the ultimate resistance of plated struc-
tures must be acknowledged if an accurate assessment of the actual reliability level is to be
achieved. These parameters are on one hand geometrical (dimensional and geometric devi-
ations) and material (yield stress, Young’s modulus and residual stresses). Several authors
have drawn their attention to the reliability analysis of plated structures (Fukumoto & Itoh
(1984), Duc et al. (2013), Rahman et al. (2017), Gaspar et al. (2015)). All of these authors per-
formed their analyses by modeling the plate isolated from the rest of the structure, applying,
therefore, idealized boundary conditions: in these studies, all edges are perfectly simply
supported.
Recently, Schillo (2017) collected several experimental results challenging the level of safety
given by the effective width methodology (based on the well-known Winter curve) as it is pre-
sented in Eurocode 3-1-5. These experimental results are from mild and high strength steel
welded class 4 box-sections. As a result of these and own experimental results,
a recommendation for higher values of the partial factor or, alternatively, a new exponential
curve to replace the Winter curve was proposed.
Considering that the studies dealing with the stochastic nature of strength of plates are
limited to perfect boundary conditions, and that there are cases where the current rules of the
effective width methodology given by EC3-1-5 is apparently unsafe, Martins et al (2019)
evaluated the reliability level of the referred methodology applied to welded I- cross-sections
under compression. For this purpose, a Monte Carlo experiment was performed where the
Latin Hypercube technique was applied to reduce the size of the sample. All simulations with
the Monte Carlo experiment are evaluated using the finite element method and a normal dis-
tribution is fitted to the obtained resistance distribution. Finally, this normal distribution is
compared to the nominal resistance provided by EC3-1-5 and partial factors are derived using
the formulation given in EC0. These results are briefly described in the following sub-sections.

3.2 Scope of the analysis


In this analysis, the following parameters are considered as having a stochastic nature: steel
properties (S460 steel was considered; statistical data from prEN1993-1-1, see Table 2) and
geometry of the cross-section (flange width and thickness, web height and thickness; statistical
data from prEN1993-1-1). The remaining parameters, i.e. shape and amplitude of initial
imperfections (where the first buckling mode and hw/200 are considered, respectively) and
residual stresses (which are not considered explicitly) are taken as deterministic. In fact, these
assumptions correspond to the equivalent imperfections option for the modeling of
imperfections.
In order to include a practical range of applications, 28 samples consisting in seven class 4
cross-sections (Table 3) with 4 values for the aspect ratio (1.00, 1.25, 1.50 and 1.75) were con-
sidered. These cross-sections are inspired in product catalogs (specifically, HE600B, HE800B,
and HL1100B). Table 4 gives the nominal resistance for each cross-section.

33
Table 1. Second-order forces.
Mode εI ðxÞ εII ðxÞ ηðxÞ fη
00
FB NðxÞ EIi ðxÞδ cr ðxÞ αcr NðxÞ
ηðxÞ
(N) AðxÞfy AðxÞfy ðαcr  1Þ EIi ðxm Þjδ00 cr ðxm Þj
FB NðxÞ Mi ðxÞ EIi ðxÞδ00 cr ðxÞ My jδcr ðxÞj αcr NðxÞ
þ ηðxÞ þ
(N+M) AðxÞfy Wi ðxÞfy AðxÞfy ðαcr  1Þ Wy ðxÞfy ðαcr  1Þ   EIi ðxm Þjδ00 cr ðxm Þj
LTB Mi ðxÞ αðλð xÞ  0:2Þf ηδflcr ðxÞ
 
(M) Wi ðxÞfy EIz ðxÞ v00 cr ðxÞ þ h2 θ00 cr ðxÞ þ θ0 cr ðxÞh0 Ncr;z;eq
LTB NðxÞ Mi ðxÞ ηðxÞ  
þ AðxÞfy ðαcr  1Þ EIi ðxm Þ v00 cr ðxm Þ þ h2 θ00 cr ðxm Þ þ θ0 cr ðxm Þh0
(N+M) AðxÞfy Wi ðxÞfy
Table 2. Statistical data of S460 structural steel.
fy [MPa] E [MPa]

Structural Steel Nominal Mean Co.V. Nominal Mean Co.V.

S460 460 529 4.5% 210 000 210 000 3.0%

Table 3. Geometrical statistical data of the analyzed cross-sections.


Welded hw [mm] bf [mm] tw [mm] tf [mm]
cross-sections
(hw × bf × tw × tf) Mean Co.V. Mean Co.V. Mean Co.V. Mean Co.V.

540 × 300 × 19.75 × 30 540 300 19.75 29.4


540 × 300 × 15.5 × 30 540 300 15.5 29.4
734 × 300 × 17.5 × 33 734 300 17.5 32.34
1028 × 300 × 20 × 36 1028 0.9% 400 0.9% 20 2.5% 35.28 2.5%
1228 × 300 × 20 × 36 1228 400 20 35.28
1528 × 300 × 20 × 36 1528 400 20 35.28
1528 × 300 × 18.82 × 36 1528 400 18.82 35.28

3.3 Analysis of results


Using a purely random sampling method leads to the need of a large number of simulations in
a Monte Carlo experiment, otherwise the convergence of the method is questionable. Several
authors suggest methods to calculate the required sample size: for example, while Mann et al.
(1974) recommends a minimum of 10000 simulations for a 95% confidence limit, Melchers
(1999) suggests, for the same confidence limit and for a failure probability equal to 10-3,
around 3000 simulations. Additionally, the number of independent variables in each simula-
tion may also influence the required sample size.
Here, in order to decrease the size of the sample, the Latin Hypercube Sampling technique
is applied. It was concluded that 5000 simulations for each sample were more than sufficient
to achieve convergence. Specifically, convergence of the mean value and standard deviation
(see example in Figure 2), as well as the probability of failure were achieved for all samples.
The probability of failure (and the reliability index) is obtained for each aspect ratio and
cross-section by fitting a normal distribution to the results and calculating the percentage of
lower results than the nominal value of the resistance (see Table 3). It should be highlighted

Figure 2. Convergence of mean value and standard deviation for cross-section 1528 × 300 × 18.82 × 36,
α=1.

35
that the resistance statistical distribution passed the Kolmogorov-Smirnov test for the normal
hypothesis. Therefore, the probability of failure is obtained by:

ð
Rd;nom
 
PF ¼ P R5Rd;nom ¼ N ðμR ; σR Þ ð4Þ
∞

where μR and σR are the mean and standard deviation, respectively, for each sample. The reli-
ability index is simply obtained by:

β ¼ F1 ðPF Þ ð5Þ

Finally, the partial factor is iteratively calculated until the reliability index of each resistance
distribution equals β=3.04, see Eq. (6).
 
  Rd;nom
F βTARGET ¼ P R  ð6Þ
γM

3.3.1 Reliability index, partial factors and web reduction factor


As already mentioned, all resistance distributions follow a normal distribution. In Figure 3a it
can be seen that the highest value for the partial factor is equal to 1.05 for the stockiest cross-
section. This conclusion indicates that only for cross-sections with lower web plate slenderness
the partial factor is underestimated by EC3-1-5. On the other hand, for the remaining cross-
sections, the values of the partial factor are significantly lower than 1. This gives the indication
that the estimative of EC3-1-5 is safe sided.
Finally, the web reduction factor may be obtained by integration of stresses at ultimate
load and calculating the equivalent width that would be yielded, or, alternatively (acknow-
ledging the fact that the flanges are always fully effective) using the following equation:
 
Aw;i :ρ þ Af ;i fy;i ¼ NFEM ð7Þ

where Aw,i is the gross area of the web, is the reduction factor, Af,i is the gross area of the
flanges, fy,i is the yield stress and NFEM is the numerically obtained resistance.

Figure 3. Partial factors and equivalent plate reduction.

36
Figure 4. Ranges of parameters in existing bridge decks (Reis et al. (2017)).

Figure 3b plots the minimum and maximum values of the reduction factor (using equation
(7)) against the Winter curve. Analyzing the results, it becomes clear the lower bound nature
of the Winter curve.

3.4 General remarks


Recently, a good amount of effort was carried to propose design rules for stiffened
curved plates. An example of this effort is the RFCS project OUTBURST. In fact,
curved steel plates in bridge decks are increasingly used in recent years (Reis et al.
(2017)), both for aesthetic and structural demands (efficiency of load carrying capacity;
high strength to weight ratio, added value from an architectural point of view, enhanced
behaviour to wind loading). However, as pointed out by some researchers (Tran et al.
(2013)), curved panels do not behave neither as a flat plate nor as a full revolution shell.
This poses a problem for designers. Martins et al. (2018), give a detail revision on the
available expressions to calculate the elastic critical stress and the ultimate strength of
curved plates.
Curved panels are usually defined by the global curvature parameter which is a non-
dimensional parameter given by the following expression

b2
Z¼ ð8Þ
Rt

where b is the width of the plate, R is the radius of curvature and t is the thickness of the
plate.
This parameter may be also used to defined the local curvature of sub-panels (in which
cases b is replaced by blocal, i.e., width of the unstiffened part of the plate). Figure 4 shows the
values of global and local curvature, as well as values of the aspect ratio of the curved plates,
of real curved plates found in bridge designs.

3.5 Elastic critical stress of unstiffened curved plates under compression and shear
The first attempt to obtain the elastic critical stress of curved panels under compression is due
to Redshaw (1935) who proposed a formula which was later modified by Stowell (1944).
Later, Timoshenko (1961) based on approximate expressions of the displacement field pro-
posed a different formula.
Concerning the shear load case, the first known studies are due to Leggett (1937), Kromm
(1939) and Timoshenko and Gere (1961). Subsequent works have been performed later by
Batdorf (1947) and Schildcrout and Stein (1949).
Based on the compilation of the work performed by these authors, in 1968 NASA (1968)
publishes a document containing curves allowing to obtain the elastic critical coefficient of

37
Table 4. Elastic buckling coefficient for short curved panels under compression (Martins et al. (2013))
(1<Z<100).
ψ = σ1/σ2 1 1>ψ>0 0 0 > ψ > -1 -1
kσ A
Bþψ
A
B þC ψþ D ψ 2

C ¼ c1 þ c2 Z þ c3 Z2
A ¼ a1 þ a2 Z þ a3 Z2
D ¼ d1 þ d2 Z þ d3 Z2
B ¼ b1 þ b2 Z þ b3 Z2

a1 ¼ 8:2 b1 ¼ 1:05 c1 ¼ 6:29 d1 ¼ 9:78


a2 ¼ 0:0704 b2 ¼ 0:0002 c2 ¼ 0:1971 d2 ¼ 0:2174
0 < Z  23
a3 ¼ 0:0163 b3 ¼ 0:0003 c3 ¼ 0:0004 d3 ¼ 0:0002

c1 ¼ 9:124 d1 ¼ 5:843
a1 ¼ 3:214 b1 ¼ 0:961
c2 ¼ 0:0646 d2 ¼ 0:0556
23 < Z  100 a2 ¼ 0:5976 b2 ¼ 0:0104
c3 ¼ 0 d3 ¼ 0:0002
a3 ¼ 0:0028 b3 ¼ 0

curved plates. These curves have been for very long time the most up-to-date methodology to
obtain the elastic critical stress of curved panels.
Recently, due the extraordinary increase in the computation capacity, the finite element
method as enabled researchers to run large parametric studies consisting in linear buckling
analysis and calibrate expressions which are more accurate.
Within the framework of the above-mentioned research project, the formulas given in
Tables 4 and 5 were developed to compute the elastic critical stress of curved panels under
compression and under shear stresses, respectively.

4 CONCLUDING REMARKS: OUTLOOK OF FUTURE VERSIONS OF


EUROCODE

The first revision of the Structural Eurocodes is currently taking place, with a planned release
in 2023. Concerning Eurocode 3 and the design rules related to the buckling resistance of

Table 5. Elastic buckling coefficient for curved panels under shear (Ljubinkovic et al.
(2019a)) (1< Z < 100).
 2
kτ ¼ A þ B α1 α≤1 α>1

A ¼ 0:214Z þ 2:88 A ¼ 0:096Z þ 5:15


C1 Z B ¼ 0:135Z þ 3:18
B ¼ 5:343 
175:6
A ¼ 0:247Z þ 2:732 A ¼ 0:124Z þ 4:94
C2 Z B ¼ 0:137Z þ 3:756
B ¼ 5:34   2
150:4 Z
A ¼ 0:2734Z þ 2:794 A¼ þ 0:349Z þ 5:424
28:86
Z  2
B ¼ 5:33  Z
C3 127:2 B¼  0:0452Z þ 2:422
37:735

38
members, plates and curved panels, prEN 1993-1-1 (2018) and prEN 1993-1-5 (2019) already
implement some of the aspects briefly presented and discussed above.
Concerning the buckling resistance of members, prEN 1993-1-1 (2018) implements the
Ayrton -Perry based developments for the buckling resistance of beams and extends the scope
of the beam-column interaction equations to monosymmetric cross-sections. It further simpli-
fies the interaction factors by adopting a single methodology for their evaluation. A more
detailed selection of methodologies is also adopted concerning the choice of global analysis
methods.
Concerning the determination of the effective width of slender plates, a distinction is intro-
duced in the buckling curves as a function of the degree of restraint along the longitudinal
boundary conditions. Finally, it is planned to extend the scope to cylindrically curved panels.
These developments, coupled with the possibility to implement design based on advanced
finite element calculations will ensure that the Structural Eurocodes remain a leading code of
practice that addresses current and future design challenges.

REFERENCES

Aguero A., Pallarés F.J., Pallarés L. (2015a). Equivalent geometric imperfection definition in steel struc-
tures sensitive to flexural and/or torsional buckling due to compression, Engineering Structures, 96,
pp. 160–177.
Aguero A., Pallarés F.J., Pallarés L. (2015b). Equivalent geometric imperfection definition in steel struc-
tures sensitive to lateral-torsional buckling due to bending moment”, Engineering Structures, 96,
pp. 41–55.
AISC (2016), American Institute of Steel Construction, Specification for Structural Steel Buildings, Chi-
cago, Illinois, USA.
Alinia M.M., Habashi H.R., Khorram A. (2009) Nonlinearity in the post-buckling behavior of thin steel
shear panels. Thin-Walled Structures 47, 412–420.
Amani, M., Edlund, B.L.O., Alinia, M. M. (2011) Buckling and post-buckling behavior of unstiffened
curved plates under uniform shear. Thin-walled Structures, 49 (8), 1017–1031.
Anonymous. Buckling of thin-walled truncated cones. NASA Space Vehicle Design Criteria (Structures),
NASA SP-8019, September 1968.
Badari B., Papp F. (2015) On design method of lateral-torsional buckling of beams: State of the art and
a new proposal for a general type of design method” Periodica Polytechnica Civil Engineering, 59, pp.
179–192.
Batdorf, S.B. (1947) A Simplified Method of Elastic Stability Analysis for Thin Cylindrical Shells.
NACA Technical Report No. 847.
Beer H., Schulz G. (1970). Bases Théoriques des Courbes Européennes de Flambement, In:
Construction Métallique, no.3, pp. 37–57.
Beg D., Kuhlamann U., Davaine L., Braun B. (2010)., Design of Plated Structures-Part 1-5: Design of
Plated Structures, ECCS.
Boissonnade N., Greiner R., Jaspart J.P., Lindner J. (2006). Rules for member stability in EN 1993-1-1,
Background documentation and design guidelines. ECCS (European Convention for Constructional
Steelwork) Publication no. 119, Brussels, 2006.
CEN (2005). EN 1993-1-1, Eurocode 3: Design of steel structures - Part 1-1: General rules and rules for
buildings CEN, Brussels.
CEN/TC250 (2017). Eurocode 3: Design of steel structures - Part 1-1: General rules and rules for build-
ings, CEN/TC 250/SC 3 N 2532 - prEN 1993- 1-1- Final draft.
Chladny´ E, Stujberová M. (2013),” Frames with unique global and local imperfection in the shape of
the elastic buckling mode (part1)”, Stahlbau, 82.8, pp. 609–617.
Chladny´ E, Stujberová M. (2013),” Frames with unique global and local imperfection in the shape of
the elastic buckling mode (part2)”, Stahlbau, 82.9, pp. 684–694.
Domb, M.M. & Leigh, R.L. (2002) Refined design curves for shear buckling of curved panels using non-
linear finite element analysis. 43rd AIAA/ASME/AHS/ASC Structures, Structural Dynamics and
Materials Conference, Denver, U.S.A, Paper #2002-1257.
ESCS Steel RDT Programme, Research Project: Lateral-torsional buckling in Steel and Composite Beams,
No. 7210-PR-183, testing of 4 Tapered Steel Beams. (2002)

39
Featherston, C.A. & Ruiz, C. (1998) Buckling of curved panels under combined shear and compression.
Proceedings of the Institution of Mechanical Engineers, Part C: Journal of Mechanical Engineering Sci-
ence, 212 (3), 183–196.
Featherston, C.A. (2000) The use of finite element analysis in the examination of instability in flat plates
and curved panels under compression and shear. International Journal of Non-Linear Mechanics 35,
515–529.
Featherston, C.A. (2003) Imperfection sensitivity of curved panels under combined compression and
shear. International Journal of Non-Linear Mechanics, 38, 225–238.
Gerard, G. (1959), Handbook of Structural Stability: Supplement to Part III – Buckling of Curved Plates
and Shells, NASA Technical Report D-163, Washington D.C.
Greiner R., Kaim P. (2003a). Comparison of LT-buckling curves with test results, Supplementary
Report, ECCS TC8, No. 2003-10, May 2003.
Greiner R., Kettler M., Lechner A., Jaspart J.P. Weynand K. Ziller, C., Örder, R. (2011). SEMI-COMP+:
Valorisation Action of Plastic Member Capacity of Semi-Compact Steel Sections – a more Economic
Design, RFS2-CT-2010-00023, Background Documentation, Research Programme of the Research
Fund for Coal and Steel – RTD.
Hoglund T. (2014). A unified method for the design of steel beam-columns. Steel Construction, 7, pp.
230–245.
Janss J., Maquoi R., Evaluation of test results on lateral torsional buckling in order to obtain strength func-
tion and suitable model factor, Background report to Eurocode 3.
Johansson B., Veljkovic M., (2001) Steel Plated Structures. Progress in Structural Engineering and Mater-
ials, 3:1 2001.
Kromm, A. (1939) The limit of stability of curved plate strip under shear and axial stresses. NACA. Tech-
nical note No.: 898.
Leggett, D.M.A. (1937) The elastic stability of a long and slightly bent plate under uniform shear. Proc.
R. Soc., A162, 62–83.
Ljubinković, F., Martins, J.P., Gervásio, H., Simões da Silva, L. (2019a), “Eigenvalue analysis of cylin-
drically curved steel panels under pure shear”. Thin-Walled Structures 141, 447–459.
Ljubinković, F., Martins, J.P., Gervásio, H., Simões da Silva, L. (2019b), “Ultimate load of cylindrically
curved steel panels under pure shear”. Thin-Walled Structures 142, 171–188.
Manterola, J. (2008) Pasarela de Peatones - Zaragoza EXPO 2008 [PowerPoint presentation].
Marques L., Taras A., Simões da Silva L., Greiner R., Rebelo, C. (2012), Development of a consistent
design procedure for tapered columns, Journal of Constructional Steel Research, 72, pp. 61–74.
Marques L., Simões da Silva L., Greiner R., Rebelo C., Taras, A. (2013). Development of a consistent
design procedure for lateral-torsional buckling of tapered beams, Journal of Constructional Steel
Research, 89, pp. 213–235.
Martins, J.P., Simões da Silva, L., and Reis, A. (2013) Eigenvalue analysis of cylindrically curved panels
under compressive stresses – Extension of rules from EN1993 1 5. Thin-Walled Structures, 68, pp.
183–194.
Martins J, Simões da Silva L, Reis A. (2014) Ultimate load of cylindrically curved panels under in-
plane compression and bending – extension of rules from EN 1993- 1-5. Thin-Walled Structures
77, 36–47.
Naumes J. (2009). Biegeknicken und Biegedrillknicken von Stäben und Stabsystemen auf einheitlicher
Grundlage, PhD thesis, RWTH Aachen, Germany.
Papp F. (2016). Buckling assessment of steel members through overall imperfection method. Engineering
Structures, 106, pp.124–136.
Pope G.G. (1965), “On the axial compression of long, slightly curved panel ”, Technical Report British
Aeronautical Research Council; Ministry of Aviation, Reports and Memoranda No. 3392.
Rebelo C., Lopes N., Simões da Silva L., Nethercot D., Vila Real P.M.M. (2009). Statistical Evaluation
of the Lateral-Torsional Buckling Resistance of Steel I-beams, Part 1: Variability of the Eurocode 3
resistance model, Journal of Constructional Steel Research., 65, pp. 818–831.
Reis, A. Pedro, J.O., Graça, A.B., Hendy, C., Romoli, P., Simões da Silva, L., Martins, J.P. (2017)
Report on the characterization of relevant parameters of curved plated bridge structures and identification
of bridge cases where they can be found. RFCS Research Project OUTBURST (RFCS-2015-709782):
Deliverable 2.1.
Redshaw, S.C. (1935), The Elastic Instability of a Thin Curved Panel Subjected to an Axial Thrust, Its
Axial and Circumferential Edges Being Simply Supported, Report and Memorandum No. 1565, British
Aeronautical Research Committee.
Salzgeber G. (2000). LT-buckling curves, ECCS TC8, Report No. 2000-001, 20 March 2000.

40
Schildcrout, M. and Stein, M. (1949) Critical combinations of shear and direct axial stress for curved rect-
angular panels. NACA. Technical Note No.: 1928.
Schillo, N. (2017) Local and global buckling of box columns made of high strength steel. Ph.D. thesis,
RWTH Aachen University.
Sekine, H., Tamate, O. (1969), “Postbuckling behavior of thin curved panels under axial compression”,
Bulletin of Japan Society of Mechanical Engineering, 12, pp. 415–420.
Sfintesco D. (1970). Fondement Expérimental des Coubres Européennes de Flambement,
Construction Métallique, no.3, pp. 5–12.
Simões da Silva L., Simões R., Gervásio H., (2016). Design of steel structures. Eurocode3: Design of steel
structures. Part-1-1 - General rules and rules for buildings. 2nd Edition. European Convention for Con-
structional Steelwork, John Wiley & Sons, 2016.
Snijder H.H., Hoenderkamp J.C.D. (2007). Buckling curves for lateral torsional buckling of unrestrained
beams, Rene Maquoi 65th birthday anniversary, 2007, Liège Belguim.
Stowell, E.Z. (1943), “Critical compressive stress for a curved sheet supported along all edges and elastically
restrained against rotation along the unloaded edges”, NACA War Report L-691.
Strating J., Vos H. (1973). Simulation sur Ordinateur de la Coubre C.E.E.M de Flambement á l‘аide de
la Méthode de Monte-Carlo, Construction Métallique, no.2, pp. 23–39.
Szalai J. (2017). Complete generalization of the Ayrton-Perry formula for beam-column buckling
problems. Engineering Structures, 153, pp. 205–223.
Tankova, T, Simões da Silva, L., Marques, L., Rebelo, C. and Taras, A. (2014). Towards a standardized
procedure for the safety assessment of stability design rules, Journal of Constructional Steel Research,
103, 290–302 (2014).
Tankova T., Marques L., Simões da Silva L., Andrade A. (2017). Development of a consistent method-
ology for the out-of-plane buckling resistance of prismatic beam-columns,: Journal of Constructional
Steel Research, 128, pp. 839–852.
Tankova, T., Simões da Silva, L., Marques, L. (2018). Buckling resistance of non-uniform steel members
based on stress utilization: general formulation, Journal of Constructional Steel Research, 149, 239–256
(2018).
Taras A. (2010). Contribution to the development of consistent stability design rules for steel members”
PhD Thesis, Technical University of Graz, Graz, Austria, 2010.
Usami, T. (1993) Effective width of locally buckled plates in compression and bending. Journal of Struc-
tural Engineering, 119 (5), 1358–1373.
Volmir, A. S. (1963), “Stability of Elastic Systems” (in Russian), Fizmatgiz, Moscow. (English transla-
tion: NASA Report AD 628508).
Von Karman T., Sechler E.E., Donnell L.H., (1932). The strength of Thin Plates in Compression, Trans-
actions of the American Society of Mechanical Engineers, vol. 54, p. 53, 1932.
Winter, G. (1947) Strength of Thin Steel Compression Flanges, Transactions of American Society of Civil
Engineers, 112, 527–554.
Winter, G. (1948) Performance of Thin Steel Compression Flanges, Preliminary Publication, 3rd Congress
IABSE, New York, N.Y., 137–148.
Ziemian R.D. (2010). Guide to Stability Design Criteria for Metal Structures, Sixth Edition. John Wiley &
Sons Inc.

41
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Advancements in the stability design of steel frames considering


general nonprismatic members and general bracing conditions

D.W. White, R. Slein & O. Toğay


Georgia Institute of Technology, Atlanta, GA, USA

ABSTRACT: This paper presents an innovative approach for design of planar steel frames
composed of prismatic and/or nonprismatic members. The method uses an inelastic eigenvalue
buckling analysis configured with column, beam and beam-column inelastic stiffness reduction
factors derived from the ANSI/AISC 360-16 Specification provisions to evaluate the member
overall buckling resistances. The resulting procedure provides a relatively rigorous evaluation of
all member strength limit states accounting for moment and axial force variations along the
member lengths, nonprismatic geometry effects, general out-of-plane bracing conditions, and
beneficial end restraint from less critical adjacent unbraced lengths and/or from end boundary
conditions. The approach uses a pre-buckling analysis based on the AISC Direct Analysis
Method to estimate the in-plane internal forces, including second-order effects. Given these forces,
a buckling solution is conducted to evaluate the overall member stability. Other limit states are
addressed by cross-section strength checks given the computed internal second-order analysis
forces. Calculations from this approach are compared with results from recent experimental tests.

1 INTRODUCTION

In recent years, much progress has been achieved in the application of AISC and AASHTO
design criteria toward the efficient design of steel frames using nonprismatic members. The cur-
rent state of the art is captured in the second edition of AISC Design Guide 25 (DG25) (White
& Jeong 2019). In addition to discussing more traditional elastic design methods and their asso-
ciated “manual” calculations, DG25 provides guidance for application of inelastic nonlinear
buckling analysis (INBA) procedures to isolated member unbraced lengths. However, further
advantages can be realized by applying INBA tools to the assessment of entire planar frame
structural systems. This paper provides an overview of the INBA calculations and illustrates the
benefits of this “high end” application of the INBA procedures. Recommended INBA calcula-
tions are applied to isolated critical unbraced lengths as well as to the full test members from
recent experimental tests conducted by Smith et al. (2013).
Specifically, the recommended INBA approach accounts for the effects of:
– Double- and single-symmetry of member cross-sections,
– Single and multiple linear web taper, as well as general continuous variations in the cross-
section dimensions along the member lengths,
– Steps in the cross-section geometry, associated with changes in plate dimensions,
– Any combination of compact, noncompact and slender flanges and/or webs, pertaining to
member flexural resistance,
– Any combination of slender and/or nonslender cross-section plate elements, pertaining to
member axial resistance,
– Any combination of equal or unequal spacing of out-of-plane lateral bracing on one or
both flanges, as well as torsional bracing such as from diagonal members framed between
the inside flanges of frame members and outset girts or purlins of wall or roof systems,

42
– End restraint in critical unbraced lengths due to continuity with adjacent less-critical
unbraced lengths and/or due to physical boundary conditions,
– The combined influence of flexure and axial loading, and
– Load height of transverse loads applied along the member lengths.
These INBA capabilities are implemented within the software system SABRE2 V2 (White
et al. 2019). Tools such as SABRE2 eliminate the need for tedious and relatively approximate
manual calculations of Cb factors, accounting for moment gradient and load height effects,
and effective length factors, K, accounting for column and beam end restraint effects.

2 INBA METHODOLOGY

The following sections explain the net stiffness reduction factors (SRF) employed within the
recommended INBA approach. These factors, derived from the AISC member resistance equa-
tions, are summarized for the cases of axial compression only, flexure only, and combined flex-
ure and axial compression. The corresponding equations are presented in the context of AISC
Load and Resistance Factor Design (LRFD). These SRFs are applied cross section by cross sec-
tion within a general-purpose frame finite element based on thin-walled open-section beam
theory. The frame element has seven dofs per node – three translations, three rotations and one
warping dof – and is formulated to address the influence of nonprismatic geometry (Jeong &
White 2015). The reader is referred to White et al. (2016), Toğay & White (2019) and White &
Jeong (2019) for further calculation details.

2.1 Stiffness reduction factor for axial compression only


The stiffness reduction factor implicit within the AISC (2016) Chapter E axial compression
strength curve may be written as

SRF ¼ 0:877ϕc τa Ae =Ag ð1Þ

where ϕc is the resistance factor for axial compression, taken as 0.9 in AISC LRFD,

 
Pu Pu Pu
τa ¼ 2:724 ln for > 0:390 ð2Þ
ϕc Pye ϕc Pye ϕc Pye

and

τa ¼ 1:0 otherwise ð3Þ

In these equations, Pu is a multiple of the member required LRFD axial resistance Pu, ϕc Pye is
the factored yield strength of the effective cross-section under axial compression, Ae is the effective
cross-section area based on the internal axial force ΓPu, and Ag is the cross-section gross area.
As shown by White et al. (2016) and White & Jeong (2019), when the SRF given by Equation
1 is applied to the section rigidities and a buckling solution is obtained at a multiple of the
applied load Γ, ΓPu for a member subjected to pure axial compression is in effect a rigorous
calculation of the AISC factored design capacity ϕcPn. These solutions include basic prismatic
simply supported columns, where ΓPu = ϕcPn reproduces the exact result from the column resist-
ance equations with K = 1. In addition, they include more sophisticated solutions involving gen-
eral end restraint conditions, continuity with less-critical adjacent unbraced lengths, variations
in internal axial force along the member length, and any type of lateral and/or torsional bracing
such as lateral bracing offset from the centroidal or shear center axis.

43
2.2 Stiffness reduction factor for flexure only
The stiffness reduction factor implicit within the AISC Chapter F I-section flexural resistance
equations may be written as

SRF ¼ ϕb Rpg τltb ð4Þ

where ϕb is the resistance factor for flexure (0.9 in AISC LRFD), Rpg is the bend buckling factor
for slender-web members, equal to 1.0 if the web is compact or noncompact, and τltb is the base
lateral-torsional buckling (LTB) stiffness reduction factor. The factor τltb may be expressed as
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u
u Y 4X 2 F
τltb ¼ th  2 i when m4 L ð5Þ
6:76X 2 Fyc =E m2 þ 2Y 2 F yc

for compact- and noncompact-web I-sections, where m = ΓMu/Myc, and where


2  3
1  Rmpc     
Lr Lp Lp 5 Fyc 1
Y ¼ m4  þ ð6Þ
1  RpcFLFyc rt rt rt E 1:95

and

X 2 ¼ Sxc ho =J ð7Þ

In the above equations, FL is the stress beyond which the inelastic LTB limit state applies under
uniform bending, and Fyc is the yield strength of the flange in flexural compression. In addition,
ΓMu is a given multiple of the required LRFD moment Mu, and Myc is the yield moment to the
compression flange. The following terms are as defined in the AISC Specification: J = the cross-
section St. Venant torsion constant; Lr = the limiting unbraced length for inelastic LTB under
uniform bending; Lp = the limiting unbraced length corresponding to the LTB “plateau” strength
(i.e., the compression flange yielding (CFY) strength) under uniform bending; Rpc = the web plas-
tification factor; Sxc = the elastic section modulus to the compression flange; ho = the distance
between the centroids of the I-section flanges; and rt = the effective radius of gyration for LTB.
For slender-web I-sections, the following simpler form applies for the base LTB factor:

2 32
R  m 1 rffiffiffiffiffiffiffi 
m 4 h Rpg Fyc c c5 Rpg FL
τltb ¼   þ when m4 ð8Þ
Rpg Rh  Fyc
FL F L π π Fyc

where c is the coefficient in the equation for Lp, equal 1.1 for general welded I-section mem-
bers in the current AISC Specification.
For all I-section members, when m ≤ RpgFL/Fyc, the base LTB stiffness reduction factor is

τltb ¼ 1:0 ð9Þ

As demonstrated by White et al. (2016) and White & Jeong (2019), when the SRF given by
Equation 4 is applied to the section rigidities EIy, ECw and GJ and a buckling solution is obtained
at a multiple of the applied load Γ, ΓMu for a member subjected only to bending is in effect
a rigorous calculation of the AISC ϕcMn for LTB. This includes basic prismatic simply-supported
beams subjected to uniform bending, where the buckling solution reproduces the exact result from
the AISC LTB resistance equations for Cb = 1. In addition, it includes more advanced solutions
involving general nonprismatic geometry, complex end restraint conditions, continuity with less-

44
critical adjacent unbraced lengths, general variations in internal moment along the member
length, transversely applied loads at a specified height within the cross-section, and any type or
combination of lateral and/or torsional bracing.
In Equations 5, 6 and 8, the term m may be expressed as

M M Mmax
m¼ ¼ Rpg Rpc ð10Þ
ϕb Myc ϕb Mmax Mmax:CFY

where

Mmax ¼ min ðMn:CFY ; Mn:FLB ; Mn:TFY Þ ð11Þ

is the maximum possible cross-section resistance based on the separate limit states of com-
pression-flange yielding (CFY), flange local buckling (FLB), or tension flange yielding (TFY).
The SABRE2 software (White et al. 2019) implements the cross section based CFY, FLB and
TFY yielding checks from the Specification, in addition to the fundamental LTB resistance
checks, which are captured via inelastic buckling analysis. If the maximum cross-section resistance
ϕbMmax is reached at any location prior to the onset of LTB, the available member resistance is
limited by this cross-section resistance. With the exception of checking ϕbMmax, the INBA algo-
rithm for beams is essentially the same as that for the calculation of the column buckling load in
Section 2.1.

2.3 Stiffness reduction factors for combined axial tension or compression and bending
As discussed in the introduction, INBA methods can be applied to assess the strength of any type
of I-section member subjected to in-plane bending and axial load, accounting for member overall
stability limit states and their potential interaction with cross-section based limit states. The INBA
procedures accomplish this in a more rigorous manner than can be achieved by routine applica-
tion of Specification resistance equations. For members subjected to combined axial loading and
flexure, this is achieved by a straightforward interpolation between the SRFs for axial loading dis-
cussed in Section 2.1 and the SRFs associated the AISC flexural resistance equations in Section
2.2. The equation for the interpolated beam-column inelastic stiffness reduction factor is
   
ζ Ae ζ
SRF ¼ 0:877ϕc τa þ 1  o ϕb Rpg τltb ð12Þ
90o Ag 90

where
 
Pu =ϕc Pye
ζ ¼ atan ð13Þ
Mu =ϕb Mmax

Furthermore, in the equations for τa and τltb, the unity check (UC) value from the following
cross-sectional strength interaction equations is substituted for ΓPu/ϕcPye and ΓMu/ϕbMmax:
– For cross sections in which all the plates are nonslender under axial compression and com-
pact under flexural compression:

Pu 8 Mu Pu


UC ¼ þ for  0:2 ð14Þ
ϕc Pye 9 ϕb Mmax ϕc Pye

Pu Mu
UC ¼ þ otherwise ð15Þ
2ϕc Pye ϕb Mmax

45
– For cross sections with slender plates under axial compression, and/or with noncompact or
slender plates under flexural compression:

Pu Mu
UC ¼ þ ð16Þ
ϕc Pye ϕb Mmax

The net stiffness reduction factor from Equation 12 is applied to the rigidities EIy, ECw and
GJ on a cross_section-by-cross_section basis. Toğay & White (2019) demonstrate the accuracy
of the above interpolation for a comprehensive suite of prismatic I-section members. Further
details regarding the corresponding INBA calculations are discussed in (White et al. 2016).

2.4 Rationale for the specific recommended INBA approach


There are numerous ways to characterize the stiffness of steel structures for inelastic nonlinear
buckling analysis (INBA). These range from refined plastic zone analysis, in which the detailed
spread of plasticity is tracked through the member cross sections and along their lengths as the
loads are increased, including consideration of residual stress and geometric imperfection effects,
to other phenomenological approximations comparable to the SRFs discussed above.
The INBA calculations using the above SRFs provide results that are fully consistent with the
application of the AISC Direct Analysis Method and the Specification member resistance equa-
tions for basic prismatic members, and they extend the application of the AISC provisions to gen-
eral member geometries, loadings, end restraints, and bracing conditions. Solutions employing
refined plastic zone analysis arguably have the greatest level of rigor due to their ability to directly
capture the influence of any specified member cross-section geometry, residual stresses and geo-
metric imperfections. However, appropriate nominal residual stresses and geometric imperfections
must be specified, and the results from plastic zone analysis never match precisely with predictions
from the Specification equations in cases where a close or exact match might be expected or
desired. The AISC Specification equations are a “codified” fit to member strengths considering
these effects for a general range of steel structures. The above SRF values capture this fit for basic
cases, and allow extension of the Specification rules to more general structures.

3 POTENTIAL IMPROVEMENTS IN AISC SPECIFICATION RESISTANCE


EQUATIONS

The AISC Specification resistance equations (AISC 2016) have many excellent qualities in
terms of their ability to represent the strength limit states of steel I-section members and
frames. However, potential improvements to these provisions may further enhance their abil-
ity to capture these strength limit states. These improvements are summarized below. Add-
itional details are explained in Subramanian et al. (2018) and Toğay & White (2019). Since the
SRFs within the INBA approach depend on the underlying Specification resistance equations,
these potential improvements are important in demonstrating the merits of the approach.

3.1 Lateral-torsional buckling strength improvements


For major-axis bending of welded I-section members, Subramanian et al. (2018) have demon-
strated that the reliability index, estimated based on existing experimental data, is somewhat lower
than the target value of β = 2.6. They recommend that the parameter FL, defined as the stress limit
beyond which the inelastic LTB limit state applies under uniform bending, should be taken as

FL ¼ 0:5Fyc ð17Þ

for these member types. In addition, they recommend that the limiting unbraced length corres-
ponding to the “plateau” strength under uniform bending should be taken as

46
pffiffiffiffiffiffiffiffiffiffiffiffiffi
Lp ¼ 0:8rt E=Fyc ð18Þ

for routine design of welded I-sections utilizing the AISC resistance equations and assuming
warping free and lateral-bending free conditions at the ends of the unbraced lengths, and that
pffiffiffiffiffiffiffiffiffiffiffiffiffi
Lp ¼ 0:63rt E=Fyc ð19Þ

should be employed when end restraint is considered explicitly in the LTB evaluation. Equations
17 and 18 are effectively the same as the original recommendations for slender-web plate girders
by Cooper et al. (1978). It should be noted that in the recommended FLB calculations discussed
below, the corresponding FL should still be taken as 0.7Fyc as in the current AISC Specification.

3.2 Web bend buckling strength improvements


Subramanian et al. (2018) have also recommended that the noncompact-web limit in the
AISC Specification, λrw, which establishes the transition between noncompact and slender
web behavior, and influences the calculated values for the web plastification factor, Rpc, and
the web bend buckling factor, Rpg, should be modified to
pffiffiffiffiffiffiffiffiffiffiffiffiffi
λrw ¼ crw E=Fyc ð20Þ

where crw = 3.1 + 5/aw, but not less than 4.6 nor larger than 5.7, aw = 2Dcy tw/bfc tfc, bfc and tfc
are the width and thickness of compression flange respectively, Dcy is the depth of the web in
compression at the nominal onset of compression flange yielding, and tw is the thickness of the
web. Equation 20 recognizes that I-section members with relatively small compression flanges
tend to exhibit a reduction in the effective noncompact web limit.

3.3 Improved characterization of compression flange local buckling resistance


The AISC FLB provisions tend to underestimate I-section member flexural resistances when the
compression flange becomes increasingly slender. This is because the AISC equations do not
account for the reserve local post-buckling capacity. The following calculations consider an effect-
ive width of the compression flange to account for its local post-buckling strength.
For sections with a slender compression flange in flexure:
1) The flange effective width is calculated directly given the flange elastic buckling stress

0:9Ekc
Fel ¼ ð21Þ
ðbfc =2tfc Þ2

and taking the compression flange stress within the effective width as Fyc at the flexural
strength limit, where kc is the flange local buckling coefficient defined by the AISC Specifica-
tion. The terms Feℓ and Fyc are substituted into Winter’s unified effective width equation
 qffiffiffiffiffi qffiffiffiffiffi
be ¼ bf 1  0:22 FFyce‘ Fe‘
Fyc ð22Þ

2) The location of the effective cross section’s neutral axis at nominal initial yielding of the
compression flange, measured relative to the inside of the compression flange, Dcye, and
the corresponding yield moment, Myce, are determined.

47
3) The FLB resistance, considering the flange local post-buckling strength, is calculated as
RpgMyce, where Rpg is less than 1.0 for slender-web sections but is equal to 1.0 for compact-
and noncompact-web sections.
For sections having a noncompact flange in flexure:
1) The effective width reduction based on the noncompact flange slenderness limit, λrf, is
applied to the compression flange, regardless of the actual flange slenderness, and the corres-
ponding resistance Myce(λr) = RpgMyce(λr) is determined using the procedure explained
above. This establishes an “anchor point” corresponding to λf = λrf.
2) A linear interpolation is then employed between (λpf, Mmax.FLB) and (λrf, Myce(λr)),
where Mmax.FLB is the plateau resistance for FLB, equal to Mp for a compact-web section,
RpcMyc for a noncompact-web section, and RpgMyc for a slender-web section, where Myc is
the yield moment to the compression flange for the gross cross-section.

3.4 Improved handling of tension flange yielding


When a singly-symmetric section with the larger flange in compression is subjected to flexure, the
current AISC flexural resistance may be governed by tension flange yielding (TFY). If the section
has a slender web, the TFY resistance is equal to the moment at the first nominal yielding of the
tension flange, Myt. This estimate can be quite conservative. Sections with Myt < Myc, can have
substantial inelastic reserve strength associated with distributed yielding in flexural tension. The
conservative TFY calculation can be eliminated, and the Specification can be substantially short-
ened, by calculating Myc and Myce as the “true” yield moments to the compression flange, consid-
ering the early yielding in tension for these section types. It is recommended that these true yield
moments to the compression flange should be used in the limit state calculations of the Specifica-
tion. In addition, it is recommended that the depth of web in compression at the first nominal
yielding of the compression flange, based on the gross-cross-section, Dcy, be used in calculating
the slenderness of the web. That is, λw is defined as 2Dcy/tw. Figure 1 shows an example stress
distribution at Myc for a homogeneous I-section of this type. For homogeneous cross-sections,
relatively simple closed-form equations are available for Dcy and the “true” Myc.

3.5 Calculation of flexural resistance for members with unequal flange and web yield strengths
Measured yield strengths generally can be different for both flanges and for the web in experimen-
tal tests. Measured yield strengths on thinner web material are often larger than the flange yield
strengths. The measured yield strengths should be employed when comparing strength predictions
to experimental test results. In addition, in bridge construction, it is common to use “hybrid” I-gir-
ders, having a lower-grade steel for the top flange and web combined with a higher-grade steel for
the bottom flange. To accommodate all of these considerations, it is important to define the calcu-
lation of the flexural resistance for any combination of plate yield strengths.
The recommended extensions to the AISC I-section member provisions are as follows:
– The compression flange yield strength, Fyc, should be employed for Fy in the AISC provisions
everywhere Fy appears either within the context of the compression flange, or within the con-
text of assessing any aspects related to structural stability. This is an established precedent in
(AASHTO 2017) and elsewhere. It should be noted that the flange in flexural compression
depends on the sign of the bending moment.
– The actual or specified yield strengths of the compression and tension flanges, Fyc and Fyt, and
of the web, Fyw, should be employed in calculating the plastic moment, Mp, regardless of the
relative magnitude of the different strengths, except Fyw should not be taken larger than 1.43
min (Fyc, Fyt). This is based on AASHTO (2017) Article 6.10.1.3 and is intended to avoid
counting on web yield strengths beyond the limits that have been evaluated experimentally.
– The “true” yield moments to the compression flange, Myc and Myce as applicable, should
be calculated from a strain-compatibility analysis including any early yielding in the web or

48
Figure 1. Stress distribution associated with the “true” yield moment Myc and the corresponding depth
of the web in compression, Dcy, for a homogeneous cross-section with Myt < Myc.

tension flange. The above 1.43 min (Fyc, Fyt) limit on the useable Fyw also should be applied
in this calculation. These moments should be employed where the corresponding “true”
yield moments appear within the calculations discussed above. Evaluation of the true Myc
and Myce values is straightforward to program, and SABRE2 (White et al. 2019) imple-
ments this calculation. The algorithm sets the strain at the extreme fiber of the compression
flange to Fyc/E and the section curvature is varied until a stress distribution is obtained for
which the total cross-section axial force is zero.

4 COMPARISON OF INBA PREDICTONS TO EXPERIMENTAL RESULTS

Smith et al. (2013) have conducted 10 experimental tests evaluating the LTB behavior of a range
of web-tapered I-section members. The primary aim of these tests was to gain a better understand-
ing of the cyclic LTB behavior of these types of members. However, all of the members were
loaded past their flexural capacity within an initial monotonic half-cycle of the loading; therefore,
these tests are also valuable for evaluating static monotonic strength predictions. Smith et al. pro-
vide an overall positive assessment of the ability of the first edition of DG25 (Kaehler et al. 2011)
to predict the LTB resistance under static monotonic loading, contingent upon the consideration
of end restraint effects from support conditions and less-critical adjacent unbraced lengths using
elastic eigenvalue buckling calculations. The following discussions complement the assessments by
Smith et al. by comparing INBA calculations based on the current AISC Specification, as well as
the AISC provisions with the improvements discussed in Section 3, to the experimental results.
The overall configuration of the experimental tests conducted by Smith et al. (2013) is illus-
trated in Figure 2. The specimens were tested in a horizontal orientation, simulating the rafter of
a metal building frame, with moment applied at the north end of the specimen via an end-plate
connection to a vertical loading column. The south end of the specimens was flexurally and tor-
sionally simply supported, i.e. major- and minor-axis bending rotations, warping of the flanges,
and longitudinal displacements were unrestrained, but torsional rotation and vertical and out-of-
plane lateral displacements were prevented. Minor-axis bending and torsional rotations, warping
of the flanges, and out-of-plane lateral displacements were effectively prevented at the north end
of the specimens at the end plate connection to the loading column, and longitudinal and vertical
displacements were restrained by a pin support below the knee at the bottom of the column.
Flange-level out-of-plane lateral bracing was provided at different locations along the top and
bottom flanges of the specimens. A typical conceptual arrangement of these lateral braces is indi-
cated by the x symbols on the drawing. Two of the 10 tests included a constant axial load applied

49
Figure 2. Test configuration, adapted from (Smith et al. 2013).

to the specimens. This was accomplished by tensioning of rods between the north side of the
column at the knee of the frame and the south end of the specimen.
Table 1 summarizes all the pertinent geometry and material attributes of the test specimens.
Three groups of tests were conducted as denoted by the test names:
1) The CF tests had constant taper throughout the test length and the critical unbraced length
for LTB was the first unbraced length adjacent to the column.
2) The CS tests had constant taper throughout the test length, but the critical unbraced length
was the second unbraced length from the column.
3) The PF tests had a pinch point within the test length, and the critical unbraced length was
the first unbraced length adjacent to the column. Test PF1 had a pinch point at the south
end of its critical unbraced length while test PF2 had a pinch point at an intermediate loca-
tion within its critical unbraced length.
Clearly there is substantial complexity in the combined overall configuration of the member
geometry and plate yield strengths, and the bracing and end restraint conditions in these tests.
Table 2 summarizes the test to predicted strength ratios and the flexural failure modes identified
from INBA solutions conducted using SABRE2. This table shows the analysis results using the
AISC (2016) provisions as well as the AISC provisions with the potential improvements defined
in Section 3. (The different plate yield strengths are included in the current AISC calculations as
specified in Section 3.5, except that the AASHTO (2017) hybrid cross-section factor, Rh, is
employed along with the current calculations as a commonly employed approximation, rather
than calculate the “true” yield moments; the reader is referred to (AASHTO, 2017) for the specific
equations.) In addition, the INBA calculations are performed in two ways:
1) The entire test specimen is modeled. This captures the influence of end restraint on the crit-
ical unbraced length from the less-critical adjacent lengths due to continuity across the
braced points. In addition, the specified end conditions at the end-plate connection to the
column (minor-axis bending and flange warping fixed) are modeled in these solutions.
2) Only the critical unbraced length is modeled. In this case, the common design assumption
of torsionally simply-supported end conditions (minor-axis bending and flange warping
free at both ends of the unbraced length) is employed in the SABRE2 solutions. This is the
inherent assumption associated with the common implicit use of a LTB effective length
factor K = 1, and the use of just the unbraced length Lb rather than a KLb < Lb for the
critical unbraced length in design practice.

50
Table 1. Summary of specimens tested by Smith et al. (2013).
bf tf h1# tw α* Lb♣ * Fyf Fyw
Test^ (mm) (mm) (mm) (mm) (°) Flg. Web (m) (MPa) (MPa)

CF1 152 9.37 305 4.70 4.58 C♦ S♦ 2.13 (1.07, 1.07), 1.22, 2.35 431 427
12.65♠ 414*
CF2 & 152 9.40 305 4.72 4.58 C S 2.13 (1.07, 1.07), 1.22, 2.35 397 496
CF2-A
CS1 & 152 6.45 203 6.22 5.60 N♦ N 0.610, 3.05 (1.52, 1.52), 2.04 425 428
CS1-A
CS2 127 6.53 305 4.70 4.58 N S 0.406, 3.05 (1.52, 1.52), 2.25 481 427
CS3 152 9.37 305 4.70 4.58 C/ S 0.914, 2.74 (1.37, 1.37), 2.04 431 427
8.08 N 481
8.08** 462**
CS4 152 7.80 305 4.72 4.58 N S 0.610, 3.05 (1.52, 1.52), 2.04 376 496
4.22^^ 460^^
PF1## 152 9.37 356 4.70 9.46 C S 2.45 (1.23, 1.23), 1.05, 2.20 431 427
PF2♣♣ 127 7.87 457 4.67 14.0 C/ S 2.44 (1.23, 1.21), 1.22, 2.04 383 396
6.20 3.81 N 468 401
^
The nominal web depth at the end plate is the same, h2 = 762 mm, for all of the tests.
#
Web depth at the simple-support, at the right-hand end of the test.
*
Taper angle

The unbraced lengths for the top flange (in flexural compression) are the values not listed in
parentheses.
*
The unbraced lengths for the bottom flange (in flexural tension), are the same as those for the top flange
except for one segment where an additional intermediate brace is placed on the bottom flange. The brace
spacings for the segment containing the additional bottom flange brace are listed in parentheses just after
the corresponding top flange unbraced length. The corresponding top (compression) flange unbraced
length is the critical one for lateral-torsional buckling of the members.

CF1 is the only linearly-tapered member test that has nominally different top and bottom flange dimen-
sions; the first and second values listed correspond to the top and bottom flanges respectively (the top
flange is in flexural compression). The resulting singly-symmetric section has Myt > Myc, and therefore
the nominal onset of yielding occurs first at the top (compression) flange.

C indicates that the flange is compact within the critical unbraced length, N indicates that the corres-
ponding flange or web plate is noncompact within the critical unbraced length, and S indicates that the
web is slender within the critical unbraced length.
**
CS3 has a flange splice in both flanges at 2.29 m from the end plate; the second and third reported
values correspond to the top and bottom flange plates to the right of the flange splice.
^^
CS4 has a web splice at 0.610 m from the end plate; the second reported value corresponds to the web
to the right of the web splice.
##
PF1 has a linear taper from the end plate down to a pinch point at the brace location at 2.44 m from
the end plate, then a constant web depth of 356 mm to the right of that location; the flange and web
plates are the same on each side of this pinch point.
♣♣
PF2 has a linear taper from the end plate down to a pinch point at the brace location at 1.22 m from
the end plate, then a constant web depth of 457 mm to the right of that location. The top flange plate is
thicker within the tapered length of the member, resulting in Myt being (slightly) less than Myc and cor-
responding minor early yielding in flexural tension. Also, the web plate thickness is reduced to the right
of the pinch point. The second value listed for the flange plate thickness corresponds to the flange plates
other than the thicker top flange plate within the critical unbraced length, and the second value for the
web plate thickness corresponds to the web plate to the right of the pinch point.

The INBA solution considering the entire specimen and using the AISC provisions with the
recommended potential improvements provides the best predictions of the test results, giving
an average strength ratio of 1.07 along with the smallest coefficient of variation (COV = 0.06)
and a minimum strength ratio of 0.98. In addition, this solution identifies the governing failure
mode clearly as LTB in all of the tests except PF2. In PF2, FLB governs for the thinner flange
just to the south of the pinch point. PF2 exhibited a local failure at the pinch point within the
experimental test, due to the lack of sufficient capacity of the web to resist the concentrated

51
Table 2. Test to predicted strength ratios, flexural failure modes identified by SABRE2, and moment capacities for the tests conducted by Smith et al. (2013).
Entire member modeled Only critical unbraced length modeled

Current AISC AISC w/recom. Current AISC AISC w/recom.

Test/Sum. Test/Pred. Failure Test/Pred. Failure Test/Pred. Failure Test/Pred. Failure Moment Capacity
Stats. Strength Mode Strength Mode Strength Mode Strength Mode Mtest♣ (kN m)

CF1 0.99 CFY 1.05 LTB 1.10 LTB 1.26 LTB 633.4
CF2 1.10 CFY 1.15 LTB 1.21 LTB 1.37 LTB 639.3
CF2-A* 1.10 CFY 1.16 LTB 1.22 LTB 1.38 LTB 611.9
CS1 0.99 FLB 1.02 LTB 1.26 LTB 1.40 LTB 424.3
CS1-A# 1.05 FLB 1.11 LTB 1.52 LTB 1.62 LTB 429.9
CS2 0.86 LTB 1.00 LTB 1.62 LTB 1.68 LTB 333.2
CS3 1.02 FLB 1.08 LTB 1.13 LTB 1.31 LTB 536.7
CS4 1.01 FLB 1.07 LTB 1.15 LTB 1.34 LTB 413.0
PF1 0.95 CFY 0.98 LTB 1.06 LTB 1.19 LTB 477.0
PF2♠ 0.81 FLB 0.79 FLB 0.81 FLB 0.83 LTB 187.9
Avg. 1.01 1.07 1.25 1.39
COV 0.07 0.06 0.15 0.11
Max. 1.10 1.16 1.62 1.68
Min. 0.86 0.98 1.06 1.19
*
CF2-A had a constant axial compression of 126 kN applied throughout the experimental testing. This load was applied at 318 mm above the bottom of the web at the left face
of the column, and at 140 mm above the bottom of the web at the simply-supported end of the specimen. This corresponds approximately to the centroidal depth of the cross-
section along the test length.
#
CS1-A had a constant axial compression of 185 kN applied throughout the experimental testing. The stated intent was to apply this load at the centroidal depth of the cross-
section along the test length. However, the information provided by Smith et al. (2013) indicates that the resultant of the axial load was only at 50.8 mm above the bottom of the
web at the simply-supported end of the test in CS1-A. The load was applied at 318 mm above the bottom of the web at the left face of the column.

The moment capacities from the experimental tests reported here are the values at the deepest end of the critical unbraced length as identified by Smith et al. (2013) and by the
SABRE2 full member solutions based on the AISC provisions with recommended improvements. The corresponding locations in the test specimens are at the deepest end of the
top flange unbraced length adjacent to the column for the CF tests, and at the deepest end of the second top flange unbraced length from the column in the CS tests. In the PF
tests, the “controlling cross-section” was identified by Smith et al. (2013) as the cross-section at the small end of the critical unbraced length for test PF1, and the cross-section on
the thin-web side of the pinch point splice at the middle of the critical unbraced length for test PF2. In the SABRE2 solutions, these are the cross sections that have the largest
internal moment relative to the cross-section flexural strength, and the corresponding smallest values of the SRF.

The results for test PF2 are not included in the summary statistics since the strength in this test was governed by a local failure at the pinch point.
transverse force caused by the change in angle of the top flange plates. It is expected that if
a partial-depth bearing stiffener had been provided at the pinch point, the predictions would
be accurate for PF2.
The INBA solutions using the current AISC provisions give the best prediction of the tests
on average, with a mean test/predicted strength of 1.01; however, they have a larger COV of
0.07 and they give a relatively small test/predicted strength of only 0.86 for test CS2. In add-
ition, the INBA solution using the current AISC provisions predicts LTB as the governing
failure mode only for test CS2. Based on the experimental results, LTB was clearly the domin-
ant failure mode for all the tests, with the exception of PF2, as discussed above.
The solutions based on both the current AISC provisions as well as the AISC provisions with
the recommended potential improvements exhibit significant conservatism when applied only to
the critical unbraced lengths, assuming torsionally simply-supported boundary conditions at the
ends of these lengths. The current AISC provisions actually give the more accurate predictions in
these cases, due to their tendency to predict larger LTB strengths in general for these tests.
Figure 3 shows the buckling modes obtained from SABRE2 for tests CS2 and PF1, using the
AISC provisions with the potential improvements. The darker arrows in the figure indicate the
constraints from the end conditions and the intermediate lateral bracing. The light shaded circular
arrow at the left-hand (north) end of the models shows the applied moment from the loading
column. One can observe the influence of the warping and minor-axis bending restraint at the left-
hand ends, as well as the effect of the close spacing of the first set of intermediate braces from the
left-hand end in CS2 (preventing out-of-plane displacement and twist) on the buckled shape.
Figure 4 plots the SRF and cross-section unity check (UC) values along the normalized speci-
men lengths for tests CS2 and PF1. The nonlinear variation of the UC for CS2, corresponding to
the noncritical FLB limit state in this test, is due to the linear taper of the web depth. The smallest
SRF values for CS2 are at the left-hand end where the cross-section is deepest. However, these
SRF values occur within the short 0.406 m unbraced length adjacent to the loading column. The
CS2 specimen is critical for LTB within the second unbraced length from the column. Due to
sharper web taper between the left-hand support and the pinch point at the first brace from the
column on the top flange, the SRF values for PF1 reduce to a minimum at the pinch point. The
UC is equal to 1.0 at the pinch point in PF1, indicating that the CFY limit state (i.e. the plateau
strength for LTB) is close to being reached at this location. The maximum UC and minimum
SRF values tend to occur at the “governing cross-sections” in DG25 (White and Jeong 2019)
strength checks based on more routine elastic LTB solutions (Slein and White 2019). However,
the INBA solutions of the entire test members account for the “true” restraint from adjacent less-
critical unbraced lengths with better accuracy than can be achieved using elastic LTB solutions.
The predicted capacities for the experimental tests discussed in this paper are affected only
a minor extent by the improved handling of the FLB and TFY limit states discussed in Sections
3.3 and 3.4. Toğay and White (2019) show test simulation solutions that highlight the benefit of
these improvements. Furthermore, there is evidence from these and other experimental tests that
the LTB resistance of I-section members fabricated with minimal single-side welding of the flanges
to the webs may be somewhat larger than characterized by the more generally applicable

Figure 3. Lateral-torsional buckling modes for tests CS2 and PF1 obtained from INBA solution based
on the AISC Specification with the potential improvements discussed in Section 3.

53
Figure 4. Cross-section unity check (UC) and stiffness reduction factor (SRF) values versus the normal-
ized position along the length at the LTB strength limit for tests CS2 and PF1, AISC Specification based
INBA solution including potential improvements discussed in Section 3, ϕc and ϕb taken equal to 1.0.

recommendations by Subramanian et al. (2018) listed in Section 3.1. Additional research is needed
to further evaluate rules for characterization of the LTB resistance of welded I-section members.

5 CONCLUSIONS

Overall member stability can be influenced significantly by restraint from adjacent less-critical
unbraced lengths as well as general member end restraint, bracing, and loading conditions that
are difficult to capture within Specification resistance equations themselves. By implementing the
Specification resistance equations as corresponding stiffness reduction factors within a general-
purpose buckling analysis based on thin-walled open-section beam theory, the impact of the
above attributes can be captured by direct modeling of the actual structural conditions.

REFERENCES

AISC 2016. Specification for structural steel buildings, ANSI/AISC 360-16, Chicago: American Institute
of Steel Construction.
AASHTO 2017. AASHTO LRFD bridge design specifications, 8th Ed., Washington: American Associ-
ation of State Highway and Transportation Officials.
Cooper, P.B., Galambos, T.V. & Ravindra, M.K. 1978. “LRFD criteria for plate girders,” Journal of the
Structural Division, ASCE, 104(ST9): 1389–1407.
Jeong, W.Y. & White, D.W. 2015. “General nonprismatic frame finite element based on thin-walled
open-section beam theory,” Research Report, School of Civil and Environmental Engineering, Geor-
gia Institute of Technology, Atlanta, GA.
Kaehler, R.C., White, D.W. and Kim, Y.D. 2011. Frame design using web-tapered members, Steel design
guide 25, 1st Ed., Chicago: American Institute of Steel Construction.
Slein, R. and White, D.W. 2019. “Streamlined design of nonprismatic I-section members,” Proceedings,
Annual Stability Conference, Structural Stability Research Council, St. Louis, MO.
Smith, M. D., Turner, A.K., & Uang, C.M. 2013. “Experimental study of cyclic lateral-torsional buckling
of web-tapered I-beams”. Final Report to Metal Building Manufacturers Association, Department of
Structural Engineering University of California, San Diego, La Jolla, CA.
Subramanian, L., Jeong, W.Y., Yellepeddi, R., & White, D.W. 2018. “Assessment of I-section member
LTB resistances considering experimental test data and practical inelastic buckling design
calculations,” Engineering Journal, AISC, 55(1): 15–44.
Toğay, O. & White, D.W. 2019. “Advanced design evaluation of planar steel frames composed of general
nonprismatic I-section members”. Research report, Georgia Institute of Technology, Atlanta, GA.
White, D.W., Jeong, W.Y. & Toğay, O. 2016. “Comprehensive stability design of planar steel members
and framing systems via inelastic buckling analysis,” International Journal of Steel Structures, 16(4):
1029–1042.
White D.W & Jeong, W.Y. 2019. Frame design using nonprismatic members, Steel design guide 25, 2nd
Ed., Chicago: American Institute of Steel Construction (to appear).
White, D.W., Toğay, O., Slein, R. & Jeong, W.Y. 2019. “SABRE2-V2”. <white.ce.gatech.edu/sabre>.

54
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Design by advanced analysis–2016 AISC specification

R.D. Ziemian
Bucknell University, Lewisburg, Pennsylvania, USA

Y. Wang
Cornell University, Ithaca, New York, USA

ABSTRACT: At the heart of the provisions for assessing structural stability within the
American Institute of Steel Construction’s Specification for Structural Steel Buildings is the
direct analysis method. The fundamental concept is that the more behavior that is explicitly
modeled within the analysis, the simpler it is to define the Specification’s design requirements.
In other words, the direct analysis method consists of calculating strength demands and avail-
able strengths according to a range of well-defined and fairly detailed analysis requirements.
This paper begins with an overview of two logical extensions to AISC’s direct analysis
method, both of which are now provided in the Specification’s Appendix 1 – Design by
Advanced Analysis. In establishing these approaches, many systems were investigated and it
was found that systems with beam-columns subject to minor-axis bending appeared to deserve
additional attention. This paper presents a detailed study that investigates such members.

1 INTRODUCTION

For the past sixty years, the Effective Length Method (ELM) has been a widely employed stability
design method (Ziemian, 2010). By scaling actual unbraced lengths to effective lengths when cal-
culating the available strengths of compression members, the effective length K-factor is assumed
to account for most factors known to impact the stability of structural systems, including geomet-
ric system imperfections, stiffness reduction due to inelasticity, and to a much lesser degree uncer-
tainty in strength and stiffness (AISC, 2016a). In 2005, design by the Direct Analysis Method
(DM) first appeared in the American Institute of Steel Construction’s (AISC’s) Specification for
Structural Steel Buildings. In DM, the available strengths of compression members are based
simply on the unbraced length (K = 1), as long as system imperfections (but not member imper-
fections) and stiffness reduction due to inelasticity are represented in the structural analysis. Since
then, many in the structural design profession have moved from employing ELM to DM. As
a result, DM was relocated in AISC’s 2010 Specification from Appendix 7 to Chapter C, while
ELM was relocated from Chapter C to Appendix 7.
Both design methods rely on establishing the unbraced lengths of compression members, which
in some cases may be difficult, if not impossible, to define. Examples include, but are not limited
to arches, tree columns, and Vierendeel trusses. In response to this predicament, AISC introduced
a Design by Advanced Elastic Analysis Method that appears in Appendix 1 of their 2016 Specifica-
tion. In addition to the analysis modeling requirements of DM, the method further requires the
direct modeling of member imperfections and, therefore the method is often represented by the
acronym DMMI. In applying this approach, engineers can avoid the complexities of defining
unbraced lengths, thereby being permitted to compute the nominal strengths of compression
members as their axial cross-sectional strengths. This paper reports on an ongoing study to com-
plement previous studies on systems (Nwe Nwe, 2014, Giesen-Loo, 2016) to evaluate the perform-
ance of DMMI, especially with an eye towards members that are subject to the combination of
compression and minor-axis bending. Using AISC’s Design by Advanced Inelastic Analysis

55
Method, which is based on employing a rigorous geometric and material nonlinear analysis with
imperfections (GMNIA), the accuracy of DMMI is assessed and further compared with the more
traditional ELM and DM design methods. Additionally, the significances of thermal residual
stresses, which are a consequence of uneven cooling of rolled cross-sections, and the axis of bend-
ing (minor versus major) are also explored.
The paper begins by providing an overview of AISC’s ELM, DM, DMMI, and GMNIA
methods along with details of the analysis method and interaction equation employed in each.
Results of the study are then presented primarily in tabular format, which are followed by
discussions of the effects residual stresses, axis of bending, and design method employed.

2 OVERVIEW OF DESIGN METHODS

In this study, the ends of simply supported columns of various slenderness ratios are subjected
to a wide range of combinations of applied axial force and end bending moments that are of
equal magnitude and opposite direction (in the absence of axial force such moments would
produce a uniform moment distribution). In all cases, the members are assumed to be fully
braced out-of-the-plane of bending. To assess the LRFD strength of beam-columns based on
an elastic analysis, the following interaction equation is provided in AISC’s Specification:
 
Pu 8 Mux Muy
þ þ  1:0 for Pu =Pn  0:2 ð1Þ
Pn 9 Mnx Mny
 
Pu Mux Muy
þ þ  1:0 for Pu =Pn 50:2 ð2Þ
2Pn Mnx Mny

where  ¼ 0:9, Pu is the required axial strength, Mu is the required bending strength, Pn is the
nominal axial strength, and Mn is the nominal bending moment about either the major x- or
minor y-axis. The analysis for the required axial strength Pu and flexural strength Mu should
include second-order (geometric nonlinear) effects.
The following design methods, including ELM, DM, and DMMI, are represented by
Equation 1 with terms defined per that specific method. In all cases, the controlling
combinations of axial force and bending moment are determined for each of these elastic
design methods by iteratively solving for the maximum value of Mu for a given value of
Pu that will satisfy Equation 1. For reference, Figure 1 shows the deflected shape of the
beam-column. Equilibrium on the deformed shape is given by:

Mu ðx; PÞ þ P  vðxÞ þ M ¼ 0 ð3Þ

where vðxÞ is the total lateral deflection as a function


  of span length location x, and equals the
sum of a geometric imperfection v0 ðxÞ ¼ δ0  sin πx L , and deflection vPM ðxÞ due to the applied
combination of P and M.

2.1 Effective Length Method (ELM)


In computing the nominal axial strength Pn from AISC’s column curve, the effective length
factor of a simply supported beam-column is K = 1. In determining the required flexural
strength Mu , equilibrium equations are defined on the deflected shape to account for second-
order effects. For a structural analysis associated with ELM, the beam-column is assumed
geometrically straight, v0 ðxÞ ¼ 0, prior to any applied loading (AISC’s column curve accounts
for member out-of-straightness). As a result, the P  δ effect in this method need only account
for the interaction between the applied axial load and bending moments, and thereby is not
influenced by the presence of an initial member imperfection.

56
Figure 1. Deflected shape of beam-column with second-order effects due to applied loading and geometric
imperfection.

In establishing the design adequacy of this member, the required moment Mu ðx; PÞ is
a maximum at mid-span because vPM ðxÞ takes on a maximum value when x ¼ L=2. Thus, the
interaction equation only needs to be checked at mid-span, where the required strengths (terms
in the numerators) are at a maximum. For an elastic analysis of a simply supported, originally
perfectly straight, and prismatic member, the deflection and required flexural strength Mu at
mid-span, which includes second-order effects, can be calculated as a function of the applied
force P and moment M by the following “exact” equation (McGuire et al., 2000):

1
jMu mid j ¼  pffiffiffiffiffiffiffiffiffiffiffi M ð4Þ
π
cos 2 P=Pe

where Pe ¼ Euler buckling strength of the beam-column.


With Pu ¼ P at mid-span, substitution of these terms for Pu and Mu ¼ jMu mid j in Equation 1
results in an interaction equation for ELM defined by:

 p1 ffiffiffiffiffiffiffiffi M
P 8 cos π2 P=Pe
þ   1:0 ð5Þ
fPn 9 fMn

In which specific to ELM, Pn ¼ Fcr Ag , Pe ¼ π2 EI=L2 , Mn ¼ Fy Z; where, Fcr is the critical


buckling stress as defined by AISC’s column curve with K = 1 for the simply supported end
conditions being investigated in this study, Ag is the gross area of the cross-section, E is the
elastic modulus of the material, I is the moment of inertia, L is the unsupported length of the
beam-column, Fy is the material yield stress, and Z is the plastic section modulus. In comput-
ing Mn , it is important to note that only members with compact sections are investigated, and
any members subject to major-axis bending are assumed fully braced out-of-plane.

2.2 Direct Analysis Method (DM)


Although DM permits the use of the unbraced length (K = 1), this provides no advantage over
ELM for the specific end support conditions of the single beam-column investigated in this
study. In fact, DM is somewhat penalized in this case by its required use of a stiffness reduc-
tion factor within the structural analysis. Although the equilibrium analysis is of the same
form as that given for ELM, the Euler buckling strength Pe used in the analysis of the
member is modified to represent the inelastic buckling strength of the member. As a result,

57
interaction equation Equation 1 for DM can be written as Equation 5, except Pe is defined as
the inelastic buckling strength. Hence, Pn ¼ Fcr Ag , with Fcr defined by AISC’s column curve
with no 0:8τb stiffness reduction on E, Pe ¼ π2 ð0:8τb E ÞI=L2 , and Mn ¼ Fy Z. According to the
AISC Specification

 and given that all sections are compact, τb is calculated as
 (2016a)
τb ¼ 4 P=Py 1  P=Py for P=Py 40:5, and τb ¼ 1:0 for P=Py  0:5 where Py ¼ Fy Ag .

2.3 Design by Advanced Elastic Analysis Method (DMMI)


As described earlier, DMMI is an alternative design method that may be particularly useful
for more complex structures in which the unbraced length is not discernable. By directly mod-
eling member out-of-straightness and representing potential inelasticity through the use of the
stiffness reduction strategy employed in DM, the nominal axial strength Pn of the member
may be taken as its cross-section strength. The artificial and dramatic increase in axial
strength Pn that appears in the interaction equation is compensated for by a larger required
flexural strength Mu , which is obtained from an advanced elastic structural analysis that
accounts for initial system and member imperfections, second-order (geometric nonlinear)
effects, and stiffness reduction due to inelasticity.
In contrast to the above analysis for determining strengths for ELM and DM, the analysis
for DMMI must also include the direct modeling of member out-of-straightness. In this study,
the shape of the initial imperfection is assumed a sine wave with an amplitude at mid-span of
δ0 ¼ L=1000 per the AISC’s Code of Standard Practice for Steel Buildings and Bridges (AISC,
2016b). As such, the second-order P  δ effect needs to include both the impact of the applied
axial force and bending moment as well as the initial imperfection.
The solution to the governing differential equation (Eq. 3) at mid-span is given by:
qffiffiffiffi qffiffiffiffi
  P π
  Pe π
P
L δ0 M 2sin Pe 2 sin
v ¼ þ   qffiffiffiffi ð6Þ
2 1  PPe P sin P
π
Pe

 
With v L2 , equilibrium on the deformed shape at mid-span will result in a required moment
strength of:
 
L
jMu mid j ¼ M þ P  v ð7Þ
2

Similar to DM, a stiffness reduction factor of 0.8τb should be applied to all the members of
the system, which in this study means that all EI terms (within Pe) in the above equations
should be 0.8τb EI. With values of Pu ¼ P and Mu ¼ jMu mid j as defined above, the interaction
equation (Eq. 1) is expressed for DMMI as:
 
P 8 M þ P  v L2
þ   1:0 ð8Þ
fPn 9 fMn
 
In which specific to DMMI, v L2 is given by Equation 6, with δ0 ¼ L=1000, and Pe and τb as
defined for DM, Pn ¼ Fy Ag , and Mn ¼ Fy Z.

2.4 Design by Advanced Inelastic Analysis Method (GMNIA)


Since 2010, the Design by Advanced Inelastic Analysis Method has been provided in Appendix 1
of the AISC Specification. Given that this design method is based on a geometric and materially
nonlinear analysis, it will be referenced by the acronym GMNIA. The second-order inelastic ana-
lysis routines used in this study are included in the finite element analysis software FE++ (Alem-
dar, 2001), in which a distributed plasticity model is employed. Each beam-column is modeled by

58
eight line elements, thereby permitting a sine wave member out-of-straightness of δ0 ¼ L=1000 to
be directly modeled in the analysis. Residual stresses are represented by pre-stressing (compression
or tension) the cross-section fibers that define the cross-section. The applied axial force P and
bending moments M are applied simultaneously, and an incremental-iterative arc length solution
scheme is employed until a limit point is achieved. Because of the relatively high accuracy of this
analysis, the below error analyses of the above elastic design methods is based on the combin-
ations of P and M that this inelastic design method would permit and still satisfy the provisions
of Appendix 1 of AISC’s Specification.
It is well known that partial yielding of the cross-section can have a significant effect on the
stability of beam-columns. In cases where member out-of-straightness is not removed by pro-
cesses, such as rotary straightening, this partial yielding can be accentuated by the presence of
residual stresses. On the other hand, the use of such straightening processes can be shown to
alleviate or even eliminate the presence of residual stresses (Ge and Yura, 2019). As result,
ultimate strength combinations were determined for cases in which residual stresses are and
are not included in the analysis. When residual stresses are taken into account, the Galambos
and Ketter (1959) residual stress distribution was employed with a maximum compressive
stress at the flange tips of Fy 0:3Fy . Additionally, the material elastic modulus E and yield
stress Fy Fy are reduced by a factor of 0.9, per the requirements of Appendix 1 of the AISC
Specification. An elastic-perfectly plastic material model is employed.

3 NORMALIZED P-M INTERACTION CURVES AND ERROR CALCULATIONS

To compare the accuracy of each of the design methods, with special attention on DMMI,
normalized P-M interaction curves of ELM, DM, DMMI, and GMNIA are first plotted.
Data points are obtained by determining the maximum combination of axial load P and bend-
ing moments M that can be applied at the member ends such that the strength requirements
of the design method would just be satisfied. Calculation of error values in the curves are then
computed using the GMNIA curve as a basis. To further allow the errors to be comparable
for the wide range of member slenderness ratios investigated, all axial forces and moments
were normalized by the maximum GMNIA values, with PGMNIA being the maximum axial
strength when the applied moment is M = 0, and with M GMNIA being the maximum moment
strength when the applied axial force is P = 0 (which would equal 0.9FyZ for all members in
this study). As an example, Figure 2 shows the normalized P-M interaction curves and a plot
of the radial errors for a W12X120 member with an L/r = 90 that is subjected to minor-axis
bending and with residual stresses included.

Using radial lines at 10 increments measured clockwise from the normalized P-axis to
the M-axis, the intersections of the radial lines and the P-M curves are determined. It is noted
that values at intersection points that lay between computed data points are obtained from
a parabolic interpolation between the adjacent three data points. The percent errors of the
design methods are then established by comparing their radial R-distances from the origin to
the interaction curves according to:

RXXX  RGMNIA
percent radial error ¼  100% ð9Þ
RGMNIA

where RXXX is the radial distance of the P-M curves for the elastic design methods (ELM,
DM, and DMMI), and RGMNIA is the radial distance to the GMNIA P-M curve. As a result,
error plots (Figure 2b) at different radial angles represent a comprehensive range of different
combinations of applied axial force and moment. Points with positive percent errors are indi-
cative of situations in which the elastic design method (ELM, DM, or DMMI) are unconser-
vative when compared to design strengths determined by GMNIA.
The legend within the rightward radial error graph (Figure 2b) contains information import-
ant to this study. Working from the top downward, rows within this legend represent results

59
Figure 2. For a W12X120 member with an L/r = 90 subject to minor-axis bending and with residual
stresses included, (a) normalized P-M interaction curves of the four design methods, and (b) plots
of percent radial errors.

Table 1. W-shapes studied.


W14 W14X730 W14X665 W14X605 W14X550 W14X500 W14X455 W14X426
W14X398 W14X370 W14X342 W14X311 W14X282 W14X257 W14X233
W14X211 W14X193 W14X176 W14X159 W14X145 W14X132 W14X120
W14X109 W14X82 W14X74 W14X68 W14X61 W14X53 W14X48
W12 W12X336 W12X305 W12X279 W12X252 W12X230 W12X210 W12X190
W12X170 W12X152 W12X136 W12X120 W12X106 W12X96 W12X87
W12X79 W12X72 W12X58 W12X53 W12X50 W12X45 W12X40
W10 W10X112 W10X100 W10X88 W10X77 W10X68 W10X60 W10X54
W10X49 W10X45 W10X39 W10X33
W8 W8X67 W8X58 W8X48 W8X40 W8X35

for the ELM, DM, and DMMI methods, respectively. The first two numbers in each row rep-
resent the error of each design method with an angle (θÞ that corresponds to where the DMMI
error is at its maximum. The second two numbers correspond to the maximum error of each
design method and the angle (θÞ where this maximum occurs. Points with positive percent
errors are indicative of situations in which the elastic design method (ELM, DM, or DMMI)
are unconservative when compared to design strengths determined by GMNIA.

4 CROSS SECTIONS INVESTIGATED

As indicated in Table 1, this study investigated 65 wide flange W-shapes of A992 steel (E = 200
GPa and Fy = 345 MPa). These shapes are all of the compact sections that appear in the column
design portion of the AISC Manual (AISC, 2016c), and their depth to width ratios are all less
than 1.5.

5 RESULTS

Interaction curves and plots of percent radial errors that correspond to the four different
design methods (ELM, DM, DMMI, and GMNIA) were prepared (see for example, Figure 2

60
and Appendices 1 and 2) for all 65 W-shapes
pffiffiffiffiffiffiffiffiffi over a range of member slenderness L/r ratios of
30, 60, 90, 120, and 150, with r ¼ I=A. With four cases, including minor- or major-axis
bending and with or without residual stresses, this study evaluates 1,300 conditions, which are
represented by a total of 57,200 analysis data points.
A summary of the results for all members is provided in Table 2, in which the maximum,
average, and median of all of the individual member maximum percent radial errors are
reported. In general, the percent radial errors reported for the three design methods are fairly
similar. The largest percent radial errors are always for the DMMI method, and the
smallest percent radial errors are for the DM method. Given that the ELM and DM methods
are essentially the same, except that DM requires the analysis to include the stiffness reduction
0:8τb , it is expected (and confirmed in Table 2) that DM will be more conservative (smaller
radial errors) than ELM for all slenderness ratios. It is further noted that larger unconservative
errors for DMMI for sections with residual stresses consistently occur when the applied loading
combination is predominately axial force (θ = 10°), where in contrast the larger unconservative
errors for ELM and DM occur when the loading is primarily bending (θ = 80°).

6 DISCUSSION

As would be expected, not including a residual stress distribution increases the design capaci-
ties of the beam-columns per the GMNIA design method. As a consequence, and given that
the GMNIA results form the basis for the error analysis, the unconservative percent radial
errors for all three of the elastic design methods (ELM, DM, and DMMI) are significantly
reduced. A representative example of this is shown in Figure 3, where the performance of the
DMMI design method is significantly improved with much better agreement, smaller radial
errors, with GMNIA.
This increase in accuracy, however, is relatively pronounced where θ is small, when
the axial load is more significant than the bending moment, and is less obvious when θ
is large, a combination of a larger bending moment and a smaller axial force. Of course,
this is expected because it is well known that such residual stresses rarely impact the
strength of a member primarily subjected to a loading combination that is predominately
bending (again noting that all members in this study are either subject to minor-axis
bending or laterally braced when subject to major-axis flexure). The trend observed in
Figure 3 is consistent for all shapes and design methods investigated in this study,
regardless of the slenderness ratio or the axis of bending investigated. In general, the
reduction in DMMI errors for sections without residual stresses is largest when the slen-
derness ratio is L/r = 60 for minor-axis bending and L/r = 90 for major-axis bending.
The change in error is the smallest at the extreme slenderness ratios investigated, includ-
ing the least-slender (L/r = 30) and most-slender (L/r = 150) members. It is further
noted that the ELM and DM design methods are significantly more conservative when
residual stresses are not present.
With the exception of more-stocky members (L/r = 30), the percent radial errors for all
three design methods, especially DMMI, are reduced when members are subject to major-axis
bending instead of minor-axis bending.
As further shown in Table 2, all three elastic design methods will produce unconservative
errors when compared with GMNIA-based design. For the reasons given earlier, DM will
always provide smaller percent radial errors when compared with ELM. It is important to
note that this applies only for the simply-supported member explored in this study – for sys-
tems comprised of members with effective length K-factors exceeding 1.0, this will not neces-
sarily be the case (Martinez-Garcia, 2006).
The results for DMMI and ELM are not significantly different, with the largest differences
occurring for members subject to minor-axis bending in the low- to mid-slenderness (L/r = 60
to 90). Knowing that ELM has been a well-established design method that has performed well
in the U.S. since the early 1960’s, it is the authors’ opinion that the unconservative errors

61
reported in Table 2 for all three elastic design methods may not be reason for significant
concern.

Table 2. Summary of percent radial errors for minor- and major-axis bending with and without residual
stresses included in the GMNIA design.
Minor-axis Minor-axis Major-axis Major-axis
with residual without residual with residual without residual
stresses stresses stresses stresses

L/r = 30 DMMI Max = 3.0% Max = 1.8% Max = 7.0% Max = 5.9%
Ave = 2.2% Ave = 0.5% Ave = 6.5% Ave = 5.0%
Median = 2.2% Median = 0.4% Median = 6.6% Median = 5.0%
ELM Max = 3.2% Max = 2.5% Max = 6.9% Max = 5.8%
Ave = 2.1% Ave = 1.1% Ave = 6.1% Ave = 4.7%
Median = 2.0% Median = 1.1% Median = 6.2% Median = 4.6%
DM Max = 2.6% Max = 1.9% Max = 6.0% Max = 4.9%
Ave = 1.5% Ave = 0.6% Ave = 5.1% Ave = 3.8%
Median = 1.5% Median = 0.5% Median = 5.2% Median = 3.6%
L/r = 60 DMMI Max = 14.8% Max = 7.3% Max = 10.5% Max = 7.5%
Ave = 13.7% Ave = 6.1% Ave = 10.0% Ave = 6.7%
Median = 13.9% Median = 6.1% Median = 10.0% Median = 6.6%
ELM Max = 9.7% Max = 8.8% Max = 9.2% Max = 6.1%
Ave = 8.4% Ave = 7.5% Ave = 8.5% Ave = 5.2%
Median = 8.4% Median = 7.6% Median = 8.6% Median = 5.3%
DM Max = 8.2% Max = 6.7% Max = 6.4% Max = 3.6%
Ave = 7.3% Ave = 5.5% Ave = 5.7% Ave = 2.9%
Median = 7.3% Median = 5.5% Median = 5.8% Median = 3.0%
L/r = 90 DMMI Max = 15.8% Max = 9.7% Max = 10.0% Max = 5.4%
Ave = 14.8% Ave = 8.2% Ave = 9.2% Ave = 4.7%
Median = 14.8% Median = 8.2% Median = 9.2% Median = 4.7%
ELM Max = 13.0% Max = 11.3% Max = 7.6% Max = 4.5%
Ave = 11.1% Ave = 9.8% Ave = 6.9% Ave = 3.5%
Median = 11.1% Median = 9.8% Median = 6.9% Median = 3.5%
DM Max = 11.2% Max = 8.2% Max = 3.9% Max = 1.5%
Ave = 9.6% Ave = 6.7% Ave = 3.2% Ave = 0.6%
Median = 9.6% Median = 6.7% Median = 3.3% Median = 0.6%
L/r = 120 DMMI Max = 15.3 % Max = 11.0% Max = 7.1% Max = 2.9%
Ave = 14.2% Ave = 9.6% Ave =6.2% Ave = 2.2%
Median = 14.1% Median = 9.6% Median = 6.2% Median = 2.2%
ELM Max = 12.7% Max = 11.4% Max = 5.8% Max = 2.7%
Ave = 11.3% Ave = 9.9% Ave = 4.6% Ave = 1.8%
Median = 11.3% Median = 9.9% Median = 4.6% Median = 1.8%
DM* Max = 9.5% Max = 8.1% Max = 2.6% Max = n/a
Ave = 8.0% Ave = 6.6% Ave = 1.5% Ave = n/a
Median = 8.0% Median = 6.6% Median = 1.4% Median = n/a
L/r = 150 DMMI Max = 14.0% Max = 11.8% Max = 5.6% Max = 2.1%
Ave = 12.6% Ave = 10.4% Ave = 4.8% Ave = 1.2%
Median = 12.6% Median = 10.4% Median = 4.7% Median = 1.2%
ELM Max = 12.4% Max = 11.2% Max = 4.4% Max = 1.4%
Ave = 10.9% Ave = 9.8% Ave = 3.4% Ave = 0.5%
Median = 10.9% Median = 9.8% Median = 3.4% Median = 0.5%
DM* Max = 9.0% Max = 7.8% Max = 1.1% Max = n/a
Ave = 7.6% Ave = 6.4% Ave = 0.2% Ave = n/a
Median = 7.6% Median = 6.4% Median = 0.1% Median = n/a

DM* no unconservative errors are observed as indicated by n/a

62
Figure 3. Percent radial errors for a member of an L/r = 60 subjected to minor-axis bending comprised
of a W12X120 section that (a) includes and (b) excludes residual stresses in the GMNIA-based design.

7 SUMMARY AND CONCLUSIONS

This study evaluates three elastic design methods (ELM, DM, and DMMI) appearing in the
2016 AISC Specification by making comparisons with a fourth method (GMNIA) that some
may consider the most “exact” and titled Design by Advanced Inelastic Analysis Method,
which also appears in this Specification. With 1,300 conditions studied that required a total of
57,200 analyses, simply-supported beam-columns comprised of a fairly wide range of column
W-sections and slenderness ratios are investigated for conditions of minor- or major-axis flex-
ure that include or exclude the presence of residual stresses.
In general, all three elastic design methods provide fairly similar results, with AISC’s rela-
tively new Design by Advanced Elastic Analysis Method consistently indicating more strength
(1% to 5%) than AISC’s Effective Length. Conditions of major-axis bending significantly
improved the performance of all three elastic design methods. Regardless of the axis of bend-
ing, results are always improved when residual stresses are not present, a condition that is
often the consequence of rotary straightening during the rolling process.

REFERENCES

AISC 2016a. ANSI/AISC 360-16 Specification for Structural Steel Buildings, American Institute of Steel
Construction, Chicago, IL.
AISC 2016b. Code of Standard Practice for Steel Buildings and Bridges, American Institute of Steel Con-
struction, Chicago, IL.
AISC 2016c. Steel Construction Manual, Fifteenth Edition, American Institute of Steel Construction,
Chicago, IL.
Alemdar, B.N. 2001. Distributed Plasticity Analysis of Steel Building Structural Systems. PhD disserta-
tion, Georgia Institute of Technology, Atlanta, GA.
Galambos, T.V. & Ketter, R.L. 1959. Columns Under Combined Bending and Thrust. Journal of the
Engineering Mechanics Division, ASCE, 85(EM2): 135–152.
Ge, X. and Yura, J. 2019. The Strength of Rotary-Straightened Steel Columns, Proceedings-Annual Sta-
bility Conference, SSRC, St. Louis, MO: 425–442.
Giesen-Loo, E. 2016. Design of Steel Structures by Advanced 2nd-Order Elastic Analysis –Background
Studies. Honors Thesis, Bucknell University, Lewisburg, PA.
Martinez-Garcia, J.M. & Ziemian, R.D. 2006. Benchmark Studies to Compare Frame Stability Provisions.
Proceedings-Annual Technical Session and Meeting, SSRC, San Antonio, TX: 425–442.

63
McGuire, W., Gallagher, R., & Ziemian, R. 2000. Matrix Structural Analysis, John Wiley & Sons, Inc.,
New York, NY.
Nwe Nwe, M.T. 2014. The Modified Direct Analysis Method: An Extension of the Direct Analysis
Method. Honors Thesis, Bucknell University, Lewisburg, PA.
Ziemian, R.D. (ed.) 2010. Guide to Stability Design Criteria for Metal Structures. John Wiley & Sons,
Inc., Hoboken, NJ.

Appendices
A. Plots of interaction curves and percent radial errors
As a complement to Figure 2, which is for L/r = 90, the remaining normalized P-M interaction
curves and corresponding plots of percent radial errors that were studied for the specific case of
a W12X120 member that includes residual stresses and subjected to minor-axis bending are
provided.

Figure A1. L/r = 30.

Figure A2. L/r = 60.

64
Figure A3. L/r = 120.

Figure A4. L/r = 150.

B. Data for plots of percent radial errors


The following tables provide numerical values for the data points appearing in the percent
radial error plots given in Figure 2 and Appendix A.

Table B1. L/r = 30 (values are percent radial errors).


Major-axis bending
Minor-axis bending Minor-axis bending Major-axis bending without residual
with residual stresses without residual stresses with residual stresses stresses

θ DMMI ELM DM DMMI ELM DM DMMI ELM DM DMMI ELM DM



0 2.1 0.6 0.6 -6.0 -7.4 -7.4 2.1 -0.2 -0.2 -0.8 -3.1 -3.1

10 -2.2 -2.8 -4.1 -6.7 -7.2 -8.4 3.8 2.6 1.3 -1.1 -2.2 -3.5

20 -5.3 -5.9 -7.0 -6.9 -7.6 -8.6 5.1 3.8 2.7 1.0 -0.2 -1.3

30 -8.2 -8.8 -9.7 -9.6 -10.2 -11.1 5.8 4.6 3.6 3.3 2.2 1.2

40 -10.4 -10.8 -11.7 -12.1 -12.5 -13.4 6.2 5.4 4.3 3.3 2.4 1.4

50 -10.6 -10.8 -11.7 -12.9 -13.2 -14.1 6.5 5.9 4.8 4.6 4.0 3.0

60 -8.5 -8.5 -9.4 -11.1 -11.2 -12.1 6.3 6.0 5.0 5.0 4.6 3.6

70 -3.5 -3.4 -4.2 -6.0 -5.9 -6.7 5.9 5.8 5.0 4.5 4.4 3.5

80 1.1 1.8 1.2 0.4 1.1 0.4 2.9 3.4 2.8 2.2 2.8 2.2

90 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

65
Table B2. L/r = 60 (values are percent radial errors).
Minor-axis bending Major-axis bending
Minor-axis bending without residual Major-axis bending without residual
with residual stresses stresses with residual stresses stresses

θ DMMI ELM DM DMMI ELM DM DMMI ELM DM DMMI ELM DM



0 12.3 7.1 7.1 -4.7 -9.1 -9.1 3.8 -2.6 -2.6 -6.5 -12.3 -12.3

10 13.7 8.0 4.6 1.0 -4.2 -7.0 6.8 0.3 -2.7 -2.0 -8.0 -10.7

20 10.9 6.3 2.6 1.1 -3.4 -6.7 8.1 2.5 -1.0 1.0 -4.4 -7.7

30 8.0 4.5 0.8 -0.2 -3.8 -7.2 8.5 4.0 0.3 3.2 -1.2 -4.8

40 6.6 3.8 0.2 -0.7 -3.6 -7.0 9.0 5.3 1.6 4.5 0.8 -2.8

50 6.1 3.8 0.4 -0.4 -2.8 -6.0 9.8 6.8 3.3 5.3 2.2 -1.3

60 6.5 4.7 1.7 1.2 -0.7 -3.7 10.5 8.1 4.9 6.2 3.8 0.5

70 7.4 7.1 4.5 4.2 3.2 0.5 9.5 8.5 5.7 7.0 5.5 2.7

80 6.9 8.3 6.4 6.3 7.7 5.7 5.1 6.3 4.3 3.8 5.0 3.0

90 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

Table B3. L/r = 90 (values are percent radial errors).


Major-axis bending
Minor-axis bending Minor-axis bending Major-axis bending without residual
with residual stresses without residual stresses with residual stresses stresses

θ DMMI ELM DM DMMI ELM DM DMMI ELM DM DMMI ELM DM



0 9.7 10.5 10.5 -2.7 -2.0 -2.0 0.4 -1.0 -1.0 -6.7 -8.0 -8.0

10 14.7 9.1 2.2 2.2 -2.8 -8.8 4.1 -2.1 -8.1 -0.8 -6.7 -12.4

20 14.9 8.9 2.1 3.3 -2.1 -8.2 5.4 -0.9 -7.1 0.5 -5.6 -11.5

30 13.9 8.2 1.8 3.7 -1.7 -7.4 6.6 0.5 -5.4 2.6 -3.4 -9.2

40 13.2 7.8 1.9 4.2 -0.9 -6.5 8.0 2.1 -3.6 3.6 -2.1 -7.7

50 12.8 8.0 2.5 5.1 0.4 -4.8 9.2 3.9 -1.5 4.0 -1.2 -6.4

60 12.1 9.0 4.2 6.5 2.9 -1.8 9.4 5.4 0.5 4.6 0.4 -4.3

70 11.0 11.0 6.9 7.6 6.7 2.5 8.0 6.9 2.7 4.4 2.8 -1.3

80 9.5 10.8 8.0 8.2 9.7 6.7 4.8 6.1 3.2 1.9 3.2 0.2

90 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

Table B4. L/r = 120 (values are percent radial errors).


Major-axis bending
Minor-axis bending Minor-axis bending Major-axis bending without residual
with residual stresses without residual stresses with residual stresses stresses

θ DMMI ELM DM DMMI ELM DM DMMI ELM DM DMMI ELM DM



0 -0.3 5.5 5.5 -5.6 -0.1 -0.1 -5.7 -1.4 -1.4 -7.1 -3.0 -3.0

10 11.1 3.0 -4.9 0.9 -6.3 -13.4 1.0 -7.0 -14.0 -1.3 -9.1 -15.9

20 13.8 3.9 -3.5 3.2 -5.7 -12.5 3.6 -6.0 -12.8 0.3 -8.9 -15.6

30 14.2 4.2 -2.8 4.9 -4.5 -10.8 5.4 -4.6 -10.9 1.0 -8.6 -14.7

40 13.7 4.8 -1.6 5.8 -2.7 -8.7 6.0 -3.0 -9.0 1.4 -7.3 -13.1

50 13.1 6.0 0.0 6.7 -0.2 -5.9 6.3 -1.1 -6.8 2.1 -5.0 -10.5

60 12.6 7.8 2.5 7.7 2.8 -2.4 6.2 0.9 -4.2 2.0 -3.2 -8.1

70 12.0 10.4 5.9 8.7 6.8 2.3 5.6 3.4 -1.0 1.6 -0.6 -4.9

80 10.8 11.1 7.9 9.4 9.8 6.5 4.2 4.4 1.2 1.4 1.6 -1.5

90 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

66
Table B5. L/r = 150 (values are percent radial errors).
Major-axis bending
Minor-axis bending Minor-axis bending Major-axis bending without residual
with residual stresses without residual stresses with residual stresses stresses

θ DMMI ELM DM DMMI ELM DM DMMI ELM DM DMMI ELM DM



0 -3.0 0.5 0.5 -6.5 -3.2 -3.2 -7.4 -4.9 -4.9 -9.5 -7.2 -7.2

10 8.6 -2.4 -9.8 -0.2 -10.2 -17.0 -0.8 -11.3 -18.0 -4.8 -14.9 -21.2

20 11.6 -0.8 -8.0 2.4 -9.0 -15.6 1.6 -10.3 -16.8 -2.0 -13.4 -19.7

30 12.5 0.3 -6.4 4.2 -7.2 -13.4 2.8 -8.9 -15.0 -0.1 -11.5 -17.5

40 12.7 1.6 -4.7 5.7 -4.8 -10.8 3.6 -7.1 -13.0 0.6 -9.9 -15.7

50 12.6 3.3 -2.6 7.1 -2.0 -7.7 4.3 -4.9 -10.4 0.0 -9.0 -14.3

60 12.6 5.7 0.4 8.5 1.5 -3.6 4.7 -2.4 -7.4 0.4 -6.5 -11.4

70 12.5 9.0 4.4 9.7 6.0 1.5 4.7 0.8 -3.5 0.9 -2.9 -7.2

80 11.5 10.9 7.6 10.2 9.6 6.2 3.9 3.3 0.1 1.2 0.5 -2.7

90 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

67
Regular papers
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

On the modal buckling of longitudinally stiffened plates

S. Adany & M.Z. Haffar


Budapest University of Technology and Economics, Budapest, Hungary

ABSTRACT: In this paper the elastic buckling of simple longitudinally stiffened plates
are discussed. The special focus is on the buckling of the stiffeners that are assumed to
have trapezoidal shapes. Stiffener buckling is frequently handled as a combination of
plate buckling and column buckling in the design of orthotropic stiffened plates. On the
other hand, stiffener buckling is handled as distortional buckling in the design of cold-
formed steel members. Here the buckling of stiffened plates is discussed by using the
global, distortional and local deformation modes, by using the recently introduced con-
strained finite element method.

1 INTRODUCTION

It is a usual engineering practice to apply stiffeners in thin plates to increase the resistance
against plate buckling. In cold-formed steel structural members almost exclusively longitu-
dinal stiffeners are applied. Since the stiffened plates are typically slender, buckling is always
crucially important. It is also important to properly distinguish between the buckling of the
plate panels between the stiffeners, and the buckling of the stiffeners, since these two kinds of
buckling might have significantly different post-buckling reserve.
When longitudinal stiffeners in cold-formed steel members are discussed, the buckling of
longitudinal stiffener, i.e., when flexural deformations of the stiffener plus plate-like deform-
ations of the plates are combined, is typically called distortional buckling. In design codes for
cold-formed steel members, such as the North-American direct strength method, DSM (2006),
or the European design method, CEN (2006a), distortional buckling is clearly distinguished
from local-plate buckling and from global buckling, and though the design methods are differ-
ent in the various design codes, some reduction factor is calculated from the elastic critical
load to each of the buckling types.
In the case of welded plate girders longitudinal stiffeners are widely employed, too, but
their buckling is unusual to describe as distortional buckling. A possible design approach,
which is included in the relevant part of the Eurocode, see CEN (2006b), is to interpret the
stiffener buckling as a combination of so-called plate-like behavior and column-like behavior,
and therefore to calculate reduction factor by interpolating from reduction factors defined for
plate-like and column-like behavior.
In the actual paper the buckling of simple, longitudinally stiffened plates is studied, how-
ever, by adopting the worldview of cold-formed steel design, by interpreting the deform-
ations as the combination of global (G), distortional (D), local (L) and other (O)
deformation modes. Linear elastic buckling behavior is discussed, considering various plate
and stiffener geometries. The calculations are conducted by the constraint finite element
method (cFEM), as in Ádány (2018a, 2018b) and Ádány et al. (2018), which can readily be
applied for stiffened plates and can easily and objectively separate the global, distortional
and local deformations.

71
2 THE CONSTRAINED FINITE ELEMENT METHOD

The constrained finite element method (cFEM) is essentially a shell finite element calculation,
but the method is developed so that modal decomposition would be possible .[In order to
maintain constraining ability, the longitudinal shape functions of the employed finite elements
are specially selected, but the shell elements can be used as any regular flat shell element. Sep-
aration of the behavior modes is realized by applying mechanical constraints. When a member
is constrained into a deformation mode, it is enforced to deform in accordance with some
mechanical criteria, characterizing for the intended deformation mode. The subsequent appli-
cation of the properly selected mechanical criteria finally leads to an alternative basis system
of the displacement field defined by the finite element nodal degrees of freedom: the practic-
ally useful feature of this alternative basis system is that the deformation modes are separated.
An important feature of the finite element analysis by cFEM is that transverse and longitu-
dinal directions are strictly distinguished. When deformations are constrained or decomposed,
these are the cross-section deformations (i.e., transverse deformations and/or displacements)
that are manipulated, practically independently of the employed longitudinal shape functions.
Therefore the cFEM modes can also be interpreted, and will be referred here, as cross-section
modes, similarly as it is done in the generalized beam theory.
The modal decomposition might beneficially be used in two tasks. One is the calculation in
a reduced (i.e., constrained) deformation space. For example, the linear buckling analysis can
be done for the G deformations only, meaning that all the buckled shapes will satisfy the
mechanical criteria of G deformations; in simple words, all the buckled shapes will be global
(e.g., flexural, or lateral-torsional buckling). The calculation in a constrained deformation
space is realized by selecting only those modal basis vectors that belong to the desired deform-
ation mode space.
The other task of cFEM is modal identification. The objective is to determine the participa-
tion of the characteristic deformations modes (G, D, L, etc.) in a deformed shape of the
member. The deformed shape might come from a linear static analysis, or linear buckling ana-
lysis, etc. A special basis system is necessary in which the various deformation modes are sep-
arated. Such basis system is naturally provided by the cFEM basis functions. Mathematically
this means that any deformation, defined by the displacement vector, can be expressed as the
linear combination of the basis vectors. Once the combination factors are calculated, the mag-
nitudes of the combination factors give the relative importance of the basis vectors in the dis-
placement vector to be identified. Since the basis vectors are separated into G, D, L, etc. mode
spaces, the relative importance of the G, D, L, etc. deformations can be calculated. For more
information, see Ádány (2018a).

3 DEFORMATION MODES OF PLATES WITH TRAPEZOIDAL STIFFENERS

In Figure 1 some characteristic deformation modes are shown for plates with three longitu-
dinal trapezoidal stiffeners. The deformation modes can be interpreted as the modes of
a beam or column member, where the direction of the stiffener identifies the longitudinal axis
of the “beam”. As always, the deformation modes are independent of the supports, defined
solely by the member (i.e., stiffened plate) geometry. In Figure 1 the deformation modes are
shown by axonometric figures, by using a single half-sine-wave for the longitudinal distribu-
tion, but it is to emphasize that the longitudinal displacement distribution has no real effect
on the displacement modes, so the half-sine-waves are used here just for illustration.
The nature of the deformation modes is primarily determined by the number of stiffeners,
though the exact shapes, or the order of the deformation modes are dependent on the geometric
proportions (and in some cases on the discretization, too). Thus, the deformation modes pre-
sented here can be regarded as qualitatively representative for the given number of stiffeners.
Global modes are characterized by rigid-body cross-section displacements. Usually 4 global
cross-section modes are distinguished: axial mode, two bending modes in two perpendicular
direction (e.g., principal axes, marked as GB1 and GB2), and torsional mode that involves

72
Figure 1. Deformation mode samples for a plate with 3 stiffeners.

twisting rotation of the cross-section. In the case of plates with trapezoidal stiffeners the
“cross-section” of the member has one or multiple closed cells (i.e., the stiffeners themselves),
hence global torsion mode does not exist, since global modes are defined to be free of in-plane
shear, and rigid-body torsion is not possible for closed cells without in-plane shear
deformations.
The number of D deformation modes is strictly determined by the number of stiffeners. If
this latter one is denoted by nst, the number of distortional modes is 2 × nst-1. It is found that,
from practical aspect, there are two major types of D modes. There are modes when the stiff-
eners are not (or hardly) distorted, but are displacing perpendicularly to the plate. Other
D modes involve (mostly) the distortion of the stiffeners. When there is one single stiffener
only, the D mode involves stiffener distortion. In the case of two stiffeners, the first D mode
shows stiffener translation, the other two D modes show stiffener distortion. In the case of 3

73
stiffeners (Figure 1), the first two D modes show stiffener translation, while the remaining
three D modes show stiffener distortion. For higher number of stiffeners roughly half of the
D modes involve stiffener translation, the other half stiffener distortion. These are always the
modes with stiffener translation that come first, meaning that the modes with stiffener distor-
tion are associated with higher energy content, hence these seem to have smaller importance
in practical problems.
The number of primary L deformation modes (LP) is also defined by the number of stiff-
eners: it is 4 × nst + 4. The number of secondary L modes (LS) is dependent on the discret-
ization. Figure 1 shows the first 12 LP modes for the plate with three stiffeners.

4 BUCKLING RESULTS

In this Section buckling results are presented for one single plate, but with various numbers of
stiffeners. In all these examples the plate is square with side length of 3000 mm. The plate
thickness is 12 mm. The stiffener size is constant, with 75 mm depth and 6 mm thickness,
while the widths are 60 and 90 mm. The material is steel, with Young’s modulus of 210000
MPa and Poisson’s ratio of 0.3. In all the examples the plate is supported along all its 4 edges
by simple supports. These are realized by supporting the edges against translation perpendicu-
larly to the plate only, while the edges of the stiffeners are left free. This can be considered as
the possible least restrictive pinned support. First linear buckling analysis is performed, with-
out any constraints, from which the first few critical stress values and corresponding buckling
shapes are determined. The buckled shapes then are identified, by calculating the participa-
tions from the various deformations modes.

4.1 Plates in compression


In Figure 2 selected results are shown when the plate is in uniaxial (i.e., longitudinal) compres-
sion. The load is transmitted to the structure as uniformly distributed over the end sections
(including the stiffeners, too). As the results show, for the actual plate parameters the first
mode always involve the buckling of the stiffener, which – from modal decomposition
aspect – is always a combination of L and G/D deformations. The importance of G modes is
remarkable, since global buckling alone is not possible due to the edge supports. The partici-
pations of G, D and L modes in the first buckling modes are shown in Figure 3 in the function
of the stiffener number. The plot suggests that the importance of D modes is increasing, while
the importance of G and L modes are slightly decreasing as the number of stiffeners is
increasing.
Some higher buckling modes are also shown in Figure 2. Among these higher modes there
are modes that are dominantly L, e.g., mode #2 in the case of 1 stiffener, or mode #3 in the
case of 3 stiffeners. Not surprisingly, these quasi-pure local-plate modes appear higher and
higher as the distance between two neighboring stiffeners is getting smaller.

4.2 Plates in bending


In Figure 4 selected results are shown when the plate is in bending, i.e., linearly varying dis-
tributed loads are applied over the end sections (including the stiffeners, too), varying along
the plate width. In all the cases the compressed part of the plate buckles. As the results show,
for the actual plate parameters, the first mode might be dominantly local-plate buckling when
the number of stiffeners is small), or stiffener buckling (when the number of stiffeners is large
enough). The participations of G, D and L modes in the first buckling modes are shown in
Figure 5 in the function of the stiffener number.
The identification results suggest that even if the buckling visually seems to be local-
plate buckling, there is non-negligible participation from the G modes, mostly due to the
rigid-body-like twisting of certain segments of the member, which displacements are

74
Figure 2. Buckling results samples: plate subjected to compression.

assigned to the G modes. Independently of the number of the stiffeners, there always exists
a buckling mode with “classic” stiffener buckling, such as mode #4 for plates with 1 or 2
stiffeners, or mode #1 for plates with 3 or more stiffeners. In these cases the buckled shape
is identified as the combination of L and G/D modes. As the number of stiffeners is
increasing, the importance of D modes is increasing while that of G and L modes are
slightly decreasing. However, the D modes have somewhat higher participations in gen-
eral, compared to the compressed plate.

75
Figure 3. G, D and L participations in the first buckling modes: plates subjected to compression.

4.3 Plates in shear


Plates in shear has also been studied. The pure shear is realized by applying constant distributed
load along all the 4 plate edges, parallel with the relevant edges. (The stiffeners are not loaded.)
Selected results are shown in Figure 6. The participations of G, D and L modes in the first
buckling modes are shown in Figure 7 in the function of the stiffener number. The plot shows
tendencies similar to those observed in plates in compression and bending, however, the import-
ance of D modes now is considerably higher, while the importance of G modes is smaller.
As the results show, even in the case of pure shear, it is possible to have buckling modes with-
out significant stiffener deformations: these modes are characterized by high L participations
(though with the here-assumed geometric parameters pure L buckling is atypical, best approxi-
mated by mode #3 of the plate with 1 stiffener). For the actual examples the majority of shear
buckling modes involves significant stiffener buckling, as a combination of L and G/D modes.

5 SUMMARY

In this paper some examples for the modal buckling analysis of longitudinally stiffened plates
are presented, by considering trapezoidal, i.e., hollow stiffeners. Examples are shown and dis-
cussed: unconstrained buckling problems are solved, the buckling shapes are determined, then
the shapes are identified. The results prove that cFEM can successfully solve the modal
decomposition for longitudinally stiffened plates: the modes can be determined by following
the usual cFEM procedure, then the modes can be used as in any modal decomposition
method. The here presented results show that modal decomposition leads to meaningful
results even for plate-like members.
Further aspects have also been studied, which are not reported here in detail. For example,
plates with various widths have been considered. From modal identification aspect it is found
that the wider the plate, the more importance the D modes might have. This is partially due to
the fact that wider plates might include larger number of stiffeners, which increases the
number of D modes, and finally the D participations might be higher.
Other geometric parameters have also some influence, such as the thickness of the plate, or
the size of the stiffener. The tendencies are mostly as-expected, e.g., thinner plate are more
susceptible to local-plate buckling, or larger-size stiffeners increase the potential of having
nearly-pure local-plate buckling.
Moreover, cFEM can easily solve the buckling problem in a reduced deformation space,
which can lead to pure global, pure distortional, or pure local buckling. Though for the con-
sidered edge supports pure global buckling does not exist, and pure distortional buckling is
usually associated with fairly large critical stress, pure L, as well as G+D or G+D+L con-
straints lead to meaningful critical stresses and deformed shapes.
It is also to mention that the support conditions have non-negligible effect. In the above-
presented examples the simple (i.e., pinned) restraints were realized in the possible least restrictive
form. However, pinned supports could be slightly more restrictive, too, e.g., by supporting the

76
Figure 4. Buckling results samples: plate subjected to bending.

edges of the stiffeners, too, or by supporting the plate edges against certain rotations or transla-
tion in the plane of the plate. Obviously, more restrictive pinned supports lead to higher critical
stresses, but they influence the modal identification results, too. Though the effect of various sup-
ports were not shown in this paper, we add here just one comment. It was observed that the here-
applied pinned supports allow some in-plane movement of the plate edges, which is reflected in
a few percentage of participation from other (i.e., mostly in-plane shear) modes; this contribution
of other modes is reduced by the more restrictive pinned supports.

77
Figure 5. G, D and L participations in the first buckling modes: plate in bending,

Figure 6. Buckling results samples: plate subjected to shear.

78
Figure 7. G, D and L participations in the first buckling modes: plate in shear,

It is believed that the introduced modal worldview can contribute in the understanding of
the buckling behavior of stiffened plates. Moreover, modal decomposition technique can
potentially lead to computerized design of stiffened plates, by developing an appropriate
design method similar to the direct strength method.

REFERENCES

Ádány, S. 2018a. Modal identification of thin-walled members with and without holes by using CFEM,
Proceedings of the ICTWS 2018 conference, July 24–27, 2018. Lisbon, Portugal.
Ádány, S. 2018b. Constrained shell Finite Element Method for thin-walled members, Part1: constraints
for a single band of finite elements, Thin-Walled Structures, Vol 128, July 2018, pp. 43–55.
Ádány, S., Visy, D. & Nagy, R. 2018. Constrained shell Finite Element Method, Part 2: application to
linear buckling analysis of thin-walled members, Thin-Walled Structures, Vol 128, July 2018, pp. 56–70.
CEN. 2006a. EN 1993-1-3:2006 - Eurocode 3: Design of steel structures - Part 1-3: Supplementary rules
for cold-formed members and sheeting, Brussels, Belgium.
CEN. 2006b. EN 1993-1-5:2006 - Eurocode 3: Design of steel structures - Part 1-5: Part 1-5: Plated Struc-
tural elements, Brussels, Belgium.
DSM. 2006. Direct strength method design guide. American Iron and Steel Institute. Washington DC, USA.

79
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Strength characterisation of a CFS section with initial geometric


imperfections

Hashmi S.S. Ahmed


Department of Civil Engineering, Marathwada Institute of Technology, Aurangabad, India

Siddhartha Ghosh
Structural Safety, Risk & Reliability Lab, Department of Civil Engineering, Indian Institute of Technology
Bombay, Mumbai, India

ABSTRACT: The strength of a cold-formed steel CFS member gets highly affected by the pres-
ence of any geometric imperfection (local or global). Moreover, these imperfections are the major
source of uncertainty in the prediction of buckling strength (Pn) of a CFS member. Previous
works on strength characterization of CFS members have shown the effects of considering the
uncertainty in local imperfection for a member. However, the effect of uncertainties in global
imperfections is still not addressed. The prediction of Pn in current design codes is based on some
deterministic value of the global imperfection. Considering this, the current paper aims at the stat-
istical characterisation of Pn taking into account uncertainties in both local and global imperfec-
tions. For this reason, a CFS member undergoing both local (Type 1) and global (flexural-
torsional) buckling is selected. However, the probabilistic assessment of Pn, in conjunction with
finite element (FE) simulations becomes computationally expensive. Thus, to address this high
computational cost, a metamodelling technique approximating the original simulation model with
a simplified mathematical model is proposed. Characteristic strength (Pd) values are recom-
mended for a selected range of non-dimensional slenderness and these values are then compared
with current design recommendations for Pn. It is observed that at the higher range of non-
dimensional slenderness (λc ≥ 1), where the flexural-torsional buckling mode governs the perform-
ance of CFS members, the AISI prediction becomes conservative, while at the lower range (λc < 1),
the AISI predictions are found to be close to uncertainty analysis based recommendations.

1 INTRODUCTION

Inherent slenderness in the cross-section of cold-formed steel (CFS) members makes it suscep-
tible to geometric imperfections. These geometric imperfections affect the strength of a (CFS)
members significantly. The geometric imperfections (GI) can take any possible shape in
a member. However, researchers have idealised them into two categories, namely i) sectional
and ii) global imperfection. The sectional imperfection is again classified by Schafer & Peköz
(1998) into two subcategories:
1. Type 1: Maximum local imperfection in a stiffened element, such as a web (and can be rep-
resented by Figure 1 (a))
2. Type 2: Maximum deviation from straightness for a lip-stiffened or an unstiffened flange
(and can be represented by Figure 1 (b))
Moreover, the global imperfections are further classified into three different groups by Zei-
noddini & Schafer (2011):
1. Bow: Sweep about the minor axis of a member
2. Camber: Sweep about the major axis of a member
3. Twist: Sweep about the longitudinal axis of a member

80
Figure 1. (a) Type 1 imperfection (b) Type 2 imperfection [Schafer & Peköz, 1998].

Figure 2. Deformed shapes associated with different global buckling modes in a CFS channel section,
(a) bow, (b) camber and (c) twist mode.

Figure 2 shows the typical shape for bow, camber and twist in a CFS member.
Importantly, these geometric imperfections (GI) are very random in nature resulting in large
uncertainty in the prediction of buckling strength (Pn ). Thus, a probabilistic estimation of Pn
becomes essential, as recommended by Schafer & Peköz (1998). In this line, Ahmed et al. (2017)
presented a work on the statistical assessment of buckling strength of a CFS section with local
Type 1 imperfection. In their work, they have shown the effects of considering the uncertainty
into the local imperfection for a member undergoing local buckling. For this purpose, they
have used statistical data on local geometric imperfection provided by Schafer & Peköz (1998).
Similar to local GI, global GI also carries uncertainty in their magnitude and shape. How-
ever, the uncertainty imparted due to global imperfection is still not addressed. In fact, the
buckling strength predictions in current design codes such as the AISI code are based on some
deterministic value of the global imperfection. Many of the researchers including Zeinoddini
& Schafer (2011), reported this deterministic value as L=960, where L represents the length of
a member. However, a single GI amplitude does not give the realistic picture of a CFS
member. Thus, considering the uncertainty involved in both local and global imperfections
becomes very imperative in complete characterisation of Pn .
The primary goal of the work presented here is a statistical characterisation of Pn consider-
ing uncertainties in both local and global imperfections. For this reason, a CFS member
undergoing local (Type 1) and global (flexural-torsional) buckling is selected, as this kind of
behaviour is typically exhibited by most commonly used CFS sections. However, the probabil-
istic assessment of Pn entails multiple simulations of a high fidelity FE model, making the pro-
cess computationally expensive. To address this high computational cost, metamodelling
technique in the form polynomial chaos expansion is employed in this work.

2 MODELLING OF GEOMETRIC IMPERFECTIONS

While assessing the effects of GI on the performance of a CFS member, modelling of the GI
becomes very important issue (Dinis & Camotim 2011). To this end, the conventional

81
approach of using critical buckling mode is found to be the simplest but effective method
indeed (Schafer et al. 1998). Thus, in this work critical (first) governing buckling mode for
that member, is used to perturb the perfect geometry. However, for the execution of probabil-
istic investigation, the statistical characterisation of GI is also essential. To this end,
a probabilistic model of local GI was available, proposed by Ahmed et al. (2017). This prob-
abilistic model on local GI can be expressed as,

d1 =t LN ð0:50; 0:66Þ ð1Þ

where d1 = magnitude of local GI; t = thickness of a CFS section.


Using Equation 1 the magnitude of local GI are obtained. The statistics of global GI (G1 ,
G2 and G3 ) in the form of cumulative distribution function (CDF) are adopted from the work
of Zeinoddini & Schafer (2012), as presented in Table 1.
Based on the Kolmogorov-Smirnov goodness-of-fit test (K-S test), the lognormal model for
G3 is found to be a better fit for the given statistics. The comparison between the two prob-
ability models with respect to the original data is provided through their empirical cumulative
distribution functions (CDF) in Figure 3. Equation 2 shows the proposed probabilistic model
for global GI.

Table 1. Statistics of global geometric imperfection.


CDF Bow, G1 Camber, G2 Twist, G3
(L=δ0 ) (L=δ0 ) (Deg/m)

0.25 4755 6295 0.20


0.50 2909 4010 0.30
0.75 1659 2887 0.49
0.75 845 1472 0.85
0.75 753 1215 0.95
Mean 2242 3477 0.36
Std. dev. 3054 5643 0.23

Figure 3. CDF of the observed data and trial distributions for normalised imperfection.

82
θ ðdeg=mÞ LN ð0:36; 0:23Þ ð2Þ

The scale factors (SF) are generated to scale the global ( flexural-torsional) mode using Equa-
tion 2. Using these factors, the desired twist imperfection is applied (about the shear centre) in
the section along with the associated flexural component (camber) of the imperfection. In fact,
it is observed that this camber part has an insignificant effect on the performance of the buck-
ling strength of the selected section.

3 FINITE ELEMENT SIMULATION

3.1 Selection of the specimen


A lipped channel section is selected for this work, which is designated as 400S200-68. This sec-
tion has the following dimensions: depth of the web = 101.6 mm, width of a flange =
50.80 mm, thickness = 1.72 mm. The selection of section is done based on a preliminary buck-
ling analysis using the software CUFSM (Schafer & Adany 2006), such that the global buck-
ling mode ( flexural-torsional) governs the member behaviour at higher non-dimensional
slenderness (i.e. λc  1), and the local buckling mode (Type 1) governs at lower non-
dimensional slenderness (i.e. λc 5 1). For this reason, five different lengths of the member are
considered (1.42 m, 2.13 m, 2.82 m, 4.12 m and 5.38 m), which give non-dimensional slender-
ness (λc ) in the range of 0.65 to 2.20.

3.2 Finite element analysis of the imperfect member


The general purpose finite element package Abaqus is used to find the critical buckling
strength (Pn ) of the selected lipped CFS channel member. “Fixed-fixed” end boundary con-
dition is applied by restraining all the degree of freedoms at both ends, except for the trans-
national degree of freedom in the axial direction at the ‘forcing end’. These boundary
conditions as well as the axial compression force are applied at the centroid of the end cross-
sections. These are transferred to the member by connecting (all degrees of freedom of)
the centroid with the nodes at the end section through “kinematic couplings”, as shown in
Figure 4. The S4R element is used to model the member, with an elastic-perfectly plastic
material model: elastic modulus ¼ 203 GPa; Poissons ratio = 0.3; and yield stress =
388 MPa. In order to understand the post-(local) buckling behaviour of the CFS section the
Static, Riks iterative scheme in Abaqus is adopted.
In order to have an imperfect member, the initial geometry of the section is perturbed
with the governing buckling mode scaled to the desired amplitude of the GI. Figure 5
shows the members at failure at two different λc values (0.65 and 1.25). It is observed
that the failure is predominantly due to the local buckling of the member at lower λc ,
while for member with higher λc the failure is totally governed by global buckling

Figure 4. FE model of the CFS member with “kinematic couplings”.

83
Figure 5. Deformations after buckling of section 400S200-68.

(flexural-torsional). The incremental axial compressive load and the axial deformation of
the member are monitored throughout the incremental analysis. Figure 6 shows sample
load-deformation curves for λc = 0.65 and λc = 1.25. The load-deformation curves pre-
sented here are for the CFS sections with and without geometric imperfection. For these
cases, a magnitude of d1 =t = 3.676 corresponding to the local buckling mode and magni-
tude of θ = 1.351 =m corresponding to the flexural-torsional mode was used as the ini-
tial geometric imperfection to obtain the buckling load. A significant weakening effect of
geometric imperfections is observed in both cases.

Figure 6. Load deformation curves for the section 400S200-68 with d1 =t ¼ 3:676 and θ ¼ 1:351 =m:

84
4 STRENGTH CHARACTERISATION USING METAMODELLING APPROACH

4.1 Formulation of metamodel


The probabilistic assessment of Pn by means of Monte Carlo simulation (MCS), in conjunc-
tion with finite element (FE) simulations becomes computationally expensive, since it involves
a large number of FE model runs (Schenk & Schuëller 2003, Hashmi et al. 2014). To overcome
this issue, the metamodelling technique of polynomial chaos expansion (PCE) is used in this
work. The PCE based metamodel describes the uncertainty in a structural response parameter,
using a set of orthogonal polynomial bases and associated coefficients (Spanos & Ghanem
1989). The general form of PCE can be represented as (Blatman & Sudret 2011)
X
Y ¼ MðξÞ ¼ aα α ðξÞ ð3Þ
α

where, α are the multivariate orthonormal polynomials, and α are the multi-indices that map
the multivariate α to their corresponding bases, which are denoted as the PCE coefficients
aα . These associated coefficients in the PCE based metamodel can be interpreted in the form
of mean, variance, and higher moments of the random output parameter of interest.
Individual PCE based metamodels are created for each λc of the selected section. The metamo-
del contains polynomial based equations defining the response Pn as a function of the geometric
imperfection magnitudes, which is treated as stochastic input variables. For this purpose, the PCE
module of the MATLAB toolbox UQLab (Marelli & Sudret 2014) is been used, which facilitates
the application of state-of-the-art algorithms for a non-intrusive approach of PCE metamodelling.
As the imperfection magnitudes (both local and global) follow lognormal distribution, the Her-
mite polynomial basis is selected for the PCE metamodelling (Xiu & Karniadakis 2002). The coef-
ficients aα are calculated, from the results obtained using a deterministic structural analysis
model, at the selected experimental design points (i.e. λc ), using a non-intrusive approach. For
this reason, least angle regression (LARS) method is adopted to calculate the PCE coefficients
{a0 . . . a4 }. Based on the findings of Hashmi et al. (2017) only 50 runs of FE simulations are used
to formulate a fourth-order (N ¼ 4) PCE metamodel. Hashmi et al. (2017) have shown that PCE
response (with only 50 FE runs) for Pn of a CFS member with GI are sufficient to emulate (com-
putationally-expensive) MCS response (of 1000 FE runs) with desirable accuracy. The truncated
form of fourth order PCE metamodel (MPCE ) can be expressed as

Pn ’ M PCE ðξÞ ¼ a0 þ a1 ξ þ a2 ðξp1Þ


ffiffi þ a3 ðξ p3ξÞ
ffiffi þ a4 ðξ  6ξ 2 þ 3Þ
2 3 4
pffiffi ð4Þ
2 6 2

These coefficients, due to the prudent formulation of the (sparse) PCE, can be interpreted as
the mean (μPCE ) and the variance (σ2 ) as follows (Blatman & Sudret 2011):

μPCE ¼ E½MPCE ðXÞ


¼ a0 ð5Þ
X
σ2 ¼ a2α ð6Þ
α

The PCE coefficients interpreted in the form of these two statistical parameters for the selected
section are provided in the Table 2.

4.2 Strength characterisation based on PCE metamodel output


In order to perform the strength characterisation of an imperfect CFS section, PCE based
metamodels are created and the response statistics are obtained at each λc . A total of 250
“true model” simulations are performed, comprising of 50 different geometric imperfection
realisations for each λc . PCE based metamodels are employed, however restricted to reduce
the computational efforts only, to obtain the statistical parameters ( μPCE and σ2 ). These

85
Table 2. PCE coefficients (in kN).
P
a2α ¼ σ2
Design point a0 ¼ μPCE α

λc = 0.65 122.3 5.035


λc = 0.96 116.1 5.787
λc = 1.25 88.19 3.385
λc = 1.76 51.28 0.5984
λc = 2.20 37.05 0.1793

Table 3. PCE coefficients (in kN).


λc Pd (kN) Pn (kN) Pd =Pn

0.65 112.4 112.9 0.9956


0.96 104.8 94.99 1.103
1.25 81.56 74.72 1.092
1.76 50.11 43.98 1.139
2.20 36.69 28.53 1.286

statistical parameters are reported in Table 2. The 5-percentile values are also obtained for
each λc , which represents the characteristic value (Pd ) for the flexural buckling strength from
a design specification perspective. This characteristic value (Pd ) includes the effect of geomet-
ric imperfection and associated uncertainty. Pd for a selected range of non-dimensional slen-
derness are compared with the current (AISI) design recommendations for Pn and are
represented in Table 3. The Pn values recommended in the AISI specification are calculated
for each value of λc as,
Pn ¼ Fn Ae ð7Þ

where,

Fn ¼ 0:658λc Fy
2
for λc  1:5

0:877 ð8Þ
¼ Fy for λc 41:5
λc 2

where, Ae = effective area of the cross-section and Fy = yield stress of the material.
Pd =Pn value in Table 3 shows the effect of consideration of uncertainty in geometric imper-
fections. At lower values of non-dimensional slenderness (λc 51), where failure is mainly gov-
erned by local buckling (Type 1), the Pd =Pn value is slightly less than 1, highlighting the fact
that the AISI value is slightly unconservative, however negligible, at this level. Thus, the effect
of uncertainty is found to be very low or almost negligible at this level. However, for the
higher range of non-dimensional slenderness (λc  1), where failure is governed by global
buckling ( flexural-torsional), this effect becomes more significant, as the Pd =Pn values are
higher than 1.

5 CONCLUDING REMARKS

The effect of uncertain (local and global) geometric imperfections on the strength of a lipped
channel under compression is investigated here. This study involves a consideration of geo-
metric imperfection and associated uncertainties for both the local and global GI.
A framework of FE simulations within the loop of the uncertainty quantification is employed

86
in this work. Metamodelling technique in the form of polynomial chaos expansions is used
here to avert the use of computation heavy Monte Carlo simulation. Eventually, the statistical
estimate of buckling strength at different non-dimensional slenderness is obtained for the
selected CFS section, using this approach.
Probabilistic modelling of the global imperfection (twist component or G3 ) shows that
the lognormal distribution is suitable fit for the given statistics. Based on the numerical
investigations, it is observed that at the higher range of non-dimensional slenderness
(λc  1), where the global ( flexural-torsional) buckling mode governs the performance of
CFS members, a large and noticeable variation exist between Pd and Pn values. This
shows that the AISI recommendation for design is very conservative at the higher range
of non-dimensional slenderness (λc  1). In fact at very high value of non-dimensional
slenderness (i.e. λc ¼ 2:2) the AISI prediction can be conservative up to 28.6%. While at
the lower range (λc 51), the AISI recommendation is found to be close to the recom-
mendations based on the uncertainty analysis.

REFERENCES

Ahmed, H.S.S., Ghosh, S. & Mangal, M. 2017. Probabilistic estimation of the buckling strength of
a CFS lipped-channel section with Type 1 imperfection. In Thin-Walled Structures 119: 447–456.
AISI. AISI-S100, North American specification for the design of cold-formed steel structural members
2007. American Iron and Steel Institute, Washington, DC: USA.
Blatman G, & Sudret B. 2011. Adaptive sparse polynomial chaos expansion based on least angle
regression. In Journal of Computational Physics 230(6): 2345–2367.
Hashmi, S., Sable, S. & Ghosh, S. 2014. Probabilistic Capacity Estimation of CFS Channel Sections with
Type-1 Imperfection. In The Twelfth International Conference on Computational Structures Technol-
ogy. 106. Naples: Italy.
Hashmi, SA., Subhamoy, S., & Ghosh, S. 2017. Prediction of the buckling strength of cfs members with
local geometric imperfection using stochastic kriging. In The Twelfth International Conference on
Structural Safety and Reliability. Vienna: Austria.
Marelli, S. & Sudret, B. 2014. UQLab: a framework for uncertainty quantification in MATLAB. In
Beer M, Au SK & Hall JW, (editors), Vulnerability, Risk Analysis and Management (ICVRAM2014).
USA: American Society of Civil Engineers. 2554–2563.
Schafer, B.W. & Adany, S. 2006. Buckling analysis of cold-formed steel members using cufsm: Conven-
tional and constrained finite strip methods. In Eighteenth International Specialty Conference on Cold-
Formed Steel Structures. 39–54. Orlando: Florida.
Schafer, B.W., Grigoriu, M. & Pekz T. 1998. A probabilistic examination of the ultimate strength of
cold-formed steel elements. In Thin-walled Structures 31(4): 271–288.
Schafer, B.W. & Pekz, T. 1998. Computational modeling of cold-formed steel: Characterizing geometric
imperfections and residual stresses. In Journal of Constructional Steel Research 47(3):193–210.
Schenk, C. & Schuller, G. 2003. Buckling analysis of cylindrical shells with random geometric
imperfections. In International Journal of Non-Linear Mechanics 38(7): 1119–1132.
Simulia 2013. Abaqus FEA Users Manual, Version 6.13. Vlizy-Villacoublay, France.
Spanos, PD. & Ghanem, R.S 1989. Stochastic finite element expansion for random media. Journal of
Engineering Mechanics 115(5): 103553.
Xiu, D. & Karniadakis, GE 202. The wiener-askey polynomial chaos for stochastic differential
equations. SIAM Journal on Scientific Computing 24(2): 619–644.
Zeinoddini, V. & Schafer, B. 2011. Global imperfections and dimensional variations in cold-formed steel
members. In International Journal of Structural Stability and Dynamics 11(05): 829854.
Zeinoddini, V. & Schafer, B. 2012. Simulation of geometric imperfections in cold-formed steel members
using spectral representation approach. Thin-Walled Structures 60: 105–117.

87
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Shear behaviour of sandwich panel fasteners in fire

Tesfamariam Arha, Kamila Cábová, Nikola Lišková & František Wald


Czech Technical University in Prague, Czech Republic

ABSTRACT: In recent years the usage of sandwich panels as wall cladding and roofing has
increased significantly. It has been shown that by using sandwich panels and trapezoidal sheet-
ing as a stabilizing members, a considerable amount of savings of steel can be achieved for
structural members at ambient temperature. Previous researchers has not covered this topic
under elevated temperature and these stabilising effects may also help to achieve similar sav-
ings in case of fire. The behaviour of sandwich panel fasteners in fire is very important in
order to predict and investigate the whole structure. Therefore, an experimental investigation
was conducted to study the shear behaviour of sandwich panel connections in fire when
loaded in shear under the diaphragm action. In this paper the experimental results of 16 sand-
wich panel fastener tests are presented which are carried out under the RFCS research project
STABFI which is performing a research currently on the stabilization of a structural building
using the cladding systems in fire. The results of the tests show that bearing failure of the
inner steel sheet was the main failure mode. There was no failure of screws for all the tests at
ambient and elevated temperatures except one. The results of the tests provided experimental
data for the sandwich fasteners related to building stabilization in fire through the cladding
systems which is under investigation of RFCS project STABFI.

1 INTRODUCTION

Sandwich panels are increasingly being used as structural and non-structural components in
buildings, such as wall and floor assemblies. Usually these comprise two layers of thin steel
faces and a thick core layer of lightweight insulation material. The four most commonly used
insulation materials are polyurethane foams (PUR), polyisocyanurate foams (PIR), extruded
polystyrene, and mineral wool (MW) based products. The main advantages of such assemblies
are the lightness, easy assembly and the highly efficient insulation characteristics.
In recent years, the use of light-weight sandwich panel construction, for both industrial and
civil buildings, is becoming more and more popular. When sandwich panels are used as
a cladding wall, they contribute to structural behaviour of the building in addition to the aes-
thetic effects [1]. Recent RFCS projects [2,3,4,5] have shown that by using sandwich panels
and trapezoidal sheeting as a stabilizing members, a considerable amount of savings of steel
can be achieved for structural members at ambient temperature. These projects have not con-
sidered the fire limit state and the aim of this ongoing RFCS project STABFI is to produce
new information to cover the fire situation.
In the tests the translational displacement of the fastener was measured when the specimen
was exposed to fire and then shear loaded mechanically using the displacement controlled
machine. Also, the temperatures of the inner and outer steel sheet, insulation core and the sup-
porting steel were measured. In addition to that, the failure mode of the specimen was also
monitored and recorded.

88
2 TEST PROGRAMME

The main objective of the experimental testing was to determine translational stiffness of
a sandwich panel connection with a supporting steel member at ambient and elevated temper-
atures. Testing was carried out at laboratory of Faculty of Civil Engineering, Czech Technical
University in Prague, Czech Republic.
The testing arrangement consists of a sandwich specimen of size 300 mm × 500 mm
(width × length) connected to a steel plates of thickness 8 mm using a self-tapping stainless
steel screw (SxC14–S19–5.5xL) of diameter 5.5 mm. The steel plates are considered equiva-
lent to the flange thicknesses of the steel profiles to which the sandwich panels are connected
in practice. Two types of sandwich panels: SPA E panels with mineral wool (MW) core and
KS1000RW panels with PIR core were used for the tests. The tests include sandwich panel
thicknesses of 100 mm (MW & PIR), 160 mm (PIR) and 230 mm (MW). The tests were
conducted at elevated temperatures of 250°C, 300°C and 450°C for PIR panels and 300°C,
450°C and 600°C for MW panels. The material properties of the specimens are given
in Table 1.

3 TEST SETUP

Figure 1 shows the general scheme of the translational stiffness test. The sandwich panel was
connected to the steel plates, which is equivalent to the flange thickness of the steel frames on
which the sandwich panels are connected in reality. The boundary conditions and stiffness of
the steel plates are considered to be similar to the flange of the steel supporting frames. The
sandwich panel is connected using self-tapping screws from both ends. The lower end of the
specimen was fixed on the testing machine with a pin while the upper end was free to move in
the direction of loading (upward). For tests at elevated temperatures heating was applied
using manning’s heating pads on one side of the sandwich panel and the supporting steel
member close to the lower fastener, see Figure 2b. The sandwich panel was connected with
a single screw at the lower connection where the elongation was measured using the optic
extensometer, while it was connected with three screws on the other end in order to minimize
the elongation of the fastener at that location.
Displacement-controlled testing machine was used for the experiment. The displacement
was increased monotonically until failure. A loading rate of 1 mm/min was used for the experi-
ments according to the ECCS recommendation “The testing of connections with mechanical
fasteners in steel sheeting and sections, ECCS No.124” [6]. The displacement of the fastening

Table 1. Material properties and dimensions of the sandwich panel connection.


Yield strength Thickness
Parts Material [MPa] [mm]

SPA E panel (MW)


Inner sheet S280GD+Z 280 0.5
Outer sheet S280GD+Z 280 0.6
Insulation Mineral wool 0.086 (Tension) 100/230
SPA E panel (MW)
Inner sheet S250GD 250 0.4
Outer sheet S250GD 250 0.5
Insulation PIR 0.05 (Tension) 100/160
Screw
(SxC14–S19–5.5xL) Stainless steel 450 Ø5.5
A2/AISI 304
Supporting steel S355 355 8

89
Figure 1. Scheme & experimental setup for shear resistance of sandwich panel joint.

Figure 2. Experimental setup, a) Placement of optic sensors and b) Placement of manning’s heating pads.

was measured using an optical extensometer with optic sensors placed at the central axis of
the connection as shown in Figures 1b and 2a.
At elevated temperatures, the tests were carried out in two steps. Firstly, the specimens were
heated to the designed temperatures by manning’s heating pads at the rate of 16.67°C
per minute. Then the specimen was loaded at a rate of 1 mm/min and the temperature was
kept constant until the end of the experiment. According to the “Preliminary European
Recommendations for the Testing and Design of Fastenings for Sandwich Panels, ECCS
No.127” [7] the maximum load is reached at a deformation of 3 mm for serviceability limit
state. However, fire being an accidental action large deformation is an alternative mechanism
to prevent the failure of the structure. In addition, catenary action of sheeting occurred in the

90
large deformation. Thus, the tests were stopped when a displacement of 20 mm was reached
both at room temperature and at elevated temperatures.

4 RESULTS AND DISCUSSIONS

4.1 Sandwich connection failure modes


Figure 3 shows a typical failure modes observed for the sandwich panel connection in
the experimental tests both at ambient and elevated temperatures. For all the experi-
ments the major failure was a bearing failure of the inner face of the sandwich panel
near the hole of the screw connection. For most of the experiments at ambient tempera-
ture the tearing of the inner sheet was in a narrow path approximately equivalent to the
screw diameter with a small stacking (folding) of the steel sheet during tearing off. How-
ever, for the experiments at elevated temperature the steel sheet teared off in a wider
area and also experiencing stacking (folding) of the steel sheet near the end of the open-
ing. This folding of the steel sheet strengthens the steel sheet further after the initial
peak and enables it to carry more load which can be represented by the further peak
strength values after the first peak, which can be seen in Figures 6 and 7. From all the
16 experiments there was only one specimen in which the screw failed in shear in add-
ition to the bearing failure of the sheet, see Figure 4b. Generally the screws after the
test showed no significant deformation for the PIR sandwich panels while it showed
a small bending deformation with MW sandwich panels, Figure 4.

4.2 Heating and temperature distribution on a sandwich panel


The specimens were heated to the designed temperatures by a system of ceramic heating
pads at a rate of 16.67°C per minute. Then, the temperature was kept constant until
failure. The mechanical loading of the specimen was started when the required tempera-
ture of the steel sheet of the sandwich panel was reached. During the experiment, tem-
perature of different components of the testing specimen was recorded by coated
thermocouples of type K diameter 2 mm. As a sample, the temperature distribution for
the experiments at 300°C are given in Figures 5 and 6 for both SPA E (MW) and
KS1000RW (PIR).
When the obtained temperatures for the two core materials with the same thickness are
compared, it was found that the temperature for PIR core was higher than that of MW core
due to the fact that PIR has greater thermal conductivity than MW.

Figure 3. Bearing failure of the inner steel sheet for MW_230 mm at: a) 20°C, b) 300°C, c) 450°C, d) 600°C.

91
Figure 4. Deformation of screws after the test for PIR and MW sandwich panels of 100 mm thickness.

Figure 5. Temperature distribution at 300°C for SPA E (MW), a) 100 mm thick and b) 230 mm thick.

4.3 Force-displacement curves


Displacement of the connection was measured using an optical extensometer in which the two
sensors were placed on the central axis of the specimen close to the connection. It measures
the relative movement of the two sensors using a beam of laser light. Because fire is an acci-
dental action, the design requires that the structure will not collapse in case of fire. In add-
ition, catenary action of sheeting occurred in large deformation. From these two points, the
displacement was done up to 20 mm in the experiments.The applied force was also recorded
from the testing machine, therefore the force-displacement curve was plotted for all the experi-
ments in order to determine the initial translational stiffness and the ultimate shear strength.
The force-displacement curves of the connections at ambient and elevated temperatures are
shown in Figures 6 and 7.
It can be seen that with the increase of temperature, both strength and stiffness of the con-
nections are reduced. The stiffness of connection starts to reduce when the slip is initiated at

92
Figure 6. Temperature distribution at 300°C for KS1000RW (PIR), a) 100 mm thick, b) 160 mm thick.

Figure 7. F-D curves at 300°C for KS1000RW (PIR), a) 100 mm thickness and b) 160 mm thickness.

inner steel sheet because of the spreading of yielding zone. At elevated temperatures, the
stiffness reduced further due to the degradation of the material. It can be observed that the
folding of the steel sheet after the first peak strengthens the steel sheet further and enables it
to carry more load which can be represented by another peaks after the first one as shown in
Figures 7 and 8.

Figure 8. F-D curves at 300°C for SPA E (MW) a) 100 mm thickness and b) 230 mm thickness.

93
4.4 Shear stiffness and resistance
The maximum load and the shear displacement corresponding to the maximum load in
the fastening shall be measured in order to determine the shear stiffness of the screw
connection. According to the ECCS recommendation [7] the ultimate load is defined to
be the smallest of:
– the load corresponding to a displacement of 3 mm, if this occurs on the rising part of the
load - deflection curve.
– the maximum load recorded during the test.
– the load at which the first decrease in load is observed in the load - deflection curve.
According to ECCS recommendation [6], the shear flexibility (ch ) of a fastening should
be determined by Eq. (1). The shear stiffness is obtained from the inverse of its shear
flexibility.
P
1 ah
ch ¼ : ð1Þ
Rd= n
γ1
Rk
Rd ¼ ð2Þ
γ2

Where:
ah ¼ the slip of the fastening at a load equivalent to Rd=γ .
1
Rd ¼ design resistance of the fastening
γ1 ¼ an appropriate factor (1.5 for fastening sheets to substructure with the one-fastener test)
n ¼ number of test specimens
Rk ¼ is the characteristic resistance
γ2 ¼ partial factor for resistance (1.25 according to EN 1993-1-3[8])

The recorded maximum load and its corresponding displacement at elevated temperatures are
shown in Table 2. The initial shear stiffness calculated according to Eq. (1) is also given there.

Table 2. Initial stiffness and recorded maximum load during the test with its corresponding
displacement.
Maximum load Maximum Elongation Initial stiffness from Eq. (1)
Test Identification No. [KN] [mm] [KN/mm]

MW_100mm_8mm_20°C 1.616 1.308 2.15


MW_100mm_8mm_300°C 1.708 1.48 1.34
MW_100mm_8mm_450°C 1.532 2.176 0.862
MW_100mm_8mm_600°C 0.692 2.792 0.488
MW_230mm_8mm_20°C 2.324 4.38 0.863
MW_230mm_8mm_300°C 1.28 4.116 1.076
MW_230mm_8mm_450°C 2.196 8.032 0.527
MW_230mm_8mm_600°C 0.888 9.616 0.405
PIR_100mm_8mm_20°C 1.948 6.132 0.667
PIR_100mm_8mm_250°C 1.312 7.48 0.96
PIR_100mm_8mm_300°C 0.924 2.664 0.322
PIR_100mm_8mm_450°C 0.808 7.84 0.195
PIR_160mm_8mm_20°C 1.8 2.44 1.23
PIR_160mm_8mm_200°C 1.6 3.684 0.745
PIR_160mm_8mm_300°C 1.128 6.516 0.19
PIR_160mm_8mm_450°C 0.725 5.014 0.17

94
5 CONCLUSIONS

The STABFI project is investigating the stabilising effect of sandwich panels and trapezoidal
sheets for structural buildings in fire. Under the mentioned project this paper presented the
experimental results on the shear behaviour of sandwich panel fasteners both at ambient and
elevated temperature.
The major failure was a bearing of the inner face of the sandwich panel. The bearing of the
steel sheet at ambient temperature was narrow with small folding of sheet while at elevated tem-
perature it teared off in a wider area with pronounced folding of the sheet. This folding of the
steel sheet strengthens the steel sheet and enables to carry more load further. There was no
shear failure of screws for all the 16 tests at ambient and elevated temperatures except one. With
the increase of temperature, both the strength and stiffness of the connections are reduced.
The obtained initial shear stiffness of the fasteners from the experiments will be used in the
analysis of a whole building for further studies.

REFERENCES

[1] De Mattes. G, Landolfo. R. Structural behaviour of sandwich panel shear walls. An experimental
analysis. Materials and Structures, Vol. 32, 1999, pp. 331–341.
[2] EASIE. Ensuring advancement in sandwich construction through innovation and exploitation. Pro-
ject final report, 2013.
[3] ECCS/CIB. European recommendations on the stabilization of steel structures by sandwich panels.
Publication 379, ECCS TC7 and CIB W056, 2013.
[4] Hedman-Petursson, E. Column buckling with restraints from sandwich wall elements. Doctoral
thesis, Steel Structure Division, Department of Civil and Mining Engineering. Lulea, Sweden, 2001.
[5] Misiek, T., Käpplein, S., Saal, H. and Ummenhofer, T. Stabilization of beams by sandwich panels –
Lateral and torsional restraint. EuroSteel, August 31-September 2, 2011. Budapest, Hungary.
[6] ECCS. The Testing of Connections with mechanical Fasteners in Steel Sheeting and Sections, ECCS
publication No. 124, 2009.
[7] Preliminary European Recommendations for the Testing and Design of Fastenings for Sandwich Panels -
ECCS publication No. 127, 2009
[8] EN1993-1-2, Eurocode3. Design of steelstructures.Part1-2: General rules. Structural fire design, 2005.

95
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Bracing details for trapezoidal steel box girders

S.V. Armijos-Moya, Y. Wang, T. Helwig, M. Engelhardt, E. Williamson & P. Clayton


The University of Texas at Austin, Austin, Texas, USA

ABSTRACT: Steel trapezoidal girders (tub girders) with a cast in-place concrete deck on top
are a popular alternative for straight and horizontally curved bridges due to their high torsional
stiffness and aesthetic appearance. However, steel tub girders possess a relatively low torsional
stiffness during construction due to the thin-walled open section that is susceptible to stability
issues. Top flange lateral bracing, in the form of a horizontal truss, is installed along the entire
length of the steel tub girder to increase the torsional stiffness of the girder. Internal K-frames are
placed to control cross-sectional distortion. This paper provides a summary of a research study
focused on improving the efficiency of steel tub girders by investigating the impact of bracing
details on the behavior of the girders. The study includes large-scale experimental tests and para-
metric finite element analytical studies. The goal of the study is to propose efficient details for trap-
ezoidal steel girders to make them more cost-effective without undermining their structural
performance.

1 INTRODUCTION

Steel trapezoidal box girders have become a popular alternative for straight and curved
bridges. The girders, often referred to as “tub girders”, consist of a single bottom flange, two
sloping webs and two top flanges. The smooth profile of the girder provides an aesthetically
appealing bridge that also possesses several structural advantages compared to other girder
types. As a result of the large torsional stiffness, the girders are a popular choice in horizon-
tally curved systems where the bridge geometry leads to large torsional moments. However,
during construction the girders are an open section and generally require extensive bracing.
The primary bracing systems consist of plate diaphragms at the supports, a top flange lateral
truss, and intermediate internal and external K-frames (Figure 1).
Though tub girders have mainly been used on horizontally curved bridges where concrete
girders are not viable due to the longer span lengths or due to the curvature, steel tub girders
have also been shown to be feasible for straight bridges with span lengths normally reserved
for concrete girder systems. Relatively shallow straight steel tub girders were recently used by
the Texas Department of Transportation in the Waco District in the U.S.A, what demon-
strates that steel trapezoidal box girders offer a viable alternative for a wider variety of bridge
applications. To augment the viability of the tub girders, improved girder geometries and bra-
cing details may lead to improved economy and structural efficiency. Details investigated in
this research study include the spacing between internal K-frames, the layout of the top lateral
truss, and the cross-sectional geometry of the steel tub girders. Common geometrical practices
for the tub girders consist of a 4V:1H web slope and the top flanges centered over the webs.
A flatter web slope can lead to increased lateral coverage of a single girder and may eliminate
a girder line, thereby improving economy. In addition, offsetting the top flanges towards the
inside of the tub girder can provide increased efficiency with respect to connections to the bra-
cing systems. To study the aforementioned-proposed details, three tub girders were fabricated
for the experimental program.
The experimental studies included loading the girders in vertical bending as well as in com-
bined bending and torsion. Part of the experimental results are summarized in this paper.

96
Figure 1. Bracing systems in twin tub girder during construction.

Finally, an analytical study to evaluate the torsional response in a continuous bridge under
construction loads is presented. This paper focuses on the impact of improved bracing details
on the torsional response of tub girders under demands expected on straight and horizontally
curved systems.

2 TEST OF LARGE SCALE SPECIMENS

2.1 Description of specimens


Three steel tub girders were designed and fabricated for the experimental stage of the study.
First, the baseline girder has a web slope of 4V:1H with the flanges centered over the webs.
An additional specimen also has a 4V:1H web slope with the top flanges offset towards the
inside of the girders (offset top flange girder), while the final specimen has a web slope of
approximately 2.5V: 1H and top flanges centered over the web (flatter web girder). All of the
internal K-frames and top lateral truss members were bolted to facilitate variations of the bra-
cing between tests. The baseline girder was designed and fabricated according to current
engineering practices in the U.S.A. for straight and curved tub girders. The other two speci-
mens were sized by conducting preliminary finite element analyses so that the girders are able
to reach global lateral torsional buckling before any type of local buckling.
The focus of this study is on both straight and horizontally curved girders. Though the
research team considered fabricating horizontally curved girders, laboratory space limitations
as well as the limitation of being able to test a single girder curvature was not desirable.
Instead, the research team focused on a setup that allowed eccentric loading that can simulate
the torsion from the horizontal curvature of the girder. With the ability to offset the load to
achieve a torque, girder geometries from straight to a simulated curvature of approximately
600 ft. were possible.

2.1.1 Tub girder geometries


The proportions of the girders were selected so that the girders would remain elastic during mul-
tiple bending and combined bending and torsion tests. The clear span L of the simply supported
specimen was selected to be 84 ft., while the girder depth D was defined as 3 ft. (L/D = 28).
A distance W equal to 5 ft. and 3 in. was selected as the separation of the top of the sloped webs
(Figure 2). The resulting width-to-depth ratio (W/D) was 1.75, which is similar to values
observed in current practice. The major difference between specimens is the thickness of the
cross-section plates, the location of the top flanges with respect to the webs, and the slope of the
webs. All the flanges and webs were fabricated with steel AASHTO M270 (ASTM A709),
grade 50W.

97
Figure 2. Specimen cross-section: a) baseline girder, b) top flange offset girder, c) flatter web girder.

The baseline steel tub girder was sized with webs sloped to 4V:1H (Figure 2a). The thickness
of webs and flanges was set equal to 7/16 in., what is considerably smaller than commonly util-
ized in current bridge practice (1 in). However, this thickness was deemed necessary to obtain
the elastic-buckling response of the system based upon finite element studies. This base line tub
girder was built with two 12” wide top flanges which were centered to the center line of the
sloped webs. The offset top flange girder was built with two 13” wide top flanges which were
connected to the sloped webs at 1” from the edges, leaving 12” of unstiffened plate (Figure 2b).
Finite element analyses were carried out to define the thickness of the top flanges for
this second specimen to assure an elastic behavior of the girder during the tests. The top flange
thickness was set equal to 9/16”. The bottom flange and sloped webs were sized with 7/16” thick
plates. The flatter web girder, on the other hand, was fabricated with web slopes equal to
approximately 2.5V:1H (Figure 2c), which exceeds the limits of AASHTO 2014. Similar to the
baseline tub girder, the flatter web girder was built with top flanges centered over the webs.
Webs and flanges were 7/16 in. thick.

2.1.2 Bracing geometry


The spacing of the top lateral truss panel points was defined as 7 ft., generating 12 panels
along the length of the beam (Figure 3). The internal K-frames and top lateral diagonals were
able to be installed or removed as desired to study the behavior of the girders as the bracing is
varied. In the cases where the internal K-frames or top lateral truss diagonals were removed,
top lateral struts between the two top flanges were kept at a 7 ft. spacing to control separation
of the top flanges.
The single-diagonal type (SD-type) top truss was used as the top lateral system. This system
is formed by single diagonals and struts connected to the tub girder top flanges. The top truss
diagonals were WT5 x 22.5 members designed to be connected directly underneath the top
flanges through three 3/4 in. high strength bolts. The strut cross-section was sized as a 2 in
diameter x-strong pipe (2.375 in. outside diameter and 0.218 in. wall thickness). The struts
were connected to a stiffener welded to the web of the tub girder through bolted connections
made of 1/2 in. thick steel plates (ASTM A-36) and 7/8 in. high strength bolts. The vertical

Figure 3. Example of bracing layout - half of baseline steel tub girder specimen - Plan view.

98
eccentricity between the top flange and the centerline of the strut is 3.75 in. which is an accept-
able value (Helwig and Yura 2012). The diagonals and pipes were designed and fabricated
with steel ASTM A705, Grade 50 and ASTM A53, Grade B, respectively. Three diagonals
were installed at each end of the steel tub girders to simulate partial lateral bracing of the top
flange. Different cases of partial top lateral bracing were tested by removing diagonal mem-
bers of the horizontal truss at each end (4 layouts).
One strut (which is part of the top lateral truss) and two diagonals formed the internal
K-frames. The section of the strut was sized for the top lateral bracing system, and the same sec-
tion has been adopted for the K-frame diagonals (2 in. x-strong) for facility during fabrication.
The K-frame bracing elements were fabricated with ASTM A53 – Grade B steel. Three different
arrangements of internal K-frames were tested for each configuration of top lateral bracing.
K-frame bracing at every 2, 4 and 6 panel points were evaluated during the experimental program.

2.2 Description of test setup


The test setup (Figure 4a) consisted of two steel supports 84 ft. apart over which each specimen
was tested as simply-supported straight girder under both pure positive bending and torsional
loading conditions. Each steel support consists of three 12 ft. long W36 x 135 rolled beams stacked
vertically so as to raise the elevation of the test girders above the loading system. The support
located on the south side of the laboratory floor was supported laterally with two diagonal braces
to stiffen the test setup and simulate “pinned conditions”. The opposing support consisted only of
the stacked W36 x 135 sections and allowed some flexibility to simulate a “roller”.
Two gravity load simulators (GLS) shown in Figure 4b were used to apply either pure bend-
ing or bending with torsion. Each GLS is able to apply vertical loads up to 160 kips, and to
keep the load vertical even if the ram moved laterally up to 6 in. Consequently, the GLS pro-
vides minimal lateral restraint and essentially “simulates gravity load”.

2.3 Instrumentation and initial imperfections


Two 100-kip load cells were used to measure the loads applied with the two GLS. Horizontal and
vertical deflections of the specimens were measured at the third points along the tub length (28 ft.
and 56 ft.) and at mid-span (42 ft.). The deflections at the third points were obtained with four
string potentiometers; while at mid-span they were collected with two infrared cameras that were
able to monitor the signal from LED markers attached to the tub girder with relatively high accur-
acy (error of approximately 0.01 mm). Rotations were calculated from the measured deflections.

Figure 4. a) Test setup, b) Gravity load simulator (GLS).

99
Prior to testing, initial imperfections of each steel tub girder were measured. Two piano wires
were extended between the test setup supports located 6 in. from both edges of the bottom flange.
The taut wires served as reference point to measure lateral and vertical out-of-straightness of the
tub girder. The baseline, top flange offset, and flatter web girders had an initial twist at midspan
of 1.30, 1.60, and 2.30 degrees, respectively; and a maximum out-of-straightness on top flange of
about L/1300 towards the east, L/750 towards the west, and L/500 towards the east, respectively.

2.4 Testing procedure


Elastic-buckling tests were carried out by limiting the maximum loads applied to the specimen
to keep stresses below 60% of nominal yield stress (30 ksi) to consider the impact of residual
stresses and initial imperfections in the response, and to ensure that the girders remained elastic.
Two types of loading conditions were studied: vertical positive bending and combined bending
and torsion due to vertical eccentric loads (to simulate horizontal curvature). Two vertical loads
were applied with gravity load simulators at approximately quarter points of the specimen (loca-
tion denoted as “Pa” on Figure 3). Henceforth, the load on each GLS will be referred to as load
“P”. The combined vertical bending and torsional demands were obtained by applying vertical
eccentric loads at 8 in. and 16 in. from the shear center of the girders to simulate demands pro-
duced by curvature in tub girders with radii of curvature equal to 1260 and 630 ft., respectively.

2.5 Bracing configuration


To measure the impact of bracing in the response of the specimens, different bracing layouts were
tested on the each tub girder under the same loading conditions. Four configurations of top lateral
bracing (zero, one, two, and three diagonals on each end of the simple supported girder) and three
layouts of internal K-frames (frames at every two, four, and six panel points) were tested, which
resulted in a total of 12 tests. These 12 configurations of top lateral and K-frame bracing were
evaluated for the three cases of vertical loads (concentric, eccentric at 8 in., and eccentric at 16 in.)
producing a total of 36 elastic tests performed on each specimen. The impact of each bracing
layout in the response of the specimens was evaluated and is summarized in the next sections.

2.6 Experimental results

2.6.1 Impact of partial top lateral bracing distribution in stiffness


To simulate the demands on straight tub girders, the GLSs were used to apply vertical concen-
tric loads near quarter points. Figure 5 shows the total vertical load applied (2P) versus the
twist angle of the baseline specimen at midspan (β), when the specimen was tested with zero,
one, two, and three bracing diagonals at each end. K-frames were kept every 2 panel points.
The tub girder without top lateral bracing presented a deformation curve that suggested the
girder was approaching the elastic lateral torsional buckling limit during the test, which can be
observed by the significant nonlinear response of the load versus deflection curve. The torsional
stiffness of the specimen reduced significantly as the girder approached the lateral torsional
buckling limit. The capacity to resist LTB is significantly improved with the addition of diag-
onals at the ends of the girder. The system without diagonals had a twist at midspan of
1.88 deg. at 70 kips of total load; while the specimen with 1 diagonal per end had a twist at
midspan of 0.30 deg. at the same load step. This is a reduction of about 85% in the rotation of
the cross-section and indicates that the torsional stiffness is highly improved with a single diag-
onal at each end. The baseline specimen with two truss diagonals per end presented a twist
angle of 0.09 deg. showing a clear improvement in the torsional stiffness of the specimen. The
tub girder with three diagonals per end did not show a significant improvement in torsional
stiffness with respect to the previous case. Instead, the three diagonals per end produced a shift
in the direction of lateral movement (shift of mode shape). The first diagonal on each end of the
specimen produced the most significant improvement in the resistance to LTB, while additional

100
Figure 5. Total GLS load vs twist angle at midspan – Baseline specimen (concentric loading).

diagonals were not as effective at improving the behavior. The experimental results shown that
the effectiveness of the top lateral truss is lower with increasing distance from the girder ends.
When applying eccentric vertical loads, similar trends to the concentric cases was observed.
The specimens showed poor torsional resistance when no top lateral diagonals were installed.
The torsional stiffness of the girders was enhanced when bracing diagonals were installed at
each end because they restricted the warping deformations on the girders.
Figure 6 plots the absolute values of torsional response of the baseline girder with zero and
three diagonals per end subjected to concentric and eccentric loads. Similar to the concentric
loading case, K-frames were maintained at every 2 panel points. As expected, in regards to the
unbraced cases, the torsional stiffness of the baseline girder goes down when the torsional

Figure 6. Total GLS load vs twist angle at midspan – Baseline specimen with 0 and 3 diagonals.

101
demands increase. The initial torsional stiffness of the unbraced concentric case is about 4 and
12 times higher than the cases with loads applied at 8 in. and 16 in. of eccentricity, respectively.
Regarding the braced cases, with three diagonals, the specimen shows lower differences in the
torsional stiffness between the three loading cases. The stiffness observed during the concentric
test is about 1.5 and 4 times higher than the results obtained with eccentric loading at 8 in. and
16 in., respectively. As previously discussed, the addition of top bracing diagonals produced
a high increment in the torsional stiffness. The torsional stiffness increased about 10 to 30 times
after installing partial top lateral bracing with three diagonals. The horizontal truss diagonals
are more effective at the supports where warping deformations are higher. Similar type of
impact on the torsional behavior was observed in the other two specimens.

2.6.2 Impact of internal K-frame distribution in stiffness


Figure 7 shows the total vertical load (2P) versus the twist at midspan (β) of the baseline girder
with no top lateral bracing and three different configurations of internal K-frames subjected to
concentric vertical loads. The tub girder with internal bracing at every four and six panels show
the same response with no major variation in the torsional stiffness. Torsional demands in straight
tub girders are small which implies that distortional effects are low. Even though the specimen
with K-frames every 2 panels presents higher torsional stiffness for lower load levels, the impact
on the response tended to be similar to the aforementioned two cases at higher loads. Elastic lat-
eral torsional buckling was observed during the three tests. The relative insensitivity of girder
response to the internal K-frame spacing is similar to previous observations in the case of the
Marcy Pedestrian Bridge failure (Yura and Widianto 2005), as well as the system-buckling mode
of narrow I-girder systems (Yura et al. 2008). Consequently, the K-frame bracing system becomes
less effective in straight steel tub girders under pure positive bending, and there is no major change
in its torsional behavior when the number of internal braces is reduced. Similar effect was observed
in the other two specimens with partial top lateral bracing, including two and three diagonals at
each end.

3 ANALYTICAL STUDY

Parametric studies were conducted to further confirm the experimental findings and evaluate the
effects of partial top lateral bracing on the stiffness of steel tub girders with different girder geom-
etries and configurations. As part of the parametric study, the section in Figure 8 was evaluated
with different amounts of partial-length top lateral bracing for both simply supported and con-
tinuous girder configurations. The simply supported tub girder was modeled as straight girder

Figure 7. Total load vs midspan twist angle - Different K-frame layout (No TLB) – Baseline girder.

102
Figure 8. Prototype tub girder for analytical study.

with a clear span of 216 ft. (L/D ≈ 30), while the continuous tub girder was modeled as 2-equal-
span horizontally curved system (radius of curvature = 2500 ft) with clear spans of 216 ft. as well.
The cross-section dimensions of each prototype section were re-proportioned accordingly to sat-
isfy the load demand, and the corresponding plate thicknesses for both systems are presented in
Table 1. The number of top lateral truss panels for each girder was set at 16 panels per span with
a panel length of 13.50 ft. Following current engineering practices, K-frames were placed every
panel point for all the cases presented in this section. Different amounts of partial top lateral bra-
cing were considered for both tub girders. The minimum levels of bracing for adequate perform-
ance during construction for both girder systems was then evaluated.
Lateral displacements of the top flanges (δ) were examined to evaluate the load-deflection
response of the prototype steel tub girders in the analytical study. The results were normalized
by maximum allowable displacement and section twist defined for this study. In this study,
a limitation of L/1000 is defined for lateral displacement of the top flange based on the max-
imum out-of-straightness fabrication tolerance specified in the AISC Code of Standard Prac-
tice (2005) for straight compression members.
FEAs were first carried out on the straight girder case with various amount of top lateral
bracing to identify the minimum required top lateral bracing to satisfy the admissible require-
ments described above. The layout of top lateral bracing varies from non-braced to fully
braced cases. Figure 9 shows the sample lateral displacement responses of the simply sup-
ported girder with seven different layouts of top lateral bracing. (i.e., three diagonals (0.38 L)
at each end, four diagonals (0.5 L) at each end and a fully braced system (1.0 L)). The X-axis
of the plot shows the lateral displacement of top flange normalized by L/1000 (δmax = 2.60
in), while the Y-axis shows the bending moment normalized by an estimated moment pro-
duced by construction loads. In the plot, the different cases are labelled based on the percent-
age of braced length with respect to the girder span length. In Figure 9, the minimum amount
of bracing required not to exceed the admissible lateral displacement (1.0 in the X-axis) is

Table 1. Plate thicknesses for prototype girder.


Girder System bf (top) tf (top) tf (bot) tw

in in in in

Simply Supported 20” 2.5” 2.0” 0.81”


Continuous (2 Span) 30” 3.0” 2.0” 0.81”

103
Figure 9. Bending moment vs lateral displacement of top flange – simply-supported girder.

50% of bracing (0.5 L) under the assumed construction demands (1.0 in the y-axis). Beyond
this case, the girder starts to exhibit larger deformations.
Similar FEAs were performed to examine the possibility of using partial top lateral trusses
on horizontally curved tub girders. It was found that a radius of curvature equal to 2500 ft.
was adequate to comply with the aforementioned requirements. Additionally, the top flanges
within the unbraced length had to be resized because the lateral bending stresses induced in
top flanges were higher that the yielding stress. The requirements were satisfied after resizing
the top flanges for the unbraced length and choosing a curvature equal to 2500 ft. To resize
the top flanges for the unbraced length, the lateral bending stresses in the top flanges were
calculated considering the unbraced length of the top flanges between partial top lateral bra-
cing with warping restrain at the transition points (panel point where the tub girder changes
from top laterally braced to unbraced).
After resizing the top flanges, the girders in this study were able to satisfy the response
requirements for horizontal curvatures of 2500 ft or higher. Figure 10 shows the bending
moment versus lateral displacement on the top flange of the critical span for the two span con-
tinuous girder with radius of curvature equal to 2500 ft. The most critical construction loading
scenario was obtained when the construction loads were applied over the positive moment
region of one span.
These analyses indicate that 50% partial top lateral bracing is a reasonable minimum
amount of bracing to keep maintain admissible girder deformations during construction for
straight and mildly horizontally curved steel tub girders with minimum radius of curvature of
2500 ft. The horizontally curved girders in this study showed acceptable response up to lengths
of 216 ft. with 50% partial bracing. Longer lengths required additional top lateral bracing
(60%) to maintain acceptable behavior.
Additionally, it was observed during the finite element analyses that when partial top lateral
bracing is used, a K-frame has to be installed in the transition panel point (between the girder
with top lateral diagonals and the unbraced girder) in order to avoid detrimental impact in the
torsional stiffness of the girder. K-frames should be placed at the transition points and at mid-
span. It is recommended to use internal K-frames every 2 panel points and at transition zones
for adequate torsional performance.

104
Figure 10. Bending moment vs lateral displacement of top flange – continuous girder.

4 CONCLUSIONS

– Experimental tests and analytical analyses showed that the top flange lateral bracing systems
are more effective in the region near to the supports of straight girders where shear deform-
ations are at the maximum. The LTB capacity of the straight tub girders was significantly
improved by adding 50% of top truss diagonals at supports. The inclusion of subsequent
diagonals resulted in significantly smaller increments in the performance as the distance to
the supports increased. Thus, top lateral diagonals located near mid-span add little to no
benefit in the LTB behavior and likely at increasing the torsional stiffness of the girder.
– Internal K-frames provide minimal contribution to resist LTB in straight tub girders in
comparison to top lateral bracing. Due to lower torsional demands, internal K-frames are
less effective along straight tub girders. Thus, the number of K-frames, and their distribu-
tion along the straight girder, did not shown a significant impact in the torsional response
of the girder.
– When partial top lateral bracing is used, a K-frame has to be provided in the transition zone
between braced and unbraced girder in order to maintain adequate torsional response.

REFERENCES

American Association of State Highway Transportation Officials (AASHTO) (2014). “AASHTO LRFD
Bridge Design Specifications, 6th Ed.” American Association of State Highway and Transportation
Officials, Washington, D.C.
Helwig, T. and J. Yura (2012). “Steel Bridge Design Handbook: Bracing System Design”, U.S. Depart-
ment of Transportation Federal Highway Administration. 13.
Yura, J. A., and Widianto, (2005), “Lateral buckling and bracing of beams – A re-evaluation after the
Marcy bridge collapse”, Proc., Structural Stability Research Council, Montreal, April 7–9, pp 277–294
Yura, J., Helwig, T.A., Herman, R., and Zhou, C. (2008), “Global Lateral Buckling of I-Shaped Girder
Systems,” ASCE Journal of Structural Engineering, Vol. 134, No. 9, pp. 1487–1494, September.

105
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Behaviour of slender plates in case of fire of different stainless


steel grades

F. Arrais, N. Lopes, P. Vila Real & C. Couto


RISCO—Civil Engineering Department, University of Aveiro, Portugal

ABSTRACT: Stainless steel has countless desirable characteristics for a structural material.
Although initially more expensive than conventional carbon steel, stainless steel structures can
be competitive because of their smaller or none need for thermal protection material and
lower life-cycle cost, thus contributing to a more sustainable construction. Regarding struc-
tural fire resistance, in order to have a comprehensive understanding of the overall members’
resistance, it is important to first analyse the cross-section resistance, directly affected by local
instabilities occurrence on the composed thin plates. This work presents a numerical study on
the behaviour of isolated plates at elevated temperatures, corresponded to the web (internal
element) and flanges (outstand element) of I-cross sections, comparing the numerically
obtained ultimate load bearing capacities with simplified calculation formulae for the applica-
tion of the effective width method. Comparisons between the numerical results and the EC3
formulae for determining the effective area of thin plates is also presented.

1 INTRODUCTION

The application of stainless steel as a structural material has been increasing, due to a number
of desirable qualities such as its durability, resistance to corrosion and aesthetic appearance
(Gardner, 2005 & Euro Inox, 2006). Despite having a high initial cost, stainless steel can be
a competitive material if life cycle cost analysis is considered, due to its low maintenance
needs. Moreover, it has a higher fire resistance when compared to carbon steel (CEN, 2005b)
allowing in some cases the absence of thermal protection.
The austenitic stainless steels are generally the most used groups for structural applica-
tions but some interest has being recently shown for increasing the use of ferritic and
austenitic-ferritic (Duplex) steels for structural purposes due to specific advantages. Some
of those advantages are the very good resistance to wear and stress corrosion cracking of
the duplex grade and the lower percentage of Nickel of the ferritic grade, which reduces
its price.
Regarding structural fire resistance, in order to have a comprehensive understanding of the
overall members’ resistance, it is important to first analyse the cross-section resistance, which
is directly affected by local instabilities occurrence on the composed thin plates.
For structural design purposes, Eurocode 3 (EC3) (CEN, 2006a) considers that the walls
slenderness determine the cross-section classification (from Class 1 - stocky sections to
Class 4 - slender sections). Subsequently, the cross-section resistance is calculated consid-
ering plastic section properties for Classes 1 and 2 sections, elastic section properties for
Class 3 sections and effective section properties, applying the effective width method, for
Class 4 sections. In addition, for cross-section of Classes 1, 2 and 3 at elevated temperat-
ures the strength at 2% total strain should be considered as the yield strength and for Class
4 cross-sections it should be applied the 0.2% proof strength (CEN, 2005b).
Although the subject of local buckling at elevated temperatures has been studied by differ-
ent authors (Couto et al., 2014, Couto et al., 2015, FIDESC4, 2014, Knobloch & Fontana,
2006, Maraveas et at., 2017, Quiel & Garlock, 2010), the mentioned research works only

106
address carbon steel elements and research of the local buckling effect on stainless steel sec-
tions at elevated temperatures is scarce and mostly focus on the member behaviour.
According to Part 1-2 of EC3 (CEN, 2005b) design rules, stainless steel stress-strain rela-
tionships at elevated temperatures are characterized by having an always non-linear behaviour
with an extensive hardening phase, when compared with carbon steel constitutive law. As
existing fire design guidelines for stainless steel, such as in EN 1993-1-2 (CEN, 2005b), are
based on the formulations developed for carbon steel members (CEN, 2005a, CEN, 2006b), in
spite of their different material behaviour, it is still necessary to develop knowledge on stain-
less steel structural behaviour at elevated temperatures.
This research work has the main objective of analysing the accuracy of EC3 present calcula-
tion proposals for stainless steel cross-sections in case of fire, subjected to compression or
bending, by means of Geometrical and Material Non-linear Analysis with Imperfections
applying the Finite Element software SAFIR (Franssen & Gernay, 2017). Plates behaviour at
elevated temperatures is analysed considering compression or bending and different boundary
conditions for modelling isolated outstand elements (flanges) and internal elements (webs),
following the methodology used for the development of carbon steel design approaches
(Couto et al., 2014, FIDESC4, 2014, CEN, 2006b). In this parametric study, as different stain-
less steel grades exibit different stress-strain relationships behaviours at elevated temperatures
(CEN, 2005b), the following grades were considered: i) 1.4301 (Austenitic grade); ii) 1.4003
(Ferritic grade); iii) 1.4462 (duplex).
Comparisons between the obtained numerical results, the EC3 design methods and a recent
proposal for Class 4 carbon steel sections (Couto et al., 2015), are made, being concluded that
new design expressions should be developed for the effective with method application on
stainless steel I-sections subjected to fire.

2 SIMPLIFIED DESIGN RULES

2.1 Eurocode3
According to EN 1993-1-2 (CEN, 2005b), the section resistance of a stainless steel member in
case of fire is calculated in the same way as for carbon steel, changing only the mechanical
properties of the material to consider uniform elevated temperatures in the section.
Regarding the cross-section classification, Equation 1 was used to determine the factor ε,
a parameter necessary for the determination of the EC3 classification limits (Franssen & Vila
Real, 2015).

0:5
235 E
εθ ¼ 0:85 ð1Þ
fy 210000

The design resistance value of axially compressed members of Class 1, 2 or 3 cross-sections


with a uniform temperature θa is determined from Equation 2.

Nfi;t;Rd ¼ A fy;θ =γM;fi ð2Þ

For Class 4 sections, according to Annex E of EN 1993-1-2, the effective area (Aeff ), obtained
from EN 1993-1-5 (CEN, 2006b), should be considered instead of the gross cross-section area A.
In a fire situation higher strains are acceptable when compared to normal temperature design,
therefore, instead of 0.2% proof strength usually considered at normal temperature, for cross-
section of classes 1, 2 and 3 at elevated temperatures the stress corresponding to 2% of total strain
should be adopted as the yield strength (CEN, 2005b).

107
fy;θ ¼ f2%;θ ¼ k2%;θ fy ð3Þ

However, for Class 4 cross-sections, according to Annex E of EN 1993-1-2, the proof strength
at 0.2% strain should be used, thus

fy;θ ¼ f0:2p;θ ¼ k0:2p;θ fy ð4Þ

The mentioned reduction factors are given on Annex C of EN 1993-1-2 for stainless steel at
high temperatures for the different analysed stainless steel grades.
In beams, the design value of the bending moment resistance of a cross-section with
a uniform temperature θa is determined from:

h i
Mfi;θ;Rd ¼ ky;θ γM;0 =γM;fi Mc;Rd ð5Þ

Being Mc; Rd for Classes 1 and 2 the plastic bending moment capacity, for Class 3 the elastic
bending moment capacity, and for Class 4 sections the effective bending moment capacity, at
normal temperature, determined with the effective section properties obtained from EN 1993-
1-5. The effective area and effective section modulus (Weff ;y ) are determined through the appli-
cation of the effective width method, considering the reduction of resistance due to local buck-
ling effects (CEN, 2006b). On this regard, the EN 1993-1-4 (CEN, 2006a) provides specific
equations for the determination of the plate reduction factors (ρ) to the width of elements
composing the stainless steel sections, as presented in Equation 6 and Table 1. It can be
observed that the reduction factor for internal elements do not depend on the stress distribu-
tion as proposed in carbon steel plates (CEN, 2005a).

beff ¼ ρ:b ð6Þ

The plate slenderness – λp – value is determined with Equation 7.

sffiffiffiffiffiffi 
fy b t
λp ¼ ¼ pffiffiffiffiffi ð7Þ
σcr 28:4ε kσ

2.2 Proposal for class 4 carbon steel sections at elevated temperatures


As mentioned before, recent research works (Couto et al., 2015, Knobloch & Fontana, 2006)
proposed the use of the stress corresponding to 2% of total strain as the steel yield strength
also for Class 4 cross-sections at elevated temperatures, as it is done for the remaining

Table 1. Reduction factor for stainless steel sections


elements.
Cross-section elements Reduction factor

Welded outstand elements ρ ¼ λ1p  0:242


λ2
1
p

Welded internal elements ρ¼ 0:772


 0:125 1
λp 2
λ p

108
Table 2. Reduction factor proposed for carbon steel sections
elements in case of fire (Couto et al., 2015).
Cross-section elements Reduction factor
ðλp þ 1:10:52
ε Þ 0:188
1:2
Outstand elements ρ¼  1:0
ðλp þ 1:10:52ε Þ
2:4

ðλp þ 0:90:26 ε Þ 0:055ð3þψÞ


1:5

Internal elements ρ¼  1:0


ðλp þ 0:90:26ε Þ
3

sections, providing the plate reduction factors would be calculated as presented in Table 2.
The accuracy of the application of this proposal for stainless steel sections is tested in this
paper.

3 PLATES BEHAVIOUR

3.1 Numerical modelling


Members composed of different cross-section shapes may exibit diferent plates behaviour. For
instance in I-shape sections subjected to compression have both flanges and web in compres-
sion, whereas when the members are subjected to bending in the strong axis, a flange is in
compression while the web is subjected to bending. Rectangular hollow sections subjected to
compression will have only internal elements subjected to compression and when subjected to
bending will have an internal element subjected to compression and others internal elements
subjected to bending.
To determine the ultimate load of rectangular plates the program SAFIR was used. Each
shell element has four nodes with six degrees of freedom (three translations and three rota-
tions). Simply supported conditions where applied to the plates by restraining the vertical dis-
placements, in addition the rotations at the edges of the plate were also restrained to simulate
the web-flange continuity. For the outstand elements, the vertical displacements were
restrained on three sides while for the internal elements the vertical displacements were
restrained in all four sides, this methodology follows the same principles as in Couto et al.,
(2014). Figure 1 presents the obtained deformed shapes of an outstand element subjected to
compression, an internal element subjected to compression and an internal element subjected
to bending.
Geometric imperfections were introduced into the numerical model by changing the nodal
coordinates affine to the buckling mode shapes obtained with the program CAST3M (CEA,
2012) and applying the interface RUBY (Couto et al., 2013). For the amplitude of the imper-
fections, it was considered 80% of b/50 for outstand elements and 80% of b/100 for internal
elements, following the recommendations of EN 1090-2 (CEN, 2011). Plates of the stainless
steel grade 1.4301, 1.4003 and 1.4462 (CEN, 2006a) subjected to four temperatures were con-
sidered 350ºC, 400ºC, 450ºC and 500ºC (common critical temperatures in slender sections).
The nominal values applied of yield strength, ultimate strength and elastic modulus of stain-
less steel in the numerical models are presented in Table 3.
These mechanical properties are reduced at elevated temperatures as presented in Figure 2,
which vary for each grade.

3.2 Plates subjected to compression


The results obtained for outstand and internal plate elements subjected to compression are
here presented. At elevated temperatures, the equation to determine the reduction factor has
to be adapted due to the transition that occurs from Class 3 to Class 4 section because of the
change on the limit strength, leading to a discontinuity in the curve, as presented in Equa-
tion 8.

109
Figure 1. Deformed shapes (x5): a) outstand element subjected to compression; b) internal element sub-
jected to compression; c) internal element subjected to bending.

Table 3. Nominal values for different stainless steel grades (CEN, 2006a).
Type Grade Yield strength fy (MPa) Ultimate strength fu (MPa) Elastic modulus E (GPa)

Austenitic 1.4301 210 520 200


Ferritic 1.4003 280 450 220
Duplex 1.4462 460 660 200

Figure 2. Mechanical properties reduction at elevated temperatures (CEN, 2005b): a) yield strength
retention; b) young modulus reduction.

N c;Rd fy;θ
ρθ ¼ ¼ρ ð8Þ
N Rd f2;θ

110
Figure 3. Results for a) outstand elements and b) internal elements subjected to compression for austen-
itic stainless steel at elevated temperatures.

Figures 3 to 5 presents the comparisons between the ultimate load bearing capacities, for
outstand and internal plate elements subjected to compression, obtained with EC3, the new
proposal for Class 4 carbon steel elements (“CS New Proposal” in the chart) and SAFIR.
For both outstand and internal elements subjected to compression, the proposal for carbon
steel Class 4 sections (Couto et al., 2014) eliminates the un-conservative nature given by the
plateau of EC3 for austenitic (Figure 3) and ferritic stainless steel (Figure 4). The results for
austenitic-ferritic stainless steel (Figure 5) revealed that the rules are over conservatives. None-
theless, the results highlight the need of improved design equations specifically developed for
stainless steel plates subjected to compression at elevated temperatures, considering the stain-
less steel grade.

Figure 4. Results for a) outstand elements and b) internal elements subjected to compression for ferritic
stainless steel at elevated temperatures.

Figure 5. Results for a) outstand elements and b) internal elements subjected to compression for
austenitic-ferritic (Duplex) stainless steel at elevated temperatures.

111
3.3 Plates subjected to bending
The obtained results for internal plate elements subjected to bending are presented in Figure 6.
The ultimate bending moments obtained in each plate for all methods were divided by the plas-
tic bending moments.
The different plateaus, in this figure, observed in Eurocode procedure of EN 1993-1-4
correspond to the transitions between Class 2 and Class 3 sections (from plastic to elas-
tic resistance) and from Class 3 to Class 4 where at elevated temperatures the yield
strength changes, as mentioned before. The curves from both proposals are over conser-
vative when compared with the numerical results, which leads to conclude that specific
formulae for stainless steel plates should be developed, which can be observed specific-
ally for austenitic-ferritic stainless steel.

3.4 Statistical analysis


The average value (μ) and the standard deviation (s) are important values to take into
account in the statistical analysis for the different methodologies of EC3 and different
Proposals. For each stainless steel grade and for each analysed curve it is possible to
evaluate the ratio between the analytical value and the corresponding SAFIR (Figures 7
and 8).
From Figures 7 and 8 and Table 4 it is possible to observe that the different design
rules are not adapted for stainless steel thin plates with low average values and high
standard deviations. The number of unsafe results is also relevant for the accuracy of
these design rules.

Figure 6. Results for internal elements subjected to bending for a) austenitic, b) ferritic and c)
austenitic-ferritic (Duplex) stainless steel at elevated temperatures.

112
Figure 7. Comparison between EN 1993-1-4 and SAFIR – Grades a) 1.4301, b) 1.4003 and c) 1.4462.

Figure 8. Comparison between CS New Proposal and SAFIR – Grades a) 1.4301, b) 1.4003, c) 1.4462.

Table 4. Statistical evaluation for different stainless steel grades at high temperatures
No. of Average Standard
Steel Grade Design rule simulations (n) value (μ) deviation (s) % Unsafe

EN 1993-1-4 283 0.8916 0.1564 14.1%


1.4301
CS New Proposal 283 0.7568 0.1798 0.0%
EN 1993-1-4 302 0.9025 0.1029 17.2%
1.4003
CS New Proposal 302 0.7391 0.1915 1.0%
EN 1993-1-4 276 0.8223 0.1283 1.4%
1.4462 CS New Proposal 276 0.7576 0.1262 1.4%

4 CONCLUSIONS

This work presented a numerical study regarding the plates’ behaviour of different stainless
steel grades (austenitic, ferritic and austenitic-ferritic (Duplex) stainless steel), composing the
cross-sections of members in fire situation.
In order to better understand the behaviour of these stainless steel sections, thin plates at
elevated temperatures were analysed. This study, on compressed outstand elements, com-
pressed internal elements and internal elements subjected to bending, concluded that EC3
does not provide accurate and safe approximations to their numerically obtained counterparts
regarding the ultimate load bearing capacities. Following this conclusion, a recent proposal
for carbon steel plates (Couto et al., 2014) was also investigated. It was observed that using

113
this proposal allowed to overcome the unsafety that was previously observed for Class 3 sec-
tions, but results remained too conservative for Class 4 sections.
In summary, the prediction of the resistance of stainless steel members for the case of fire is
still not completely understood, thus motivating and justifying the development of more stud-
ies with the objective to achieve more precise and safe formulations for these members.

ACKNOWLEDGEMENTS

This research work was performed within the framework of the project “Fire design of stainless
steel members” - StaSteFi - POCI-01-0145-FEDER-030655, supported by the Operational Pro-
gram “Competividade e Internacionalização”, in its FEDER/FNR component, and the Portu-
guese Foundation for Science and Technology (FCT), in its State Budget component (OE).

REFERENCES

CEA. 2012. CAST 3M research FEM environment. development sponsored by the French Atomic
Energy Commission http://www-cast3m.cea.fr/.
CEN European Committee for Standardisation. 2005a. EN 1993–1–1, Eurocode 3: Design of steel Struc-
tures – Part 1–1: General rules and rules for buildings. Belgium.
CEN European Committee for Standardisation. 2005b. EN 1993–1–2, Eurocode 3, Design of Steel Struc-
tures – Part 1–2: General rules – Structural fire design. Belgium.
CEN European Committee for Standardisation. 2006a. EN 1993–1–4, Eurocode 3: Design of steel Struc-
tures – Part 1–4: General rules – Supplementary Rules for Stainless steels. Belgium.
CEN European Committee for Standardisation. 2006b. EN 1993–1–5, Eurocode 3: Design of steel Struc-
tures – Part 1–5: Plated structural elements. Belgium.
CEN European Committee for Standardisation. 2011. EN 1090–2, Technical requirements for the execu-
tion of steel structures. Belgium.
Couto, C., Vila Real, P. & Lopes, N. 2013. RUBY an interface software for running a buckling analysis
of SAFIR models using Cast3M. University of Aveiro.
Couto, C., Vila Real, P., Lopes, N. & Zhao, B. 2014. Effective width method to account for the local
buckling of steel thin plates at elevated temperatures. Thin Walled Structures, 84, 134–149.
Couto, C., Vila Real, P., Lopes, N. & Zhao, B. 2015. Resistance of steel cross-sections with local buckling
at elevated temperatures. Journal of Constructional Steel Research, 109, pp. 101–114.
Euro Inox, SCI, Steel Construction Institute. 2006. Design Manual for Structural Stainless Steel. 3rd ed.
FIDESC4. 2014. Fire Design of Steel Members with Welded or Hot-Rolled Class 4 Cross-Section.
RFCS-CT-2011-2014, Technical Report No. 5.
Franssen, J-M. & Vila Real, P. 2015. Fire Design of Steel Structures. ECCS; Ernst & Sohn, a Wiley Com-
pany, 2nd edition.
Franssen, J-M. & Gernay, T. 2017. Modelling structures in fire with SAFIR®: theoretical background
and capabilities. Journal of Structural Fire Engineering.
Knobloch, M., & Fontana, M. 2006. Strain-based approach to local buckling of steel sections subjected
to fire. Journal of Constructional Steel Research, 62(1–2), 44–67.
Gardner, L. 2005. The use of stainless steel in structures. Progress in Structural Engineering and Mater-
ials, vol 7, pp 45–55.
Maraveas, C., Gernay, T. & Franssen, J-M. 2017. Amplitude of local imperfections for the analysis of
thin-walled steel members at elevated temperatures, Applications of Structural Fire Engineering
(ASFE’17), Manchester, UK.
Quiel, S.E., & Garlock, M.E.M. 2010. Calculating the buckling strength of steel plates exposed to fire.
Thin-Walled Structures, 48(9), 684–695.

114
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Numerical modelling of cold-formed steel members at elevated


temperatures

F. Arrais, N. Lopes & P.Vila Real


RISCO — Civil Engineering Department, University of Aveiro, Portugal

M. Jandera
Czech Technical University, Prague, Czech Republic

ABSTRACT: Steel structural elements composed of cold-formed thin-walled sections have


a high susceptibility to the occurrence of different buckling phenomena, particularly at ele-
vated temperatures, such as local, global and distortional buckling. Due to the high costs asso-
ciated with fire experimental tests, the evaluation of those cold-formed steel members’
resistance and behaviour has mainly been carried out through numerical studies, which should
be duly validated. Hence, this work presents numerical simulations of experimental works per-
formed by different authors, considering cold-formed simply supported beams and columns
under fire situation, contributing to the establishment of corroborated procedures.

1 INTRODUCTION

The cold-formed steel profiles can be applied to almost all existing buildings typologies. Cur-
rently, cold-formed profiles are commonly used in buildings due to its lightness and ability to
support large spans, being quite common as roof or walls support elements (Silvestre and
Camotim, 2010a). Besides the mentioned characteristics, other advantages are provided such
as: the high strength and stiffness; the faster manufacturing process and with relatively light
loads; the easy prefabrication and mass production; the favourable strength-to-weight ratios;
the cross-section’s shape allowing the compact packaging; the economy in transportation,
handling and the sustainability on construction (Yu, 2000; Silvestre et al., 2013).
On one hand, compared with other materials, the combination of previous qualities can
result in cost saving in construction. On the other hand, structural steel elements with thin-
walled sections are characterized by being subjected to the possibility of different failure
modes occurrence such as local, distortional and global (for example, lateral-torsional buck-
ling on elements under bending and flexural buckling on elements under compression).
The instability phenomena and its influence on the ultimate strength at normal temperature
have been widely studied (Gonçalves and Camotim, 2007; Silvestre and Camotim, 2010b), but
the corresponded behaviour under fire still requires further research.
Thus, cold-formed steel profiles behaviour in fire has recently received more attention
(Laím et al., 2015, Arrais et al., 2016, Muftah et al., 2016). In fact, the fire resistance evalu-
ation of cold-formed profiles may have a major role in the design of these elements. According
to Sidey and Teague (1988), the cold-formed steel strength can suffer a reduction of 10 to 20%
higher when compared to hot rolled profiles due to the metallurgical composition. The thin
walls of these members, together with the steel’s high thermal conductivity, are the reason for
the great loss of strength and stiffness of these structural elements in fire situation.
Due to the high costs, there is still a small number of fire experimental tests on cold-formed
steel profiles, and the evaluation of their resistance has mainly been carried out through
numerical studies, which should be duly validated.

115
This work aims at contributing to the validation of the FEM numerical models, which is
made against different experimental works performed by different authors, analysing dif-
ferent parameters that can be important on the numerical models conception, such as con-
stitutive law at elevated temperatures and geometrical imperfections. Hence, experimental
tests to members with lipped-channel (C), sigma (Σ) and zed (Z) cross-sections, from the
literature, are here numerically modelled and analysed. Laterally unrestrained beams
under bending moment about the strong axis (subjected to lateral torsional buckling) and
axially compression columns, taken into account the buckling around the minor axis, are
studied.

2 NUMERICAL MODELLING

In this section, the main and general parameters considered for the numerical studies pre-
sented in this paper are described. Table 1 shows the chosen case studies, which were obtained
from experimental fire tests found in the literature.
In the finite element model, rectangular shell finite elements with four notes and six degrees
of freedom (3 translations and 3 rotations) were used to reproduce possible local buckling
phenomena, due to the walls high slenderness. The numerical analysis is performed using the
software SAFIR (Franssen and Gernay, 2017).
Local, distortional and global instability modes obtained in CAST3M (CEA, 2012) were
used to define the geometrical imperfection shapes by applying the interface with SAFIR,
RUBY (developed at the University of Aveiro, in Portugal) (Couto et al., 2013).
According to Annex E of Part 1–2 of EC3, the steel stress-strain relationships at elevated
temperatures of thin-walled cold-formed profiles should be the same as the one proposed for
hot rolled sections (CEN, 2005b). However, the French National (FN) Annex of the same
Part 1–2 of EC3 (CEN, 2007) proposes different reduction factors for the steel constitutive
laws at elevated temperatures for cold-formed profiles. The FN Annex proposes lower values
for these reduction factors, of the yield strength and Young’s modulus at high temperatures,
for cold-formed profiles. The comparison of the different constitutive laws of Annex E and
FN Annex from Part 1–2 of EC3 is presented in Figure 1. This last constitutive law was the
one considered in the present research work for presenting better approximations to the
experimental behaviour.
The elevated temperatures were considered uniform throughout the cross-section due to
the reduced thickness of the cross-sections’ walls. In the absence of information from the
authors, the geometric imperfections adopted in all models were in accordance with
EN 1090–2 (CEN, 2011). The values correspond to 80% of the geometric fabrication toler-
ances, following the recommendations from Annex C of EN1993–1–5 (CEN, 2006b), and
described in section D.1 from Annex D of EN 1090–2+A1. The initial imperfections were
combined considering EN1993–1–5 recommendations. This part states that in combining
imperfections, a leading imperfection should be chosen and the accompanying imperfec-
tions may have their value reduced to 70%. The leading imperfection was chosen in
function of the achieved lower resistances. Residual stresses and corner enhancement were
not considered.

Table 1. List of analysed experimental tests.


Element type Cross-section Reference

Columns C Mesquita et al. (2014)


Beams Σ Laím et al. (2016)
C Laím et al. (2016)
Z Jandera and Schwarz (2014)

116
Figure 1. Constitutive law for elevated temperatures according to EC3.

3 RESULTS AND DISCUSSION

3.1 Axially compressed columns


The research work by Mesquita et al. (2014) presents an experimental study on simply sup-
ported cold-formed steel columns with lipped-channel cross-sections, comparing obtained fire
resistance with Eurocode simplified models.
From de different cases analysed by Mesquita et al. (2014), the lipped-channel element
C_150×51×20×(1.5), reference C17 under ISO 834 fire curve and 1000 mm length, was con-
sidered. A transient condition was applied, the procedure of the fire test starts with the increasing
of the load until a given load level, being afterwards kept constant during the fire action applied
according to the ISO 834 fire curve. Other experimental data such as imperfections amplitude and
mechanical properties can be found in Mesquita et al. (2014). The same procedure was applied on
the numerical model considering the FN Annex constitutive law. The result from both experimen-
tal and numerical analyses are presented on Table 2 and in Figure 2. Figure 2 also shows the
numerical model and the respective deformed shape at the collapse instance.
The collapse temperature corresponds to the value obtained since the furnace was switched
on until the collapse occurrence. From Table 2 and Figure 2, it is possible to observe differ-
ences of almost 12%, which is a reasonable approximation of the numerical model to the
experimental results, providing also reasonable similar behaviour during the fire test.

3.2 Beams
Laím et al. (2016) presented the results of an experimental study about cold-formed steel
beams with different cross-section under fire conditions. Different cases of this research work,
C_250 × 43 × 15 × (2.5) and Σ_255 × 70 × 25 × (2.5) with 3000 mm length, were considered.
Again, the test was performed under transient test conditions. The maximum load applied was
50% of the design value of the load-bearing capacity of the element at normal temperature.
When reached the required load, the element was subjected to the ISO 834 fire curve. Other
experimental data such as imperfections amplitude and mechanical properties can be found in
Laím et al. (2016).

Table 2. Experimental vs. Numerical results for C_150×51×20×(1.5).


Experimental Numerical Difference

Collapse Temperature 484 °C 429 °C 11.4%

117
Figure 2. a) Shortening–time chart experimental and numerically obtained for the C_150×51×20×(1.5)
model b) before and c) after the SAFIR analysis at high temperatures.

The temperatures obtained from the real fire test were applied on the numerical analysis.
Again, the constitutive law from the FN Annex of EC3 Part 1-2 was used.
The results from both analyses, experimental and numerical, and for the different models,
are presented in Table 3, and Figure 3 presents the displacement-time chart for lipped-channel
and sigma cross-sections with the experimental and the numerical analysis.

Table 3. Experimental vs. Numerical results for high temperatures.


Collapse temperatures
Section type Experimental Numerical Differences

C_250×43×15×(2.5) 718 ºC 709 ºC 1.3 %


Σ_255×70×25×(2.5) 681 ºC 686 ºC 0.8 %

Figure 3. a) Lipped-channel section and b) Sigma section displacement–time chart (experimental and
numerical) at high temperatures.

118
From Table 3 and Figure 3, it is possible to observe a good agreement between both ana-
lysis with differences less than 2%, with similar element behaviour during the fire test as
shown in Figure 4 and 5.
The sigma element fire test demonstrated to be in a better agreement with the numerical
results and behaviour.
Finally, experimental test on cold-formed Z purlins, performed in the Fire testing labora-
tory PAVUS by the Czech Technical University in Prague, in a horizontal furnace is here
modelled. The tested profile had a Z section with a height of 200 mm and thickness of
1.5 mm. Different Z profiles were spanned for 6 m, inside the furnace, with an overhanging
part (cantilever) of 2.5 mm on one side, outside the furnace (see Figure 6).
The aim of the overhanging part (cantilever) was to simulate an internal support of
a continuous beam. A sleeve system was adopted in this connection. Inside the furnace one
element was used for the span, outside the furnace for the cantilevered part, and for the

Figure 4. a) Numerical and b) experimental (Laím et al., 2016) C-beam after the analysis and the test.

Figure 5. a) Numerical and b) experimental (Laím et al., 2016) Σ-beam after the analysis and the test.

Figure 6. Template of the tested Z purlin (Jandera and Schwarz, 2014).

119
connection a thicker element (Z purlin height 200 mm and thickness of 2.0 mm) was intercon-
necting both elements at the hogging moment area. All sections are S350GD steel grade. In
the experimental test the connections were insulated.
The loads were chosen in order to represent a realistic example of a roof structural element
(purlins), based on the ultimate limit state combination, at normal temperatures. The dead
weight of a sandwich panel and the snow load representing more than 90% use of the element
in normal conditions were considered. This led to a load of 0.5 kN/m for the member inside
the furnace, imposed by attaching it to the web in each 1 m of the span length, and a 1 kN at
the end of the element outside the furnace was also imposed, at the top of the member.
Comparisons between the standard curve and the measured temperature inside the furnace
are presented on Figure 7. The element temperature was measured with welded thermocouples
in three parts (upper flange, web and lower flange) of two different cross-sections (mid-span
and at 500 mm from the support). The average temperatures at mid-span and at the support
are presented on Figure 7.
According to EN 1363–1 (CEN, 2012), for flexural loaded elements, the following criteria
are presented (Table 4) where L is the span of the element and d is the distance from the
extreme fibre of the cold design compression zone to the extreme fibre of the cold design ten-
sion zone of the structural section.
However, the rate of the deflection criteria is not applied in the first 10 min of the fire test,
since relatively fast deflection can occur until stable conditions are reached.
Global geometric imperfections were taken into account in the numerical model. Local and
distortional imperfections where not taken into account due to the long length of the element.
In these long beams local and distortional imperfections have much lower influence on the
results, as shown in Arrais (2016).
The temperature considered in the Z purlin on the SAFIR numerical model was the one
obtained on the experimental test (Figure 7). The constitutive law applied was the one pro-
posed by the FN Annex of Part 1–2 of EC3. The numerical model with all the previous con-
siderations is presented on Figure 8 where it is possible to compare the design drawing with
the final numerical model adopted for the present research work.

Figure 7. Comparison between the ISO 834 curve and the gas temperature, and element temperature.

Table 4. Criteria for flexural loaded elements (CEN, 2012) and respective values
for the present fire test.
Limiting deflection Limiting rate of deflection

L2 60002 dD L2 60002
D¼ ¼ ¼ 450 mm ¼ ¼ ¼ 20 mm  min1
400d 400  200 dt 9000d 9000  200

120
Figure 8. Numerical model from the Z purlin adopted for the SAFIR numerical model analysis.

Figure 9. Criteria for flexural loaded elements reached in experimental and numerical tests (mid-span).

The obtained results from the experimental fire test and from the numerical analysis, con-
sidering the global imperfections, the FN Annex constitutive law and the loads applied on
plates (Figure 8), are presented in Figure 9.
From Figure 9 it is possible to observe the evolution of the deflection from the Z purlin, in
function of time, at mid-span. The first limiting criterion to be reached was the limiting rate of
deflection, after 17 min by the experimental fire test and after 22 min by the numerical ana-
lysis. As mentioned before, according to EN 1363–1, the rate of the deflection criterion is not
applied in the first 10 min of the fire test, since relatively fast deflection can occur until stable
conditions are reached as observed at the minute 6 of the fire resistance test. The second limit-
ing criterion was reached by the fire test after 42 min and by the numerical analysis after
59 min. After 15 min it is possible to observe the lateral-torsional phenomenon occurring
before the first criterion reached. Around the 20 min of fire test and 22 min of numerical ana-
lysis the phenomenon of catenary action is starting to develop. The Z purlin behaviour is simi-
lar to a tensile element (catenary action).
The end of the experimental test and numerical analysis was considered after the catenary
action occurs for a high deflection as observed. In Figure 10 and 11, it is visible the deformed
shape of the entire element after the fire test and the numerical analysis, and the detailed buck-
ling phenomenon observed next to the support. Considering the present numerical model it is
visible a good agreement between the numerical analysis and the experimental test, in terms of
resistance and overall element behaviour.

121
Figure 10. a) Experimental (Jandera and Schwarz, 2014) and b) numerical purlin after the test.

Figure 11. a) Experimental (Jandera and Schwarz, 2014) and b) numerical purlin support after the test.

4 CONCLUSIONS

In this work it is presented the numerical modelling of experimental tests to cold-formed steel
profiles under fire exposure, from the literature.
The consideration of initial geometric imperfections with the shapes of the critical buckling
modes and the constitutive law at elevated temperatures proposed in FN Annex of Part 1–2 of
EC3, resulted on reasonable approximations to the experimental results.
In general, from the analysis, and in spite of the possible occurrence of different complex
instabilities phenomena in cold-formed steel profiles (such as local, distortional and global
buckling), a good agreement between the numerical analysis and the experimental tests, in
terms of resistance capacities and structural behaviour in fire, could be observed.

REFERENCES

Arrais, F., Lopes, N., Vila Real, P. 2016. Behaviour and resistance of cold-formed steel beams with
lipped channel sections under fire conditions, Journal of Structural Fire Engineering, Emerald Group
Publishing Ltd, ISSN: 2040–2317, volume 7/4, pp. 365–387.
CEA. 2012. CAST 3M research FEM environment. development sponsored by the French Atomic
Energy Commission http://www-cast3m.cea.fr/.
CEN European Committee for Standardisation. 2005a. EN 1993–1–1, Eurocode 3: Design of steel Struc-
tures – Part 1–1: General rules and rules for buildings. Belgium.
CEN European Committee for Standardisation. 2005b. EN 1993–1–2, Eurocode 3, Design of Steel Struc-
tures – Part 1–2: General rules – Structural fire design. Belgium.
CEN, European Committee for Standardisation. 2006a. EN 1993–1–3, Eurocode 3: Design of Steel Struc-
tures – Part 1–3: General rules - Supplementary rules for cold-formed members and sheeting. Belgium.
CEN European Committee for Standardisation. 2006b. EN 1993–1–5, Eurocode 3: Design of steel Struc-
tures – Part 1–5: Plated structural elements. Belgium.

122
CEN, European Committee for Standardisation. 2007. NF EN 1993–1–2, Eurocode 3: Calcul des struc-
tures en acier – Annexe Nationale à la NF EN 1993–1–2: Calcul du comportment au feu. Belgium.
CEN European Committee for Standardisation. 2011. EN 1090–2, Technical requirements for the execu-
tion of steel structures. Belgium.
CEN, European Committee for Standardisation. 2012. NP EN 1363-1, Ensaios de resistência ao fogo -
Parte 1: Requisitos gerais, Belgium.
Couto, C., Vila Real, P. & Lopes, N. 2013. RUBY an interface software for running a buckling analysis
of SAFIR models using Cast3M. University of Aveiro.
Franssen, J-M. & Gernay, T. 2017. Modelling structures in fire with SAFIR®: theoretical background
and capabilities. Journal of Structural Fire Engineering.
Gonçalves, R.; Camotim, D. 2007. Thin-walled member plastic bifurcation analysis using generalised
beam theory, Advances in Engineering Software. Elsevier. Vol. 38, n.º 8–9, p. 637–646.
Jandera, M., Schwarz, I. 2014. Structural fire behaviour of Z purlins: Eurosteel 2014: 7th European Con-
ference on Steel and Composite Structures, Naples, Italy.
Laím, L., Rodrigues, J.P.C., Craveiro, H.D. 2015. Flexural behaviour of beams made of cold-formed
steel sigma-shaped sections at ambient and fire conditions, Thin-Walled Structures. Vol. 87, p. 53–65.
Laím, L., Rodrigues, J.P.C., Craveiro, H.D. 2016. Flexural behaviour of axially and rotationally restrained
cold-formed steel beams subjected to fire, Thin-Walled Structures. Vol. 98, Part A, p. 39–47.
Mesquita, L., Mendonça, M., Ramos, R., Barreira, L., Piloto, P. 2014. Thermomechanical analysis of
cold-formed steel sections: 9º Congresso Nacional de Mecânica Experimental, Aveiro, Portugal.
Muftah, F., Sani, M., Mohammad, S., Ngian, S., Tahir, M. 2016. Temperature development of
cold-formed steel column channel section under standard fire, AIP Conf. Proceedings. Vol. 1774, n.º 1.
Sidey, Teague. 1988. Elevated temperature data for structural grades of Galvanised steel: British Steel,
Welsh Laboratories.
Silvestre, N.; Camotim, D. 2010a. Construção em aço leve, Revista da Associação Portuguesa de Con-
strução Metálica e Mista, ano 11, n.º 20, Março.
Silvestre, N.; Camotim, D. 2010b. On the mechanics of distortion in thin-walled open sections, Thin-
Walled Structures - Elsevier. Vol. 48, n.º 7, p. 469–481.
Silvestre, N.; Pires, J.; Santos, A. 2013. Manual de Conceção de Estruturas e Edifícios em LSF - Light
Steel Framing: Associação Portuguesa de Construção Metálica e Mista. ISBN 9789899560581.
Yu, W. W. 2000. Cold-Formed Steel Design - 3rd Edition: John Wiley & Sons.

123
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Experimental study on the general behaviour of stainless


steel frames

I. Arrayago, E. Real, E. Mirambell & I. González de León


Department of Civil and Environmental Engineering, Universitat Politècnica de Catalunya, Barcelona,
Spain

ABSTRACT: In recent years, a considerable am ount of research has been devoted to the
understanding of the structural performance of single isolated stainless steel members. Not-
withstanding, advances related to the analysis of more complex stainless steel structures, such
as frames, are scarce. On this basis, a comprehensive experimental programme on sway and
non-sway austenitic stainless steel frames with slender and stocky rectangular hollow sections
subjected to static loading was carried out at Universitat Politècnica de Catalunya. The paper
presents the preliminary experimental results on one non-sway and one sway frame as illustra-
tive of the experimental programme. A more in deep analysis of these frames will contribute
to the appraisal of existing design rules for stainless steel, which are based on carbon steel, in
terms of predicted ultimate capacities, plastic design and second order effects, considering the
particular non-linear behavior and strain hardening for stainless steels.

1 INTRODUCTION

Mechanical properties such as high ductility, adequate toughness, considerable strain harden-
ing and good fire resistance make stainless steel an excellent construction material for struc-
tures, especially those required to withstand accidental loading (Baddoo, 2008). During last
decades, significant amount of research has been focused to the understanding of the struc-
tural performance of single isolated stainless steel members recently (Arrayago & Real, 2016;
Afshan & Gardner, 2013; Arrayago et al., 2016; Huang & Young, 2013), but advances related
to the analysis of more complex stainless steel structures, such as frames, are scarce (Arrayago
et al., 2017; Walport et al., 2019). As a matter of fact, EN1993-1-4+A1 (2015) does not estab-
lish specific design rules associated with the global analysis of stainless steel frames and hence,
provisions given for carbon steel in EN 1993-1-1 (2005) need to be adopted.
Recent research on the behaviour of stainless steel frames by Walport et al., 2019 demon-
strated that the degradation of stiffness – due to the nonlinear material response – consider-
ably affects the characteristics of the structural system, causing greater deformations and
increasing second order effects. This has a direct influence on the definition of non-sway –
structures for which second order effects are negligible – and sway – structures for which
internal force amplifications are relevant – frames. Thus, it was recommended that material
nonlinearity should be considered in the global analysis of stainless steel frames, especially
when the susceptibility to second order effects (through the αcr critical load factor) is con-
sidered. Moreover, the lack of guidance on plastic design in general and design of frames in
particular is an obstacle to the optimal design of stainless steel structures considering the
remarkable differences in this behaviour compared with carbon steel.
On this basis, an extensive experimental programme on sway and non-sway austenitic stain-
less steel frames with slender and stocky Rectangular Hollow Sections (RHS) subjected to
static loading has been conducted at the laboratory of the Department of Civil and Environ-
mental Engineering “Luis Agulló” of the Universitat Politècnica de Catalunya. The experi-
mental programme comprises several sub-programmes in which the performance of these

124
structures has been investigated at different levels – material characterization, cross-sections,
members and frames (Arrayago et al., 2019a; Arrayago et al., 2019b). The final objective of
this experimental programme and subsequent research is to assess the currently existing design
rules, in terms of predicted ultimate capacities and second order effects, for stainless steel
structures.
In this context, this paper consists in a brief resume of the adopted experimental set-up for
frame tests, as well as presenting the preliminary results on two of the tested austenitic stain-
less steel frames –including one non-sway and one sway frames. The comprehensive analysis
of the tests through the measured data will be carefully analyzed in detail in the future.

2 FRAME DESCRIPTION

Frame tests have been performed on single span and single height austenitic stainless steel
frames with RHS, with a mid-section height (h) of 2 m and a span between columns (L) equal
to 4 m. The connection between the beam and the columns were performed by welding an
auxiliary steel plate with an inclination of 45º. Likewise, for the connections at supports, add-
itional steel plates, which were provided with perforations to be screwed, were welded at the
bottom of the columns. Hence, both fixed and pinned boundary conditions were allowed with
the same general configuration.
The experimental programme included a total of four frame tests from EN 1.4301 austenitic
stainless steel with the same general geometry, but with different cross-sections. Table 1 sum-
marizes the general definition of frame specimens, in which the overall geometries, the bound-
ary conditions and cross-section shapes are reported. H corresponds to the height of the RHS,
while B is the width and t is the wall thickness.
In addition, αcr parameter values, which indicate the susceptibility of the frames to second
order effects are also reported, along with the estimated column slenderness λc , based on the
effective length calculations for sway and non-sway frames. Likewise, the local slenderness of
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
the cross-sections λp ¼ σ0:2 =σcr;l is reported, with σ0:2 and σcr;l standing for the 0.2% proof
stress and the local critical buckling stress, respectively. Note that, since according to EN1993-
1-4+A1 (2015) local buckling effects appear beyond a slenderness value of λp ¼ 0:65, Frames 1
and 2 correspond to stocky cross-sections in pure compression, while Frames 3 and 4 corres-
pond to slender cross-sections. A preliminary finite numerical model was conducted by means
for the advanced software ABAQUS and the critical buckling behavior of the frames was pre-
dicted according to the defined loading scheme. This preliminary analysis showed that Frame 1
and Frame 2 can be considered as non-sway frames, while Frames 3 and 4 are sway frames.
Although the experimental programme comprised four different frame tests, in this paper
only results corresponding to Frame 1 –a non-sway frame with a stocky sections- and Frame
3–a sway frame with a considerably slender section- are presented.

2.1 Preliminary tests and measurements: Material and cross-section response, imperfections
For a correct analysis of the frame tests, several preliminary tests for the material character-
ization and for the determination of the cross-section and member resistances were conducted

Table 1. General definition of frame specimens (nominal properties).


Boundary Column Local
H B t
h [mm] L[mm] conditions slenderness slenderness
λc [mm] [mm] [mm] λp
(supports) αcr

Frame 1 – S1 2000 4000 Fix-end 11.7 0.60 120 80 6 0.50


Frame 2 – S2 2000 4000 Fix-end 11.8 0.60 100 80 4 0.60
Frame 3 – S3 2000 4000 Pin-end 3.4 2.53 120 40 4 0.77
Frame 4 –S4 2000 4000 Pin-end 7.6 1.15 200 100 3 1.64

125
previously. Since a more in detail description of such tests is available in Arrayago et al.,
2019a, only a brief summary of the most relevant parameters is provided in this paper.
The stress-strain behavior of coupons extracted from flat (F) and corner (Co) sections of
the RHS was determined by means of tensile tests. Key material parameters for sections S1
and S3 (comprising Frame 1 and Frame 3) are reported in Table 2, in which E is the Young’s
modulus, α0.2 is the yield strength (proof stress for 0.2% plastic strain), αu and εu are the tensile
strength and corresponding strain, and n, m are the optimized strain hardening parameters.
Likewise, Table 3 summarizes the cross-section resistances to compression Nu and to
bending Mu corresponding to sections S1 and S3, obtained from stub column and four-point
bending tests, together with the cross-section classifications according to EN1993-1-4+A1
(2015).
The actual initial geometry of the frames was carefully measured prior to testing by
means of theodolites. The actual geometry of the columns was characterized through five
different points along their height, while the position of additional five points along the
beam length were also measured. This will allow introducing accurate initial imperfections
into the future numerical studies, as well as evaluating the influence of local and global
initial imperfections on the general behavior of the frames. In addition, several points of
the frames were monitored during the tests and the movements were recorded using
a Lidar system (see Figure 1).

Table 2. Key material characterization parameters from tensile coupon tests.


Specimen E [MPa] α0.2 [MPa] αu [MPa] εu [mm/mm] n m

S1-F 159642 495 715 0.50 7.3 2.6


S3-F 210966 601 769 0.29 6.2 3.9
S1-Co 187795 643 840 0.38 4.7 6.8
S3-Co 186153 637 856 0.20 4.5 6.1

Table 3. Key experimental results for stub column tests and 4 point bending tests.
Cross-section class Cross-section class
Specimen in compression Nu [kN] in bending Mu [kNm]

S1 1 1197.6 1 57.6
S3 4 552.3 1 26.1

Figure 1. General view of theodolite and Lidar system.

126
3 FRAME TEST SET-UP

In this paper the preliminary results of the experimental programme on austenitic stainless steel
frames subject to static loading are presented. The loading scheme was carefully defined after an
exhaustive study of different alternatives and options for vertical and horizontal loading schemes,
since sway displacements difficult vertical loads to remain vertical (Wilkinson & Hancock, 1999,
Blum & Rasmussen, 2018, Zhang et al., 2016, Avery & Mahendran, 2000).
The adopted final loading to allow reproducing vertical – gravity- and horizontal loads con-
sisted on a two-step equivalent loading scheme. During the first step, the frames were loaded
vertically through two jacks applied on the beams up to a load value corresponding to the 60-
70% of the maximum vertical resistance of the frames. In the second step, while the vertical
load was kept constant, a horizontal displacement was imposed into the column supports
through a horizontal jack and a rigid beam. During these two steps, the top-right corner of
the frames was tied to a reaction wall, restraining the in-plane horizontal movement of the
beam, and thus, allowing the proposed loading scheme to work. A more exhaustive study of
the loading options, adopted scheme and employed auxiliary elements is provided in Arrayago
et al., 2019b, although the most relevant aspects are summarized in the following section, as
well as the instrumentation considered during the tests.

3.1 General test layout


The layout of the general test arrangement is presented in Figure 2, in which the most relevant
parts are highlighted and described below: vertical loading sections, horizontal loading rigid
beam, load cells, column supports and fixed horizontal point.

3.1.1 Vertical loading sections


Vertical loads were introduced at a distance of 800 mm from each column through the use of
two jacks, which were connected to guarantee that two identical point loads and were applied
simultaneously while allowing different vertical displacements at these points once the hori-
zontal loads were applied. In order to prevent local web failure at loading supports, guarantee
a correct distribution of the loads and to contribute to the lateral stability of the frame, auxil-
iary elements provided by neoprene and Teflon plates and adjustable to the different cross-
sections were fabricated (see Figure 3).

3.1.2 Horizontal loading rigid beam and sliding supports


As explained previously, horizontal loads were introduced by imposing a displacement at
column supports once the vertical loads were applied. To ensure a uniform distribution

Figure 2. General layout of stainless steel frame tests.

127
Figure 3. Detail of the vertical loading section.

of displacement at both supports, a rigid beam was prepared. This beam was supported
on two specially fabricated 500 mm long sliding supports, which allow a frictionless
movement in-plane thanks to Teflon plates. In addition, the beam included a load cell
that measured the horizontal reactions at the left supports, which together with the total
applied load obtained from the horizontal jack, allowed a complete characterization of
both supports.

3.1.3 Load cells and column supports


Vertical load and bending moment reactions at supports were measured through two specially
fabricated load cells, which consisted in four steel studs instrumented with strain gauges, and
welded to two steel plates enabling the connection to the frame columns and the horizontal
rigid beam (see Figure 4). These cells were calibrated prior to using them in the tests, as
described in Arrayago et al., 2019b.
In order to guarantee both fixed and pinned boundary conditions for the tested frames,
a bolted connection through the columns and the load cells was adopted. While fix-ended
boundary conditions were achieved by using 12 bolts between the steel plates welded to the
columns and the load cell, for pin-ended conditions only four bolts were used.

3.1.4 Fixed horizontal point


To restrain horizontal displacement of the frames, the top of the right column was tied to
a reaction wall by two Macalloy bars, which also contributed to the lateral stability of the
frames (Figure 1).

3.1.5 Adopted loading protocols


The loading scheme adopted for the frame tests subject to static loading consisted of two load-
ing steps, as described previously. Different vertical and horizontal loading rates were defined
for each of the frame tests to ensure safety and a reasonable duration of the tests, which
showed an approximate duration of the tests of 90 minutes. Both loading steps were per-
formed under displacement control, although the hold in the vertical load during the horizon-
tal loading step was load-controlled. Table 4 reports the adopted test rates for Frame 1 and
Frame 3, as well as the values of the maximum applied total vertical loads.

128
Figure 4. Detail of the horizontal jack, loading cell and column supports.

Table 4. Adopted test rates for vertical and horizontal loading steps in frame tests.
Step 1: Vertical loading Step 2: Horizontal loading
(30 min approx.) (60 min approx.)
Test rate Fv,u Test rate
Proportion of Fv,max
Specimen [mm/min] [kN] [mm/min]

Frame 1 2.00 155.0 65% 3.3


Frame 3 1.27 50 60% 2.5

3.2 Instrumentation
In this sub-section the instrumentation adopted in the frame tests is described. The measure-
ment of different parameters is key for the accurate characterization of the performance of
stainless steel frames and for guaranteeing the correct development of the test (see Figure 1).
Actuators directly recorded the total applied vertical Fv,tot and horizontal Fh,tot loads,
as well as their displacements. However, additional deflection measurement devices were
adopted for the loading sections and the horizontal displacement of the rigid beam.
Moreover, two laser devices were placed on the top of both columns to measure possible
horizontal and out-of-plane displacements in an effort of tracking the adequacy and
safety of the conducted tests.
In addition, two inclinometers were placed to measure in-plane and out-of-plane rota-
tions close to the support sections at frame columns and the strain gauges of the load
cells provided the relevant information on the distribution of loads and bending moments
in both supports.
Strains at the most relevant sections of the frames were recorded by means of strain
gauges, which will be used to estimate stress distributions and bending moments at these
representative points. Finally, an alternative Digital Image Correlation (DIC) system was
also employed to record the behaviour of the upper area of the right column in all
frames, as these sections were identified as critical in the preliminary numerical analysis.

129
Nevertheless, these measurements are still being processed and thus no preliminary results
are available.

4 TEST RESULTS

This section presents the preliminary test results from the conducted frame tests. As examples,
results corresponding to one non-sway frame –Frame 1– and one sway frame –Frame 3– are
provided.
The recorded load-displacement paths for Frame 1 and Frame 3 are presented in Figures 5
and 6: while the responses corresponding to the vertical loading step are shown in Figure 5,
the horizontal behaviour of such frames is illustrated in Figure 6. Figures show that the non
linear behavior is more pronounced in the non-sway stocky frame (Frame 1).
Achieved ultimate load and deflections for each of the loading steps are summarized in
Table 5 for Frame 1 and Frame 3. Fv,u is the maximum vertical load applied and dv,u the verti-
cal displacement when the vertical load is stopped and starts horizontal loading. Fh,v is the

Figure 5. Vertical load-displacement.

Figure 6. Horizontal load-displacement.

130
Table 5. Summary of ultimate load and displacements.
Specimen Fv,u [kN] dv,u [mm] d*v,u [mm] F,h,u [kN] dh,u [mm]

Frame 1 157,7 71 81 39,3 83


Frame 3 56,9 48 55 18,4 82

Figure 7. Deformed frames after vertical and horizontal loading.

maximum horizontal loading and dh,u the maximum horizontal displacement. Notice that ver-
tical displacements also increase after the hold on the vertical loading and during the horizon-
tal loading (d*v,u).
Likewise, Figure 7 shows the final deformed configurations for Frame 1 and Frame 3 once
the tests were completed.

5 CONCLUSIONS

This paper presents a series of tests on sway and non-sway austenitic stainless steel Rectangu-
lar Hollow Sections frames as part of a vaster experimental programme devoted to the investi-
gation of the performance of stainless steel structures under static loading.
Experimental results corresponding to deflections of the beam caused by vertical loading
and displacement of the supports caused by horizontal loading are presented for one stocky
non-sway frame and for one slender sway frame.
Furthermore, experimental results highlighted the non linear behavior of the frames and the
more pronounced effect on the stocky non-sway frame.

131
Once the experimental programme on sway and non-sway frames is finalized, efforts will be
focused on analyzing in detail the obtained data to advance in the knowledge of stainless steel
behaviour, which is currently scarce in plastic performance and frame performance, and to
propose a specific guidance for the optimal design of stainless steel structures.

ACKNOWLEDGMENTS

This experimental programme was developed in the frame of the Project BIA2016-75678-R,
AEI/FEDER, UE “Comportamiento estructural de pórticos de acero inoxidable. Seguridad
estructural a acciones accidentales de sismo y fuego”, funded by MINECO (Spain). The last
author would also like to thank the financial support provided by the Spanish Ministry for
Science, Innovation and Universities for the deveolpemnt of her PhD.
Authors would like to acknowledge the support, advice and suggestions by Professor
K. Rasmussen from the University of Sydney, Professor Ben Young from the Hong Kong
University and Professor Leroy Gardner from Imperial College London in the preparation of
this experimental programme. Finally, the time and efforts from the staff at the Structures
Laboratory at the Universitat Politecnica de Catalunya is much appreciated.

REFERENCES

Afshan, S. & Gardner, L. 2013. Experimental study of cold-formed ferritic stainless steel hollow sections.
Journal of Structural Engineering (ASCE) 139(5), 717–728.
Arrayago, I. & Real, E. 2016. Experimental study on ferritic stainless steel simply supported and continu-
ous beams. Journal of Constructional Steel Research 119, 50–62.
Arrayago, I., Real, E. & Chacón, R. 2019a. Experimental performance of austenitic stainless steel beams
and columns. Proceedings of the 9th International Conference on Steel and Aluminium Structures
(ICSAS-19). Bradford, UK.
Arrayago, I., Real, E. & Mirambell, E. 2019b. Preliminary study and tests arrangements for experimental
programme on stainless steel frames. Proceedings of the 9th International Conference on Steel and
Aluminium Structures (ICSAS-19). Bradford, UK.
Arrayago, I., Real, E. & Mirambell, E. 2016. Experimental study on ferritic stainless steel RHS and SHS
beam-columns. Thin-Walled Structures 100, 93–104.
Arrayago, I., Real, E., Mirambell, E. and Chacón, R. 2017. Global plastic design of stainless steel
frames. Proceedings of the 8th European Conference on Steel and Composite Structures (Eurosteel
2017). Copenhagen, Denmark.
Avery, P. & Mahendran, M. 2000. Large-scale testing of steel frame structures comprising non-compact
sections. Engineering Structures 22, 920–936.
Baddoo, N.R. 2008. Stainless steel in construction: A review of research, applications, challenges and
opportunities. Journal of Constructional Steel Research 64(11),1199–1206.
Blum, H.B. and Rasmussen, K.J.R. 2018. Elastic buckling of columns with a discrete elastic torsional
restraint. Thin-Walled Structures 129, 502–511.
EN 1993-1-1:2005. Design of steel structures. Part 1-1: General rules – General rules and rules for build-
ings. European Committee for Standardization Eurocode 3. Brussels, Belgium.
EN 1993-1-4:2006 + A1:2015. Design of steel structures. Part 1–4: General rules. Supplementary rules for
stainless steels. European Committee for Standardization Eurocode 3. Brussels, Belgium.
Huang, Y. & Young, B. 2013. Tests of pin-ended cold-formed lean duplex stainless steel columns. Journal
of Constructional Steel Research 82, 203–215.
Walport, F., Gardner, L., Real, E., Arrayago, I. & Nethercot, D.A. 2019. Effects of material nonlinearity
on the global analysis and stability of stainless steel frames. Journal of Constructional Steel Research
152, 173–182.
Wilkinson, T. & Hancock, G.J. 1999. Tests of cold-formed rectangular hollow section portal frames.
Research Report No. R783. Sydney: The University of Sydney.
Young, B. & Rasmussen, K.J.R. 2003. Measurement techniques in the testing of thin-walled structural
members. Experimental Mechanics 43, 32–38.
Zhang, X., Rasmussen, K.J.R. & Zhang, H. 2016. Experimental investigation of locally and distortion-
ally buckled portal frames. Journal of Constructional Steel Research 122, 571–583.

132
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Experimental investigation of flexural buckling of sandwich panels


with steel facings

I. Balázs & J. Melcher


Institute of Metal and Timber Structures, Faculty of Civil Engineering, Brno University of Technology,
Brno, Czech Republic

ABSTRACT: Sandwich panels are widely used in building industry particularly as members of
cladding. Although they primarily resist transversal loads, axial forces may arise e.g. due to stabil-
izing function of panels that may prevent buckling of supporting members of steel load-bearing
structure and transfer stabilizing forces. The paper focuses on problem of flexural buckling of
sandwich panels with thin steel facings. The problem of stability of axially loaded sandwich
panels is outlined. To verify the actual behavior of sandwich panels under axial load, a series of
full-scale tests of flexural buckling of selected type of sandwich panels with thin steel facings and
soft core was performed. Evaluation of the test results was performed using Southwell plot and
statistical methods. The paper presents the utilized test setup, procedure of testing, failure modes
and summarizes selected results of the tests. The findings obtained from the tests are discussed.

1 INTRODUCTION

The sandwich panels are widely used in civil engineering as members of roof and wall cladding.
They usually consist of thin metal facings and soft insulation core (various types of foams in
most cases). It results from purpose of the panels that they are primarily loaded by transversal
loads. The panels may also prevent buckling of supporting members of load-bearing structure
(e.g. purlins) which results in stabilizing (axial) forces to be transferred by the panels. In some
cases of small buildings, the load-bearing structure itself may consist of sandwich panels (with
no substructure) which was investigated recently (Käpplein, Misiek, 2010). In that cases, the
sandwich panels should be designed to resist axial forces. The assessment of the load-bearing
resistance becomes complex especially in case of compressive axial force and related stability
problems resulting from slenderness of the member. For sandwich panels, various materials of
the facings and the core can be utilized. There is a wide range of types of these structural mem-
bers available for building industry. Experimental analysis can be an effective way of investiga-
tion of the actual behavior of axially loaded sandwich members made of specific materials.

2 BUCKLING RESISTANCE OF AXIALLY LOADED MEMBERS

2.1 Slender metal members


The behavior of slender members in compression was subject of number of theoretical and
experimental studies in the past and now is good understood. It is necessary to distinguish
between ideal member with no imperfections (structural, geometrical, constructional) and
actual member. The problem of buckling of an ideal member in compression was first studied
by Euler who established basis for mathematical concept of the theory of stability of slender
members. The buckling (bifurcation of equilibrium) of an ideal member occurs when the value
of the so called critical load Ncr is reached (Euler, 1744). The problem was defined using
homogenous differential equation of second order and the critical load was found using solu-
tion of the eigenvalue problem of the equation. The theory was significantly developed and

133
extended by Vlasov who formulated general theory of slender members (Vlasov, 1962). In
case of buckling of axially loaded members, the Vlasov’s theory takes into account also tor-
sional and warping stiffness of the member.
In case of actual member with unavoidable initial imperfections, the lateral deflection of the
member increases with load (as opposed to an ideal member) and the bifurcation in terms of
theory of stability of ideal member does not occur. The resistance of the member is therefore
lower than the magnitude of critical load. Nevertheless, the value of the critical load is an
important value necessary for evaluation of resistance of actual members. Within the frame of
analysis and structural design of an actual member in compression, all types of imperfections
are usually replaced by initial deflection (initial bow imperfection) of the member with the
amplitude e0. A simplified comparison between behavior of ideal and actual member in form
of relationship between load N and lateral deflection f is in Figure 1.

2.2 Sandwich panels


The actual behavior of sandwich members under axial loads is very complex due to possible
occurrence of stability problems (local buckling of thin facings and global buckling of the panel
and their combination) and due to significant difference of material properties of the facings
and the core. The theoretical problem of stability of sandwich structures was extensively devel-
oped on theoretical basis by Kovařík et al. (Kovařík, Šlapák, 1973). The work based on complex
mathematical solutions of eigenvalue problems provided, among others, formulas for evaluation
of global buckling of sandwich plates with rigid or soft core and also solution of local buckling
of sandwich plates (wrinkling). Mechanical properties of both facings and core of the sandwich
plate are taken into account within the calculation procedure. The theory of stability of sand-
wich members was later dealt with e.g. by Davies. Formula of the critical load of sandwich
member was developed and results of few tests of axially loaded panels (with eccentricities of
the applied normal force) were presented. The significant influence of the eccentricity of the
applied load on the test results was highlighted (Davies, 1987). An extensive research of behav-
ior of sandwich panels with metal facings and selected types of insulation core was performed in
the frame of EASIE project (Käpplein, Misiek, 2010). Axially loaded panels were a subject of
theoretical, experimental and numerical investigation. Within the project, design procedures for
sandwich panels were developed as well.

3 EXPERIMENTAL INVESTIGATION

3.1 Object of the tests and preparation of the test setup


To study the actual behavior of sandwich panels under axial compressive load, a series of full-
scale tests was proposed as an effective method of investigation. It is assumed that response of the
panels to applied load is influenced by large number of factors. The experimental analysis of such
structural members using full-scale tests can bring valuable results and can be used for prospective

Figure 1. Load-deflection relationship: ideal and actual member.

134
verification of theoretical models (Melcher, 1997). A total of fourteen tests was completed in the
testing laboratory. The boundary conditions of the tested specimens were proposed to comply
with simply supported member as it is an elementary case when investigating buckling behavior of
structural members. In the frame of four tests, the specimens failed by local collapse at the sup-
ports with no observation of global buckling. These tests were not considered for final evaluation
of global buckling. For the following ten tests, the support area of the panels was therefore slightly
modified to prevent this type of failure. The final test setup used for the test program is described
below.

3.2 Specimens
For experimental verification of buckling behavior, sandwich panels with slightly profiled steel
facings of nominal thickness of 0.5 mm and polyisocyanurate (PIR) insulation core were used.
Polyisocyanurate foam is one of the number of materials that can be used as insulation core of
sandwich members. The foam itself is brittle (Gilbert, 2016). The benefit of use of polyisocyanu-
rate is higher fire resistance in comparison with other types of foams, e.g. polyurethane. It can
resist temperatures up to 150°C (Brydson, 1999).
The length of the tested panels was 4.0 m, width 1.0 m and thickness of the core 80 mm.

3.3 Test setup and procedure of testing


A special test device was prepared to perform the tests of buckling of sandwich panels. Simply
supported member was assumed in the frame of the planning of the test program. The specimen
was attached to a testing frame at both ends by pinned connections which enabled rotation of
the specimen at the supports out of plane to comply with the assumed boundary conditions.
The pins were greased to reduce friction. At one end of the specimen, hydraulic cylinder was
attached to apply axial force. At the ends of the specimen, plywood plates were attached to pre-
vent local failure at the supports. The plates did not prevent rotation of the specimen. The sup-
ports enabled lateral adjustment of the position of the specimen and therefore its precise
positioning on the testing device. The test setup is in Figure 2.
During each test, continuous measurement of applied force was performed. The out-of-
plane displacement was measured using draw-wire sensors at midspan of the specimen. Strain
gauges were attached at midspan of the specimen on both facings. Before applying the normal
force using hydraulic cylinder, additional lateral force caused by a 120 kg concrete block
(attached to the specimen via steel cable and pulley) was applied to the specimen at its mid-
span to cause an initial deflection. The deflection f0 caused by this lateral force was recorded.
The specimen was then loaded by the hydraulic cylinder until the failure was reached.

Figure 2. Test setup.

135
3.4 Test results and their evaluation
The evaluation of the test results focuses on specimens failed by global (flexural) buckling
which was caused by failure of the compressed face of the members. The tests with local fail-
ure mode were excluded from further evaluation. Failure of the specimens by global buckling
can be observed in Figure 3. Typical relationship between load and deflection (recorded
during one of the tests) is in Figure 4. Initial deflection is not displayed in the chart (but was
carefully recorded during each test).
The evaluation of the tests primarily focused on experimental determination of the critical
load of the tested panels. Measured data (axial force and lateral displacement) was used for
evaluation. As the term “critical load” applies only for ideal members (with no initial imper-
fections), the critical load cannot be determined directly. For the evaluation, Southwell plot
was used (Southwell, 1932). This method is suitable for evaluation of elastic critical load of
compressed members using a chart with lateral displacement of the member f on horizontal
axis and ratio between lateral displacement and actual applied force f/N on the vertical axis. It

Figure 3. Failure of the specimen.

Figure 4. Typical load-deflection curve.

136
has been found that this experimentally obtained relationship between f and f/N is linear only
in the medium range (Březina, 1962). The critical load is then obtained as cotangent of the
slope of the line passing through the medium range of the graphical relationship.
Typical graphical interpretation of evaluation of this relationship is presented in Figure 5
(relationship between displacement and ratio between the displacement and applied force for
one of the performed tests). The deflection starts at zero value (initial deflection is not con-
sidered in this graphical interpretation). The line is constructed in the linear range of the rela-
tionship. The slope of this line is quantified and used for calculation of the critical load.
The initial deflection f0 caused by the additional lateral force at midspan (concrete block)
was used to calculate bending stiffness of the panel using a simple formula for deflection
caused by a point load (Equation 1). This bending stiffness (E·I)panel is used to calculate Euler
critical load according to Equation 2.
Table 1 summarizes results of the tests and calculations: initial deflection f0 caused by lat-
eral force, bending stiffness of the panels (E·I)panel, maximum axial load Nmax applied using
the hydraulic cylinder at failure of the specimen, appropriate displacement f (initial deflection
is included), critical load determined using Southwell plot and Euler critical load determined
using experimentally obtained bending stiffness for each test according to Equation 2.
F L3
ðE  I Þpanel ¼ ð1Þ
48 f0

π2  ðE I Þpanel
Ncr ¼ ð2Þ
L2

Figure 5. Southwell plot.

Table 1. Selected results of the tests and calculations.


f0 (E·I)panel Nmax f Ncr (Southwell) Ncr (Euler)

Test No mm 109 N·mm2 kN mm kN kN

01 8.32 1.922 84.55 20.28 135.14 118.56


02 7.67 2.085 88.11 20.64 119.05 128.62
03 8.23 1.944 83.39 17.97 133.33 119.90
04 7.95 2.013 77.53 13.31 163.93 124.19
05 8.15 1.963 65.00 29.70 87.72 121.06
06 8.01 1.998 67.91 29.82 116.28 123.28
07 7.60 2.104 61.87 18.27 161.29 129.84
08 7.94 2.014 60.41 18.31 116.28 124.26
09 7.95 2.013 71.49 20.85 151.52 124.18
10 8.11 1.973 68.46 28.92 98.04 121.69

137
The magnitudes of maximum loads at failure were statistically assessed in terms of EN
1990, Annex D (Eurocode, 2003). Assessment via the characteristic value is performed pro-
vided there is no prior knowledge of coefficient of variation of the tested quantities. The char-
acteristic value of the axial load resistance is 53.83 kN for the tested specimens.

4 CONCLUSION

The paper summarizes results of series of full-scale tests of global buckling of sandwich panels
with steel facings and polyisocyanurate insulation core. The fundamental part of the paper is
evaluation of experiments, especially experimental determination of critical load of the tested
specimens. Comparison with Euler critical load determined using experimentally obtained
values of bending stiffness of the specimens provided good agreement between results of both
approaches. The research will be followed by experimental determination of mechanical proper-
ties of the facings and the core of the panels. The experimentally determined material properties
can be then used within the frame of theoretical and numerical analysis of stability of sandwich
member. They will enable accurate comparison between results of experiments and results of
calculation of critical load according to theory of sandwich structures.

ACKNOWLEDGMENT

The paper was elaborated within the frame of the research program No FAST-S-18-5550 of the
Faculty of Civil Engineering, Brno University of Technology and within the project No LO1408
AdMaS UP – Advanced Materials, Structures and Technologies, supported by the National Sus-
tainability Program I of the Ministry of Education, Youth and Sports of the Czech Republic.

REFERENCES

Brydson, J.A. 1999. Plastics Materials. Boston: Elsevier. ISBN 978-0-7506-4132-6.


Březina, V. 1962. Vzpěrná únosnost kovových prutů a nosníků (Buckling resistance of metal members).
Prague: Czechoslovak Academy of Sciences Publishing.
Davies, J.M. 1987. Axially loaded sandwich panels. Journal of Structural Engineering (United States) 113
(11): p. 2212–2230. ISSN 0733-9445. DOI: 10.1061/(ASCE)0733-9445(1987)113:11(2212).
EN 1990 Eurocode: Basis of structural design. 2003.
Euler, L. 1744. De curvis elasticis. Cited in: Březina, V. 1962. Vzpěrná únosnost kovových prutů a nosníků
(Buckling resistance of metal members). Prague: Czechoslovak Academy of Sciences Publishing.
Gilbert, M. 2016. Brydson’s Plastics Materials. Boston: Elsevier. ISBN 978-0-323-35824-8.
Käpplein, S. & Misiek, T. 2010. EASIE – Ensuring Advancement in Sandwich Construction Through
Innovation and Exploitation. Report No.: D3-2 – part 3 Tests on axially loaded sandwich panels. Karls-
ruhe: Karlsruher Institut für Technologie.
Käpplein, S. & Ummenhofer, T. 2010. Axial beanspruchte Sandwichelemente in rahmenlosen Konstruk-
tionen (Axially loaded sandwich panels in frameless buildings). Stahlbau 79 (10): p. 761–770. ISSN
0038-9145. DOI: 10.1002/stab.201001367.
Kovařík, V. & Šlapák, P. 1973. Stabilita a kmitání sendvičových desek (Stability and vibration of sandwich
plates). Prague: Academia – Czechoslovak Academy of Sciences Publishing.
Melcher, J. 1997. Full-Scale Testing of Steel and Timber Structures: Examples and Experience. In
Virdi et al. (ed.), Structural Assessment: The role of large and full-scale testing. London: E & FN
Spon, p. 301–308. ISBN 0-419-22490-4.
Southwell, R.V. 1932. On the Analysis of Experimental Observations in Problems of Elastic Stability.
Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences 135 (828):
p. 601–616. ISSN 1364-5021. DOI: 10.1098/rspa.1932.0055.
Vlasov, V.Z. 1962. Tenkostěnné pružné pruty (Thin-Walled Elastic Bars). Prague: State Publishing of
Technical Literature.

138
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Nonlinear behavior and instability of deployable arches

Ana Beatriz G. Barcellos, Murillo V.B. Santana & Paulo B. Gonçalves


Department of Civil and Environmental Engineering, Pontifical Catholic University of Rio de Janeiro, Rio
de Janeiro, Brazil

ABSTRACT: Deployable structures consist of a group of structures capable of modifying


their shape and volume in order to meet a range of conditions and needs. They are usually
prefabricated structures consisting of straight or curved bars linked together in a compact
bundle, which can then be unfolded into large-span, load bearing structural shapes. These
structures have dual functionality since they act as mechanisms during its deployment and
become immovable structures capable of supporting external loads during the service phase.
In addition, they should be lightweight and compact to be easily transported and simple and
quick to deploy. All these restrictions make it difficult to choose the best parameters regarding
the shape and material of the structure, since many analyzes must be performed in order to
find parameters that give lowest weight, highest stiffness, and that allow the structure to per-
form its two functions and ensure its reuse. Among the types of folding structure, those made
of pantographic elements (scissors) have attracted great interest from engineers and architects
in recent years. This study evaluates the geometric nonlinear behavior of plane arches consti-
tuted by two classic type of pantographic element, namely: polar and translational. For this,
a detailed nonlinear geometrical analysis is conducted through a tailored corotational finite
element software in order to evaluate the influence of the structure’s geometrical parameters,
type of scissor units and supports on the nonlinear behavior and stability of the structure. The
results obtained by our analyzes reveal, in most cases, a characteristic non-linear behavior of
these structures with the nonlinear equilibrium path exhibiting several load and displacement
limit points where jumps to remote and undesirable configurations may occur. Based on
them, the influence of system parameters on the load carrying capacity of the arch is
investigated.

1 INTRODUCTION

Deployable structures are a class of structures that can be rapidly erected and easily folded for
reuse. Moreover, they can change their topology in response to varying conditions and needs.
Several transformable structures can be found in nature, like extensible worms, deployable
leaves and certain insect wings [1], which serve as inspiration for engineering structures,
including many civil engineering applications such as emergency shelters, exhibition halls, rec-
reational structures and temporary buildings, among others. For this purpose, preassembled
deployable scissor structures are highly effective: besides being transportable, they have the
advantage of quick and ease deployment, while offering a huge volume expansion. The scissor
mechanism was first presented in 1960 by the Spanish architect Emilio Perez Piñero. This
mechanism uses the pantograph principle and consists of a bundle of scissor-like elements.
These elements are comprised of two beams that are crosswise interconnected by a revolute
joint, referred to as a pivotal connection, which introduces constraints of rotation normal to
their common plane, as exemplified in Figure 1 where different scissor configurations are dis-
played. The units are connected to each other by joints at their end nodes that allow for in-
plane rotation. The upper and lower points of a scissor unit are connected by what is called
unit lines [2].

139
Figure 1. A scissor unit during deployment: (a) Completely deformed (b) partially deformed and (c)
fully closed.

After Piñero, deployable scissor structures became more popular and attracted the interest
of researchers who tried to improve their design and understand their mechanical behavior.
Escrig [3], Gantes [4], Hoberman [5] and You and Pellegrino [6], among others, determined
the geometric relations for different scissor configurations, proposed new types of scissor units
and studied their structural response during deployment and under service loads in order to
enhance their performance. New types of units were then obtained by changing the position of
the pivot and the size and form of the beams, which have a profound influence on the final
shape of the space enclosure and its deployment behavior. In spite of numerous new scissor
elements configurations, two types of scissor units are more commonly used and are investi-
gated in this work: translational units and polar units. Translational units are composed of
two straight beams of same length with the pivot positioned in the same position in both
beams, as illustrated in Figure 2(a), so that their axes remain parallel during deployment. In
polar units, the pivot is shifted from the middle with a certain eccentricity which makes their
axes intersect at a variable angle during deployment with respect to a fixed central point, as
illustrated in Figure 2(b) [7].
Zeigler [8] improved Piñeros systems by introducing geometric incompatibilities between the
member lengths of a scissor element thus creating self-locking and self-stiffening bistable struc-
tures which didn’t require any additional locking system. These bistable scissor structures are
ideally geometrically compatible before and after deployment, but, during deployment, the large
displacements combined with the geometric incompatibilities result in intended bending of some
of the members. Consequently, these structures exhibit a snap-through behavior at sufficiently
small loads during deployment to achieve a stable post-buckling configuration. The self-locking
phenomenon can thus also be beneficial and desired because of the ease and speed of deployment.

Figure 2. (a) Plane translational scissor unit; (b) plane polar scissor unit.

140
However, the structural response of these structures is complex, involving large displacements
and rotations and a fully nonlinear analysis is necessary to evaluate the degree of nonlinearity
and load carrying capacity. Plastic material behavior during deployment and use should be
avoided since it would result in reduced load-bearing capacity in the deployed configuration and
permanent strains. Due to this complexity, a nonlinear analysis requires the use of a set of
advanced numerical tools. This research presents a detailed parametric analysis to evaluate the
influence of the type of scissor unit (translational or polar), number of units and type of support
on the structural behavior of scissor arches. For this, a tailored finite element software is used to
conduct the geometric nonlinear analysis and stability of these structures [13].

2 DESIGN

The realization and design of scissor like structures is hindered by their dual functionality:
during deployment, it acts as a mechanism and afterwards, during the service phase, it becomes
an immovable load-bearing structure. This increases the complexity of the design process where
the relation between geometric design parameters, kinematic properties and structural response
is crucial and important to take into account. Moreover, the designer has a large freedom when
choosing design parameters for a scissor structure. With more structural insight into the geo-
metrically nonlinear response of the scissor system, the engineer can make legitimate decisions
related to these parameters in the design process. Langbecker and Albermani [9] and Mira et al.
[10] investigated the effect of several design parameters on the structural behavior of barrel
vault structures during its service phase when subjected to static loading. In addition, Gantes
[11] and Arnouts [12] point out the importance of also evaluating the nonlinear behavior of scis-
sor-like structures during their deployment and service life. The scissor elements are made of
aluminum class A, EN-AW 6060, with modulus of elasticity equals to E = 70GPa, the same
used by Mira et al. [10]. The bars of the translational units have square profiles of 100 x 3mm
while the polar units have square profiles of 90 x 4mm. An arch with a span of 6m and height
of 3m (height-to-span ratio of 0.5) when completely unfolded is considered and the number of
units varies from 6 to 18. Five different types of support are considered, namely (a) pinned sup-
ports at one node of each end unit, (b) pinned supports at both nodes of the two end units, (c)
pinned supports at the pivot of each end unit, (d) a fixed support at one node of each end unit
and (e) fixed supports at both nodes of the two end units. The final configuration of the
deployed arch depends on the number of units [1-12]. The design of the structures follow the
methodology described in [7], where the lengths of the bars are calculated through a numerical
method, the lengths varying with the number of elements. While in structures of polar units all
the bars have the same length, in the ones with translational units the bars have different
lengths, as illustrated for a structure with 10 units in Table 1.

Table 1. Length of the bars of the arch with 10


units (translational and polar).
Bar lenght (m)

Bar Number Tralational units Polar units

1/20 1,48 0,98


2/19 0,54 0,98
3/18 1,42 0,98
4/17 0,6 0,98
5/16 1,3 0,98
6/15 0,7 0,98
7/14 1,2 0,98
8/13 0,82 0,98
9/12 1,06 0,98
10/11 0,94 0,98

141
3 RESULTS

Due to their importance in the structural behavior, self-weight is taken into account in all ana-
lyses. In order to understand the arch nonlinear response, first the behavior of an arch com-
posed of polar units with pinned supports at the end units of bars 1 and 20 (see Table 1)
subjected to a concentrated load at the top none is investigated. Figure 3(a), where the load is
plotted as a function of the vertical displacement of the top node, shows the nonlinear
response of the arch for an increasing number of polar units. Structures composed of a small
number of units (see results for 6 to 9 units) exhibit a strongly nonlinear behavior with several
load and displacement limit points. Figure 3(b) shows in detail the nonlinear response for the
structure composed of 7 units together with the deformed shape of the arch at each limit
point. As observed here, the initial stiffness of the structure is very low leading to large dis-
placements for small load levels. As the number of units increase, the nonlinearity decreases,
as shown in Figure 3(c) for the structures composed by 10 and 11 units. In this case, the initial
response is practically linear up to about 30kN. As the load increases beyond this point, the
effective stiffness decreases and becomes zero at the first limit point where the structure jumps
to a new stable position associated with large deflections, as illustrated in Figure 3(d) for the
structure with 11 units, where the configuration of the structure before and after snap-through
are displayed. This first limit point controls the load capacity of the deployed structure.
Although not shown here for lack of space, structures composed of translational units exhibit
larger deformations for the same number of units. The high flexibility of the structure is due
to the type of support used in the model.

Figure 3. (a) Nonlinear equilibrium paths for structures composed by polar units pinned at one
node of the end unit; (b) load-deflection curve for the structure with 7 units; (c) detail of nonlinear
equilibrium paths for the structures composed of 10 and 11 units; (d) configuration of the structure
composed by 11 units before and after snap-through.

142
In order to evaluate the influence of the supports and type of unit on the arch response,
Figure 4(a) shows the nonlinear response for increasing number of polar units, while Figure 4
(c) shows the corresponding nonlinear response considering translational units. Figures 4(b)
and (d) show a zoom of the initial behavior of the arch up to a vertical displacement of the top
node displacement equal to 0.06m (arch span/100) for, respectively, the polar and translational
units. Here, in both cases, the arch is pinned at the pivot node of the two end units, increasing
thus its stiffness. By comparing the results in Figure 3(a) and Figure 4(a), both for polar units,
the same overall behavior is observed. However, the load capacity of the structure is much
higher when pinned at the pivot extreme nodes. There is no direct relation between the number
of units and the first limit point load. First, for n=6 and 7 large deflections is observed for small
load levels and no limit point is observed. For n=8 the path exhibits a flat region, indicating the
appearance of a limit point. For higher number of unit, a limit point load appears and increases
with the number of units. The highest load level is attained with 14 polar units, which presents
a linear response up to a load level of 80kN. For n>14 the limit point load decreases with the
number of units. By comparing the results in Figure 4(a, b) and Figure 4(c, d), the higher stiff-
ness and load capacity of the arch with polar units becomes evident. As observed in Figures 4
(b) and (d), for the same vertical displacement, the load capacity of structures composed by
polar units is at least five times higher than that for the structure composed by translational
units. In addition, structures composed by translational units begin to display a nonlinear
behavior at small load levels and vertical deflections, as shown in Figure 4(c, d).
In order to understand the influence of the arch topology on its load capacity, Tables 2
shows the load level corresponding to a vertical displacement of the top node equal to 0.06m
for the five types of support, considering an increasing number of polar units. Table 3 shows
the same results for translational units. In all cases the highest load level occurs for the arch
with clamped supports on both nodes of the two end units, as expected. The highest load level
for the arch composed by polar units occurs with n=10 (71.03kN) and for the arch composed
by translational units it occurs with n=8 (67.25kN), a decrease of about 5%. Having as

Figure 4. Nonlinear equilibrium paths for arches pinned on the pivot node of the end units: (a) arch
composed by polar units; (b) zoom of Figure 4 (a); (c) arch composed by translational units; (d) zoom of
Figure 4 (c).

143
Table 2. Load at a displacement equal to 0.06m for each support type and
number of units of structures composed of polar units.
Polar units
Load at 0.06m (kN)
Number Support Support Support Support Support
of units type (a) type (b) type (c) type (d) type (e)

n= 7 10.86 38.56 14.24 16.73 33.11


n= 8 15.63 59.70 23.28 25.85 49.45
n= 9 16.88 49.18 20.7 25.36 30.44
n= 10 24.84 70.95 34.44 38.21 71.03
n= 11 23.43 51.27 32.63 31.7 53.19
n= 12 31.43 69.84 44.32 48.84 70.03
n= 13 28.72 47.18 40.91 39.33 49.06
n= 14 38.27 58.55 52.32 51.6 51.99
n= 15 29.11 38.75 35.73 35.51 37.69

Table 3. Load at a displacement equal to 0.06m for each support type and
number of units of structures composed of translational units.
Translational units
Load at 0.06m (kN)
Number Support Support Support Support Support
of units type (a) type (b) type (c) type (d) type (e)

n= 7 8.05 7.93 40.99 10.12 41.06


n= 8 9.15 8.08 66.95 12.92 67.25
n= 9 7.09 6.27 43.66 11.16 41.48
n= 10 7.59 5.56 64.65 12.95 63.89
n= 11 6.25 4.29 35.69 11.19 35.29
n= 12 6.02 3.17 48.91 12.07 49.31
n= 13 4.99 1,00 27.82 5.91 27.67
n= 14 0.9 -0.2 33.86 5.19 33.60
n= 15 0.01 0,00 20.36 4.11 20.56

a design criterion the maximum deflection, no direct relation is observed between the load
level and the type and number of units. Also the load capacity varies highly with the type of
support for the same number of units. So, to obtain an optimal shape either a careful paramet-
ric analysis or an optimization algorithm should be employed by the designer.
Figure 5 shows the nonlinear equilibrium paths for the arch with polar units and clamped
supports at both nodes of the end units. Comparing with Figures 3(a) and 4(a), the more effi-
cient behavior of this type of support becomes clear, corroborating their influence on the non-
linear response and load capacity. In the present case, the highest load level occurs for n=12
(around 225kN).
Now the arch response under distributed load on the top nodes, as depicted in Figure 6, is
investigated. Figure 7 shows the nonlinear equilibrium paths of the arch with clamped ends
(case e) for selected values of scissor units considering both polar and translational units,
while Figure 8 shows the results for the arch with pinned end (case c). The typical nonlinear
behavior of a bistable structure is observed in both cases for most values of n, displaying an
upper and lower limit point. Again the higher load capacity of the arch with polar units is
observed for both types of support. For the clamped arch with polar units, the best structural
response is observed for n=11. For n=8 a higher limit point load is observed but it is

144
Figure 5. Nonlinear equilibrium paths for arch composed of polar units and clamped supports at both
nodes of the two end units.

Figure 6. Arch with loads distributed along the top nodes.

Figure 7. Nonlinear equilibrium paths of the arch with loads distributed along the top nodes and
clamped ends as a function of the number of scissor units. Load parameter versus central deflection.

associated with large deflections. Similar results are obtained for the pinned arch, as observed
in Figure 8(b). For polar units, the highest load level at a deflection of 0.06m is the same for
the two types of support and occurs for n=10 (35.3kN). For translational units, the load for
clamped support is higher than that for pinned supports and occurs for n=8 (37.46kN).

145
Figure 8. Nonlinear equilibrium paths of the arch with loads distributed along the top nodes and
clamped ends as a function of the number of scissor units. Load parameter versus central deflection.

4 CONCLUSION

This work presents a detailed nonlinear analysis of deployable arches using a tailored finite
element formulation. The influence of the type of scissor units (polar or translational),
number of units and type of support is investigated. The results show that deployable arches
present a highly nonlinear response with load and displacement limit points. The degree of
nonlinearity and the load carrying capacity in terms of the first limit point load or
a prescribed maximum deflection depends highly on the number and type of scissor units
which also defines the form of the deployed arch. As expected, the arch load capacity is also
highly dependent on the type of support. In all cases arches with polar units shows a higher
load carrying capacity than those with translational units. So, to obtain an optimal shape
either a careful parametric nonlinear analysis or an optimization algorithm should be
employed by the designer to obtain the best structural configuration.

REFERENCES

[1] S. Pellegrino (Ed.). Deployable structures (Vol. 412). Springer, 2014.


[2] K. Roovers and N. De Temmerman, “Deployable scissor grids consisting of translational units,”
Int. J. Solids Struct., vol. 121, pp. 45–61, 2017.
[3] F. Escríg, “Estructuras espaciales desplegables curvas,” Inf. la construcción, 1988.
[4] C. Gantes, “A Design Metodology for Deployable Structures.” Ph. D. Thesis, Massachusetts Insti-
tute of Technology, 1991.
[5] C. Hoberman, “Reversibly Expandable Doubly-Curved Truss Structures,” U.S. Patent
n. 4,942,700, 1990.
[6] Z. You and S. Pellegrino, “Foldable bar structures,” Int. J. Solids Struct., vol. 34, no. 15, pp.
1825–1847, 1997.
[7] N. De Temmerman, “Design and analysis of deployable bar structures for mobile architectural
applications,” Vrije Univ. Brussel. Ph. D. Thesis., 2007.
[8] T.R. Zeigler, “Collapsible Self-Supporting Structures”. U.S. Patent n. 3,968,808, 1976.
[9] T. Langbecker and F. Albermani, “Kinematic and non-linear analysis of foldable barrel vaults,”
Eng. Struct., vol. 23, no. 2, pp. 158–171, 2001.
[10] L.A. Mira, R.F. Coelho, A.P. Thrall, and N. De Temmerman, “Parametric evaluation of deploy-
able scissor arches,” Eng. Struct., vol. 99, pp. 479–491, 2015.
[11] C.Gantes, J.J. Connor, and R.D. Logcher, “Combining numerical analysis and engineering judg-
ment to design deployable structures,” Comput. Struct., vol. 40, no. 2, pp. 431–440, 1991.
[12] L.I.W. Arnouts, “Computational Investigation of the Structural Response of Bistable Scissor
Structures,” M. Sc. Dissertation, ULB, Brussels, Belgique, 2017
[13] M.V.B. Santana, “Tailored Corotational Formulations for the Nonlinear Static and Dynamic Ana-
lysis of Bistable Structures”, Ph. D. Thesis, ULB & PUC-Rio, 2019.

146
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Numerical advanced analysis of steel-concrete composite beams


and columns under fire

R.C. Barros, R.A.M. Silveira & P.A.S. Rocha


Federal University of Ouro Preto, Ouro Preto, MG, Brazil

D. Pires
Federal University of São João del-Rei, Ouro Branco, MG, Brazil

Í.J.M. Lemes
Federal University of Lavras, Lavras, MG, Brazil

ABSTRACT: It is well known that high temperature causes changes in physical properties
and mechanical strength of materials. In both steel and concrete, material properties deteriorate
during the exposure to fire, and a steel, RC or composite structure load capacity and stiffness
are reduced significantly as a consequence. The use of computational models as methodology of
analysis/design has as advantage of capturing the strength and stability of the structural system
and its members directly. This work shows the efficiency of the CS-ASA/FSA FE code in the
analysis of insulated steel-concrete composite structural elements, such as beams and columns,
in a fire situation. This approach performs the inelastic second-order analysis of composite
structures under fire considering the refined plastic hinge method coupled to the strain
compatibility method. Several composite structures in fire situation are analyzed using the
developed tool.

1 INTRODUCTION

The use of composite steel-concrete structures is increasing because of the many advantages
that this combination of materials provides. This constructive system allows the two materials
to be used together in beams, columns, frames and slabs in order to obtain a structure with
high structural performance, geometric precision and low on-site waste.
Despite of the significant increase in the use of steel-concrete composite structures, research
on the behavior of these structures, especially in a fire situation, is still modest in Brazil. Thus,
experimental studies and numerical models capable of simulating the behavior of structures in
fire situation becomes extremely relevant.
Espinos et al. (2016) presented the results of a numerical investigation on strategies for
enhancing the fire behavior of concrete-filled steel tubular (CFST) columns by using
inner steel profiles as circular hollow sections (CHS), HEB profiles or embedded steel
core profiles. The ABAQUS commercial software was adopted, and a 3D finite element
that allowed to compute different sections and nonlinear materials behavior under high
temperatures was considered. Parametric studies were developed with the purpose of
investigating the cross-section geometry influence on the resistant capacity of the columns
under fire.
Shallal et al. (2018) investigated the influence of high temperatures on continuous steel-
concrete composite beams through an inelastic analysis. Their numerical models were based
on an one-dimensional element of 4 nodes, with 3 degrees of freedom per node, including the
partial interaction between the steel profile and the concrete slab. Good agreement was
observed between their numerical results and the data from literature. Other recent works

147
involving composite structures numerical analysis in fire situations can be cited, i.e.: Zofrea
(2014), Lai et al. (2017) and Pak et al. (2018).
The objective of this contribution is to apply the in-house developed CS-ASA/FSA compu-
tational module (Computational System for Advanced Structural Analysis/Fire Structural Ana-
lysis) to the analysis of insulated steel-concrete composite structural elements, such as beams
and columns, in a fire situation. This numerical module was developed for the
inelastic second-order analysis of structures submitted to high temperatures. An approach
based on the Strain Compatibility Method (SCM) is proposed to evaluate the cross-section
strength level and the axial and bending stiffness of the composite structures under high tem-
peratures. The construction of the moment-curvature relationship is essential for this evalu-
ation. The tangent of the moment-curvature relationship, the stiffness, depends only on the
chosen material behavior of the constituents. This methodology is coupled to the Refined
Plastic Hinge Method (RPHM), in which the plasticity is evaluated only in the element nodal
points using generalized stiffness parameters (Lemes 2018).

2 THERMAL ANALYSIS

The thermal analysis is performed exclusively in the cross-sectional plane through finite element
heat transfer models that allow the temperature distribution determination at different points in
the cross-section. A time integration strategy based on the Finite Difference Method (FDM) is
adopted for solving the discrete differential equations. The CS-ASA/FA module (Fire Analysis)
also has two solvers: simple incremental and incremental-iterative (Picard or Newton-Raphson).
It is important to mention that the steel and concrete thermal and mechanical properties in a fire
situation are adapted according to the normative prescription of EN 1994-1-2 (2005). Further
details of this computational module are available in Barros et al. (2018) and Pires (2018).

3 STRUCTURAL ANALYSIS

The steel-concrete composite structures inelastic behavior in fire condition is captured using
the RPHM and SCM coupling (Lemes 2018). The thermal action effects on the structure, i.e.,
the material stiffness and yield strength degradations, as well as the influence of the thermal
strain on the element cross-section were considered within these numerical approaches. The
following sections bring a summary of the FE formulation via RPHM, the SCM and how the
bending moment-curvatures relationships are derived. Additional details of the structural
solution strategy can be found in Pires (2018) and Barros et al. (2018).

3.1 Finite element formulation via RPHM


The objective of the RPHM is to capture the evolution of the plastification at the nodes of the
element, from the beginning of the yielding to its total plastification with a plastic hinge for-
mation. The following main assumptions are considered in the model: all finite elements are
initially straight and prismatic and their cross-sections remain plane after deformation; the
steel profiles are compact; rigid body large displacements and rotations are represented; the
shear deformation effects are ignored. The developed finite element is delimited by nodal
points i and j, as illustrated in Figure 1. P, Mi, Mj and δ, θi, θj are the internal forces and
associated displacements in the co-rotational system, respectively.

Figure 1. Beam-column finite element.

148
The incremental equilibrium relationship of the finite element illustrated in Figure 1 is
given by:
8 9 2 38 9
< DP = k11 0 0 < Dδ =
DMi ¼ 4 0 k22 k23 5 Dθi ð1Þ
: ; : ;
DMj 0 k32 k33 Dθj

in which Δ denotes the increments of each quantity. The terms related to flexural stiffness in
the matrix depend on geometric nonlinearity. A simplified second-order formulation, pre-
sented by Yang & Kuo (1994) was adopted here. Implementation details are available in
Lemes et al. (2017) and Lemes (2018).

3.2 Strain compatibility method (SCM)


When under external loads, a structure deforms to reach equilibrium. On the cross-section
level, once the generalized internal forces equal the generalized external forces, the deform-
ation stops (Lemes et al. 2017). This cross-section deformation is addressed in the SCM. For
the application of this method, it is assumed that the strain field is linear and the section
remains plane after deformation (Figure 2). This method couples the cross-section deformed
configuration to the constitutive relationship of the material composing it.
The evaluation of the axial and flexural stiffnesses is derived from the moment-curvature
relationship and depends on the modulus of elasticity, which is obtained from the material
(steel and concrete) constitutive relations (Pires 2018).

3.3 Moment-curvature relationship


A discretization of the cross-section into fibers is performed with the objective of describing
the deformation distribution using the axial strain (εi) in the plastic centroid (PC) of each
fiber. From this the stress (σi) in each fiber is computed through the constitutive relationships
of the materials. The axial strain in the ith fiber is given by:

εi ¼ ε0 þ ϕyi þ εth;a þ ϕth yi ð2Þ

where yi is the distance between the plastic centroids of the fiber analyzed and the cross-
section; ε0 is the axial strain at the PC of the section; ϕ is the respective curvature; εth,a is the
thermal axial strain; and ϕth is the curvature from the thermal strain, which is determined
according to Pires (2018).

Figure 2. Linear strain field.

149
The Newton-Raphson method is used at the cross-section level in order to obtain the
moment-curvature relationship (M - ϕ). In matrix notation, the variables ε0 and ϕ are compo-
nents of the generalized strain vector X = [ε0 ϕ]T. For computing the axial and flexural gener-
alized stiffness, X = 0 is adopted in the first iteration, so that convergence is reached quickly
(Chiorean 2013). It can be said that the cross-section equilibrium is obtained when the follow-
ing equation is satisfied:

Next Nint
FðXÞ ¼ f ext  f int ¼  5tol ð3Þ
Mext Mint

where fext is the external force vector given by the axial force Next and bending moment Mext; the
terms Nint and Mint are the components of the internal force vector, fint and tol is a tolerance.
The internal forces are obtained from the deformed configuration of the cross-section
through the expressions:

ðð ðð ðð X
nfib;a X
nfib;c X
nfib;b
Nint ¼ σa dA þ σc dA þ σb dA ffi σai Aai þ σci Aci þ σbi Abi ð4Þ
i¼1 i¼1 i¼1
Aa Ac Ab

ðð ðð ðð X
nfib;a X
nfib;c X
nfib;b
Mint ¼ σa ydA þ σc ydA þ σb ydA ffi σai Aai yai þ σci Aci yci þ σbi Abi ybi ð5Þ
i¼1 i¼1 i¼1
Aa Ac Ab

where nfib,a is the number of fibers; σa, σc and σb are the stress in steel, concrete and steel
rebar, respectively; Ai and yai are the ith fiber area and its positions in relation to the Plastic
Neutral Axis (PNA), respectively.
For the following iteration, k+1, the strain vector is calculated according to (Chiorean 2013):

 1  
Xkþ1 ¼ Xk þ F0 Xk F Xk ð6Þ

where F0 is the Jacobian matrix for the non-linear problems, i.e.:

  " ∂Nint ∂Nint


#
0 ∂F ∂ε0 ∂ϕ
F ¼ ¼ ∂Mint ∂Mint
ð7Þ
∂x ∂ε0 ∂ϕ

This numerical procedure is adapted and used for obtaining the N-M interaction curves as
well. For a given axial force, the limit moment from the relationship of the moment-curvature
is obtained, corresponding to the cross-section total plastification. This pair of forces define
a point on the N-M interaction diagram. Noteworthy is the fact that the interaction curves are
obtained independently the structural analysis, i.e. computed beforehand, in order to acceler-
ate the execution of the structural simulations. More details on the computation of interaction
curves, the structural analysis, as well as the thermal structural problem, can be found in Pires
(2018), Barros et al. (2018) and Lemes (2018).

4 NUMERICAL ANALYSIS

4.1 Steel-concrete composite beams


Caldas (2008) presented the numerical results of two simply supported steel-concrete compos-
ite beams in a fire situation. These beams were initially analysed by Huang et al. (1999) using
the VULCAN software and experimental results were presented by Wainman & Kirby (1988).
The beams cross-section is composed of a steel I profile 254 × 136 mm × 43 kg/m and

150
Figure 3. Room temperature structural analysis results of the composite beams. a) Test 15: deflection
x temperature. b) Test 16: deflection x temperature.

a concrete slab with 624 × 130 mm. The reinforcing steel bars on the slab were neglected in the
CS-ASA/FSA numerical simulations. The steel profile yield strength was considered equal to
255 MPa and the concrete had a compressive strength of 30 MPa. As shown in Figure 3a,
four loads were applied in the experimental tests: 32.47 kN (Test 15) and 62.36 kN (Test 16).
The room temperature structural analysis results, seeking to determine the beam critical load,
as well as the interaction diagram between axial force and bending moment (N-M) for t = 0,
are shown in Figure 3b.
The steel-concrete composite beams were then exposed to standard fire curve (ISO 834-1
1999) on the steel profile bottom flange and the thermal properties were considered according
to EC4 Pt.1.2 (EN 1994-1-2 2005). Eight linear finite elements were used to discretize the
structural system and 346 linear quadrilateral finite elements (Q4) for the cross-section.
Figures 4a and 4b present the mid-span deflection versus temperature of the bottom flange
of the beam for the two tests, together with the numerical results of Caldas (2008) and Huang
et al. (1999), and the Wainman & Kirby (1988) experimental results.

4.2 Steel-concrete composite columns


Huang et al. (2008) presented a numerical study on fire resistance of embedded I-section com-
posite columns through a FE program named FEMFAN-3D. The objective was to examine
the effects of cross-sectional dimension and load level on the column fire resistance. Four
groups of columns consisting of square cross-section were chosen for the study. Their cross-
sectional dimensions range from 250 × 250 to 400 × 400 mm2, respectively. These columns
were subjected to axial compression forces and four-face uniform heating according to the
standard fire curve (ISO 834-1 1999). All columns had 3 m length and were named as SZ1 to
SZ4. The typical cross-section of columns with embedded I-sections and further details on the
section’s dimensions are presented on Figure 5a. As in the previous example, Figure 5b shows
the results of the structural analysis at room temperature and the interaction diagrams for the
4 cross-sections.
Each column was modeled considering a mid-height initial deflection of L/300, where L is
the length of the columns, and they were subjected to a centered axial load level of 0.5Py. The
corner rebars were considered with yield strength equal to 460 MPa. For the concrete,
a compressive strength of 42 MPa was adopted, while the steel profile yield stress was 330
MPa. Thermal and material properties of steel, concrete and rebar were taken from EC4

151
Figure 4. Fire structural analysis results of composite beams.

Figure 5. Room temperature structural analysis results of the composite columns. a) Cross-section
geometry. b) Equilibrium path and interaction diagram.

Pt.1.2 (EN 1994-1-2 2005). Concrete moisture content was set at 130 kg/m3 and relative emis-
sivity was set at 0.5. The structural system was discretized with 10 linear finite elements and
the cross-section with 240 linear quadrilateral finite elements.
The thermo-structural analysis of the four columns are shown in Figure 6a, and Figure 6b
brings the members’ last deformed configuration with the respective time intervals.

152
Figure 6. Fire structural analysis results of composite columns. a) Lateral displacement x time. b) Last
deformed configuration of columns.

5 FINAL REMARKS

The present study applied the CS-ASA/FSA computational module to steel-concrete compos-
ite structural elements in fire situation. This module efficiency had already been tested with
steel and reinforced concrete structural elements in fire situations, where good results in both
cases were observed (Barros et al. 2018, Pires 2018).
The first numerical analysis consisted of composite beams subjected to thermal gradients.
In general, the results obtained by CS-ASA/FSA showed good agreement with those found in
literature, including experimental results.
The numerical results presented for the steel-concrete composite columns were satisfactory
as well. In this example it is worth pointing out a difference in numerical results and literature
in the interval between 40 and 70 min. It may be associated to the representation of the deg-
radation of the parameters of steel strength and stiffness, since a similar observation was
made by Barros et al. (2018).
As a final conclusion, it is possible to affirm that the computational implementations, spe-
cific for the thermo-structural analysis of steel-concrete composite structural elements under
fire were successfully performed and yield satisfactory results, capturing well the behavior of
the composite structures at high temperatures.

ACKNOWLEDGMENT

The authors would like to thank CAPES and CNPq (Federal Research Agencies), FAPE-
MIG (Minas Gerais Research Agency), Batir/ULB, Gorceix Foundation and UFOP/
PROPP for their support during the preparation of this research work. They also thank
the prof. Péter Berke’ comments for the enhancement of this manuscript.

153
REFERENCES

Barros, R.C., Pires, D., Silveira, R.A.M., Lemes, I.J.M., Rocha, P.A.S., 2018. Advanced inelastic ana-
lysis of steel structures at elevated temperatures by SCM/RPHM coupling. Journal of Constructional
Steel Research, vol. 145, pp. 368–385.
Caldas, R.B., 2008. Análise numérica de estruturas de aço, concreto e mistas em situação de incêndio. Tese
de Doutorado, Programa de Pós-Graduação em Engenharia de Estruturas, EE/ UFMG,Belo
Horizonte, MG, Brasil (in Portuguese).
Chiorean, C.G., 2013. A computer method for nonlinear inelastic analysis of 3D composite steel-concrete
frame structures. Engineering Structures, vol. 57, pp. 125–152.
Espinos, A., Romero, M.L., Lam, D., 2016. Fire performance of innovative steel-concrete composite col-
umns using high strength steels. Thin-Walled Structures, vol. 106, pp. 113–128.
European Committee for Standardization - EN 1994-1-2:2005. Eurocode 4: Design of composite steel and
concrete structures, Part 1-2: General rules, Structural Fire Design.
Huang, Z., Burgess, I.W., Plank, R.J., 1999. The influence of shear connectors on the behavior of com-
posite steel-framed buildings in fire. Journal of Constructional Steel Research, 51, 219–237.
Huang, Z.F, Tan, K.H., Toh, W.S., Phng, G.H., 2008. Fire resistance of composite columns with embed-
ded I-section steel - Effects of section size and load level. Journal of Constructional Steel Research, 64,
312–325.
ISO 834-1, 1999. Fire resistance tests - elements of buildings construction, Part 1: General requirements.
ISO - International Organization for Standardization. Geneva.
Lai, Z., Varma, A.H., Agarwal, A., 2017. Analysis of rectangular CFT columns subjected to elevated
temperature. Proceedings of the Annual Stability: Conference Structural Stability Research Council,
San Antonio, Texas, USA.
Lemes, Í.J.M., 2018. Estudo numérico avançado de estruturas de aço, concreto e mistas. Tese de Doutor-
ado, Programa de Pós-Graduação em Engenharia Civil, Deciv/EM/UFOP, Ouro Preto, MG, Brasil
(in Portuguese).
Lemes, Í.J.M., Silva, A.R.D., Silveira, R.A.M., Rocha, P.A.S., 2017. Nonlinear analysis of
two-dimensional steel, reinforced concrete and composite steel concrete structures via coupling SCM/
RPHM. Engineering Structures, vol. 147, pp. 12–26.
Pak, H., Kang, M.S., Kang, J.W., Kee, S.H., Choi, B.J., 2018. A numerical study on the
thermo-mechanical response of a composite beam exposed to fire. International Journal of Steel Struc-
tures, vol. 18(4), pp. 1177–1190.
Pires, D., 2018. Análise numérica avançada de estruturas de aço e de concreto armado em situação de incên-
dio. Tese de Doutorado, Programa de Pós-Graduação em Engenharia Civil, Deciv/EM/UFOP, Ouro
Preto, MG, Brasil (in Portuguese).
Shallal, M.A., Almusawi, A.M., 2018. Non-linear analysis of continuous composite beams subjected to
fire. International Journal of Civil Engineering and Technology, vol. 9, pp. 521–532.
Wainman, D.E, Kirby, B.R., 1988. Compendium of UK Standard Fire Test Data, Unprotected Struc-
tural Steel – 1, Rotherham (UF): Swinden Laboratories, British Steel Corporation, No. RS/RSC/
S10328/1/87/B.
Yang, Y.B., Kuo, S.B., 1994. Theory and Analysis of Nonlinear Framed Structures. Prentice Hall.
Zofrea, M., 2014. Behavior, analysis and design of concrete filled double skin tubular columns under fire.
Tesi di laurea. Dipartimento ICEA, Corso di Laurea Magistrale in Ingegneria Civile, Padova, Italy.

154
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Buckling of spatial laced columns composed of built-up


cold-formed channel members

C.C.D.O. Bastos & E.M. Batista


Civil Engineering Program, COPPE, Federal University of Rio de Janeiro, R.J., Brazil

ABSTRACT: This paper presents results of full-scale experimental tests carried out in
spatial laced built-up columns, designed with lipped channel cold-formed steel members.
The laced column chords, composed of two lipped channel members connected with self-
drilling screws, were previously analyzed using Generalized Beam Theory to identify its
buckling loads and modes. Elastic buckling and nonlinear FEM analyses were performed
with 3D bar elements for the laced column as a whole. Two analytical methods were
applied to obtain global buckling load. The obtained results indicated the tested columns
perform little influence of shear effect and critical buckling load may be accessed with the
help of available analytical equations. Laced columns experimental performance and
strength were recorded during the tests and compared with numerical results. Column col-
lapse mechanism and ultimate strength proved to be strongly affected by local buckling
effects. The obtained results indicate acceptable comparison between computed and experi-
mental data.

1 INTRODUCTION

Spatial laced columns composed of steel cold-formed steel members (CFS) are investigated
in the present research, with the four chords connected with diagonal and bracing elem-
ents, conducting to trussed-type structural behavior. Hashemi and Jafari (2009, 2012),
Bonab et al (2013), Kalochairetis et al (2014), Hashemi and Bonab (2013), have investi-
gated the elastic critical load and strength of laced (or batten) columns through laboratory
tests, but in all cases there were only two hot rolled chord members, connected by lacing
bars.
The behavior of a laced built-up column depends on its bending and shear stiffness, as well
as on the connections stiffness. The effect of shear deformations may cause distortion of laced
panels. Engesser (1891) was the first to consider the shear deformation effect on the elastic
critical load of columns. Bleich (1952), Timoshenko and Gere (1963), and several other
researchers followed the original findings of Engesser, such as: Gjelsvik (1991), Paul (1995),
Aslani and Goel (1991) and Razdolsky (2005, 2010, 2011, 2014a and 2014b).
The main objective of the present article is to present the results of full-scale experimental
tests carried out in spatial laced built-up columns, designed with lipped channel cold-formed
steel members. Four spatial laced columns were tested, with 12200 and 16200 mm length, 0.8
and 1.25 mm plate thickness, and 400 x 400 mm cross-section shown in Figure 1(a). The laced
column longitudinal members (chords), composed of two lipped channel members connected
with self-drilling screws, were previously analyzed by the generalized beam theory (GBT), in
order to identify its buckling loads and modes. For the built-up column as a whole, elastic
buckling and nonlinear FEM analyses were performed, with 3D bar elements. Two analytical
methods were applied to obtain the global buckling load including shear effect, for pin-ended
condition: Timoshenko (1963) and Eurocode 3 (CEN, 2006b).

155
Figure 1. (a) Four chords laced built-up column cross-section (centroid of each chord is indicated
with “+C”); (b) single lipped channel 88x86x40x42x12mm; (c) chords 2U88x86x40x42x12mm.

2 LACED BUILT-UP COLUMNS DESCRIPTION

All tested columns have the same 400 x 400 mm cross-section shown in Figure 1(a). The laced
column chords shown in Figure 1(c) were composed of two lipped channel members con-
nected with 4.8 mm diameter self-drilling screws spaced at each 220 mm. Chords, diagonals
and bracings are composed of lipped channel CFS with the measured dimensions (average) of
single lipped channel cross-section 88 x 86 x 40 x 42 x 12 mm in Figure 1(b). Diagonals and bra-
cings were connected to the chords with self-drilling screws. In order to evaluate the effect of
global buckling, two lengths of columns were tested: 12200 and 16200 mm, as shown in
Table 1. Laced columns were manufactured with two steel plate thicknesses t, for each column
length: 0.8 or 1.25 mm. The chord’s edge stiffeners were cut off in the region of the arrival of
diagonals and bracings, in order to enable connection. The laced columns were tested in hori-
zontal position, as illustrated in Figure 2.

3 EXPERIMENTAL ANALYSIS OF THE LACED COLUMNS

The built-up laced columns were tested in horizontal position, with compression load applied
by a servo controlled hydraulic actuator. The actuator was placed at one end of the column,
as shown in Figure 3, and a Dywidag steel rod crosses through the column length and is
attached to a reaction stiff plate placed at the opposite end of the laced column. Loading con-
dition was programed to smooth displacement control at 0.02 mm/s rate.
The end condition for axial compression test was actually a semi-rigid type for flexural rota-
tions (not perfect pined-pined). Rigid plate combined with spherical hinge was adopted at end
section as shown in Figure 3. Semi-rigid behavior was imposed, since the spherical hinge is not
able to allow actual free rotations for low loading condition.
Displacement transducers (DT) were applied for minor and major axis deflection measure-
ments, at mid-length and quarter-lengths of the column and close to the fixed end support.
The minor inertia plane was placed in horizontal position for the tests.
The strain gauge distribution included four sensors at each column end, placed externally in
the web element of the chords. At the mid-length were placed eight strain gages, four in the
internal side of the webs, thus performing four couples of strain gauges addressed to record
the onset of local buckling deformation.
The laced columns were manufactured with structural steel (commercial identification
ZAR345) with nominal yield stress fy = 345 MPa. The steel mechanical properties were meas-
ured through standard tensile tests of 14 coupons extracted from the chord members (7 coupons
for each thickness). The geometry of the coupons and the testing procedure were based on the
guidance provided by ABNT ISO 6892 (2013) standard. The average result of yield stress fy,
ultimate stress fu, Young modulus E and residual strain after failure εr are given in Table 2.

156
Figure 2. Lateral views of the laced built-up column placed in horizontal position for the test.

Table 1. Built-up laced column IDs.


Column ID L (mm) CFS

T12 x 0.8 12200 U88 x 86 x 40 x 42 x 12 mm lipped channel CFS, t = 0.8 mm


T12 x 1.25 12200 U88 x 86 x 40 x 42 x 12 mm lipped channel CFS, t = 1.25 mm
T16 x 0.8 16200 U88 x 86 x 40 x 42 x 12 mm lipped channel CFS, t = 0.8 mm
T16 x 1.25 16200 U88 x 86 x 40 x 42 x 12 mm lipped channel CFS, t = 1.25 mm

During the tests of the laced columns, it was possible to observe the development of typical
local buckling deformation mode along the webs of the chords before collapse. Although
observed in almost all the tested columns, it was more visible for the case of 0.8 mm thick
built-up laced columns, due to more slender CFS. Figure 4(a) shows the local buckling mode
developing along the chords. Laced columns with nominal lengths 12200 and 16200 mm devel-
oped clear global buckling deformation with collapse mechanism at the mid length. It was
possible to observe interaction between local (at the chord members) and flexural global buck-
ling modes.
Figure 5a) shows the horizontal displacements w recorded by displacement transducer at
mid-length of column T16 x 1.25, indicating almost null flexural displacements until (aprox.)
110 kN. As referred, the spherical hinges were not able to allow free rotations for low loading

Figure 3. Load application schematic arrangement.

157
Table 2. Average results of the standard tensile tests.
Thickness Yield Stress Ultimate Stress Young Modulus εr
t (mm) fy (MPa) fu (MPa) E (GPa) %

0.8 370 477 198 20


1.25 375 479 215 18

Figure 4. Column T16 x 0.8, nominal length 16200 mm: (a) Local buckling deformation along the
chords during the test, (b) collapse mechanism at mid length.

condition. Higher loading developed flexural buckling behavior which forced and liberated
rotations at the ends of the laced column.
The strain gauges registered linear (compression) behavior until the onset of local buckling.
Figure 5(b) shows the strain gauges results for the built-up laced column T16 x1.25, where
nearly linear behavior can be observed until (aprox.) 85 kN applied load. The beginning of
local buckling effect was recorded by couple of strain gauges 12RI/12RE and 10RI/10RE. The
registered collapse load was 196 kN.
Finally, Table 3 shows a summary of the results for the tested built-up laced columns,
including the records of (i) the collapse load (Puexp), (ii) the mid-length flexural displacement
at the collapse load level and (iii) the loading level for which the onset of local buckling was
identified (with the help of strain gage measurements).

Figure 5. Experimental results of column T16 x 1.25: (a) Load vs. horizontal displacement recorded at
the column mid-length, (b) Load vs. strain measurements (με) at mid-length section of the laced column.

158
Table 3. Summary of the test results of the built-up laced columns.
Buit-up laced Mid-length Local buckling load
Collapse load Puexp (kN)
column displacement (mm) (kN)*

T12 x 1.25 233 31 80


T12 x 0.8 89 45 25
T16 x 1.25 196 71 85
T16 x 0.8 85 59 20

* Applied load related to the local buckling onset, according with strain gage couples measurements.

5 ELASTIC BUCKLING ANALYSIS OF THE LACED COLUMN

In order to evaluate the global buckling of laced columns, numerical model was developed
with the help of the finite element method with beam elements (BFEM). For this, SAP2000
(2011) computational program was achieved. The laced columns were taken as pined-pined
and they were described by the centroid axis of the chords, as indicated in Figure 1(a). The
first two buckling modes are flexural buckling around X and Y axes, the third buckling mode
is torsional and the forth one is the second flexural buckling mode around X-axis. Buckling
loads from numerical BFEM model are presented forward in the Table 4.

6 ANALYTICAL METHODS FOR GLOBAL BUCKLING

In this section, the analytical values of the elastic critical loads for the built-up laced columns are
computed according to methods proposed by Timoshenko (1963) and Eurocode (CEN, 2006b).
The built-up laced columns were considered as pined-pined end condition. As previously reported,
the steel properties are E = 198GPa and 215GPa, respectively for 0.8 and 1.25 mm steel plates.
According with Timoshenko (1963), the critical load for laced column with “N” lacing,
hinged ends and shear effect, is given by Equation 1:

π2 EI 1
Pcr ¼   ð1Þ
l2 π2 EI 1 b
1þ 2 þ
l Ad Esen cos2  aAb E

where f is the angle between the diagonals and the batten bracings; Ad is the cross-section area of
two diagonal members (one at each face of the laced column); Ab is the cross-sectional area of
two batten bracings (one at each face of the laced column); b is the length of batten bracings and
I is the moment of inertia of the cross-section of the laced column (considering the four chords).

Table 4. Global buckling critical loads: X- axis and Y- axis (Figure 10e) (kN).
X - axis Y - axis

Column ID Eurocode Timoshenko BFEM Eurocode Timoshenko BFEM

T6 x 1.25 1454 (0.90) 1584 (0.98) 1618 2586 (0.96) 2532 (0.94) 2698
T6 x 0.8 862 (0.90) 965 (1.01) 959 1533 (0.96) 1584 (0.99) 1600
T12 x 1.25 410 (0.93) 450 (1.02) 442 749 (0.96) 762 (0.98) 777
T12 x 0.8 243 (0.92) 269 (1.02) 263 444 (0.95) 459 (0.99) 465
T16 x 1.25 236 (0.93) 260 (1.02) 254 434 (0.95) 445 (0.98) 456
T16 x 0.8 140 (0.93) 155 (1.03) 151 257 (0.95) 266 (0.99) 270

159
Eurocode EC3 Part 1-3 (CEN, 2006a), dedicated to cold-formed steel structures, does not
present guidance to the design built-up laced columns. Therefore, Eurocode 3 Part 1-1 (CEN,
2006b) design procedure for this type of structural system will be considered. In this condition,
the critical load for laced columns is given by Equation 2:

π2 EIeff
Pcr ¼ Ieff ¼ 0:5Ach h20 ð2Þ
L2

Ieff is the effective moment of inertia of laced built-up members; h0 is the distance between the
centroids of the chords and Ach is the cross-section area of one chord.
Equation 3 gives the effective critical load of built-up column considering shear effect (Pcr,v):

1 nEAd ah0 2
Pcr;v ¼ Sv ¼ ð3Þ
1 1 Ad h30
þ d3 1 þ
Pcr Sv Av d 3

Sv is the shear stiffness of built-up columns with “N” lacing arrangement; “n” is the number
of lacing plans; Ad is the diagonal cross-sectional area and Av is the horizontal bracings cross-
sectional area, of a single plan; “a” is the length of one panel; “d” is the diagonal length.
Table 4 shows the computed results of the critical flexural buckling loads, for the laced col-
umns, according with the proposed procedures from Timoshenko and Eurocode. Buckling
loads from numerical BFEM model and the ratio between critical buckling loads related to
BFEM results (inside parenthesis) are also included, from which one may observe quite simi-
lar results for both X and Y-axis. These results reveal the analytical method is able to provide
accurate results in the present case, as well as the Eurocode procedure gives lower results if
compared to analytical solution and numerical BFEM analysis.

7 COMPARISON BETWEEN BFEM MODEL AND EXPERIMENTAL


COLLAPSE LOADS

The laced column global buckling mode from the numerical buckling analysis was used as the
initial geometric imperfection for the BFEM elastic geometrical nonlinear analysis. As recom-
mended in Eurocode 3 part 1-1 (CEN, 2006b), the initial geometrical imperfection was
adopted as L/500. The collapse of the built-up laced columns is recognized when one of the
chord members reaches its axial compressive strength. So, the chord members compression
forces, obtained from the BFEM elastic non-linear analysis, must be compared with the
strength of one chord all along the loading path of the laced column.
Additional experimental and numerical investigation of the axially compressed behavior of the
built-up lipped channel chords with 480mm length (Figure 1(c)) was performed, in order to obtain
strength of one chord. The results indicated that the direct strength method (DSM) proposed by
Schafer and Peköz (1998) is able to estimate the ultimate strength of this type of built-up lipped
channel section, for fully-composite and non-composite CFS condition. In the fully-composite sec-
tion, the actual screws connection condition between the CFS members was admitted as fully
effective and transformed in double thickness (2t) plate element in the portion of contact between
web and flange of the channel elements, as shown at Figure 6(a). The non-composite condition
considers isolated single channel section, as shown at Figura 6(c). As the members of the chords
are connected with screws at discrete points, the actual condition is in between the non-composite
and fully-composite values. Although the fully-composite CFS presented good agreement, the
better results of the computed column strength was obtained for the non-composite built-up chan-
nel CFS.
The analysis of the buckling modes and the corresponding critical buckling load Pcr is the
first step for the design calculation of the chord using the DSM. Buckling analysis was per-
formed taking into account fixed end condition, using generalized beam theory with the help

160
Figure 6. a) GBTUL fully-composite section (double thickness in the contact portion); b) local buckling
deformation mode for fully-composite section; c) GBTUL non-composite section (members isolated); d)
local buckling deformation mode for non-composite section.

of GBTUL computational program from Bebiano et al. (2018). Non-composite condition was
admitted, considering the single section (members computed isolated), as it presented better
experimental results. Both lipped and unlipped channel CFS chord member were considered,
due to localized suppression of edge stiffeners at the connections. The Young modulus was
assumed as E = 198 and 215 GPa, respectively for 0.8 and 1.25 mm steel plates, according with
standard tensile test results of specimens obtained from the laced columns (Table 2). Poisson
ration is admitted as 0.3. Figure 6 (a) and (b) shows the GBTUL fully-composite section and
its local buckling mode, with double thickness at the contact. Figure 6 (c) and (d) shows the
GBTUL non-composite section (members computed separately) and its local buckling mode.
As the local buckling was predominant in the present case, Table 5 shows the values of PcrL
for non-composite section, for single U. As the laced column has 400 mm modulation (see
Figure 2), the compression strength of the chord was computed for constrained end condition
and length L = 400 mm. Table 4 also presents the collapse load Pn obtained from DSM, for
single U, for one chord (2U’s) and for four chords (8U’s). It was considered the measured
yield stress from tensile tests fy = 370 and 375 MPa, for t = 0.8 and 1.25 mm respectively.
Table 6 shows the experimental and numerical collapse loads for the tested built-up laced
columns, considering the chord cross-sections with built-up lipped or unlipped channel mem-
bers from Table 5. It is also included the ratio between numerical and experimental results,
inside parenthesis. It is observed that the numerical results for unlipped channel chords pre-
sents adequate comparison with the experimental results. These results are sustained by the
evidence of collapse mechanism developed at the cross-section with suppression of edge stiff-
eners. In addition, the last column of Table 6 displays the results of the laced column strength
without global buckling effect, taking the axial compression strength of the chords from
Table 5 (for built-up unlipped channel option). As expected, this solution gives very much
inaccurate comparison with the experimental data.

Table 5. DSM column strength of the built-up chord members with constrained end condition, length
L = 400 mm, considering non-composite behavior.
Thickness GBTUL DSM
(mm) Single U Single U 1 Chord (2U’s) 4 Chords (8U’s)
Built-up CFS
PcritL Pn (kN) Pn (kN) Pn (kN)

Lipped channel U 88 x 86 x 42 x 40 x 12 0.80 13.4 28.8 57.5 230.1


Unlipped channel U 88 x 86 x 42 x 40 0.80 7.0 20.6 41.1 164.5
Lipped channel U 88 x 86 x 42 x 40 x 12 1.25 54.0 63.4 126.9 507.4
Unlipped channel U 88 x 86 x 42 x 40 1.25 26.8 44.9 89.8 359.0

161
Table 6. Numerical and experimental collapse loads for the tested built-up laced columns (kN).
Built-up laced No global
Experimental BFEM Non-linear analysis
column buckling

Thickness Experimental col- Lipped Unlipped Unlipped


Column ID
(mm) lapse load channel channel channel *

T12 x 1.25 1.25 233 311 (1.34) 248 (1.07) 359 (1.54)
T12 x 0.8 0.8 89 152 (1.71) 118 (1.33) 164 (1.84)
T16 x 1.25 1.25 196 207 (1.05) 186 (0.95) 359 (1.83)
T16 x 0.8 0.8 85 115 (1.35) 98 (1.15) 164 (1.93)

* See last column of Table 4 for unlipped channel (4 chords).

8 CONCLUSIONS

Elastic buckling analysis of steel laced column was performed with analytical and numerical
solutions. The obtained results indicated the tested columns perform little influence of the
shear effect and the critical buckling load may be accessed with the help of available analytical
equations from Eurocode (CEN, 2006b) and Timoshenko (1963).
As the collapse of the built-up laced columns is recognized when one of the chord members
reaches its axial compressive strength, GBTUL and DSM were used to obtain the strength of
the built-up CFS chord member. The DSM rules are able to estimate the ultimate strength of
this type of built-up lipped channel section, adopted as chords of the laced columns. The
better results of the computed chord strength were obtained for the non-composite built-up
channel CFS.
Experimental results clearly indicate the presence of global and local buckling during the
tests of laced columns. Large elastic deformation developed before plastic localized collapse at
the unlipped channel section (close to diagonal-chord connection).
The tested columns, with nominal length of 12200 and 16200mm, developed important geo-
metric nonlinear behavior from the flexural buckling mode, which was computed with the
help of BFEM analysis. The laced column strength was finally computed taking the influence
of the local buckling in the chord members, with the help of DSM equations. The obtained
results indicate acceptable comparison between computed and experimental data with
Pu/Puexp ranging from 0.95 to 1.33 (Table 6). It must be observed that the better results of the
computed laced column strength was obtained for the unlipped built-up channel CFS, follow-
ing experimental evidence of the collapse mechanism developed at these sections.

ACKNOWLEDGEMENTS

This study was financed in part by the Coordenação de Aperfeiçoamento de Pessoal de Nível
Superior- Brasil (CAPES) - Finance Code 001 and the National Council for Research and
Technology, CNPq (Process 161975/2015-1). In addition, the authors would like to thank
GYPSTEEL Company for the supply of the built-up laced columns.

REFERENCES

ABNT, 2013. ISO-6892: Materiais metálicos – ensaios de tração a temperatura ambiente. Associação
Brasileira de Normas Técnicas, Rio de Janeiro, Brasil (in Portuguese).
Aslani, F., Goel, S. C., 1991. An analytical criterion for buckling strenght of built-up compression mem-
bers. Fourth quarter, Engineering Journal, AISC.
Bleich, F., 1952. Buckling strength of metal structures, McGraw Hill, New York. pp. 174.

162
Bonab, A.P., Hashemi, B.H., Hosseini, M., 2013. Experimental evaluation of the elastic buckling and
compressive capacity of laced columns. Journal of Constructional Steel Research, vol. 86, 66–73.
Bebiano, R., Camotim, D., Gonçalves, R., 2018. GBTUL 2.0 – a second-generation code for the GBT-
based buckling and vibration analysis of thin-walled members, Thin-Walled Structures, 124, 235–253.
CEN, 2006a. EN 1993-1-3:2006 - Eurocode 3: Design of steel structures. Part 1-3: General rules, supple-
mentary rules for cold-formed thin gauge members and shetting. European Committee for Standard-
ization, Bruxelas, Bélgica.
CEN, 2006b. EN 1993-1-1:2006 - Eurocode 3: design of steel structures. Part 1-1: general rules and rules
for buildings. European Committee for Standardization, Bruxelas, Bélgica.
Engesser, F., 1891. Die knickfestigkeit gerader stabe (in german). Zentralbl bauverwaltung, Germany.
Gjelsvik, A., 1991. Stability of built-up columns. ASCE Journal of Engineering Mechanics, vol. 117,
1331–1345.
Hashemi, B.H., Jafari, M.A., 2009. Experimental evaluation of elastic critical load in batten columns.
Journal of Constructional Steel Research; vol. 65, n. 1, 125–131.
Hashemi, B.H., Jafari, M.A., 2012. Evaluation of Ayrton–Perry formula to predict the compressive
strength of batten columns. Journal of Constructional Steel Research, vol. 68, n. 1, 89–96.
Hashemi, B.H., Bonab, A.P., 2013. Experimental investigation of the behavior of laced columns under
constant axial load and cyclic lateral load. Engineering Structures, vol. 57, 536–543.
Kalochairetis, K.E., Gantes, C.J., Lignos, X.A., 2014. Experimental and numerical investigation of
eccentrically loaded laced built-up steel columns. Journal of Constructional Steel Research, vol. 101,
66–81.
Paul, M., 1995. Buckling loads for built-up columns with stay plates. ASCE, Journal of Engineering
Mechanics, vol. 121, n. 11, 1200–1208.
Razdolsky, A.G., 2005. Euler critical force calculation for laced columns. ASCE Journal of Engineering
Mechanics, vol. 131, n. 10, 997–1003.
Razdolsky, A.G., 2010. Flexural buckling of laced column with serpentine lattice. The IES Journal part
A: Civil & Structural Engineering, vol. 3, n. 1, 38–49.
Razdolsky, A.G., 2011. Calculation of slenderness ratio for laced columns with serpentine and crosswise
lattices. Journal of Constructional Steel Research, vol. 67, n. 1, 25–29.
Razdolsky, A.G., 2014a. Flexural buckling of laced column with fir-shaped lattice. Journal of construc-
tional steel research, vol. 93, 55–61.
Razdolsky, A.G., 2014b. Revision of Engesser’s approach to the problem of Euler stability for built-up
columns with batten plates. ASCE Journal of Engineering Mechanics, vol. 140, n. 3, 566–574.
SAP 2000 Version 15, 2011. Getting Started with SAP 2000 Linear and Nonlinear Static and Dynamic
Analysis and Design of Three-Dimensional Structures. Computers and Structures Inc. Berkeley, Cali-
fornia, USA.
Schafer, B.W., Peköz, T., 1998. Direct strength prediction of cold-formed steel members using numerical
elastic buckling solutions. In: 14th International Specialty Conference on cold-formed steel structures, St
Louis, Missouri, USA, 69–76.
Timoshenko, S.P., Gere, J.M., 1963. Theory of elastic stability. Mcgraw-Hill. New York, USA.

163
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Local-distortional buckling interaction of cold-formed steel


columns design approach

E.M. Batista & G.Y. Matsubara


COPPE, Civil Engineering Program, Federal University of Rio de Janeiro, Brazil

J.M.S. Franco
Urbanism and Architecture Department, Federal Rural University of Rio de Janeiro, Brazil

ABSTRACT: Local-distortional interaction buckling mode (LD) in steel cold-formed (CFS)


columns is the objective of the present research. Experimental and FEM results are the basis of
the original design proposition, based on both the direct strength method (DSM) and the Winter-
type strength equation. The present results represents improvement of solution recently published
by the authors, with modified equations bringing more simple formulation as well as keeping
acceptable structural safety condition. The proposed solution offers strength surface bridging the
widely accepted DSM solutions for local and distortional buckling. This concept is based on the
main variable of LD interaction of lipped channel columns, the ratio between distortional and
local buckling slenderness factor RλDL = λD/λL. After identification of the range of RλDL for which
columns are significantly affected by the LD interaction, the Winter-type surface equation is cali-
brated with the help of experimental data and FEM results.

1 INTRODUCTION

The design procedures of steel cold-formed section (CFS) members derives from buckling
behaviour, because of its slenderness properties. Thin-walled CFS members develop local L, dis-
tortional D and global G buckling modes identified by their critical loading and buckling shape
illustrated in Figure 1. In addition, buckling modes interaction (LG, DG, LD or LDG) pro-
duces strength reduction if compared to single mode buckling development, currently identified
as the strength erosion of thin-walled CFS member. The combination of buckling modes with
characteristic deformed shapes, as is the case of short L, long G and intermediate D semi-wave
lengths, conducts to interaction behaviours with distinct severity, obliging definition of particu-
lar solutions regarding each type of interaction.
The identification of the structural stability properties of a CFS column follows its “signa-
ture curve”, obtained with the help of numerical analysis, in order to identify the buckling
loads associated to the column semi-wave lengths, Pcr vs. L. The signature curve reveals more
than the half wavelength but the actual CFS buckling mode shape, allowing clear information
about the column stability. The most relevant numerical methods to obtain the signature
curve are (i) the finite strip (FSM) and (ii) the generalized beam theory (GBT) methods, avail-
able with the computational tools (free access) CUFSM, Li & Schafer (2010), and GBTUL,
Bebiano at al. (2008), respectively. The first order linear elastic buckling analysis results con-
figures the basis of the structural design of thin-walled steel columns and beams.
The closer the buckling loads PcrL, PcrD and PcrG, respectively related to L, D and G modes,
the more probable is the occurrence of buckling interaction, allowing stronger nonlinear post
buckling behaviour of the column. In addition, there is no general rule to take into account all
the possible buckling interaction effects of CFS columns, essentially because of the different
nature of the modes interaction behaviour, by combining short-long or short-medium half
wavelengths, LG and LD, for example. This is the reason LG buckling modes interaction

164
Figure 1. Buckling modes of lipped channel CFS columns: local L, distortional D, global flexural and
flexural torsional G.

received safe and accurate design procedures, since much time ago, and only more recently LD,
DG and LDG interactions are considered with more attention.
The present research aims at contributing for design rules for lipped channel CFS columns
experiencing LD interaction, considering almost null presence of global buckling. This condition is
assured for columns with PcrG sufficiently higher than PcrL and PcrD or, in terms of slenderness
coefficient, for λG lower enough than λL and λD, with λG ¼ (Py/PcrG)0.5, λL ¼ (Py/PcrL)0.5 and
λD ¼ (Py/PcrD)0.5, Py ¼ Afy is the column squash load and fy the steel yielding stress. Based on
FEM computation and available experimental data, the research methodology and results are
thorough described in Matsubara et al. (2019). In the following, the sets of numerical and experi-
mental data are resumed, the development of the design concept is explained and improved rule
and equations for the calculation of lipped channel columns strength PnDL are presented.

2 EXPERIMENTAL DATA

The following sets of experimental results of lipped channel columns displaying LD buckling
interaction were considered: (i) Kwon & Hancock (1992) (5 columns); (ii) Young & Rasmus-
sen (1998) (2 columns); (iii) Loughlan et al. (2012) (20 columns); (iv) Young et al. (2013)
(26 columns) and (v) Salles (2017) (3 columns). The range of variation of the geometrical and
material properties of the tested columns are indicated in Table 1, which includes the ratio
between flange and web widths bf /bw and edge stiffener and web widths bs/bw, the column
length L, fy and E, respectively yielding stress and Young elastic modulus of the steel and the
recorded experimental column strength load Puexp. In addition, Table 2 includes the slender-
ness factors related to L, D and G buckling modes, respectively λL, λD and λG, as well as the
ratio between distortional and local buckling slenderness factor RλDL ¼ λD/λL. Finally, only
columns with cross-section geometry inside the followings ranges were taken into consider-
ation: 0.4  bf /bw 5 1.0 and 0.1  bs/bw 5 0.3 (the only exception are the columns tested by
Kwon & Hancock (1992), with bf /bw 5 0.1). This is so because (i) these are practical and usual
geometries and (ii) out of these ranges, L and D modes exhibit inappropriate behavior indu-
cing important loss of strength.

Table 1. Ranges of the geometric and material properties and experimental strength of the tested lipped
channel columns.
Reference bf/bw bs/bw L(mm) fy (MPa) E (GPa) Puexp (kN)

Kwon & Hancock (1992) 0.75 0.04-0.06 400-800 590 210 49-55
Young & Rasmussen (1998) 0.5 0.12 616-1500 550 180 99-102
Loughlan et al. (2012) 0.35-0.63 0.07-0.12 1000-1800 209 193 27-33
Young et. al (2013) 0.45-1.00 0.1-0.3 615-2500 336-590 203-213 40-309
Salles (2017) 1.0 0.1 2529 342-348 180 29-33

165
Table 2. Ranges of variation of the slenderness factors and DL slenderness factor ratio of the tested
lipped channel columns.
Reference λL λD λG λmáx RλDL

Kwon & Hancock (1992) 2.64 - 2.70 1.87 - 3.32 0.20 - 0.40 2.64 - 3.32 0.71 - 1.23
Young & Rasmussen (1998) 1.55 - 1.60 1.24 - 1.27 0.63 - 0.94 1.55 - 1.60 0.77 - 0.82
Loughlan et al. (2012) 1.50 - 2.70 0.97 - 1.46 0.26 - 0.63 1.50 - 2.70 0.48 - 0.75
Young et. al (2013) 0.87 - 3.36 0.65 - 2.41 0.24 - 1.30 0.87 - 3.36 0.58 - 0.98
Salles (2017) 2.01 - 2.16 2.24 - 2.28 1.06 - 1.12 2.24 - 2.28 1.06 - 1.11

In particular, the experimental results from Salles (2017) were thoroughly described by Matsu-
bara et al. (2019) and one may observe clear evidence of LD buckling interaction. These are
lipped channel columns included in Tables 1 and 2, with fixed-fixed end condition, tested through
displacement control hydraulic system. The length L was designed to give close local and distor-
tional buckling loads, with the ratio between distortional and local buckling slenderness factor
RλDL ¼ λD/λL ¼ 1.1, at the same time global buckling mode is much higher, with PcrG ≈ 4.2PcrD.
Figure 2 shows local and distortional buckling development from column C2 from the tests per-
formed by Salles, the former from couples of strain gages placed in the web and the last from
displacement transducers, with the measuring devices placed at the mid length section of the
column. These results show (i) the onset of local buckling is clearly observed at (approx.) 20kN
load level (Figure 2(a)) and (ii) distortional buckling displaying typical inward deformation since
the very beginning of the test in Figure 2(c) (observe the signs of displacement transducers meas-
urements in Figure 2(d)). Buckling mode shapes L and D were visually observed and recorded
during the tests.

3 FEM MODEL RESULTS

Finite element SHELL181 from Ansys finite element package, Ansys (2009), was applied to
perform computational analysis of thin-walled lipped channel columns. SHELL181 is suitable
for thin-shell structures linear analysis, as well as geometric and material nonlinearity. The
analysis was performed in full integration method, as indicated by Ansys (2013) theory refer-
ence. The von Mises isotropic hardening plasticity model considered in the present analysis is
available in the FEM software package. FEM mesh is 5 mm width quadrilateral elements,
according with previous results from Fena (2011) and Silvestre et al. (2012). Convergence tests
confirmed the accuracy of the mesh.
Material ductility was introduced with (i) bilinear ductile steel elastoplastic development or
(ii) nonlinear ductile response based on experimental data from coupon tensile tests, from
Salles (2017). The FEM results of the column behaviour with options (i) and (ii) display very
similar results, with clear prevalence of elastic post buckling behaviour, which justify analo-
gous results with options (i) or (ii). Inelastic effect of thin-walled structural members only
arises at the very end of the loading, close to the ultimate limit load, with development of the
final collapse mechanism.
Initial geometrical imperfections were taken from the computed critical buckling mode
shape, with maximum amplitude 0.1 of the cross-section thickness t. This means that for
the FEM analysis of the experimental sets of columns described in Tables 1 and 2, the
initial geometric imperfection follows distortional or local buckling shape, respectively
for λD > λL or λL > λD. For the case of columns displaying combined LD critical buck-
ling mode, with λD ≈ λL, this is the imperfection geometric shape adopted for numerical
nonlinear analysis. Global buckling is never concerned in the sets of experimental results
(global buckling slenderness factor λG is always much lower than λD and λL). In add-
ition, Riks (1979) method was taken to access the ultimate limit loading, from now on
nominated FEM-based column strength PuFEM.

166
Figure 2. Experimental evidence of LD interaction mode from tested lipped channel column C2, Salles
(2017): (a) records of couple of strain gages E1-E2 show local buckling development, (b) couple of strain
gages E1&E2 placed in the web, mid length of the column; (c) D4-D5 displacement transducers show devel-
opment of inward distortional buckling, (d) displacement transducers D4&D5 at mid length of the column.

Column end conditions of the FEM model was fixed-fixed, according with fully restricted
displacement experimental condition, as described by the authors of the sets of tests. For this
purpose, stiff 25 mm thick plates were attached to the end sections of the columns and point
compressive loads were applied at the centroid position of both end sections of the column.
Figure 3 shows the shell finite element model as well as the cross-section typical geometry.
Figure 4 shows typical example of results from the FEM model for lipped channel columns.
Figure 4(a) displays (i) the evolution of the local buckling arising at the indicated load level I,
(ii) elastic LD interaction developing from load level II and (iii) final collapse configuration at
load level III. Figure 4(b) shows the FEM deformed shape at collapse of one of the columns
tested by Young et al. (2013), with clear evidence of LD buckling interaction mode.
Figure 5 shows complimentary comparison of FEM and experimental results from one of
the columns tested by Salles (2017). In this case, FEM and experimental results of displace-
ment transducers moving along the column length are in very good agreement, showing the
evidence of LD interaction mode by displaying combined mode shape L+D with increased
amplitudes for rising steps of the applied load.
The developed FEM model reproduced typical LD interaction buckling results for the com-
plete sets of experimental tests included in Tables 1 and 2, with very close agreement between

167
Figure 3. Ansys (Shell 181) FEM model of fixed-fixed lipped channel CFS columns.

Figure 4. Examples of FEM results of a column tested by Young et al. (2013): (a) load vs. distortional
buckling parametric displacement d/t at mid-length, complimented by the evolution of the deformed pattern
showing local (point I) to distortional buckling (point III) modes; (b) deformed configuration at collapse.

numerical and experimental ultimate loads, PuFEM and Puexp, as can be accessed in Matsubara
(2019). These results allowed additional computation of a planned set of 275 columns with 0.4
≤ bf /bw < 1.0 and 0.1 ≤ bs/bw < 0.3 (usual lipped channel sections), designed for the occurrence
of LD buckling interaction. The complete sets of experimental and numerical results permitted
to develop and test analytical solution for the design of lipped channel columns under LD
interaction, as presented in the following.

4 DIRECT STRENGTH METHOD BASED DESIGN

The direct strength method DSM, Schafer (2006), is a widely accepted procedure to deal with
the structural strength of CFS members. DSM solution for columns under L and D buckling
are based on Winter-type equation, respectively Equations 1 and 2. Global buckling is
addressed by the DSM with the help of widely accepted general solution of steel columns, by
providing the reduction strength factor χ ≤ 1.0, according with the slenderness factor λG

168
Figure 5. Experimental and FEM results of tested column from Salles (2017). Displacements measured by
transducers D2 and D4 (see Figure 2(d)): (a) FEM results and (b) experimental records of D2 confirm pres-
ence of L buckling mode; (c) FEM results and (d) experimental records of D4 shows evidence of D mode.

related to flexural or flexural torsional buckling. LG buckling interaction of CFS columns is


a well stablished procedure solved by appropriate combination of χ factor with Equation 1.
!
0:15 Py
PnL ¼ 1  0:8 ð1Þ
λL λ0:8
L
!
0:25 Py
PnD ¼ 1  1:2 ð2Þ
λD λ1:2
D

Concerning LD buckling interaction of CFS lipped channel columns, the most relevant pro-
positions found in recent literature are those from Schafer (2002), Martins et al. (2017) and
Matsubara et al. (2019). In the present research, the following conditions were considered to
enable easy to apply and accurate proposition, in close agreement with Matsubara et al.
(2019) findings.
a. Column strength for LD buckling interaction is based on the Winter-type Equation 3.
b. A and B parameters in Equation 3 are based on the main variables of the problem: (i) the
slenderness ratio RλDL ¼ λD/λL and (ii) the column slenderness λmax ¼ max (λD; λL).
c. The range of lipped channel columns affected by LD buckling was previously identified by
Martins et al. (2017) and additionally confirmed by Matsubara et al. (2019). The range of LD
interaction is defined by the slenderness ratio factor: 0.45 ≤ RλDL ≤ 1.05.
d. For slenderness ratio RλDL < 0.45 local buckling L is dominant and no LD interaction
should be considered. In this case the column strength takes into account Equation 1.
e. For slenderness ratio RλDL > 1.05 distortional buckling D is dominant and no LD interaction
should be considered. In this case the column strength takes into account Equation 2.
f. Finally, global buckling is not the case for the examined columns, with PcrG /PcrD (as well
as PcrG /PcrL) at least higher than 3.0, thus ensuring single LD buckling interaction mode.
 
A Py
PnLD ¼ 1 B B ð3Þ
λ λ

169
Expressions for the parameters A and B in Equation 3 were developed with the help of the
above described and calibrated FEM model. For this, a representative set of 275 lipped chan-
nel columns was developed, covering the following conditions: (i) LD slenderness interaction
factor RλDL ranging from 0.27 to 1.5; (ii) column slenderness λ ¼ λmax ¼ max(λD; λL) equal to
1.0, 1.5, 2.0 and 2.5; (iii) flange-web width ratio 0.4 ≤ bf /bw < 1.0; (iv) edge stiffener-web width
ratio 0.1 ≤ bs/bw < 0.3. The columns’ properties to enable such representative selection were
controlled by their cross-section geometry, column length L and material yielding stress fy.
Original solution from Matsubara et al. (2019) implies polynomial A and B parameters with
3rd and 4th degree, respectively, resulting in very accurate results. Recent results of the investi-
gation allowed improved solution with simpler equations and keeping acceptable accuracy of
the method. Equations 4 and 5 show the proposed solution for A and B parameters, to be
applied in the LD column strength Equation 3.

A ¼ 0:40RλDL  0:17 ð4Þ

B ¼ 2:26R2λDL þ 4:06RλDL  0:57 ð5Þ

The proposed solution with Equations 3, 4 and 5 ensures the continuous strength surface
illustrated in Figure 6 in parametric format PnLD /Py, based on the main variables RλDL and
λ ¼ λmax ¼ max(λD; λL) and performing natural transition between accepted DSM solutions for
single L and D buckling modes, Equations 1 and 2. The proposed surface includes usual ranges
of the lipped channel cross-section geometry (bf /bw and bs /bw) stablishing a solution that includes
acceptable dispersion of the results when compared with experimental and FEM sets. Original
solutions reported by Matsubara et al. (2019) include not only a unique solution for A and B but
also different equations for these parameters, addressed to restricted ranges of the flange-web
width ratio bf /bw. Of course, the more tailored the mathematical model the more precise the
comparison between analytical and reference results of the column strength. The choice of

Figure 6. Proposed strength surface PnLD/Py for LD interaction buckling of CFS lipped channel columns.

170
Figure 7. Comparison between the proposed rule PnLD for CFS lipped channel columns and (a) the
experimental results included in Tables 1 and 2; (b) the FEM results (those inside the range of validity of
LD interaction buckling: 0.45 ≤ RλDL ≤ 1.05).

a single surface solution take into account simplicity for design purposes and acceptable solution
for standards and codes of structural design practice.
Figure 7(a) shows the comparison between the proposed solution PnLD and the set of experi-
mental results included in Tables 1 and 2. Furthermore, Figure 7(b) illustrates the result of the
comparison between the proposed solution and the set of FEM results, and one may confirm
the accuracy of the proposition by the computed averages 1.08 and 0.99, and coefficient of
variation 0.15 and 0.10, respectively for experimental and FEM sets of lipped channel columns.

5 FINAL REMARKS

The presented results and comparisons between proposed formulation PnLD for LD buckling
interaction of CFS lipped channel columns is an extension of previous results recently pub-
lished by the authors in Matsubara et al. (2019). The proposed rule is essentially an integrated
solution for the design of CFS columns including local and distortional buckling modes, by
proposing the transition solution between DSM-based single L and D strength Equations 1
and 2. The proposed rule was tested and proved to be accurate by comparison with available
experimental results as well as when submitted to comparison with large set of FEM results.
Finally, the computation of the resistance factor on the basis of the load and resistance
factor design method (LFRD) confirmed adequate performance of the proposed approach,
with ϕ ¼ 0.83 (and γ ¼ 1/ϕ ¼ 1.21), in good agreement with the values recommended by both
the North American standard AISI (2016) and the Brazilian code ABNT (2010), respectively
ϕ ¼ 0.85 and γ = 1.2.

ACKNOWLEDGEMENTS

The first author acknowledges the financial support of CNPq, National Council for Scientific
and Technological Research, through scholarship for his Master degree research.

REFERENCES

ABNT. 2010. NBR 14762 Design of cold-formed steel structures. Associação Brasileira de Normas Téc-
nicas (in Portuguese).
AISI. 2016. North American Specification for the design of Cold-formed Steel Structural Members –
AISI S100, American Iron and Steel Institute, Washington, USA.

171
ANSYS. 2009. Version 12: SAS – Swanson Analysis Systems Inc.
ANSYS. 2013. ANSYS APDL Theory Reference, Release 15.0, Canonsburg, PA.
Bebiano, R.; Silvestre, N.; Camotim, D. 2008. GBT theoretical background. Available at http://www.civil.
ist.utl.pt/gbt/, (access on 8 February 2019).
Fena, R.P.T. 2011. Interacção Local/Distorcional em colunas de aço enformadas a frio com secção em
“Hat”. Master of Scienec dissertation. Technical University of Lisbon, Lisbon, Portugal.
Kwon, Y. B.; Hancock, G. J. 1992. Tests of cold-formed channels with local and distortional buckling.
Journal of Structural Engineering, v. 118: p. 1786–1803.
Li, Z.; Schafer, B. W. 2010. Buckling analysis of cold-formed steel members with general boundary con-
ditions using CUFSM: Conventional and constrained finite strip methods. In: Twentieth International
Specialty Conference on Cold-Formed Steel Structures, St. Louis, Missouri, USA, p. 1–15.
Loughlan, J.; Yidris, N.; Jones, K. 2012. The failure of thin-walled lipped channel compression members due
to coupled local-distortional interactions and material yielding. Thin-Walled Structures, v. 61: p. 14–21.
Martins, A.D.; Camotim, D.; Dinis, P.B. 2017. On the direct strength design of cold-formed steel col-
umns failing in local-distortional interactive modes. Thin-Walled Structures, v. 120: p. 432–445.
Matsubara, G.Y.; Batista, E.M.; Salles, G.C. 2019. Lipped channel cold-formed steel columns under
local-distortional buckling mode interaction. Thin-Walled Structures, v. 137: p. 251–270.
Riks, E. 1979. An incremental approach to the solution of snapping and buckling problems. Internacional
Journal of Solids and Structures, v. 15: p. 529–551.
Salles, G.C. 2017. Investigação analítica, numérica e experimental do modo de flambagem distorcional
em perfis formados a frio. Master of Science dissertation. COPPE, Federal University of Rio de
Janeiro, Rio de Janeiro, Brazil.
Schafer, B.W. 2002. Local, distortional and Euler buckling of thin-walled columns. Journal of Structural
Engineering, v. 128: 289–299.
Schafer, B.W. 2006. Designing cold-formed steel using the Direct Strength Method. In: Eighteenth Inter-
national Specialty Conference on Cold-Formed Steel Structures, Orlando, FL., USA.
Silvestre, N.; Camotim, D.; Dinis, P.B. 2012. Post-buckling behaviour and direct strength design of
lipped channel columns experiencing local/distortional interaction. Journal of Constructional Steel
Research, v. 73: p. 12–30.
Young, B.; Rasmussen, K.J.R. 1998. Design of lipped channel columns. Journal of Structural Engineer-
ing, v. 124: p. 140–148.
Young, B.; Silvestre, N.; Camotim, D. 2013. Cold-Formed Steel Lipped Channel Columns Influenced by
Local-Distortional Interaction: Strength and DSM Design. Journal of Structural Engineering, v. 139:
p. 1059–1074.

172
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Solutions to simplified von Karman plate equations

J. Becque
The University of Sheffield, Sheffield, UK

ABSTRACT: The non-linear von Karman equations describe the post-buckling behaviour
of thin plates, but a general solution cannot be found. In this paper some simplifying assump-
tions are made which follow from a consistent interpretation of von Karman’s effective width
concept. This reduces the coupled von Karman equations to a single equation and allows
a Winter-type equation to be derived for the ultimate capacity of single plates, accounting for
yielding and initial geometric imperfections.

1 SIMPLIFYING VON KARMAN’S EQUATIONS

The von Karman equations (sometimes referred to as the Föppl-von Karman equations) com-
prise a system of two coupled differential equations which describe the non-linear post-
buckling behaviour of thin elastic plates (Föppl 1907, von Karman 1910):

∂4 w ∂4 w ∂4 w
D þ 2 þ ¼ pz
∂x4 ∂x2 ∂y2 ∂y4
2 2 ð1Þ
∂ ’ ∂ ðw þ w0 Þ ∂2 ’ ∂2 ðw þ w0 Þ ∂2 ’ ∂2 ðw þ w0 Þ
þt  2 þ
∂y2 ∂x2 ∂y∂x ∂x∂y ∂x2 ∂y2
 2
∂4 ’ ∂4 ’ ∂4 ’ ∂2 w ∂2 w ∂2 w
þ 2 þ ¼E 
∂x4 ∂x2 ∂y2 ∂y4 ∂x∂y ∂x2 ∂y2
ð2Þ
∂2 w0 ∂2 w ∂2 w0 ∂2 w ∂2 w0 ∂2 w
þ2  
∂x∂y ∂x∂y ∂x2 ∂y2 ∂y2 ∂x2

In these equations w is the vertical plate deflection, w0 is the initial geometric imperfection, ϕ
is Airy’s stress function, E is the elastic modulus, t is the (constant) plate thickness, pz is the
lateral pressure on the plate and D is given by:

Et3
D¼ ð3Þ
12ð1   2 Þ

The x-y coordinate system is illustrated in Figure 1. The scope of this paper is limited to plates
under uniform in-plane compression.
Eqs. (1–2) account for imperfections, but not for material non-linearity. However, the biggest
impediment to their practical application is that, while solutions exist for a few special cases (e.g.
Levy 1942), a general solution is beyond reach. The von Karman equations consequently have
little relevance in practical design, which relies instead on the empirical Winter equation (1970):
 
Pu 1 0:22
¼ 1  1:0 ð4Þ
Py λ λ

173
Figure 1. Plate in axial compression.

where Pu is the ultimate capacity of the plate in compression, Py is the yield load and the slen-
derness λ is given by:
sffiffiffiffiffiffi
fy
λ¼ ð5Þ
σcr

In the above equation fy is the yield strength and σcr is the critical elastic local buckling stress.
The aim of this paper is to establish a link between the von Karman equations and
a Winter-type design equation. In order to achieve this, it is clear that some additional
assumptions leading to a simplification of Eqs. (1–2) are necessary. The inspiration for these is
provided by the effective width concept, credited to von Karman et al. (1932). This concept is
based on the observation that in the post-local buckling range the longitudinal stresses shift
towards the longitudinal edges of the plate and can thus be idealized as being carried by two
strips adjacent to those edges. The widths of these effective portions are obtained by equating
the total stress in the actual and idealized distributions (Figure 1). Failure is assumed to occur
when the effective strips yield. While this failure criterion will be employed later in the deriv-
ation, the major implication of this idealized stress distribution which proves useful for our
objectives at this stage is that the longitudinal membrane stress σx is only a function of the
transverse co-ordinate y and is constant along a ‘fibre’ in the longitudinal x-direction:

∂2 ’
σx ¼ ¼ f ð yÞ ð6Þ
∂y2

Integrating twice with respect to the y-coordinate yields the following form for Airy’s stress
function:

’ ¼ gð yÞ þ y:cðxÞ þ d ðxÞ ð7Þ

However, because of the symmetry of the problem, the mixed term in x and y has to vanish.
Consequently, the shear stresses in the plate have to be zero:

∂2 ’
τxy ¼  ¼0 ð8Þ
∂x∂y

This is consistent with the view that each longitudinal fibre acts independently, carrying
a constant stress along its length, while not partaking in any load sharing with its neighbours.
The final implication is that σy is independent of y:

174
∂2 ’
σy ¼ ¼ hðxÞ ð9Þ
∂x2

Two cases of boundary conditions are considered in this study. In both cases the loaded edges
(x = 0 and x = L) remain straight in the post-buckling range. This corresponds to the practical
case of a plate element in a long column, where ‘nodal lines’ develop in between buckled cells.
For the longitudinal edges two cases are considered: case A, where the edges are free to pull in
during the post-buckling stage (Figure 2), and case B, where the edges can move in while
remaining straight (Figure 3). Case A is most representative of a plate (e.g. a web) in an actual
column, given the limited out-of-plane bending stiffness of the adjacent plates (i.e. the flanges
in this particular example).
The elongation of a longitudinal fibre can be determined from the deflected shape as
follows:
82LU sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 3 9
 ð   ð qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi  < ð 0  2 =
1 1 1 4 ∂w
εx ¼ ds  L ¼ ðdxÞ2 þ ðdwÞ2  L ¼ 1þ dx5  L
L L L: ∂x ;
0
82LU 3 9 2 3
ð 0   ! = 1 ðL ∂w2
1 <4 1 ∂w 2
≈ 1þ dx5  L ≈ 4 dx  2U0 5
L: 2 ∂x ; 2L ∂x
0 0
ð10Þ

where U0 is the uniform end shortening, as shown in Figure 2, and s is the distance measured
along the fibre representing the mid-line of the plate. U0 can be related to the total applied
load P as follows:
with:

ðb
P ¼ t σx dy ð11Þ
0

with:
2L 3
ð 
E 4 ∂w 2
σx ¼  dx  2U0 5 ð12Þ
2L ∂x
0

The minus sign in Eq. (12) allows compressive stresses to be counted as positive.

Figure 2. Case A boundary conditions. Figure 3. Case B boundary conditions.

175
In case A, where the longitudinal edges are free to move in horizontally, no significant ten-
sile membrane stresses are expected to develop in the transverse direction (at least not in the
initial stages after buckling takes place) and σy is assumed to remain zero. Using Eqs. (12), (6),
(8) and (9), Eq. (1) becomes:
2 3
4 ðL  2  2 
∂ w ∂4 w ∂4 w 4 ∂w 5 ∂ ðw þ w0 Þ
γ þ 2 2 2 þ 4 ¼ 2U0  dx ð13Þ
∂x4 ∂x ∂y ∂y ∂x ∂x2
0

with:

2LD Lt2
γ¼ ¼ ð14Þ
Et 6ð1   2 Þ

Eq. (13) only features the plate deflections was an unknown function, while Airy’s stress func-
tion no longer appears.
When the boundary conditions are determined by case B, the transverse membrane stresses
σy are no longer negligible. Due to the assumed absence of shear stresses, however, the trans-
verse membrane stress can be related to the elongation of a transverse fibre employing
a rationale completely analogous to Eq. (10), leading to the following equation:
2 3
ðb  2
E 4 ∂w
σy ¼ 2U1  dy5 ð15Þ
2L ∂y
0

In the above equation U1 is the uniform shortening of the plate in the transverse direction, which
can be determined from the condition that, while transverse stresses are necessary along the edges
to keep the edges straight, their resultant is zero since there is no load applied in this direction:
ðL
t σy dy ¼ 0 ð16Þ
0

and thus, using Eq. (15):

ðL ðb  2
1 ∂w
U1 ¼ dxdy ð17Þ
2L ∂y
0 0

Using Eqs. (12), (15) and (8), Eq. (1) takes on the form:
2 3
4 4 ðL  2  2 
∂ w ∂ w
4
∂ w 42U0  ∂w 5 ∂ ðw þ w0 Þ
γ þ 2 þ ¼ dx
∂x4 ∂x2 ∂y2 ∂y4 ∂x ∂x2
0
2 3 ð18Þ
ðb  2  2 
∂w ∂ ðw þ w0 Þ
þ 42U1  dy5
∂y ∂y2
0

2 GEOMETRICALLY PERFECT PLATE WITH BOUNDARY CONDITIONS A

We consider the case of a square plate (b = L) without imperfections (w0 = 0) and we propose
the following approximate solution to Eq. (13):

176
πx πy
w ¼ A11 sin sin ð19Þ
L L

Eq. (19) is the solution to the classical St. Venant plate equation. In the context of Eq. (1), Eq.
(19) can be seen as the first term of a Fourier series, which, by virtue of being the solution to
the St. Venant equation, is dominant in the initial post-buckling range over the remaining
terms. Substituting Eq. (19) into Eq. (13) leads to:
  π 2 πx πy
π 4 L  π 4 πx 3 πy
4γ  2U0 sin sin ¼ A211 sin sin ð20Þ
L L L L 2 L L L

This equation can be re-arranged into:


"  2 #  
L πx πy πx πy L πx 3πy
2 3L
4γ  2U0 sin sin ¼ A11 sin sin  sin sin ð21Þ
π L L 8 L L 8 L L

It is now clear that Eq. (19) cannot be an exact solution to Eq. (13), since higher order Fourier
terms are necessary in Eq. (19) in order for the terms in sin(πx/L)sin(3πx/L) in Eq. (21) to
vanish. However, due to the orthogonality of the Fourier terms, the total coefficient of sin(πx/L)
sin(πx/L) in Eq. (21) can be set equal to zero, leading to:
"  2 #
8 L
A211 ¼ 4γ  2U0 ð22Þ
3L π

On the other hand, Eq. (11) results in:


2 3
ðL ðL  2
tE 4 ∂w
P¼ 2U0 L  dxdy5
2L ∂x
0 0
2 3
ðL ðL  π 2 πx πy
tE 4 ð23Þ
¼ 2U0 L  A211 cos2 sin2 dxdy5
2L L L L
0 0

Et π2
¼ 8U0  A211
8 L

Substituting Eq. (22) into Eq. (23) yields the load-elongation relationship in the post-buckling
range:

U0 4  π 2
P ¼ Et þ γ ð24Þ
3 3 L

In the initial pre-buckled state: P = EtU0. Consequently, Eq. (24) shows that the predicted
initial post-buckling stiffness equals 1/3 of the pre-buckling stiffness. Marguerre (1937)
reported a more exact value for this ratio, which depends on the Poisson’s ratio υ, but ranges
from 0.34 (for υ = 0.5) to 0.41 (for υ = 0). For steel plates (υ = 0.3) the value is 0.38, which
differs by 12% from our estimate.
The plate first buckles when A11 = 0, which according to Eq. (22) happens when:
 π 2
U0 ¼ 2γ ð25Þ
L

177
Substituting this value into Eq. (24) yields the expected result:

 π 2 4π2 E  t 2
Pcr ¼ 2Etγ ¼ Lt ¼ σcr A ð26Þ
L 12ð1   2 Þ L

where A is the cross-sectional area of the plate in the transverse direction.


The profile of the longitudinal membrane stresses is given by Eq. (12):
2 3
 π 2 ðL πx πy  2 
E 4 2 5 E π 2 πy
σx ¼ 2U0  2
A11 cos2
sin dx ¼ 2U0  A11
2
sin ð27Þ
2L L L L 2L 2L L
0

According to the effective width concept failure occurs when the maximum longitudinal mem-
brane stress (occurring at y = 0 and y = L according to the above equation) reaches the yield
stress:

EU0
σx;max ¼ ¼ fy ð28Þ
L

Using the load-shortening relationship Eq. (24), the above equation can be rewritten as:
 π 2
E 3Pu
fy ¼  4γ ð29Þ
L Et L

and, with the help of Eq. (26) as:

3Pu 2σcr
1¼  ð30Þ
Ltfy fy

With the yield load Py = Afy = Ltfy, Eq. (30) becomes:


 
Pu 1 2
¼ 1þ 2 ð31Þ
Py 3 λ

where the slenderness λ was previously defined in Eq. (5). Eq. (31) is reminiscent of the Winter
equation and both equations are compared in Figure 4. It is seen that Eq. (31) results in sig-
nificantly higher predictions of the plate capacity. It will be demonstrated in Section 4 that
this is mainly due to the absence of geometric imperfections.

3 GEOMETRICALLY PERFECT PLATE WITH FOUR STRAIGHT EDGES

Eq. (19) can also be substituted in Eq. (18) describing the buckling behaviour of a plate with
boundary conditions B. It is again assumed that the plate is square (b = L) and geometrically
perfect (w0 = 0). Applying the same methodology as employed in Section 2 results in the load-
shortening relationship:

π 2 U0
P ¼ Et γ þ ð32Þ
L 2

Eq. (32) shows that, for these boundary conditions, the initial stiffness in the post-buckling
range is half the pre-buckling stiffness, thus confirming the result found by Maguerre (1937).

178
Figure 4. Comparison of theoretical predictions with Winter curve.

Using failure criterion Eq. (28) can be shown to result in the following equation for the cap-
acity of the plate:
 
Pu 1 1
¼ 1þ 2 ð33Þ
Py 2 λ

Eq. (33) is also plotted in Figure 4, although it should not be directly compared with the
Winter equation, as the longitudinal edges were not kept straight in Winter’s tests.

4 GEOMETRICALLY IMPERFECT PLATE WITH BOUNDARY CONDITIONS A

A square plate with boundary conditions A is assumed to have an imperfection w0 which is


described by the following equation:
πx πy
w0 ¼ A0 sin sin ð34Þ
L L

Substituting the proposed solution Eq. (19), as well as the above expression Eq. (34) into Eq.
(13) leads to the following result:
"  2 #  2
3L 3 3L L L
A þ A0 A11 þ 4γ  2U0
2
A11  2U0 A0 ¼ 0 ð35Þ
8 11 8 π π

Eq. (23) is, of course, still valid for this case and leads to:

8L P
A211 ¼ U 0  ð36Þ
π2 Et

Substituting Eq. (36) into Eq. (35), while making use of the failure criterion Eq. (28) to elimin-
ate U0, can be shown to lead to the following expression:
2 3
pffiffiffi
1 16 α 7 Pu
¼ 41 þ q ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 5 P ffi 3  1 ð37Þ
λ2 2 1  Pu Py y

179
In the above equation α is an imperfection factor, which is given by:
 
π2 E A0 2
α¼ ð38Þ
8fy L

According to Eurocode 3: EN1993-1-5 (CEN 2006) the local imperfection of a plate supported
along all four edges may be taken as A0 = L/200. This ‘equivalent’ imperfection takes into
account geometric imperfections, as well as residual stresses. With E = 200 GPa and fy = 350
MPa, Eq. (38) then yields: α = 1/57. Figure 4 shows Eq. (37), plotted for this value of α. Very
good agreement with the Winter curve is obtained up to a slenderness value of about 2. Above
this value the predictions diverge from the Winter curve. This can be attributed to the limiting
assumptions in our model, in particular the proposition that the deflected shape is represented
by Eq. (19). This assumption results in a constant post-buckling stiffness, as indicated by Eqs.
(24) and (32). In reality this stiffness will further deteriorate as the load rises in the post-buckling
range. This effect is, of course, more important for more slender plates, which have a more
extended post-buckling range before yielding sets in. The predictions can likely be improved by
including higher-order Fourier terms in Eq. (19). However, this complicates the calculations and
this option can, unfortunately, not be explored within the page limit of this paper.

5 CONCLUSIONS

A non-linear equation describing the post-buckling behaviour of thin plates has been pre-
sented. The equation is derived by simplifying the set of coupled von Karman differential
equations while consistently interpreting the implications of the effective width concept. The
equation is able to account for imperfections and is wed to the effective width concept to pre-
dict failure. Approximate solutions are presented for the cases of a geometrically perfect
square plate with straight loaded edges and with the longitudinal edges either remaining
straight or being completely free to move in. The case of a geometrically imperfect plate with
the latter boundary conditions is also explored and it is seen that the capacity predictions
agree very well with the empirical Winter curve for slenderness values of up to 2.

REFERENCES

CEN. 2006. Eurocode 3-1.5. Eurocode 3: Design of Steel Structures, Part 1.5: Plated Structural Elements.
European Committee for Standardization, Brussels.
Föppl, A. 1907. Vorlesungen Über Technische Mechanik, Druck und Verlag von B.G. Teubner:
Levy, S. 1942. Bending of rectangular plates with large deflections. NACA Technical Note 846.
Marguerre, K. 1937. The apparent width of the plate in compression. NACA Technical Memorandum
No. 833.
Von Karman, T. 1910. Festigkeitsprobleme im Maschinenbau. In: Klein, F., Müller, C., Encyklopädie
der Mathematischen Wissenschaften, Druck und Verlag von B.G. Teubner: 311–385.
von Karman, T., Sechler, E.E. & Donnell, L.H. 1932. The Strength of Thin Plates in Compression.
ASME Applied Mechanics Transactions, 54(5), 53–70.
Winter, G. 1970. Commentary on the 1968 Edition of the Specification for the Design of Cold-Formed Steel
Structural Members, American Iron and Steel Institute, New York, USA.

180
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Simplified method for lateral torsional buckling of beams with


lateral restraints

A. Beyer & A. Bureau


CTICM, France

ABSTRACT: The present paper discusses the simplified method for the lateral torsional buck-
ling design of I shaped members with discrete lateral restraints. This popular method, also
adopted in Eurocode 3 Part 1-1:2005, consists in verifying the resistance of the compressed
“flange” of the beam subject to major-axis bending. In this case, the compressed flange is ficti-
tiously composed of the flange itself and 1/3 of the compressed part of the web. The paper sum-
marises the analytical derivation of the simplified method and its key assumptions. Then, it
presents inconsistencies arising if the simplified method is compared to the more general applic-
able design methods for lateral torsional buckling given in Eurocode 3 Part 1-1:2005. In addition,
the design methods are compared to numerical simulations revealing the unconservatism of the
simplified method in some cases. In the last part of the paper, the authors propose slight modifi-
cations of the method in order to eliminate the mentioned inconsistencies.

1 INTRODUCTION

The simplified method proposed in paragraph 6.3.2.4 of EN 1993 Part 1-1:2005 (CEN 2005) rep-
resents an alternative for the resistance to lateral torsional buckling of members subject to major-
axis bending. The approach is based on the idea that the lateral torsional buckling check of the
member may be replaced by the verification of the lateral flexural buckling resistance of the com-
pressed flange, composed by the flange itself and one third of the compressed part of the web.
If the relative slenderness of the compressed flange is less than a limit slenderness, it is sup-
posed that the member is not sensitive to lateral torsional buckling and may be designed
according to its full cross-section resistance. The condition with respect to the limit slenderness
is given in Equation 1.

kc Lc Mc;Rd
λf ¼  λc0
if ,Z λ1 My,Ed
ð1Þ
and
λc0 ¼ λLT,0 þ 0:1

where λf = relative slenderness of the compression flange; kc = correction factor depending on


the bending moment diagram (see Table 1); Lc = distance between lateral restraints; if,z = radius
of gyration about the buckling axis of the compressed flange; λ1 = 93ε; Mc,Rd = bending
moment resistance of the cross-section; My,Ed = design value of the bending moment acting in
the member; λLT,0 = 0.4 according to EN 1993-1-1 (CEN 2005).
The correction factor kc is recalled in Table 1 for three bending moment diagrams.
If the criterion given by Equation 1 is not fulfilled, the lateral torsional buckling
resistance Mb,Rd is determined with Equation 2.

181
Table 1. Factor kc
Bending moment diagram Factor kc

1.00

0.94

0.86

Mb;Rd ¼ kf l χMc;Rd but : Mb;Rd  Mc;Rd ð2Þ

where: kf l = correction factor specific for the simplified method considering the “safe-sided” char-
acter of the method (kfl = 1.1) and χ = reduction factor for buckling of the compressed flange.
It should be noted that the moment resistance Mc,Rd has to be calculated with the section
modulus referring to the compressed flange. In case of mono-symmetric class 3 or 4 cross-
sections, its value may therefore exceed the cross-section bending resistance (as its value is always
calculated with the smallest of the two section moduli). Hence, a complementary verification of
tension flange failure is necessary.
The reduction factor χ should be calculated based on the relative slenderness of the com-
pressed flange calculated according to Equation 1 and, depending on the geometry of the cross-
section, with buckling curve “d” for welded sections if h/tf < 44ε or buckling curve “c” in all
other cases. Finally, it may be noted that the factor kf l introduces an increase to the calculated
resistance as it is supposed that the simplified method is safe-sided because the torsional stiffness
of the members is neglected (see Section 2).
In Section 4, the simplified method is compared to the strength predictions obtained with the
general case of §6.3.2.2 of EN 1993-1-1:2005 which is recalled in Equations 3 and 4.


 ¼ 0:5 1 þ αLT λLT  0:2 þ λLT ð3Þ
2

1
χ ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð4Þ
 þ 2  λLT
2


where: αLT = imperfection factor, λLT = relative slenderness for lateral torsional buckling.
In case of welded sections, the imperfection factor αLT is equal to 0.49 (reduction curve c) if
the ratio h/b is lower than or equal to 2.0. If the ratio h/b is greater than 2.0, the reduction
curve d with an imperfection factor αLT of 0.76 should be used.
Comparing the simplified method presented here before with the general case, one may
notice the following points:
1) The plateau length of the simplified method is increased by 0.1 compared to the general

case (see λc0 of Equation 1);
2) The simplified method possesses a supplementary resistance reserve corresponding to the
factor kf l equal to 1.1;
3) The buckling curves applied in the simplified method (linked to the ratio h/tf) are different
from buckling curves applied in the general case (linked to the ratio h/b).
These points imply that the resistances obtained by the simplified method may be different
from the resistances obtained by the general method. However, it may not be stated at this
point whether the simplified method leads to conservative or non-conservative results. There-
fore, it seems interesting to compare both methods with each other and to compare them to
numerically obtained ultimate resistances. Before this is done, Section 2 recalls the theoretical
basis of the simplified method.

182
2 THEORETICAL BASIS OF THE SIMPLIFIED METHOD

In order to derive the simplified method for lateral torsional buckling of I sections, it is inter-
esting to start with the expression for the critical moment Mcr of a beam with double symmet-
ric I section subject to major-axis bending:
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 ffi rffiffiffiffiffi
π2 Iw GIt L2 Iw
Mcr ¼ C1 2 EIz 1þ ¼ C N
1 cr;z εt ð5Þ
L Iz EIw π2 IZ

where: E = Young’s modulus of the material; G = shear modulus of the material; C1 = factor
considering the influence of variable bending moments; L = member length; Iz = second moment
of area about the z-axis (minor-axis); Iw = warping constant; It = torsion constant; Ncr,z = critical
axial force for flexural buckling of the member about its weak axis and εt = factor representative
of the torsional stiffness of the member.
If the torsional stiffness of the member is neglected (εt = 1), the elastic critical buckling
force of the compressed flange may be expressed as a function of Mcr as shown by Equation 6.

π2 Mcr Mcr
Ncr;z;CF ¼ C1 2
EIz;CF ¼ rffiffiffiffiffi ¼ ð6Þ
L Iw hs
2
Iz

where: Ncr,z,CF = critical buckling force of the compressed flange; Iz,CF = second moment of
area of the compressed flange about the z-axis (≈1/2 Iz) and hs = distance between the cen-
troids of the flanges.
As stated above, the simplified method proposes to verify the lateral torsional buckling
resistance of the member based on the lateral buckling resistance of the compressed flange.
Supposing that the same imperfection factor α may be applied for both cases, the simplified
method yields exactly the same resistance as the method of paragraph 6.3.2.2 of (CEN 2005) if
the condition of Equation 7 is fulfilled.
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi sffiffiffiffiffiffiffiffiffiffiffi
kc Lc ACF fy Wy fy
λf ¼ ¼ ¼ λLT ¼ ð7Þ
if ;z λ1 Ncr;z;CF Mcr
sffiffiffiffi
E
λ1 ¼ π ¼ 93:9ε ð8Þ
fy
rffiffiffiffiffiffiffiffiffiffi
Iz;CF
if ;z ¼ ð9Þ
ACF

where: ACF = area of the compressed flange and Wy = is the section modulus (class 1 and 2
cross-section: Wpl; class 3 cross-section: Wel; class 4 cross-section: Weff)
In case of double symmetric cross-sections, the condition expressed by Equation 7 is ful-
filled if the ratio between the resistance of the compressed flange ACF fy and the bending
moment resistance of the cross-section Wy fy is equal to the ratio between the elastic critical
axial force and the elastic critical bending moment. It yields:

ACF Ncr;z;CF 1
¼ ¼ ð10Þ
Wy Mcr hs

The area of the compressed flange is equal to the area of the flange itself and the area associated
to a certain height x . (hs – tCF) of the web. This area may be expressed by:

183
ACF ¼ bCF tCF þ tw ðhs  tCF Þx ð11Þ

where: bCF = width of the flange; tCF = thickness of the flange; tw = thickness of the web;
x = proportion of the web taken into account for the compressed flange.
The elastic and plastic modulus of the cross-section, used in Equation 10, may be expressed
by (possible fillets are neglected):

ðhs  tCF Þ2
Wy;el ≈ bCF tCF hs þ tw ð12Þ
6

ðhs  tCF Þ2
Wy;pl ≈ bCF tCF hs þ tw ð13Þ
4

where: Wy,el = elastic section modulus and Wy,pl = plastic section modulus.
Based on the elastic section modulus the ratio x of the web depth may be obtained using
Equations 10 and 11. It yields:

ðhs  tCF Þ
x ¼ ð14Þ
6hs

If the thickness of the compressed flange is small, the value of x tends to 1/6. In case of
a double symmetric I section, it is therefore acceptable to consider 1/3 of the compressed part
of the web as indicated in paragraph 6.3.2.4 of Eurocode 3 Part 1-1. Considering the plastic
section modulus in the derivation would lead to a ratio x of 1/4 implying that half of the com-
pressed part of the web should be used for the calculation of the compression flange. At this
point, it should be noted that considering a smaller part of the web leads also to a smaller
relative slenderness (see Equation 7) and consequently to a higher value of the reduction
factor χ. A priori, it is therefore unsafe to consider only 1/3 of the compressed part of the web
in the calculation of the compression flange if one refers to the plastic cross-section resistance.
Additionally, it is to be noted that throughout the derivation of the simplified method, it has
been (implicitly) considered that the section is doubly symmetric and that the member is not
subject to a destabilizing load (for example gravity loads applied on the upper flange). In
other cases, the simplification shown in Equation 6 is not valid anymore, i.e. the ratio between
critical axial force Ncr,z,CF flange and the critical bending moment Mcr is not equal to 1/hs.
Throughout this section, it has been shown that the simplified method may be derived analyt-
ically for doubly symmetric I-sections. Yet, it has also been shown, that there are several incon-
sistencies. In particular, the simplified method does not account for the effect of destabilizing
loads or the effect of the mono-symmetry. Additionally to the discrepancies noted in Section 1,
these inconsistencies could potentially lead to unsafe predictions of the resistance in some cases.
Therefore, it appears interesting to quantify the differences between the general case and the
simplified method for the lateral torsional buckling resistance of I sections and to evaluate the
safe-sided nature of both methods by comparing to results obtained by numerical simulations.

3 STUDIED CASE

The following comparisons are based on the cross-section given in Table 2.


In order to obtain mono-symmetric cases, the parameter ψψ defined in Equation 15 and
characterizing the mono-symmetry of the section, is varied by modifying the flange width. For
doubly symmetric cross-sections, ψ is equal to 0 and for T sections ψ is equal to ±1. However,
hereafter, T-sections are excluded from this study.

Iz;CF  Iz;TF
ψ ¼ ð15Þ
Iz;CF þ Iz;TF

184
Table 2. Studied doubly symmetric cross-section.
Cross-section

Total height of the cross-section (mm) 750


Thickness of the web (mm) 12
Width of the flange (mm) 200
Thickness of the flange (mm) 20
Yield strength (N/mm²) 355

In this study, the factor ψ is varied between 0.9 and 0.0 (double symmetric section). In order
to visualize the cross-section associated to a given value of ψ, Table 3 shows the corresponding
ratios between the width of the compression flange bCF and the width of the tension flange
bTF for the studied cross-sections.
The destabilizing effect of loads not applied at the shear centre of the cross-section is charac-
terized by the parameter –z. It is equal to the ratio between the distance measured from the load
application point to the shear centre and the distance measured from the centroid of the com-
pressed flange to the shear centre of the cross-section (=hs/2 for double symmetric sections) as
shown in Figure 1 and Equation 16. Hence, if z is equal to 0, the load is applied at the shear
centre and if it is equal to 1.0, the load is applied at the compressed flange.
zF
z ¼ ð16Þ
zCF

Hereafter, the two bending moment diagrams represented in Table 4 are studied. The first
case corresponds to a simply supported member subject to a point load at mid-span (for
example the practical case of a crane girder). The second case of constant moment may, for
example, correspond to the central segment of a member with intermediate lateral restraints
subject to two point loads introduced through these restraints (for example, purlins acting as
lateral restraints of a building rafter).

Table 3. Ratio bCF/bTF for a given value of ψ for the stud-


ied cross-sections.
ѱ bCF (mm) bTF (mm) bCF/bTF

0.9 248 93 2.67


0.6 234 147 1.59
0.4 224 169 1.33
0.2 213 186 1.15
0.0 200 200 1.00

Figure 1. –
Definition of z.

185
Table 4. Load cases.
Load
Case Bending moment diagram kc
LC 1 0.86
LC 2 1.00

4 NUMERICAL SIMULATIONS

4.1 Numerical model


In order to evaluate the safe-sided nature of both design methods, GMNIA simulations are per-
formed. The numerical model includes a global member geometric imperfection affine to the
first member Eigen mode with an amplitude of L/1000 (L is the total length of the member) as
recommended in reference (CEN 2007). Additionally, the residual stress pattern proposed in ref-
erences (ECCS 1984) and (Sedlacek et al. 2004) is included as shown in Figure 2. As the studied
cross-section is of class 2 according to Eurocode 3 Part 1-1, its theoretical plastic section resist-
ance is not affected by the influence of local buckling. Therefore, local (plate) imperfections are
not accounted for in the numerical model.
Finally, it should be noted that the members are modelled with fork end supports. Add-
itionally, for load case LC 1 (point load applied at mid-span) the section is stiffened under the
applied load.

4.2 Results
Hereafter, the results obtained by both design methods are compared to the GMNIA results.
It should be noted that the recommended value of kfl (=1.1) has been applied in case of the
simplified method. Also, it should be noted that the imperfection factor αLT used in the simpli-
fied method is equal to 0.49 (buckling curve c since h/tf > 44ε) whereas for the method
described in paragraph 6.3.2.2, the imperfection factor is equal to 0.76 (buckling curve d).
Finally, one may note that the length L of the studied members has been chosen so as to
obtain a relative slenderness of approximatively 1.10 for the reference case of a doubly sym-
metric cross-section subject to a constant bending moment diagram (L = 420 cm). The result-
ing relative slenderness is given for all members and for both methods in Table 5.
First, it may be surprising that both methods do not yield the same relative slenderness for
mono-symmetric sections and sections subject to a load applied outside of the shear centre.
However, this observation is directly linked to the analytical derivation of the simplified

Figure 2. Residual stress pattern for mono-symmetric cross-sections.

186
Table 5. Comparison of GMNIA results to the simplified method and to the general method.
Bending §6.3.2.4 §6.3.2.2
moment Mult (kNm) Mult (kNm) GMNIA
– –
diagram z– ψ λF (kf lχ) λLT (χ) Mult (kNm)

– 0.9 0.84 948.2 (0.70) 0.80 786.3 (0.58) 900.0


– 0.6 0.92 973.6 (0.65) 0.89 791.0 (0.53) 894.9
– 0.4 0.97 950.9 (0.61) 0.95 762.1 (0.49) 869.7
– 0.2 1.03 903.9 (0.58) 1.02 717.7 (0.46) 831.1
– 0.0 1.11 830.4 (0.53) 1.10 656.4 (0.42) 776.4
1.0 0.9 0.79 993.5 (0.74) 1.02 616.1 (0.46) 879.9
1.0 0.6 0.86 1027.9 (0.68) 1.09 636.2 (0.42) 881.0
1.0 0.4 0.91 1008.5 (0.65) 1.15 616.6 (0.40) 850.4
1.0 0.2 0.97 963.8 (0.61) 1.21 581.7 (0.37) 803.8
1.0 0.0 1.04 891.1 (0.57) 1.30 531.4 (0.34) 739.8

method recalled in Section 2 of this paper and in particular, it is linked to the assumption that
the compression flange is composed of the flange and 1/3 of the compressed part of the web
(valid only for doubly symmetric sections and members without destabilizing loads).
By observing Table 5, it is important to note that the simplified method is in all cases
un-conservative compared to the numerical simulations. For the constant bending moment
diagram the difference attains of about 10% and for the applied point load, the un-
conservatism attains up to 25%! The observed discrepancy may be explained by the fact that
the effect of the mono-symmetry and the effect of a load applied outside the shear centre is
not considered in the derivation of the simplified method. In case of the studied section, the
fact that the torsional stiffness is neglected in the simplified method (see Section 2 – Equation
6) does not compensate the destabilizing effect generated especially by the applied load.
Besides the unconservatism of the simplified method, the huge discrepancy between this
method and the general method (up to 70% difference) seems annoying from a practical point
of view especially as the so called “simplified method” leads to higher resistances. In general,
practitioners expect simplified methods to be more safe-sided. Nonetheless, it may be admitted
that the general case may be too safe-sided in some cases, as the variation of the bending
moment diagram (in case of the applied point load) is not fully considered here. In addition, it
has been shown in the past that the attribution of the buckling curves is rather conservative
for welded sections (see for example (Taras 2011)).

4.3 Further discussions of the results


Admittedly, the study presented here only covers a very limited number of cases. Nonetheless,
the results clearly indicate that the simplified method may be highly unsafe for some configur-
ations. The unconservatism of the simplified method may be understood easily for the cases trea-
ted in the framework of this paper. In fact, since neither the mono-symmetry nor the possible
destabilizing effect of applied loads is considered in the method, the studied cross-sections have
been chosen deliberately in order to reduce the torsional constant compared to the other section
properties. Consequently, the fact that the torsion constant is neglected in the derivation of the
simplified method does not justify the value of 1.1 for the factor kfl (that is intended to cover the
“safe-sided” character of the method). This is even true for the doubly symmetric section. Yet, it
should be noted that for typical welded sections, the εt value (indicating the importance of the
torsional constant of a cross-section – see Section 2), is generally close to 1.0. Consequently, the
observed unconservatism is of practical interest.
In the framework of this paper, the beams are supposed to possess fork end supports and no
intermediate lateral restraints. It is admitted that if the simplified method is applied to members
with intermediate restraints, it may be less unsafe as the more sensitive segments of the member
are stabilized by adjacent segments that are less loaded. However, one may wonder if this effect
should be implicitly be accounted for in a simplified method or not. Anyhow, in order to

187
Table 6. Comparison of GMNIA results to modified simplified method.
– –
Bending moment z ψ λF §6.3.2.4 GMNIA Mult.6324/Mult.GMNIA
diagram Mult (kNm)(kflχ) Mult (kNm)

– 0.9 0.85 742.5 (0.55) 900.0 0.82


– 0.6 0.92 767.8 (0.51) 894.7 0.86
– 0.4 0.97 748.4 (0.48) 869.7 0.86
– 0.2 1.03 710.6 (0.45) 831.1 0.86
– 0.0 1.11 653.2 (0.42) 776.4 0.84
1.0 0.9 0.80 782.5 (0.58) 879.9 0.89
1.0 0.6 0.86 813.3 (0.54) 881.0 0.92
1.0 0.4 0.91 795.5 (0.51) 850.4 0.94
1.0 0.2 0.97 758.5 (0.48) 803.8 0.94
1.0 0.0 1.04 700.5 (0.45) 739.8 0.95

eliminate (at least partially) the unsafe nature of the simplified method and the discrepancies
compared to the general method, the authors recommend the following modifications:
1) The factor kfl should be equal to 1.0 and can therefore be omitted;
2) The imperfection factor αLT should be chosen according to the general case.
If these modifications are introduced in the simplified method, the results given in Table 6
are obtained. Table 6 shows that the simplified method becomes reasonably safe-sided com-
pared to the numerical results for the examples studied in this paper.

5 CONCLUSION

Throughout the present paper, the simplified method for lateral torsional buckling proposed in
paragraph 6.3.2.4 of EN 1993-1-1:2005 (CEN 2005) is studied. First, the analytical derivation of
this method is recalled and it is shown that the effects of mono-symmetry and destabilizing forces
are neglected. Then, the strength predictions obtained by the simplified method for welded
(mono-symmetric) sections are compared to the results of the general case (§6.3.2.2 of EN 1993-
1-1:2005) and to numerical simulations. It has been shown that the simplified method leads to
much higher resistances than the general method. Yet, it appears even more disturbing that the
results of simplified method may exceed the numerically obtained resistances by more than 25%.
Therefore, several modifications of the simplified method are proposed that reduce the discrep-
ancy between both design methods and that lead to safe-sided strength predictions compared to
numerical simulations. However, it should be noted that the present study is limited to a small
number of configurations. In particular, it seems worthy to study the simplified method for mem-
bers with several intermediate restraints. In this case, segments of the member that are less loaded
have a favourable effect on the stability of the most loaded (central) segment of the member.

REFERENCES

CEN, 2005. “EN 1993-1-1: Eurocode 3 Design of steel structures – Part 1-1: General rules and rules for
buildings”, Brussels.
CEN, 2007. “EN 1993-1-5: Eurocode 3 Design of steel structures – Part 1-5: Plated structural elements”,
Brussels.
ECCS TC8, 1984. “ECCS – Ultimate Limit State Calculation of Sway Frames with Rigid Joints”, Brussels.
Sedlacek, G., Stangenberg, H., Lindner, J., Glitsch, T., Kindmann, R.,Wolf, C., 2004. “FOSTA P554:
Untersuchungen zum Einfluss der Torsionseffekte auf die plastische Querschnittstragfähigkeit und die
Bauteiltragfähigkeit von Stahlprofilen“.
Taras, A., 2011. “Contribution to the Development of Consistent Stability Design Rules for Steel
Members, Ph.D Thesis, Graz University of Technology.

188
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Buckling resistance of mono-symmetric I-/H-section members in


biaxial bending, axial compression, and torsion

A.-L. Bours, R. Winkler & M. Knobloch


Chair of Steel, Lightweight and Composite Structures, Ruhr-Universität Bochum, Germany

ABSTRACT: This paper addresses the stability behaviour of simply supported members in
bending, axial compression and torsion with mono-symmetric I-/H-sections. The scope of the
current beam-column rules and interaction formulae of EN 1993-1-1:2005 are extended based
on the results of a comprehensive numerical simulation study. The effects of the steel grade,
the member slenderness ratio, the Iz,fl,min/Iz,fl-ratio, the load case and the bending moment dis-
tribution on the member stability behaviour were analysed. The paper compares the numeric-
ally determined member capacities with the results of the proposed design rules according to
annex C.2 of prEN 1993-1-1:2018. The paper proposes amendments for the simplified assess-
ment rules to improve the accuracy for members with non-uniform bending moment distribu-
tions and semi-compact cross-sections.

1 INTRODUCTION

EN 1993-1-1:2005, Cl. 6.3.3 provides a simplified method to assess the buckling resistance of
I-/H-section and hollow section steel members in bi-axial bending and axial compression. The
method was developed within the framework of the Technical Commission 8,Structural Stability’
of the European Convention for Steel Construction ECCS (Greiner & Lindner 2003, Boissonnade
et al. 2006) and is often used in engineering practice. Within the framework of the development of
the second generation of the European structural design rules, the relevant drafting teams for EN
1993-1-1, i.e. Working Group 1 of CEN/TC250/SC3 and Project Team SC3.T1, analysed whether
the method can be extended to cover mono-symmetric cross-sections and torsion (Walter et al.
2017, Winkler et al. 2018). A simplified method for simply supported, uniform I-section members
in bending, axial compression and torsion was developed based on the member buckling rules in
annex B of the current EN 1993-1-1:2005 and the assessment method for lateral-torsional buckling
of runway beams provided in EN 1993-6:2007. The later method is based on former investigations
within a German national research project (Kindmann et al. 2004, Lindner & Glitsch 2004). The
limited scope of the recent study necessitated to limit the application range of the novel method
provided in annex C.2 of prEN 1993-1-1:2018. This paper presents additional studies on the buck-
ling behaviour of mono-symmetric I-section members in bending, axial compression and torsion
and investigates whether the method can be improved in terms of efficiency as well as user-
friendliness and the method’s application range can be extended.

2 STRUCTURAL BEHAVIOUR OF MONO-SYMMETRIC CROSS-SECTIONS

This section analyses the structural behaviour of members with mono-symmetric I-sections in
bending, axial compression and torsion. The structural behaviour is assessed on the basis of
the load-rotation behaviour, the development of internal forces during load increase and the
remain-ing elastic cross-sections as a result of the development of plastic zones. For this pur-
pose, a numerical simulation study was performed using the finite element program ABAQUS
Standard. Details of the FE-model are provided in (Winkler et al. 2018).

189
A simply supported steel member with a welded mono-symmetric I-section is examined for
a detailed analysis of the structural behaviour. To be consistent to investigation for common
double-symmetric I-/H-sections a fictitious section was used. The dimensions are equal to an
IPE 500 with a wider compression and tension flange. The simple member was subjected to
a normal force N, bending moments My and Mz and a linear torsional load mx. The following
figures present the structural behaviour in terms of the load-rotation curve at midspan and
the development of internal forces during load increase. Regarding the development of
internal forces, the bending moments my, mz, mω, mxp,Support as well as m1,z,fl and m3,z,fl are
plotted, mi = Mi/Mpl,i. m1,z,fl and m3,z,fl are the bending moments of the flanges, where the
bending moment My causes compression and tension, respectively. In addition, the cross-
sectional utilization at midspan as well as at support, determined according to the Partial
Internal Forces method (Kindmann & Frickel 2002), are included in the evaluation. Further-
more, the remaining elastic cross-sections are shown for different deformation states.
A deformation state indicates a change in the cross-section, for example if a part of the
cross-section starts yielding or if the whole part is plasticised. The plasticised parts have been
identified on the basis of the von-Mises yield criteria.
A member with an IPE 500 cross-section with one wider flange (bf = 31.75 cm) is analysed
first. The steel grade was S235 and the flexural buckling slenderness was λz = 1.0. The applied
moments caused uniform bending moment distributions according to first order theory. The
ratios of the internal forces were n = 0.8, my = 0.8, mz = 0.1 as well as mω = 0.1 (according to
first order theory). The loads were applied in two steps. The first load step included the appli-
cation of the bending moment about the weak axis mz and the warping moment mω. During
the second load step the normal force n and the bending moment about the strong axis my
were increased. In the following, the member with the wider and the member with the smaller
compressed flange are examined.
Figure 1 displays the structural and deformation behaviour as well as the remaining elastic
cross-sections of the member with the smaller compression flange. The load-rotation behav-
iour and the internal forces are only presented for the second load step. The cross-section
remained mainly elastic during the application of mz and mω. The remaining elastic cross-
section after the application of mz and mω is marked with LF1 = 1.0. However, the residual
stresses from the welding process and the bending moments caused a small plasticized area in
the middle of the lower flange.
In order to describe the structural behaviour of the member during the second load applica-
tion step, nine states (A to I) were selected. In point A, the tip of the upper left flange began to
plasticize, in point B the tip of the upper right flange. Point C shows the remaining elastic
cross-section at the time of the maximum load. So far, only parts of the upper flange were
plasticized. The utilization of the cross-section at midspan is 84.7 %, according to the Partial
Internal Forces method. These two aspects indicate that the significant failure mode was sta-
bility failure, (Winkler, Walter & Knobloch 2018). The remaining six deformation states
describe the member behaviour after reaching the maximum load. The last remaining cross-
section illustrates the situation at which the calculation was terminated as the maximum
number of increments were reached, point I.
In Figure 1 (b) the development of internal forces at midspan and at support is plotted
against the load proportionate factor of the second load application step. The normalised
normal force n (dark brown) and the bending moment my (light blue) were zero after the first
load step. During the second load application step n and my increased until the maximum
load. In the post ultimate range n and my decreased, because the maximum load was reached
and the forces, which generated n and my, decrease.
Figure 1 (b) shows, that the existing warping moment mω (yellow) decreased after exceeding
the maximum load. The warping moment was transformed into a primary torsional
moment mxp at the supports caused by the increase of the member rotation and the sustained
rate of loading, (Winkler et al. 2018). Furthermore, the graph shows that the smaller compres-
sion flange m1,z,fl received significantly larger loads at the beginning of the second load appli-
cation step than the wider tension flange m3,z,fl. After exceeding the plastic resistance of the
upper flange due to the high compressive stresses (Step C), the load is taken over by the wider

190
Figure 1. Structural behaviour of a simply supported I-section member with unequal flanges (larger ten-
sion flange) and a relative buckling slenderness ratio of λz = 1.0. Load-rotation behaviour in the second
load step (a), corresponding internal forces and moments (b) and remaining elastic cross-sections (c).

lower flange (Steps D to I). Considering bending moments about the weak axis of the partial
elements according to the Partial Internal Forces method allowed the bending moment of the
wide flange to slightly exceed the full cross-sectional utilization at midspan (m3,z,fl > PIM).
Furthermore, the consideration of hardening material behaviour allowed to exceed the plastic
cross-section capacity determined with the yield strength (Steps F to I).
Next, the structural behaviour of the member with a wider compression flange will be ana-
lysed. Figure 2 (a) evaluates the rotation of the member at midspan as a function of the
applied load. The rotation is small compared to the previously analysed member with
a smaller compression flange. The wider compression flange could carry larger compressive
forces without lateral deflections and rotations.
The structural behaviour of the member with a wider compressive flange is described by six
deformation states (A to F). Point A indicates the start of plastification at the tip of the upper
left flange. The maximum capacity is reached at point B when the yield strength at the tip of
the upper right flange is reached and the flange starts to plasticise. The cross-sectional utiliza-
tion at midspan is 81.4 % and main points of the cross-section remained in the elastic region.
The states C to F are in the post ultimate range.
Figure 2 (b) shows that the member with the larger compressive flange can carry a larger
bending moment (my = 0.3) than the member with the smaller compressive flange (my = 0.2).
This allows to exceed the load application factor of the previous member. Figure 2 (b) shows
that the existing warping moment mω is again transformed into a primary torsional
moment mxp at the supports. The stresses of the upper flange are also taken over by the lower
tension flange (Steps C to F). For small load application factors the cross-sectional utilisation
(PIM) at midspan is mainly affected by the bending moment of the upper flange, the normal
force and the bending moment about the strong axis.

191
Figure 2. Structural behaviour of a simply supported I-section member with unequal flanges (larger
compression flange) and a relative buckling slenderness ratio of λz = 1.0. Load-rotation behaviour in
the second load step (a), corresponding internal forces and moments (b) and remaining elastic
cross-sections (c).

3 SIMPLIFIED METHOD ACCORDING TO ANNEX C.2 OF PREN 1993-1-1:2018

This section examines the accuracy of the simplified design method in annex C.2 of prEN
1993-1-1:2018, which is here applied to determine the buckling resistance of steel members
with mono-symmetric cross-sections in bending, axial compression and torsion. Within
a comparative study, these resistances will be compared to the results of the numerical simula-
tion study carried out with ABAQUS Standard. Based on the results of the comparative
study, amendments are proposed.

3.1 Comparative study


The background and procedure of the design method of annex C.2 are described in (Winkler,
Walter & Knobloch 2018). This section focuses on the application of the specific rules to
mono-symmetric I-sections. It has to be considered, that the elastic bending moment
capacities Mz,el and Mω,el of class 3 cross-sections refer to the smaller flange, My,el refers to
the compression flange. The reduction factors for lateral torsional buckling χLT and torsional
buckling χTF have to be determined with the buckling curves according to prEN
1993-1-1:2018 clause 8.3.2.3(2). The elastic critical buckling moment My,cr as well as the rela-
tive lateral torsional buckling slenderness λLT correspond to the flange in compression caused
by the bending moment My, independent of whether it is the smaller or the wider flange. For
non-uniform bending moment distributions with changing signs along the beam axis, Mcr,λLT
and χLT must be calculated for the respective absolute maximum values of the bending
moments. Moreover, the equivalent moment coefficients Cmi are set equal to 1.0 for these

192
bending moment distributions. The application of the design method is limited to mono-
symmetric I-sections with a ratio of the moments of inertia of the flanges Iz,fl,min/Iz,fl that
exceeds 0.2. Furthermore, the reduction factor χz should be replaced by χTF if the moments of
inertia of the flanges differ by more than 50 %. If the bending moment causes com-ression in
the smaller flange, χTF has to be used as well. Moreover, modified interaction factors kyy for
mono-symmetric cross-sections are provided in annex C.1 of prEN 1993-1-1:2018.
In addition to the stability verification, the cross-sectional verification check should
be performed. For class 1 and 2 members, a nonlinear plastic interaction formula may
be applied. A separate cross-sectional verification allows a smooth transition of the
resistance of stocky members and lead to an efficient design. For the further proced-
ure, the load capacities were determined with the cross-section and the stability verifi-
cation. For the evaluation of the simplified method only stability failure was
considered.
The comparative study comprised different ratios of bending moment my and mz, axial
compression n and warping moment mω, with constant and linear (with and without changing
sign) bending moment distributions, respectively. The investigated Iz,fl,min/Iz,fl-ratio was
between 0.2 and 0.4. Furthermore, the influence of the slenderness was analysed by consider-
ing different member lengths. In order to cover the area of practical relevance from compact
to slender members, the lengths of the members corresponded to relative slenderness ratios λz
of 0.4, 1.0 and 3.0. The influence of the steel grade was also examined. The considered steel
grades were S235, S355, S460 and S690. For all varied parameters, members with a wider
flange in both compression and tension were analysed.
Figure 3 compares the numerical load capacities with those of the simplified method.
The results are divided depending on the cross-section class (CSC 2 - blue, CSC 3 -
orange) as well as depending on members with a wider compression (crosses) or tension
flange (circles). The determined load factor LF refers to the internal forces N and My
(second step of the load application). The graph shows that the simplified method led to
conservative results compared to the numerical results. The simplified method led to
over conservative results for members with non-uniform bending moment distributions

Figure 3. Comparison of the load capacities between the numerical results and the simplified design
method according to annex C.2 of prEN 1993-1-1:2018.

193
with changing sign (due to the equivalent moment coefficients Cmi) and class 3 members
(due to partial plastification). Thus, the paper will propose amendments to annex C.2 of
prEN 1993-1-1:2018 with regard to the application of the equivalent moment coefficients
and semi-compact cross-sections. For the class 2 cross-sections, the mean value was
0.672 with a standard deviation of 0.243. For the class 3 cross-sections, the mean value
was 0.605 and the associated standard deviation 0.125.

3.2 Amendment proposals


The simplified design method of prEN 1993-1-1:2018 annex C.2 leads to conservative results
compared to the computational results. The amendment proposals aim at improving the effi-
ciency of the simplified method and ensuring the consistency of the models for mono-symmetric
I-sections. For this purpose, two possible improvements for annex C.2 are analysed. The two
possible improvements are:
i) adopting the equivalent moment coefficients Cmi for non-uniform bending moment
distributions
ii) adopting the interaction formulae of annex B of prEN 1993-1-1:2018 for semi-
compact cross-section members in bending, axial compression and torsion
The results of the design method are up to 45 % on the conservative side for members
with a changing sign bending moment distribution. As previously mentioned, it is nor-
matively arranged in Annex C.2 for components with a mono-symmetric cross-section
under the load type, which is considered here, that in the case of a bending moment
distribution with changing signs of My or Mz, the equivalent moment coefficient Cmi
must be set to 1.0. The examinations have shown that the equation for double-
symmetric cross-sections given in prEN 1993-1-1:2018 can be adopted for mono-
symmetric cross-sections. The equivalent moment coefficients can therefore be calculated
using Table 1.
The second proposal applies the specific design models for semi-compact cross-sections
according to annex B of prEN 1993-1-1. Annex B proposes to determine the partial plastic
section modulus Wep with an interpolation between the elastic and plastic section modulus
for double-symmetric class 3 cross-sections according to Equations (1) and (2). This method
is also applied to mono-symmetric class 3 cross-sections in bending, axial compression and
torsion.

Wep;y ¼ Wpl;y  ðWpl;y  Wel;y Þ  βep;y ð1Þ

Wep;z ¼ Wpl;z  ðWpl;z  Wel;z Þ  βep;z ð2Þ

Table 1. Equivalent moment coefficients Cmi according to prEN 1993-1-1:2018.


Load case Range Cmy, Cmz resp. CmLT

1.0

1≤ ψ ≤ 1 0.6 + 0.4 ψ ≥ 0.4

194
Figure 4. Comparison of the load capacities between the numerical results and the simplified design
method according to annex C.2 after adopting two normatively arranged options of double-symmetric
cross-sections.

where

c  10ε c=  83ε !
tf t
βep;y ¼ Max ; w ; 0  1:0 ð3Þ
4ε 38ε

c  10ε !
tf
βep;z ¼ Max ; 0  1:0 ð4Þ

Figure 4 compares the design method considering the two amendment proposals with the
computational. Even with the amendment proposals, the method leads to conservative design
results. The mean value of accordance was improved from 0.672 to 0.691 for the class 2 cross-
sections and from 0.605 to 0.746 for class 3 sections. In conclusion, the two amendments can
be adopted to mono-symmetric I-sections and help keeping the consistency of the simplified
design methods for member stability.

4 CONCLUSION

In the context of the revision of EN 1993-1-1 the scope of the simplified models for the design
of beam-columns has been extended to members in bending, axial compression and torsion. The
paper has confirmed that the simplified method can be applied to mono-symmetric I-section
members. A comparative study has shown that the method can easily be amended by applying
the equivalent moment coefficients for non-uniform bending moment distributions and the rules
of annex B of prEN 1993-1-1:2018 for semi-compact cross-sections. Even if the amendments are
applied the simplified method has led to conservative results compared to the results of
a numerical simulation study. Advantages of the proposed method are that it remains consistent

195
to the stability verification method for I-/H-section and hollow section members in bending and
axial compression and it takes benefit of and is consistent to the non-linear cross-section inter-
action for stocky members in bending, compression and torsion.

REFERENCES

ABAQUS. Version 6.14. Providence, RI, USA.


Boissonnade, N.; Greiner, R.; Jaspart, J.P. 2006. Rules for member stability in EN 1993-1-1background
documentation and design guidelines. ECCS Technical Committee 8. Brussels.
EN 1993-1-1: Eurocode 3: Design of steel structures – Part 1-1: General rules and rules for buildings.
2005.
EN 1993-6: Eurocode 3: Design of steel structures – Part 6: Crane supporting structures. 2007.
Kindmann, R.; Lindner, J.; Sedlacek, G.; Wolf, C.; Beier, J.; Glitsch, T. et al. 2004. Untersuchungen zum
Einfluss der Torsionseffekte auf die plastische Querschnittstragfähigkeit und die Bauteiltragfähigkeit
von Stahlprofilen. Forschungsbericht P 554. Forschungsvereinigung Stahlanwendungen e.V.
Greiner, R. & Lindner, J. 2003. Die neuen Regelungen in der europäischen Norm EN 1993-1-1für Stäbe
unter Druck und Biegung. Stahlbau 72, Heft 3, S. 157–172.
Kindmann, R. & Frickel. J. 2002. Elastische und plastische Querschnittstragfähigkeit: Grundlagen,
Methoden, Berechnungsverfahren, Beispiele. Ernst & Sohn.
Lindner, J. & Glitsch, T. 2004. Vereinfachter Nachweis für I- und U-Träger – beansprucht durch dop-
pelte Biegung und Torsion. Stahlbau 73, Heft 9, S. 704–715.
prEN 1993-1-1: Eurocode 3: Design of steel structures – General rules and rules for buildings. 2018.
Walter, A.; Herbersagen, J.; Winkler, R.; Knobloch, M. 2017. Structural behaviour of simple steel beams
subject to axial compression, biaxial bending moments and torsion. ce/papers 1, 2-3, S. 1076–1085.
Winkler, R.; Bours, A.-L.; Walter, A.; Knobloch, M. 2018. Redistribution of internal torsional moments
caused by plastic yielding of structural steel members. Structures.
Winkler, R.; Walter, A.; Knobloch, M. 2018. Zum Stabilitätsnachweis von Stahlbauteilen aus einfach-
und doppeltsymmetrischen I-Querschnitten unter Biegung, Druck und Torsion. Stahlbau 87, Heft 5,
S. 476–490.

196
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Cyclic plastic behavior of steel material under uniaxial load paths

V. Budaházy & L. Dunai


Department of Structural Engineering, Budapest University of Technology and Economics, Budapest,
Hungary

ABSTRACT: The cyclic inelastic response of steel material fundamentally determines the
structural behavior of structures during an earthquake event. The constantly growing demand
on ever-accurate design and economy requires the comprehensive knowledge of structural steel
behavior especially under arbitrary large inelastic cyclic loading. Although several studies
described the most important components of steel cyclic plasticity several behavioral factors are
not explored. In the current research fatigue, hardening and memory behavior of the structural
steel are investigated experimentally under different loading circumstances by 31 uniaxial load
protocols up to 12% strain range. Five different steel materials are investigated with grades
S235 to S420, which cover the most commonly used steel grades in European seismic design
practice. Beside the description of the tendencies, quantified variables and empirical equations
are presented for the description of hardening and memory behavior of steel material.

1 INTRODUCTION AND LITERATURE REVIEW

The knowledge of the cyclic plastic steel behavior have been a long-standing demand. The first
experiments were performed in the 1960s to predict the low cyclic fatigue life and the ratcheting
behavior of metallic materials. The calibration of cyclic plastic material models also required
new types of cyclic experiments, and the response of materials to more complex load paths was
investigated. Popov and Petersson (1978) investigated A36 steel material using simple tension-
compression and torsion tests. Chaboche et al. (1979) investigated aluminum alloy, A316 stain-
less steel, 316L steel INCO 718 alloy, 30 NCD 16 nickel-chromium steel, and titanium alloy.
They concluded, that the stress-strain relationship for a stainless steel is highly dependent on the
maximum plastic strain, which means the steel “remembers” to the previously experienced load
paths. The denoted this phenomenon as the memory behavior of the material. Lee and Chang
(1991) focused on the Bauschinger effect, which is an important phenomenon of metallic mater-
ials. It is concluded in the paper, that the elastic stress range is gradually decreasing, and the
reduction of stress range is saturated under small strain rate (referred as saturation of Bauschin-
ger effect). They also studied strain softening, when strain range during cyclic loading was
reduced, it took place, however, in a slower rate than strain hardening.
Dusicka et al. (2007) performed an experimental program covering a wide spectrum of
strength grades: (i) structural grade steels ( fy = 345–485 MPa) and (ii) low yield point (LYP)
steels ( fy=100–225 MPa). The cycles were fully reversed, with 2 to 14% strain range. The
observed, maximum cyclic stress was significantly higher than the yield stress, especially for
low steel grade, however at lower strain amplitudes cyclic softening was observed for steel
grade HT440 after initial hardening.
Shi et al. (2011) investigated Q235B and Q345B steel grades, under various uniaxial
loading paths. They found that the monotonic and cyclic steel behavior significantly
differ from each other, the yield point and plateau observed in the monotonic cases grad-
ually disappeared.

197
Wang et al. (2015) studied the cyclic behavior of high strength steel (nominal yield strength
is no less than 460 MPa). The investigated Q345B Q460C materials showed cyclic softening in
tension and a shortly cyclic hardening followed by cyclic softening under compression.
Zhou et al. (2015) investigated the cyclic response of structural steel on large amplitude
loading history. They investigated a total of six different cyclic loading protocols, with ampli-
tude up to 20% (±10%). They concluded that the memory of maximum amplitudes is pre-
served at lower amplitudes. They also found that, the Young’s modulus generally decreases,
which introduced the damage of the material.
Zhang and Xuan (2017) studied the accelerated cyclic softening and decelerated stress relax-
ation response of fatigue-creep interaction. They performed a combined low-cyclic fatigue,
and creep tests up to 1% plastic strain. They concluded, the effect of creep reduces stress and
fatigue capacity of the specimen as well.
Silvestre et al. (2015) compared the hardening behavior of different steel families. Each steel
grade had different crystal structure, therefore they showed different plastic response under
inelastic load paths. The yield strength evolution, the Bauschinger effect and the distribution of
hardening/softening (isotropic and kinematic) were determined for a wide range of materials.
Recently the most studied area in steel material research is the ratcheting behavior, which
mean the accumulation of plastic strain under cyclic loading with mean stress, due to recogni-
tion of its high detrimental effect on fatigue life. These studies focused on the following ques-
tions: (i) characteristic of the uniaxial ratcheting, (ii) effect of temperature, (iii) connection
between the fatigue life and the ratcheting, and (iv) effect of the multiaxial stress condition. Rug-
gles and Krempl (1990) found, that there are two reasons of the ratcheting effect: the first is the
accumulation of plastic strain, which is rate independent, and the second is the viscous damping.
Other researchers observed (e.g. Mima et al., 2000), that low amplitude cycles cause tiny change
in stress level, and only minor kinematic hardening occurs. In their experiments the hysteresis
loops could have slight opening, induced by cyclic relaxation of mean stress, therefore it was
necessary to consider slight opening of hysteresis loops as well as visco-plasticity.
Based on the literature review it is clear that the understanding of cyclic plastic steel behav-
ior has made great progress during the past fifty years, however, numerous points have not
been answered so far, as follows:
– Qualitative and quantitative evaluation of memory behavior, and the components of the
hardening.
– The hardening behavior under more complex loading history, especially the effect of small
loops during a load path, preloading, fading memory behavior and smaller load cycles on
monotonic results.
– Most cyclic experiments were performed on steel manufactured in the USA or Far East,
investigations of steel material manufactured under European specifications are insufficient.
In the current experimental program presented in this paper, fatigue, hardening and memory
behavior of five different steel materials are analyzed under 31 uniaxial load protocols up to
12% strain range.
In the experiments the maximum cyclic stress is significantly larger, than the monotonic
ultimate strength, and this additional cyclic hardening develops in relatively small strain range
within a few cycles. The isotropic part of cyclic hardening depends on the maximum and accu-
mulated strain, and the kinematic hardening is less sensitive for the strain paths. The steel
material shows fading type memory behavior.

2 EXPERIMENTAL SETUP

The experimental arrangement is shown in Figure 1. The specimens are tested by a Zwick 400
uniaxial loading machine, where an optical extensometer is used for measuring the axial
deformation within the effective length. The experiments are strain controlled therefore the
images are evaluated in real time. The used video camera is a monochrome VGA camera,
which can detect the boundary of the black and white colors, therefore monochrome marks are

198
Figure 1. Testing frame with hydraulic grips and specimen.

used. The surface of the specimens is painted white, and on this whitewash background a black
permanent marker is used. The non-contacting video extensometer is a very effective tool to
evaluate displacements, the measurement setup, however, is very sensitive to environmental dis-
turbances. An appropriate slope of the marks is essential (approximately 6°), because it makes
possible to determine the line much more accurately than the pixel size. The blurred focus, the
constant light and the anti-twinkle surfaces are also important.
For the monotonic load paths it is possible to install high accuracy clip-on extensometers,
therefore a digital Zwick equipment is also installed (accuracy class 0.5), nevertheless during
the cyclic program, only the optical device is used.
Important aim of the experimental program is the cyclic investigation of European steel grades.
More than 100 specimens are investigated; the steel materials are summarized in Table 1. The
investigated conventional steel grades (S235-S460) are used in the practice of the European steel
construction society, and the steel plates came from four European countries (Germany, Italy,
Spain, and Ukraine). The effect of the rolling direction is also investigated; both rolling (X) and
perpendicular (Y) specimens are analyzed.
Testing standards do not describe specific protocols for cyclic steel investigations. The pro-
posed load paths are designed for determining all the important characters of uniaxial cyclic
behavior. Based on a sufficient experimental program, the response of the steel material to an
arbitrary inelastic load path can be predicted, and the program gives the essential background
for material model calibration. Some of the suggested load paths identical with those already
used in the literature to have a common basis for comparing the results (Dusicka et al., 2007,
Popov and Petersson, 1978, Shi et al., 2011).
The herein used protocols are summarized in Figure 2. A significant portion of the investi-
gation is the constant amplitude, symmetric, fully reversed fatigue loading (C-3 series), which

Table 1. Investigated steel grades.

Number of specimens

Monotonic Cyclic
Country
Grade of origin Grade ID X direct. Y direct. X direct. Y direct.

S235 J0 Spain (es) CX1 CY2 4 4 21 4


S355 J2+N Spain (es) CX3 4 4 7
S355 J2+N Italy (it) CX4 4 7
S355 J2+N Ukraine (ua) CX2 CY2 4 22 4
S460 NL Germany (ge) CX5 4 4 8
Total 20 12 65 8

199
Figure 2. Most important cyclic loading protocols.

is essential to explore the isotropic steel behavior under different strain levels. Considering
symmetric and asymmetric cases, the fatigue behavior of the steel material is better predictable
(C-3 and C-4), and the interaction can be obtained by superposing it on different constant
strain levels.
Increasing amplitude load paths with different cycle numbers of strain levels (C-1) give the
opportunity to compare the kinematic and isotropic proportions of the hardening, and the
interaction between the hardening and the strain levels.
The load memory effect can be predicted by preloaded protocols (C-1-5), and the compari-
son between the increasing and decreasing protocols also provide additional information to
the memory effects. Several special protocols (C-6) show the effect of the low strain amplitude
load cycles during the load path, and the fading memory effect, too. Note that the cyclic ratch-
eting and plastic creep phenomena are not studied, since these questions are widely investi-
gated fields of the steel plasticity in the technical literature.

3 EXPERIMENTAL RESULTS

3.1 Cyclic hardening under different strain levels


As the first step of the cyclic test results evaluation, the cyclic hardening under constant strain
amplitude is presented here. Most of the cyclic experiments are performed at ±2, ±4, ±6, ±8%
strain levels, and in case of S235 steel grade under ±12%, too. In Figure 3a the hardening
behavior of S235 J0 steel material is presented under different load paths (engineering stress is
used). The hardening under different strain level is distinct, in the case of larger strain level the
hardening is also larger. In Figure 3b the first load path, and the subsequent additional cyclic
hardening of different steel grades are presented, and the results of S235 J0 are highlighted.
The cyclic hardening behavior at a given strain level is similar, the 65–70% of the total
cyclic hardening appears after the first cycle, and additional 10–12% appears in the second
cycle. In the subsequent 4–6 cycles it gradually decreases, and finally the cyclic hardening
shakes down, and a constant hardening can be observed up to the fatigue fracture of the speci-
men. For large inelastic loading (8 and 12%), only the results of a few cycles are presented,
because the stress-strain curve showed decreasing behavior after some cycles caused by the
plastic cyclic buckling of the specimen.
From the results it can be concluded that under the investigated strain amplitude (εamp ≥ 2%)
the differences between the cyclic hardened and monotonic curves slightly increase, the hard-
ened stress-strain curve can be calculated by an offset from the monotonic result. The curves in

200
Figure 3. Comparative analysis of cyclic hardening under various strain levels: (a) stress-strain curves
of S235 (b) cyclic hardening under different strain levels.

Figure 3b are normalized by the yield strength to make the different steel grades compar-
able. It is observable, similarly to the monotonic hardening, that the normalized value of the
additional cyclic hardening also reduces as the grade increases. The gap between the red stars
and the monotonic curve represents the hardening of S235 J0. This curve belongs to low-
grade steels, and the additional cyclic hardening is 50% of the yield strength. The blue line and
stars describe the behavior of S460 NL, where this gap is smaller. The additional cyclic hard-
ening is only 25%.
The cyclic ultimate strength, which include the additional cyclic hardening as well results
larger resistance than it could be measured by monotonic coupon test. This effect can be
important in seismic design and also critical for numerical modelling. Based on the performed
experimental program, an approach proposed for determining the maximum cyclic hardening,
if only monotonic results are available, as shown in Figure 4.

3.2 The ratio of the isotropic and kinematic hardening


Important element of the cyclic steel behavior is the ratio of isotropic and kinematic harden-
ing, since these two types of hardening always should be considered in numerical modelling.
On the basis of test results it is concluded, that the smallest yield surface size appears at
approximately 0.5% strain level (saturation of Bauschinger effect). The further expansion of
the yield surface is equivalent with the isotropic part of the hardening thus the kinematic hard-
ening is also predictable. Figure 5 shows the isotropic and kinematic hardening components,
and the change of the yield surface, including the saturation of yielding zone. The kinematic
hardening develops within 1–2% strain level, and after this threshold, its characteristic does
not change significantly. The isotropic hardening, however, gradually increases till the limit of

Figure 4. Fitted linear curve for the prediction of cyclic ultimate stress.

201
Figure 5. Components of hardening in the case of S235 J0 steel grade under C-1-1 protocol.

the studied 8% strain level, and after 4%, the isotropic hardening cause larger elastic stress
than the original one.
The dashed lines in the figure shows the kinematic hardening tendency. For example ‘Iso+kin.
hard<E/3’ means that stress level, where the slope of the curve is less than E/3, so that stress
includes the total isotropic hardening, and a part of the kinematic hardening, until the tangential
of the stress-strain point become less or equal with E/3.

3.3 Isotropic hardening controlled by accumulated and maximum strain


Figure 6 shows the maximum stresses (σtot) as a function of strain level resulted in C-1-1 and
C-1-3 protocols. Because C-1-3 has 5 cycles at each strain level, this curve is illustrated by
5 points above each other, each point refers to one load cycle. The differences in hardening
caused by the accumulation of plastic strain and the maximum strain levels are clearly
separated.
The green and blue dashed lines in Figure 6 represent the maximum values (σtot) of con-
stant amplitude load paths under 2 and 4% strain levels (C-3-1 and C-3-2). Each point rep-
resents the total stress level after every cycle. In the figure not all the dots of the constant
curves are marked, just as many as the equality of strain accumulation with the C-1-3
protocol requires on the same strain level. Therefore in these cases, both the maximum
load level, and the accumulation of strain are identical, although the two load paths are
different. The total stresses of the two different strain paths are in a good agreement under
2 and 4% strain levels as well. Based on the above detailed explanation a model of the
isotropic hardening is developed, which has two components: (i) one is dependent on the
maximum strain level (Δσisomax), and it develops under a few load cycles; (ii) the second is
the additional isotropic hardening component (Δσaccum), which depends on the accumulated
plastic strain and always increasing.

Figure 6. Investigation of isotropic hardening caused by the accumulation of plastic strain.

202
Figure 7. Memory behavior of increasing and preloaded protocols.

3.4 Fading memory behavior


Considering the increasing load paths, the cyclic steel behavior is determined primarily by the
maximum strain range (qtot). Decreasing or preloaded protocols show that, the effect of the larger
amplitudes gradually disappears from the stress-strain relationship. It is similar to the cyclic hard-
ening but it requires much more cycles to evolve. Special protocols are designed (C-1-5, C-2-1,
C-2-2, and C-5-3, C-6 series) in order to explore these features of plasticity. C-1-5 shows the effect
of a large amplitude preloading, C-5-3 protocol explores this property of the cyclic plasticity.
The effect of the preloading can be investigated by the comparison of the C-1-2 and C-1-5
load paths, primarily by the changes of yield surface and total stress (σtot) (Figure 7.). Signifi-
cant hardening can be observed caused by the preloading of the material, which produce the
increased elastic stress and cyclic hardening, especially at smaller strain levels, too. Neverthe-
less during the cyclic loading the effect of the preloading becomes smaller, and the yielding
and hardening behavior approach the conditions of the conventional (C-1-2) specimen. The
reduction continues up to 2% strain, and after that cyclic hardening appears, although this
hardening is smaller than that of the original load path. When the steel reaches 3.5% strain,
the differences between the preloaded and un-preloaded curves are almost eliminated.
Figure 8 compares the effect of the number of these small loops in constant cyclic loading.
C-3-2 is the curve of constant large strain range (± 4%), C-6-2 represents the small strain range
behavior, C-6-3 and C-6-4 are the preloaded protocols with different number of small strain
amplitudes. Similar nonlinear tendency is observed as under cyclic hardening, however it
requires more – approximately 8 to 10 – cycles to evolve. When the amplitude ratio between
the large and small strain range is equal to 4:1 only 2/3 of the additional cyclic hardening can
disappear, so the previously experienced larger strain ranges only partially disappear.

Figure 8. Memory behavior: effect of small amplitude between larger amplitudes.

203
4 CONCLUSIONS

In the current cyclic steel material experimental program several important cyclic features are
investigated. The summary of the obtained cyclic results are summarized as follows:
– Although the monotonic hardening depends on the strain level, the additional cyclic hard-
ening increases only from 0 to 1–1.5% strain level, and after this threshold, the additional
cyclic hardening appears as a constant value added to the monotonic curve. The cyclic
hardening depends on the ultimate stress, higher steel grades have moderated monotonic
and cyclic hardening compared to lower grades.
– An approximation of the isotropic hardening is developed using two components: (i) iso-
tropic hardening component, which depends on the maximum strain level (Δσisomax), and it
develops under a few load cycles; (ii) additional isotropic hardening component (Δσaccum),
which depends on the accumulated plastic strain and always increases.
– The investigated steel materials show fading memory behavior. The effect of the maximum
strain level gradually disappears, the shakedown of the softening, requires more cycles than
the hardening. The effect of the small strain amplitudes (Δq ≤ 1%) is negligible, and when
the strain amplitudes are larger than 60% of the maximum, the effect of the whole preload-
ing can disappear.

ACKNOWLEDGEMENT

The research reported in this paper was supported by the FIKP grant of EMMI in the frame
of BME-Water sciences & Disaster Prevention (BME FIKP-VÍZ).

REFERENCES

Chaboche, J.L., Dang Van, K. & Cordier, G. 1979. Modelization of the strain memory effect on the
cyclic hardening of 316 stainless steel. Structural Mechanics in Reactor Technology. Transactions.
Dusicka, P., Itani, A.M. & Buckle, I.G. 2007. Cyclic response of plate steels under large inelastic strains.
Journal of Constructional Steel Research, 63, 156–164.
Lee, G.C. & Chang, K.C. 1991. The experimental basis of material constitutive laws of structural steel
under cyclic and non-proportional loading. in Fukumoto, Y. & Lee, G.C. Stability and Ductility of
Steel Structures under Cyclic Loading, pp. 3–13. Osaka, Japan.
Mima, Y., Abdel-Karim, M. & Ohno, N. 2000. Uniaxial ratchetting of 316FR steel at room temperature -
Part I: Experiments. Journal of Engineering Materials and Technology, 122, 29.
Popov, E.P. & Petersson, H. 1978. Cyclic metal plasticity: experiments and theory. Journal of the Engin-
eering Mechanics Division, 104, 1371–1388.
Ruggles, M.B. & Krempl, E. 1990. The interaction of cyclic hardening and ratchetting for AISI type 304
stainless steel at room temperature - I. Experiments. Journal of the Mechanics and Physics of Solids,
38, 575–585.
Shi, Y., Wang, M. & Wang, Y. 2011. Experimental and constitutive model study of structural steel under
cyclic loading. Journal of Constructional Steel Research, 67, 1185–1197.
Silvestre, E., Mendiguren, J., Galdos, L. & De Argandoña, E. S. 2015. Comparison of the hardening
behaviour of different steel families: From mild and stainless steel to advanced high strength steels.
International Journal of Mechanical Sciences, 101, 10–20.
Wang, Y.B., Li, G.Q., Cui, W., Chen, S.W. & Sun, F.F. 2015. Experimental investigation and modeling
of cyclic behavior of high strength steel. Journal of Constructional Steel Research, 104, 37–48.
Zhang, S.L. & Xuan F.Z. 2017). Interaction of cyclic softening and stress relaxation of 9–12% Cr steel
under strain-controlled fatigue-creep condition: experimental and modeling. International Journal of
Plasticity 98: 45–64.
Zhou, F., Chen, Y. & Wu, Q. 2015. Dependence of the cyclic response of structural steel on loading his-
tory under large inelastic strains. Journal of Constructional Steel Research, 104, 64–73.

204
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Simulation based imperfections and their effects


on stability resistance

V. Budaházy, D. Kollár & L.G. Vigh


Department of Structural Engineering, Budapest University of Technology and Economics, Budapest,
Hungary

ABSTRACT: The fabrication process of a steel structural member fundamentally deter-


mines its imperfections. The accuracy of cutting, the spring-back during cold forming, and the
specification of welding procedure leads to residual stresses and distortions. In design accord-
ing to the Eurocodes, engineer can use the equivalent geometric imperfections in geometrically
and materially nonlinear analysis simulation. In certain cases, it is difficult to determine the
amplitude of equivalent geometric imperfections, which can take the fabrication process into
account, and the detailed knowledge of the stress and distortion can result in a more accurate
design process. Primary aim of this paper is to illustrate the methodology of simulation based
imperfection analysis and resistance computation, as well as to discuss the applicability of the
method. In the paper, the calculation of axially compressed, cold-formed and welded members
is presented, where the total fabrication and loading process are investigated based on numer-
ical simulations. We compare the resistance of the member using equivalent geometric imper-
fections and considering the complete fabrication process.

1 INTRODUCTION

The fabrication process of a steel structural member fundamentally determines its imperfec-
tions and thus strongly influences its stability resistance. In certain cases, it is difficult to deter-
mine the amplitude of equivalent geometric imperfections, which can take the fabrication
process into account, and the detailed knowledge of the stress and distortion can result in
a more accurate design process. Primary aim of this paper is to illustrate the methodology of
simulation based imperfection analysis and resistance computation, as well as to discuss the
applicability of the method.
Thin-walled beam members, cylindrical rolls or continuous conical are usually fabricated
by cold forming. The numerical modelling of these fabrication processes are one of the most
reliable method to determine residual stresses, imperfections, and evaluate the effect of the
specification of cold forming. The detailed description of three-roll and four-roll bending
simulations can be found in the technical literature, several papers focuses on the connection
between the desired curvature and rollers position and parameter for single or multi-pass
bending. (Gandhi and Raval, 2008)
The major aim of welding simulations is to predict residual stresses and deformations that
have a large influence on the structural resistance. Finite element analysis of welding is not
commonly performed in civil engineering practice. However, several examples can be found in
the international literature focusing on virtually manufactured specimens. For instance,
Moradi & Pasternak (2017) studied the influence of various welding sequences on the strength
of square hollow section steel T-joints. Jiang et al. (2017) carried out a numerical study regard-
ing the effect of manufacturing process on the flexural buckling strength of high strength steel
columns with box sections. Kollár et al. carried out welding simulations focusing on residual
stresses of butt-welded plates and local plate buckling of box sections in 2017, residual stress
distribution in high strength steel columns with box sections in 2017 and welding of

205
trapezoidal corrugated web girders to evaluate the impact of welding variables and model
parameters on residual stress distributions in 2018. Kollár et al. (2019) also proposed a weld
process model for metal active gas welding for finite element applications.
Some of the examples listed above are dealing with the influence of welding on the load bear-
ing capacity of different geometries and configurations. However, according to the authors’
knowledge, the influence of distinct manufacturing processes, such as cold forming, thermal cut-
ting, welding, machining, etc., are not linked directly in resistance calculations. On the other
hand, standards such as EN 1993-1-5:2006 or EN 1993-1-6:2007 give recommendations on both
local and global equivalent geometric imperfections for the analysis of plated or shell structures,
respectively. It has been widely used in the civil engineering practice for advanced design or
research. The calculations are based on equivalent geometric imperfections instead of using real-
istic residual stress patterns and geometrical distortions due to manufacturing. For instance, the
recommendation of the standard for the amplitude of equivalent bowing imperfection is e0 = L/
200 for cold-formed hollow sections (buckling curve ‘c’).
The current paper takes the influence of cold forming and welding into account directly in
geometrically and materially nonlinear analyses with imperfections (GMNIA) by simulating
both manufacturing processes and the subsequent virtual tests. A numerical approach is devel-
oped in order to consider the whole manufacturing process chain. The study focuses on the
flexural buckling resistance of columns with tubular sections and the local buckling resistance
of cylindrical shell structures.

2 DEVELOPMENT OF NUMERICAL APPROACH

The primary research aim of the current paper is to develop a reliable finite element frame-
work to both simulate the manufacturing process and complete virtual experiment. Figure 1
illustrates the framework through an example on flexural buckling of a cold-formed circular
hollow member.
The figure shows von Mises residual stresses after each step of the manufacturing process,
(i) cold forming and (ii) welding. Once the imperfections due to cold forming and welding are
obtained, virtual testing for the calculation of buckling strength is completed, as illustrated in
Detail (iii) and (iv). The procedure can be used for two-dimensional and three-dimensional based
models as well, however, an additional step is necessary for two-dimensional models. Extruding
the models in the axial direction makes it requisite to map initial variables for all the cross sec-
tions. The problem is solved in one load step in order to achieve equilibrium of internal forces and
moments. The deformed configuration can be upgraded based on plasticity-based distortion ana-
lysis. Details can be found regarding the plasticity-based distortion analysis in the work of Jung &
Tsai (2004). It is important to highlight, that a different mesh can be used for virtual testing if
nodal stresses, strains and distortions are stored in arrays based on nodal coordinates. Thus, it is
possible to map these variables using interpolation based on the current mesh and nodal coordin-
ates by performing matrix operations on array parameter matrices in ANSYS (2016), a general-
purpose finite element program that is used in the simulations.

3 NUMERICAL MODELLING OF COLD-FORMING PROCESS

A typical fabrication process for creating large tubes and cylindrical steel members is the con-
tinuous three-roll bending of steel plates. The fabrication process is relatively simple, however,
it requires careful machine control to achieve the desired curvature, especially at the edges,
where undesired planar zone may appear. Typical practice of the roller bending still highly
depends upon the experience and ability skill of the operator. Using templates, or working by
trial and error, is a common practice in the industry.
The curvature can be generated by the movement and rotation of the top roll. The most
economical way to produce cylinder is to roll the plate with a single pass. High accuracy and
the knowledge of the spring-back (Gandhi & Raval, 2008) are required for the use of the one-

206
Figure 1. Steps of the developed approach for virtual testing of cold-formed and welded specimens.

pass method. Yang and Shima (1988) describe the bending process of a pass. The plate with
an initial radius is fed into the machine at the first bottom roller and exist at the second, with
a final curvature. The maximum bending moment and curvature appear at the contact point
of the top roller. The distribution of the curvature is asymmetric about the middle contact
point, and this point is not exactly on the symmetry axis of the rollers. Due to the above diffi-
culties, it is rather typical to create the shell by applying several (approximately 6–10) paths,
depending on the skill of the operator.
In order to simulate the conventional three-roll-bending fabrication process, numerical
investigations were performed. The numerical model was developed in ANSYS LS-DYNA
module. The plate is modelled with three-dimensional solid elements, while the rollers are rep-
resented by rigid elements. We used the following configurations:
– stiffness of the machine and rollers is considered infinitely rigid,
– 20-mm element size of the plate and 5 elements through the thickness of the plate are
applied,
– plastic steel material model is assigned to the plate,
– linear solid elements with 8 nodes are used.
The developed model is invoked for a parametric study to calculate geometrical imperfections
and residual stresses due to cold forming of tubular members with various radius-to-thickness
ratios. Out-of-roundness values, representing an ellipse, regarding the normalized tangent
points and residual stresses over the normalized thickness are shown for structural steel S275
with a nominal yield strength of 275 MPa. Radius-to-thickness ratio varies between 10 and
115. The obtained results are illustrated in Figure 2.

207
Figure 2. Imperfections: out-of-roundness and residual stress.

4 NUMERICAL SIMULATION OF WELDING

Welding simulation of cold-formed hollow sections and a cylindrical shell structure are carried
out. Uncoupled transient thermomechanical analysis is implemented in order to determine
residual stresses and deformations due to manufacturing. Two-dimensional models, assuming
generalized plane strain, and three-dimensional models are also investigated (Figure 3).
Uncoupled thermomechanical analysis means that calculated temperature fields due to weld-
ing are applied as nodal loads in the subsequent mechanical analysis. The thermal analysis is
performed on a perfect geometry, while residual stresses induced by cold forming are set as
initial variables and residual deformations are taken into account in the upgraded geometrical
configuration, i.e., imperfect geometry. These effects are rarely taken into account in case of
arc welding. In the current study, 4-node plane elements and 8-node solid elements are used in
the two-dimensional and three-dimensional finite element models, respectively. The welded
joint is created with filler material addition. Therefore, initial gaps and deposited material are
modelled in the simulation. The ‘birth and death’ procedure and the ‘quiet element technique’
are added in the thermal and mechanical analyses in order to deal with melting and solidifica-
tion. Thermal and mechanical material properties are based on EN 1993-1-2:2005. Verifica-
tion of welding simulation, the details of the material models and thermal boundary
conditions can be found in Kollár et al. (2019). The double ellipsoidal heat source model
introduced by Goldak et al. (1984) is applied. In the mechanical analysis, clamping conditions

Figure 3. Typical meshing of a) two-dimensional and c) three-dimensional models (a part is shown)


used in transient welding simulations.

208
Figure 4. Von Mises stresses of 2D and 3D models in different time steps in [MPa].

have an important role on the evolution of deformations and stresses. In the current paper,
solely the rigid body motion is avoided during welding. Large deflection effects are also taken
into account in the mechanical analysis.
Basically, 328 elements and 388 nodes are used for two-dimensional models and 131200 elem-
ents and 154786 nodes are used for three-dimensional models in the welding simulation (Figure 3).
A finer mesh is used in the vicinity of the weld, while a coarser mesh is adequate far from the weld
bead. Net heat input is 0.6 kJ/mm, while welding speed is 10 mm/s in the analyses. Outer diameter
is d = 100 mm, plate thickness is t = 4 mm. The length of the specimen is L = 4000 mm for the
solid model. Von Mises stresses are showed in Figure 4 for two-dimensional and three-
dimensional cases in different time steps. Stresses are presented before, during and after welding.
Vertical deformations (bowing imperfections) are solely presented for the three-dimensional
models in Figure 5 as it cannot be expressed for the two-dimensional model. Results confirm
that the simulation of the whole manufacturing process chain makes it possible to determine
overall residual stresses and distortions.

5 SIMULATION BASED RESISTANCE CALCULATION – FLEXURAL BUCKLING

In this section, flexural (global) buckling of columns with circular hollow sections is presented
to show the applicability of the approach. Analyses either with simulated real imperfections
(geometric imperfections and residual stresses) or equivalent geometric imperfections in
accordance to Eurocode are completed and compared. Bowing imperfection is especially
important when dealing with flexural buckling, while out-of-roundness and dimples are

209
Figure 5. Vertical deformations (5 × scaling) of a three-dimensional model in different time steps in [mm].

crucial for local buckling of cylindrical shells. Imperfect shapes are based on the combination
of corresponding eigenshapes using linear bifurcation analysis in the case of equivalent geo-
metric imperfections. The same nonlinear material models are used in the analyses.
The configuration of the columns or trusses with circular hollow sections are specified
based on the non-dimensional slenderness defined by EN 1993-1-1:2005. Four geometries are
modelled with the same cross sectional data (outer diameter is d = 100 mm and plate thickness
is t = 4 mm) and nominal yield strength ( fy = 355 MPa), while the lengths of the specimens are
L = 1000, 2000, 3000 and λ 4000 mm, yielding the non-dimensional slenderness values of 0.39,
0.77, 1.16 and 1.54, respectively. Simulations are displacement-controlled.
First, the difference of three-dimensional and two-dimensional model based virtual testing
is compared using the specimen with L = 4000 mm. The corresponding force-displacement
curves are showed in Figure 6. The figure confirms the high accuracy of the 2D model:
although there are some deviations at the initial elastic range, the flexural buckling strength is
practically the same at the two models (101.2 kN vs. 102.3 kN).
Hereafter, the virtual tests are performed on extruded two-dimensional models (as written in the
first paragraph of section 3). Reduction factors χ are calculated using the Eurocode (EC) formulae in
order to determine the difference between finite element (FEM) based and EC calculations. Results
are summed up in Figure 7 also indicating the standard buckling curves. Each reduction factor, that
is plotted in the figure, is designated with VT (virtual testing, including residual stresses and geomet-
ric imperfections), RS (residual stresses are included, geometric imperfections are ignored) or GI
(geometric imperfections are included, residual stresses are ignored) and a number denoting the
length of the structural member under pure compression. The simulation results allow the back-
calculation of an equivalent bowing imperfections (e0,EGI), that results in the same flexural buckling
strength as virtual testing of manufactured specimens (Figure 7). Applying solely residual stresses or
geometric imperfections makes it possible to estimate which type of imperfection is dominant is the
case of the investigated circular hollow sections. It is important to highlight, that the eccentric weld
bead results in a specific residual stress pattern and typical bowing imperfection. Therefore, accord-
ing to the reduction factors derived from FEM results, geometric imperfections are dominating for
these specimens assuming the manufacturing technology detailed in previous sections.
The maximum difference between FEM and EC based reduction factors is approximately 20%.
The interpretation of this result is two-fold: a) it can be concluded that taking realistic initial vari-
ables and residual deformations into account may result in smaller reduction factors than calcu-
lated on the basis of current Eurocode provisions, especially, for slenderer members; b) our

210
Figure 6. Two- and three-dimensional model based virtual tests for the 4000 mm long structural
member.

Figure 7. Reduction factors for different non-dimensional slenderness ratios.

numerical approach underestimates the buckling strength. Further study is needed to validate the
numerical model. Additional fixtures and straightening after welding are neglected. The manufac-
turing analysis results in larger bowing imperfections than the essential manufacturing tolerance
(L/1000) recommended by EN 1090-2:2018 for single-storey columns. Buckling curve ‘c’ is used in
the standard for hollow sections, nevertheless, curve ‘d’ may be on the safe side in some cases. The
tendency of the proposed values for the e0,EGI equivalent amplitude of bowing imperfection clearly
shows that the slenderer the structural member, the more rigorous condition is set. It is between
L/400 and L/100 for the investigated cases, while as compared with the standard, EN 1993-
1-1:2005 recommends to apply bowing imperfection with an amplitude of L/150 for linear and L/
200 for plastic analysis. Clause 5.3.2 (11) in the the standard proposes an alternative method to
determine the amplitude of bow imperfections resulting in amplitudes between L/650 and L/1200.

6 CONCLUDING REMARKS

A numerical study is carried out in the paper focusing on simulation based imperfections and
their effects on buckling strength. Results show that the developed approach makes it possible
to efficiently determine residual stresses and distortions due to whole manufacturing process
chain. Finally, virtually manufactured (cold-formed and welded) specimens are virtually
tested for buckling. The results call attention to the possible need of review of the Eurocode
provisions. The method can be further applied in other fields of stability analysis: e.g. shell
buckling analysis as shown in Figure 8. is a possible application. It is though straightforward
that further study is needed for the validation of the numerical model.

211
Figure 8. Failure modes of cylindrical shells (EGI & VT).

ACKNOWLEDGEMENT

Support of grant BME FIKP-VÍZ by EMMI is kindly acknowledged.

REFERENCES

ANSYS 2016. ANSYS 17.2 Mechanical APDL. ANSYS Inc, Pittsburgh, Pennsylvania, USA.
EN 1090-2:2018. Execution of steel structures and aluminium structures. Part 2: Technical requirements
for steel structures.
EN 1993-1-1:2005. Eurocode 3: Design of steel structures – Part 1-1: General rules and rules for buildings.
EN 1993-1-2:2005. Eurocode 3: Design of steel structures – Part 1-2: General rules –Structural fire design.
EN 1993-1-5:2006. Eurocode 3: Design of steel structures – Part 1-5: Plated structural elements.
EN 1993-1-6:2007. Eurocode 3: Design of steel structures – Part 1-6: Strength and stability of shell
structures.
Gandhi, A. & Raval, H. 2008. Analytical and empirical modeling of top roller position for three-roller
cylindrical bending of plates and its experimental verification. Journal of Materials Processing Technol-
ogy 197(1–3): 268–278.
Goldak, J.A., Chakravarti, A. & Bibby, M. 1984. A new finite element model for welding heat sources.
Metallurgical Transactions B 15: 299–305.
Jiang, J., Lee, C.K., Chiew, S.P. & Yee, P.L.T. 2017. Numerical investigation of high-strength built-up
box columns. Proceedings of the Institution of Civil Engineers: Structures and Buildings 170: 653–663.
doi:http://dx.doi.org/10.1680/jstbu.16.00111.
Jung, G.H. & Tsai, C.L. 2004. Plasticity-Based Distortion Analysis for Fillet Welded Thin Plate T-Joints.
Welding Journal 83(6): s177–s187.
Kollár, D., Kövesdi, B. & Néző, J. 2017. Numerical simulation of welding process – Application in buck-
ling analysis. Periodica Polytechnica Civil Engineering 61: 98–109. doi:http://dx.doi.org/10.3311/
PPci.9257.
Kollár, D. & Kövesdi, B. 2018. Effect of imperfections and residual stresses on the shear buckling
strength of corrugated web girders. In Camotim, D. and Silvestre, N. (eds.), Proceedings of the Eighth
International Conference on Thin-Walled Structures, Lisbon, p. 20.
Kollár, D., Kövesdi, B., Vigh L.G. & Horváth, S. 2019. Weld process model for simulating metal active
gas welding. The International Journal of Advanced Manufacturing Technology, DOI: 10.1007/s00170-
019-03302-3.
Moradi, M. & Pasternak, H. 2017. A Study on the Influence of Various Welding Sequence Schemes on
the Gain in Strength of Square Hollow Section Steel T-Joint. Journal of Welding and Joining 35:
41–50. doi:10.5781/JWJ.2017.35.4.7.
Somodi, B., Kollár, D., Kövesdi, B., Néző, J. & Dunai, L. 2017. Residual stresses in high-strength steel
welded square box sections. Proceedings of the Institution of Civil Engineers: Structures and Buildings
170: 804–812. doi:10.1680/jstbu.16.00139.
Yang, M. & Shima, S. 1988. Simulation of pyramid type three-roll bending process. International Journal
of Mechanical Sciences 30(12): 877–886.

212
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Development of an innovative multi-performance system for


LWS structures

A. Campiche
Department of Structures for Engineering and Architecture, University of Naples “Federico II”, Naples, Italy

ABSTRACT: Nowadays the key topics of the construction sector are safety and eco-efficiency.
In the last decades, the Lightweight Steel (LWS) systems made of Cold Formed Steel (CFS) pro-
files have shown high structural and environmental performances, joining perfectly the new trend.
In this perspective, the University of Naples “Federico II”, in cooperation with Lamieredil S.p.A.
Company, has recently started a new research project. The main goal of the project is the develop-
ment of a new solution with high seismic and environmental performances. From the structural
point of view, the innovation consists in a CFS profile wall, equipped with pre-tensioned Ultra
High Strength (UHS) steel braces, which are able to limit global displacement of structure and to
dissipate seismic energy by its yielding. The effectiveness of the wall system will be proved through
several tests carried out at Lab of the Department of Structures for Engineering and Architecture.
The experimental activity will include tests on materials and assemblies, monotonic and cyclic
tests on full-scale components and shake-table tests on a building mock-up. From the environ-
mental point of view, more appropriate materials for the envelope will be selected and production
and erection processes of the system will be analysed, in order to obtain higher acoustic and ther-
mal performances and reduction of waste. As a conclusion, a prototype building will be erected,
tested and monitored on the company property. The paper summarizes the research project in
detail, the experimental program and the design of the innovative system and prototype.

1 INTRODUCTION

The modern civil engineering aims to make constructions increasingly safe, sustainable and
comfortable. This goal can be easily reached through a joint synergy among different engin-
eering field. Moreover, selection of more appropriate materials and optimization of produc-
tion and erection processes, through prototyping and digital fabrication, play a key role for
the ecoefficiency of buildings. The last research activities have highlighted the ability of steel
structures (Costanzo 2017, Nastri et al. 2017, 2019, Tartaglia et al. 2018), and in particular of
Lightweight Steel (LWS) systems made of Cold Formed Steel (CFS) profiles (Macillo et al.
2014, 2017, 2018, Fiorino et al. 2014, 2016, 2017a, b, c, d, 2018a, b, 2019, Pali et al. 2018,
Terracciano et al. 2018), to exhibit high structural and environmental performances. In this
perspective, a new research project has been started at University of Naples “Federico II” in
cooperation with Lamieredil S.p.A. Company, which aims to develop innovative technological
solutions with higher structural and environmental performances.
The current paper describes the research project in detail, the experimental program, the
selected case study and the preliminary design of the new system. The procedures and the set-
ups developed to evaluate the seismic building performances will be deeply described.

2 RESEARCH PROJECT AND EXPERIMENTAL PROGRAM

The proposed research, in cooperation with Lamieredil S.p.A. Company, has a main objective of
developing a new technological solution for lightweight constructions with higher anti-seismic,

213
thermal and environmental performances, which also allows to optimize the production processes.
The innovative system mainly consists of CFS profiles equipped with pre-tensioned Ultra High
Strength (UHS) steel braces. The braces act as an anti-seismic device and it is able to limit global
displacement of structure by working as an elastic spring, and dissipating the seismic energy
through the yielding.
The main goal of the experimental activity is the evaluation of seismic response of the struc-
tural system. In particular, the lateral response of three different wall configurations equipped
with the anti-seismic device will be tested through the monotonic and quasi-static cyclic tests
on full-scale wall specimens. In order to understand the improvement given by the new wall
system to the global response, shake table tests will be carried out on a mock-up representative
of the whole building.
Another important outcome of the research project is the analysis and selection of materials
and products, generally in the form of rigid or flexible panels for forming the envelope, which
should be more appropriate to use in lightweight steel construction in order to increase their
thermal, acoustic and environmental performances. For the thermal and environmental opti-
mization of the LWS construction performances, the selected materials/products will be used
to develop new multifunctional design solutions. Moreover, to increase the environmental per-
formance, the production process will be controlled, through the use of CAD to CAM tech-
nologies and BIM software, with a significant reduction in waste.
In order to demonstrate the effectiveness of the proposed solutions, in all the aspects (struc-
tural, thermal and environmental) and the simplicity of the production and erection process,
a full-scale prototype will be erected on company premises. Then the building will be tested
and monitored to evaluate the thermal and environmental behaviour in a real environment.

3 CASE STUDY

The selected case study is a two storey building located in Sellia Marina, Calabria region (Italy). It
is representative of the full-scale prototype that will be erected on Lamieredil S.p.A. Company
property and used as exhibition building. The building covers an area of about 42 m2 with
a maximum height of 8 m. The entrance to the building is located on the south-east side at road
level. The wide openings on this front are protected by movable shielding elements in wooden
slats, which ensure the optimization of natural inputs and the control of solar radiation during the
summer months. The north-west front has small openings, useful for the exploitation of natural
ventilation (cross-ventilation). On the inclined roof, 11 m2 of photovoltaic panels and a mini-wind
turbine will be installed. A solar chimney on the roof will serve for heating purposes. Natural ven-
tilation mechanisms will allow better indoor air circulation. A heat recovery ventilation system
(MVHR) can additionally contribute to the control of the internal microclimate and energy con-
sumption. It will allow to use outdoor fresh air, filter and preheat it in winter, before feeding into
indoor environments. The building can also be equipped with a system for recovery and reuse of
rainwater and grey water, as well as an “intelligent” system able to manage the microclimatic and
lighting conditions. The vertical opaque and transparent closures shall ensure high performance
not only in relation to thermal transmittance, but also in relation to its behaviour in terms of ther-
mal inertia, phase shift and attenuation of the thermal wave. It is also important to underline the
possibility to build the enclosure using dry construction processes (structure/covering), that allows
to save money and time, with high technological and environmental performances. Figure 1 shows
building axonometric views.

4 STRUCTURAL DESIGN OF PROTOTYPE

4.1 Design for gravity loads


The building is structurally designed according to Italian building code (NTC2018) and EN1993-
1-3 (CEN 2006). The structure will be realized as a CFS solution by using stick-built system. All

214
Figure 1. Building axonometric views.

CFS elements will be made of S280GD+Z steel grade (with yield, fy, and ultimate, fu, nominal
strength equal to 280 MPa and 360 MPa, respectively).
The design under vertical and horizontal loads is conducted following the “all steel”
approach, neglecting the contribution of panels to the resistance. Floors consist of
260 × 60 × 30 × 3.0 mm (outside-to-outside web depth × outside-to-outside flange size × lip
size × thickness) lipped C section joists spaced at about 600 mm on the centre. The joists are
connected at the ends to unlipped U section floor tracks. The roof is made of CFS profiled
trusses with maximum span equal to 6 m, spaced at centre-to-centre distance of 600 mm. The
top and bottom chords are made up of 70 × 40 × 15 × 1.0 mm Ω-section, while studs and
diagonals have a C-section of dimensions equal to 40 × 30 × 10 × 1.0 mm. Above the trusses,
60 × 40 × 15 × 1.0 mm Ω-section joists are placed, with centre-to-centre distance of 1200 mm.
Floors and roof plans are showed in Figure 2. The walls consist of 150 × 60 × 20 × 1.5 mm
lipped C section studs, connected at the ends to unlipped U section wall tracks.

4.2 Design for seismic loads and preliminary design of the innovative wall
As far as horizontal loads, mainly wind and seismic loads, are concerned, they are resisted by
the new innovative wall system, proposed here. In particular, two wall systems will be placed
for each direction at both storeys.

Figure 2. Floors (a) and roof (b) plans.

215
The innovative wall system is mainly composed of (Figure 3): (1) diagonal braces; (2) devices
for pre-tensioning; (3) chord studs; (4) tracks; (5) hold-downs; (6) blocking and flat straps.
The diagonal braces are pre-tensioned dog bone shaped round bars having thread ends to
allow their connection and pre-tensioning. For the chord studs, a back-to-back C section
has been selected, while for tracks a box section has been chosen. The pre-tensioning device
is a U shape profile connected to the hold-down through a cylindrical hinge, which allows
the rotation in the plane of the wall.
A UHS steel (fy ¼ 1300 MPa, fu ¼ 1450 MPa) is adopted for the diagonal braces, while chord
studs, tracks blocking and flat straps are made of S280 GD+ Z steel grade, and devices for pre-
tensioning and hold-downs are made of S355 steel grade (fy ¼ 355 MPa, fu ¼ 470 MPa).
In order to guarantee three lateral resistance levels, three different configurations are designed:
Light wall (L), Medium wall (M), Heavy wall (H). Italian building code (NTC 2018) provides the
value of behaviour factor for the seismic design of non-dissipative systems, which must be in the
range from 1.0 to 1.5 on the basis of the structural typology of the system. Since the system under
investigation is not covered by Italian building code (NTC 2018), a behaviour factor equal to 1.0
is used in the design in order to have an assumption on the safe side. On the other end, since the
plastic behaviour of the system is matter of investigation, the walls are designed following the Cap-
acity design approach, in such a way to evaluate its dissipative capacity. Therefore, according to
Capacity design approach, the diagonal braces are the dissipative elements and act as fuse of the
system, in which a pre-tensioning equal to half times its yielding resistance is applied. Under this
assumption, diagonal braces work always in tension field until ultimate lateral capacity of the wall
is reached and the tension collapse of the diagonal brace in correspondence to the bar thread area
is avoided by reducing the bar diameter in its middle zone. In particular, the length of the bar with
reduced diameter, lb, is design to allow a lateral story drift of the wall equal to 5%. Therefore,
applying the Equation 1, a length of 1800 mm is chosen.

lb  Δlim h=εy cosα ð1Þ

where: Δlim ¼ maximum lateral drift of the wall set equal to 5%; h ¼ height of the wall; εy ¼ yield
strain of the steel; α ¼ angle of the brace with respect to horizontal.

Figure 3. Schematic drawing of the innovative wall.

216
The design yield capacity of the brace, Nt,Rd, is defined according to EN1993-1-3 (CEN
2006), as shown in Equation 2.

Nt;Rd ¼ Ag f y =γm0 ð2Þ

where Ag ¼ gross cross-sectional area of the fuse; fy ¼ nominal yield strength; γm0 ¼ 1.05 par-
tial safety factor according to the Italian building code (Norme Tecniche per le Costruzioni
2018). The design lateral resistance of the wall Hd is evaluated, considering the acting pre-
tensioning through Equation 3.

Hd ¼ Ag f ypt cosα=γm0 ð3Þ

where fypt ¼ nominal yield strength available in the bar after the pre-tensioning, assumed
equal to half time the nominal yield strength fy. The lateral resistances of the L, M and H wall
Configurations are respectively 107, 157 and 216 kN.
According to capacity design approach, all non-dissipative elements of the wall systems, i.e.
pre-tensioning devices, hold-downs, tracks and chord studs are designed for the expected
resistance of the diagonal braces. In particular, for non-dissipative elements the design resist-
ance Rd is evaluated according to EN1998-1 (CEN 2004), with the Equation 4.

Rd  1:1 γov Rf;Rd ð4Þ

where Rd ¼ resistance of the non-dissipative element; γov ¼ overstrenght factor, set equal to
1.15; Rf,Rd ¼ design plastic resistance of the connected dissipative member (brace), evaluated
on the basis of Equation 2.
The vertical and the horizontal components of the brace force must be transferred through the
chord stud and track elements. For the chord studs back-to-back C sections are considered to
avoid buckling due to additional axial forces under the action of an earthquake. Since blocking
elements and flat straps are included at middle wall height, the chord stud unbraced length is
assumed equal to an half of wall for in-plane global buckling and equal to the wall height for out-
of-plane global buckling. Studs and tracks resistance capacities are evaluated according to
EN1993-1-3 (CEN 2006). Ad hoc hold-down and device for pre-tensioning are designed and their
resistances are evaluated according Italian building code (NTC 2018) and EN1993-1-3 (CEN
2006). All the main structural elements used in the prototype are summarized in Table 1.

5 TESTS DOR SEISMIC PERFORMANCE EVALUATION

5.1 General
The general plan for the experimental assessment of the seismic response provides for the
following main activity: six monotonic tests on full-scale walls (two for each wall configur-
ation); six cyclic tests on full-scale walls (two for each wall configuration) and a series of
shake-table tests on a prototype representative of a whole structure. Ancillary tests for the
mechanical characterization of materials and products used in the wall systems will complete
the plan. All experimental activities will be carried out at the Lab of the Department of
Structures for Engineering and Architecture, University of Naples “Federico II”

5.2 In-plane wall tests


The lateral in-plane behaviour of the selected wall Configurations (L, M, H) will be investi-
gated by means of twelve physical tests, including six monotonic tests and six cyclic tests on
full-scale 2400 mm long and 2800 mm high wall specimens.

217
Table 1. Structural element dimensions (lengths in mm).
Structural element Component element Prototype

Stud C160 × 50 × 15 × 3
Track Box 150 × 50 × 4
Diagonal bar minimum diameter 19
Diagonal bar minimum diameter 21
Wall L Device for pre-tensioning U250 × 110 × 20
Blocking Box 150 × 50 × 4
Flat strap 50 × 3
Hold-down to chord stud fasteners no. 4 M20 bolts Class 10.9
Hold-down to steel beam fasteners no. 1 M24 bolt Class 10.9
Stud C160 × 50 × 15 × 4
Track Box 150 × 50 × 4
Diagonal bar maximum diameter 23
Diagonal bar minimum diameter 26
Wall M Device for pre-tensioning U250 × 110 × 25
Blocking Box 150 × 50 × 4
Flat strap 50 × 4
Hold-down to chord stud fasteners no.4 M22 bolts Class 10.9
Hold-down to steel beam fasteners no. 1 M27 bolt Class 10.9
Stud C190 × 50 × 15 × 4
Track Box 180 × 60 × 5
Diagonal bar maximum diameter 27
Diagonal bar minimum diameter 31
Wall H Device for pre-tensioning U250 × 110 × 30
Blocking Box 180 × 60 × 5
Flat strap 60 × 4
Hold-down to chord stud fasteners no.4 M27 bolts Class 10.9
Hold-down to steel beam fasteners no. 1 M30 bolt Class 12.9
Joist C260 × 60 × 30 × 3
Floor Track U section profiles
Top chord Ω70 × 40 × 15 × 1
Bottom chord Ω70 × 40 × 15 × 1
Roof Stud C40 × 30 × 10 × 1
Diagonal C40 × 30 × 10 × 1
Stud C150 × 50 × 20 × 1.5
Gravity wall Track U section profiles

An available steel frame set-up for in-plane wall tests will be used for the experimental
activity. The wall prototype will be restrained to the laboratory strong floor by the
bottom beam, which has a 300 × 180 × 30 (width × height × thickness) rectangular
hollow section. Horizontal loads will be transmitted to the wall through the loading
beam, which has a 200 × 120 × 10 mm (width × height × thickness) rectangular hollow
section. The out-of-plane displacements of the wall will be restrained by two steel portal
frames equipped with roller wheels. The tests will be performed by using a hydraulic
actuator having a stroke displacement of 500 mm and a load capacity of 500 kN. A slid-
ing-hinge will be placed between the loading actuator and the loading beam, in order to
avoid vertical load components. Monotonic tests will be carried out with displacements
imposed at a rate of 0.10 mm/s until the collapse of specimens occurred, whereas cyclic
tests will be carried out by adopting a loading protocol known as “CUREE ordinary
ground motions reversed cyclic load protocol” developed for wood walls by Krawinkler
et al. (2001). The schematic drawing of test set-up and specimen is shown in Figure 4.

218
Figure 4. Schematic drawing of test-set-up and specimen for in-plane wall tests.

5.3 Shake-table tests


The dynamic behaviour and earthquake response of the whole building will be evaluated through
a series of shake-table tests. The available shaking table available has a square plan of
3.00 × 3.00 m and its main characteristics are: two degrees of freedom, maximum payload equal
to 2000 kN, frequency range of 0 to 50 Hz, peak velocity of 1.0 m/s, and displacement in the range
of ± 250 mm. The tests will be performed on a three-dimensional mock-up representative of the
case study. In particular, two different dynamic test procedures will be carried out for shake table
tests, by applying the seismic input in only one horizontal direction, in which seismic resistant sys-
tems will be located. In order to evaluate the dynamic properties (fundamental vibration period
and damping ratio), low amplitude white noise tests will be performed, whereas earthquake tests
will be conducted in order to evaluate the seismic performance, by applying a selected natural
ground motion with different scaling factors.

6 CONCLUSIONS

The paper presents an ongoing research project at University of Naples “Federico II” in cooper-
ation with Lamieredil S.p.A. Company. The main objective is the development of new solutions
for LWS construction in order to obtain higher seismic, thermal and environmental performances.
A very extensive experimental campaign will be carried out to study the seismic response of
innovative wall systems. The main activities of the research will involve monotonic and cyclic tests
of full-scale walls and shake-table tests on a 3D mock-up, representative of a whole building.
A deep analysis and selection of materials and products to insert in the envelope of wall will be
conducted, in order to maximize the thermal, acoustic and environmental performances. The pro-
duction and erection processes will be controlled, limiting the waste thanks to the help of CAD to
CAM technologies and BIM software. To validate the effectiveness of the selected solutions a full-
scale prototype will be erected, tested and monitored on the company property.

REFERENCES

CEN. 2004. EN 1998-1 Eurocode 8: Design of Structures for earthquake resistance-Part 1: General rules,
seismic actions and rules for buildings. Brussels: European Committee for Standardization.
CEN. 2006. EN 1993- 1-3Eurocode 3: Design of steel structures-Part 1-3: General rules-Supplementary
rules for cold-formed members and sheeting. Brussels: European Committee for Standardization.
Costanzo, S. D’Aniello, M. Landolfo, R. 2017. Seismic Design Criteria For Chevron Cbfs: Proposals
For The Next Ec8 (Part-2). Journal of Constructional Steel Research, 138C: 17–37.

219
Fiorino, L. Iuorio, O., Landolfo, R. 2014. Designing CFS structures: The new school bfs in Naples. Thin-
Walled Structures, Elsevier Science. ISSN 0263-8231. Vol. 78, pp. 37–47. doi:10.1016/j.tws.2013.12.008.
Fiorino, L. Terracciano M.T. Landolfo, R. 2016. Experimental investigation of seismic behaviour of low
dissipative CFS strap-braced stud walls. Journal of Constructional Steel Research, Elsevier Science,
ISSN 0143-974X. Vol. 127, pp. 92–107. doi:10.1016/j.jcsr.2016.07.027.
Fiorino, L., Macillo, V., Landolfo, R. 2017. Experimental characterization of quick mechanical connect-
ing systems for cold-formed steel structures. Advances in Structural Engineering, Multi-Science. ISSN
1369-4332, Vol. 20, No. 7, pp. 1098–1110. doi:10.1177/1369433216671318.
Fiorino, L., Pali, T. Bucciero, B. Macillo, V. Terracciano, M.T. Landolfo, R. 2017. Experimental study on
screwed connections for sheathed CFS structures with gypsum or cement based panels. Thin-Walled Struc-
tures, Elsevier Science. ISSN 0263-8231. Vol. 116, pp. 234–249. doi: 10.1016/j.tws.2017.03.031.
Fiorino, L. Shakeel, S. Macillo, V. Landolfo, R. 2017. Behaviour factor (q) evaluation the CFS braced
structures according to FEMA P695. Journal of Constructional Steel Research, Elsevier Science, ISSN
0143-974X. Vol. 138, pp. 324–339. doi:10.1016/j.jcsr.2017.07.014.
Fiorino, L. Macillo, V. Landolfo, R. 2017. Shake table tests of a full-scale two-story sheathing-braced
cold-formed steel building. Engineering Structures, Elsevier Science. ISSN 0141-0296, Vol. 151,
pp. 633–647. doi:10.1016/j.engstruct.2017.08.056.
Fiorino, L. Shakeel, S. Macillo, V. Landolfo, R. 2018. Seismic response of CFS shear walls sheathed
with nailed gypsum panels: Numerical modelling. Thin-Walled Structures, Elsevier Science. ISSN
0263-8231. Vol. 122, pp. 359–370. doi: 10.1016/j.tws.2017.10.028.
Fiorino, L. Pali, T. Landolfo, R. 2018. Out-of-plane seismic design by testing of non-structural light-
weight steel drywall partition walls. Thin-Walled Structures, Elsevier Science. ISSN 0263-8231.
Vol. 130, pp. 213–230. doi:10.1016/j.tws.2018.03.032.
Fiorino, L. Bucciero, B., Landolfo, R. 2019. Evaluation of seismic dynamic behaviour of drywall parti-
tions, façades and ceilings through shake table testing. Engineering Structures, Elsevier Science. ISSN
0141-0296, Vol. 180, pp. 103–123. doi:10.1016/j.engstruct.2018.11.028.
Krawinkler, H. Parisi, F. Ibarra. L. Ayoub, A. Medina, R., 2001. Development of a Testing Protocol for
Woodframe Structures, CUREE.
Macillo, V., Iuorio, O., Terracciano, M.T., Fiorino, L., Landolfo, R. 2014. Seismic response of Cfs
strap-braced stud walls: Theoretical study. Thin-Walled Structures, Elsevier Science. ISSN 0263-8231.
Vol. 85, pp. 301–312. doi: 10.1016/j.tws.2014.09.006.
Macillo, V., Fiorino, L., Landolfo, R. 2017. Seismic response of CFS shear walls sheathed with nailed
gypsum panels: Experimental tests. Thin-Walled Structures, Elsevier Science. ISSN 0263-8231. Vol.
120, pp. 161–171. doi: 10.1016/j.tws.2017.08.022.
Macillo, V. Shakeel, S. Fiorino, L. Landolfo, R. 2018. Development and Calibration of a Hysteretic
Model for CFS Strap braced stud walls. Advanced Steel Construction.Vol. 14, no. 3, pp. 336–359.
Nastri, E., Vergato, M., Latour, M. 2017. Performance evaluation of a seismic retrofitted R.C. precast
industrial building. Earthquake and Structures, 12 (1), pp. 13–21.
Nastri E., D’Aniello M., Zimbru M., Streppone S., Landolfo R., Piluso V., Montuori R. 2019. Seismic
response of steel moment resisting frames equipped with friction beam-to-column joints. Soil Dynam-
ics and Earthquake Engineering. Vol. 119, pp. 144–157.
NTC. 2018. Ministero delle infrastrutture, D.M. 17/01/2018, Norme Tecniche per le Costruzioni.
Pali, T. Macillo, V. Terracciano, M.T. Bucciero, B. Fiorino, L. Landolfo, R. 2018. In-plane quasi-static cyclic
tests of nonstructural lightweight steel drywall partitions for seismic performance evaluation. Earthquake
Engineering & Structural Dynamics, John Wiley & Sons. ISSN 1096-9845. Vol. 47, pp. 1566–1588. doi:
10.1002/eqe.3031.
Tartaglia, R. D’Aniello, M. De Martino, A. Di Lorenzo, G. 2018. Influence of EC8 rules on P-Delta
effects on the design and response of steel MRF. Ingegneria Sismica: International Journal of Earth-
quake Engineering, Vol. 35, Issue 3, pp.104–120.
Terracciano, M.T. Macillo, V. Pali, T. Bucciero, B. Fiorino, L. Landolfo, R. 2018. Seismic design and
performance of low energy dissipative CFS strap-braced stud walls. Bulletin of Earthquake Engineer-
ing, Springer. ISSN 570-761X. Vol. 17, pp. 1075–1098. doi: 10.1007/s10518-018-0465-y.

220
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Seismic design criteria for CFS steel-sheathed shear walls

A. Campiche, S. Shakeel, L. Fiorino & R. Landolfo


Department of Structures for Engineering and Architecture, University of Naples “Federico II”, Naples, Italy

ABSTRACT: The current European earthquake standard EN1998-1 does not provide the
seismic design criteria for cold formed steel (CFS) steel-sheathed shear walls, limiting their use
as a lateral force resisting system in lightweight steel buildings. In order to overcome this lack
of guidelines, a specific study has been performed to extend the applicability of the Effective
Strip Method (ESM), given in North American standard AISI S400 to EN1998-1. The
method evaluates the shear resistance of steel-sheathed shear walls and it is permitted only in
USA & Mexico as an alternative to tabulated shear resistance for predefined wall configur-
ations. In order to further validate its scope of application, it was applied to additional avail-
able experimental results of walls tested in Canada. Only for some configurations, the ESM
does not give acceptable predictions. Furthermore, the ESM was applied following the Euro-
pean approach, to make its use possible in context of European seismic design methodology.
The overstrength factor for the design of non-dissipative elements was evaluated. Based on
the results, it is concluded that the ESM could be appropriate for the evaluation of resistance
of steel-sheathed shear walls in Eurocodes.

1 INTRODUCTION

The lightweight steel (LWS) constructions, made with Cold-Formed Steel (CFS) profiles, are
well known due to the high performances they exhibit in several circumstances. In the last years,
the lack of specific seismic codes and regulation has encouraged many research activities and
theoretical studies on CFS structures at University of Naples “Federico II”(Landolfo 2019). In
particular, the present study is devoted to the definition of seismic design criteria for CFS steel-
sheathed shear walls. CFS steel-sheathed shear walls consist of a CFS frame (tracks and studs),
with a main role of carrying the vertical loads, braced by steel sheets (Figure 1). Steel sheets
have a thickness from 0.48 to 0.84 mm and they are generally connected to the steel frame by
self-piercing screws. In order to improve the buckling behaviour of chord and interior studs by
reducing their unbraced length, bracing systems can be placed along the height of the wall in
different positions and configurations. The in-plane lateral resistance is offered by frame-to-steel
sheet connections and steel sheets, itself. Although in Europe CFS constructions have been built
mostly in recent decades, current version of European code for seismic design of buildings
EN1998-1 (CEN 2004) does not include specific criteria for seismic design. Currently, the
“North American Standard for Seismic Design of Cold Formed Steel Structural Systems” AISI
S400 (AISI 2015), used in USA, Mexico and Canada, is the only reference code for the seismic
design of this structural typology. According to AISI S400 (AISI 2015), steel-sheathed shear
walls are expected to withstand to seismic loads through the deformation of frame-to-steel sheet
connections and the energy-dissipating elements of the system are the frame-to-steel sheet con-
nections and the steel sheets itself. To evaluate the resistance of certain predefined wall configur-
ations grouped in two categories for USA & Mexico and Canada, respectively, AISI S400
provides the nominal shear resistance values per unit length, based on the results of several
experimental campaigns on these predefined configurations (Serrette 1997, Yu 2007, Ong-Tone
& Rogers 2009, Balh & Rogers 2010, Yu & Chen 2011, DaBreo & Rogers 2012). Contrarily, the
Effective Strip Method (ESM) by Yanagi & Yu (2014), is permitted only in USA & Mexico to

221
Figure 1. Steel-sheathed shear walls.

evaluate the nominal shear resistance of the shear walls as an alternative to the tabulated values
for already predefined wall configurations. The ESM was calibrated on experimental tests car-
ried out in USA (Serrette 1997, Yu 2007, Yu & Chen 2011) and could be used within certain
limitations on geometrical and mechanical characteristics of walls. The main goal of the pre-
sented study is to make possible the implementation of ESM in context of European design
methodology. First, the ESM’s wider scope of application through its acceptable shear resist-
ance predictions for additional available experimental results of walls tested in Canada (Balh &
Rogers 2010, DaBreo & Rogers 2012) is confirmed. Then, the current formulations according
to the AISI S100 (AISI, 2015) to estimate the frame-to-steel sheet connection resistance in the
ESM are compared with equivalent European formulations (EN1993-1-3). Lastly, an over-
strength factor (γov) is evaluated according to the European seismic design methodology, which
is used in the capacity design to ensure the dissipative nature of the main energy dissipating
components of the shear wall through an overstrength in the non-dissipative components. In
particular, the design strength of the non-dissipative components of the walls is increased by an
overstrength factor to take into account the expected strength of the energy dissipative wall
components.

2 TABULATED METHOD

For steel-sheathed walls currently covered by AISI S400 (AISI, 2015), nominal value of shear
resistance per unit length are given for certain predefined configurations. These values are based
on the results of several tests on full scale shear walls. In particular, nominal values given for
USA & Mexico are based on cyclic tests carried out in USA by Serrette (1997), Yu (2007), Yu
& Chen (2011), while nominal values given for Canada are based on monotonic and cyclic tests
carried out in Canada by Rogers and colleagues (Ong-Tone & Rogers 2009, Balh & Rogers
2010, Da Breo & Rogers 2012). The AISI S400 (AISI, 2015) covers 20 and 25 different wall
configurations, for USA & Mexico and Canada, respectively, with a minimum wall width of
610 mm and a maximum aspect ratio (wall height, h, to wall width, w, ratio, AR = h/w) equal
to 4. The configurations permitted for USA & Mexico have walls without bracing systems
(called Type A in this study) and with one bracing system placed in the middle stud height
(called Type B in this study), while for Canada the permitted configurations include Type
A walls and walls with three bracing systems equally spaced along the stud height (called Type
C in this study). Past experimental tests showed that the failure of walls was affected by the col-
lapse of frame-to-steel sheet connections located in the diagonal tension field developed in the
steel sheets (Figure 2). For the evaluation of the resistance of both wood and steel-sheathed
walls with an aspect ratio h/w > 2, the AISI S400 (AISI, 2015) provides a reduction coefficient
equal to 2(w/h), to consider the excessive deformability shown during experimental tests

222
Figure 2. Collapse mechanisms for walls (a): Type A (Yanagi & Yu 2014) (b): Type C (Da Breo 2012).

conducted by Serrette (1997). Indeed, the nominal shear resistance of walls with an aspect ratio
greater than 2 is limited to the achievement of a limit drift (Yu & Chen, 2009). The aspect ratio
should not exceed 4 in any case.

3 EFFECTIVE STRIP METHOD

For USA & Mexico, the shear resistance of the steel-sheathed walls can alternatively be evalu-
ated through the ESM proposed by Yanagi &Yu (2014). This method defines the width of the
effective strip where the tension field action can be schematized (Figure 3). Therefore, the
shear resistance of the wall, Rc, is the minimum of resistances of frame-to-steel sheet connec-
tions within the effective strip and the yielding resistance of the effective strip. The effective
strip width formula was calibrated on the basis of some available experimental results

Figure 3. Effective strip model.

223
(Serrette 1997, Yu 2007, Yu & Chen 2011), which are representative of predefined Type A and
Type B walls for USA & Mexico. However, it must be noted that the experiment results used
to calibrate the ESM always showed that the failure of wall is always governed by the frame-
to-steel sheet connections within the effective strip.
On the basis of the effective strip model, the shear resistance in a steel sheathed shear wall
can be determined using the Equation 1:

 
Rc ¼ minimum 1:33nFn cos α; 1:33weff tfy cos α ð1Þ

where: n is the number of connections within the effective width; Fn is the nominal shear
resistance of a frame-to-steel sheet connection within the effective strip width, weff; α is the
angle at which the tension force is acting; t is the sheathing thickness; fy is the sheathing yield
stress.
Fn can be evaluated according to the North American specification for the design of cold-
formed steel structural members S100 (AISI, 2015). In particular, it is the minimum of the
resistance associated to three types of possible failure mechanisms in frame-to-steel sheet con-
nection: (1) tilting and bearing, (2) end distance, (3) shear failure of screw. However, in all of
the experiments used for ESM calibration, failure of the frame-to-steel sheet connections was
always governed by the tilting of screw and its bearing against the surrounding steel layers.
Eventually, resistance of the connection corresponding to these failure modes is evaluated
according to the Equation 2a, if t1 =t  1:0 or Equation 2b, if t1 =t  2:5 or it can be linearly
interpolated between the values obtained from Equations 2a and 2b, if 1.05t1 =t52:5.
  1
If t1 =t  1:0; Fn ¼ min 2 t31 d 2 f u;1 ; 2:7tdf u ; 2:7t1 df u;1 ð2aÞ
 
If t1 =t  2:5; Fn ¼ min 2:7tdf u ; 2:7t1 df u;1 ð2bÞ

where: fu,1 and fu are the ultimate tensile strength of the framing elements and sheathing,
respectively; t1 is the thickness of stud or track; d is the nominal diameter of screws;
The effective strip width weff can be obtained according to Equation (3):

weff ¼ wmax ; if λ  0:0819 or weff ¼ ρwmax ; if λ > 0:0819 ð3Þ


0:12
where: wmax ¼ w=SinðArctanðh=wÞÞ; ρ ¼ 10:55ðλλ0:08 Þ
; λ ¼ 1:736 β α1 α2
; α1 ¼ f u =310:3;
1 β2 β3 ðh=wÞ
0:12 2

α2 ¼ f u;2 =310:3; β1 ¼ t=0:457; β2 ¼ t2 =0:457; β3 ¼ s=152:4; fu2 is the minimum of the ultimate-
tensile strength of stud or track; t2 is the minimum of the thickness of stud or track; s is the
screw spacing at panel edges.

4 APPLICATION OF EFFECTIVE STRIP METHOD TO WALLS TESTED


IN CANADA

The advantage of ESM over the predefined shear wall configurations resistance (tabulated)
values is that it has certain analytical base and could be used for the variety of wall configur-
ations, though with certain mechanical and geometrical limitations. Therefore, in order to check
its validity on available experimental results, which were not used for its calibrations, the ESM
was applied to Type A and Type C wall configurations tested in Canada. The comparison
between ESM (theoretical) and experimental results was done in terms of wall peak shear resist-
ance. The comparison shows an average theoretical to experimental ratio (Rn_t/Rn_exp) equal to
1.010 and 0.666 for Type A and Type C walls, respectively. Indeed, the ESM predicts with good
accuracy the shear resistance for Type A walls, but it underestimates by 35% the shear resistance
of Type C walls. The underestimation is due to the different collapse mechanism generated by

224
Figure 4. Ratio of the theoretical shear resistance (Rn_t) to the experimental shear resistance (Rn_exp) of
walls.

the presence of three bracing systems, which produces an increasing of the effective strip width.
The results are summarized in Figure 4. The limit of the aspect ratio (h/w ≤ 2) and the reduction
coefficient (2(w/h)) for walls with an aspect ratio h/w > 2 also apply to ESM.

5 USE OF EFFECTIVE STRIP METHOD ACCORDING TO EUROPEAN DESIGN


APPROACH

The applicability of the ESM is strictly linked to the definition of the frame-to-steel sheet con-
nection shear resistance, Fn. The ESM was calibrated to match the experimental results by the
AISI S100 (AISI, 2016) formulations given in Equations 2a and 2b. Equations 2a and 2b pre-
dicted the experimental shear resistance (Fn_exp) with a good accuracy for frame-to-steel sheet
connections used in Type A, B and C tested walls, as shown in Figure 5. Contrarily in the
European standard (EN1993-1-3), a different formula is proposed, in which the nominal shear
resistance of connections Fn is calculated through the Equation 4:

Fn ¼ αfu dt ð4Þ

where α can be evaluated as: pffiffiffiffiffiffiffi


if t1 ¼ t: α ¼ 3:2pt=dffiffiffiffiffiffiffi  2:1
if t1  2:5t and t51mm: α ¼ 3:2 t=d  2:1
if t1  2:5t and t  1mm: α ¼ 2:1
if t5 t1 52:5t α is obtained by linear interpolation
If Equation 4 was applied to the frame-to-steel sheet connections used in Type A, B and C tested
walls, the connection resistance was underestimated of about 50% in average (Figure 5). Since
Equation 4 underestimates the connection shear resistance of about 50%, the same underestima-
tion of the wall shear resistance is obtained if the ESM is applied by using the formula for connec-
tion shear resistance evaluation given by EN1993-1-3 (Equation 4) (Figure 6). Therefore, it is
reasonable to apply ESM by using the formula for connection shear resistance evaluation given by
AISI S100 (Equation 2).

6 OVERSTRENGTH FACTOR

In the capacity design approach, the overstrength factor plays a key role in order to guarantee
the desired collapse mechanism. The overstrength factor can be used to increase the design
resistance of non-dissipative elements to match the expected resistance of the dissipative

225
Figure 5. Ratio of the theoretical shear resistance (Fn_t) to the experimental shear resistance (Fn_exp) of
frame-to-steel sheet connections calculated according to AISI S100 and EN 1993-1-3 formulations.

Figure 6. Ratio of the theoretical shear resistance (Rn_t) to the experimental shear resistance (Rn_exp) of
walls calculated by using connection shear resistance evaluated according to AISI S100 and EN 1993-1-3
formulations.

element in a system. In the present study the Equation 5 was used to compute the overstrength
factor γov for the predefined wall configurations in AISI S400, which are within the geomet-
rical and mechanical limitations of ESM:

Rn exp
γov ¼ Rc;Rd ð5Þ

where: Rn_exp is the peak experimental shear resistance; RC,exp is the design wall shear resistance
obtained according to the ESM. In particular, four groups of walls are defined: Type A and
B walls with an AR ≤ 2; Type C walls with an A ≤ 2; Type A and B walls with an AR = 4; and
a group representative of all experiments i.e. the combination of previous three groups. Over-
strength has been calculated using the ESM at its current form in AISI S400 for the North
American design. Of course, it has been applied by using the formula for connection shear
resistance evaluation given by AISI S100 (Equation 2) and the design strength values take into
account the reduction coefficient 2w/h for walls with an aspect ratio h/w > 2. As a result, two
different values of γov were obtained for the wall with aspect ratio not greater than 2 and equal
to 4. The γov average values for each group are listed in Table 1. The difference in the over-
strength factors between the two aspect ratio walls is due to the reduction coefficient used for

226
Table 1. Average overstrength values.
Wall Type AR γov

North America A and B ≤2 1.31


C ≤2 2.13
A and B 4 2.79
A, B and C ≤2 and 4 1.74

the wall with the aspect ratio between 2 and 4. Moreover, a higher overstrength factor for
type C walls in North America is due to the weakness (underestimation) of the ESM to predict
the resistance for these types of walls due to the different collapse mechanism generated by the
presence of three bracing systems in them, which produces an increasing of the effective strip
width which is not captured by the ESM in the current version.

7 CONCLUSIONS AND FURTHER DEVELOPMENTS

Current European earthquake standard EN 1998-1 (CEN 2004, 2006) do not provide seismic
design criteria for CFS steel-sheathed shear walls, limiting their use as the lateral force resisting
system. A theoretical study has been performed to extend the applicability of the Effective Strip
Method (ESM), given in AISI S400 (AISI 2015), to implement the method in the future version
of Eurocodes. To check a wider scope of the ESM application, which is permitted only in
USA& Mexico, it was first applied to evaluate the shear resistance of A and C wall typologies
allowed in Canada, which were not part of reference experimental database used to calibrate the
ESM during its development. The ESM well fits the experimental results of Type A walls, cap-
turing the peak shear resistance. For the Type C walls, the ESM underestimates peak resistance
of about 35% due to the different collapse mechanism generated by the presence of three bracing
systems in type C walls, which produces an increasing of the effective strip width. To extend its
validity to Type C walls, a different formulation for the effective strip width definition
should be developed in the future. Moreover, if the ESM was applied by using resistances of
frame-to-steel sheet connections evaluated according to EN 1993-1-3 (CEN 2006), the shear
resistance of walls was underestimated of about 50% in average because formulas given by EN
1993-1-3 for the evaluation of resistances of frame-to-steel sheet connections underestimate the
experimental results by the same amount. Once defined the design resistance of steel-sheathed
shear walls with the ESM, the overstrength factor was evaluated for different groups of the
walls. Results showed an average overstrength factor ranging from 1.31 to 2.79, with bigger
values obtained for type C walls (2.13 in average) due to the underestimation of the ESM for
type C walls, and walls with aspect ratio equal to 4 (2.79 in average) due to the application of
the reduction coefficient given in AISI S400 for these walls. Based on the results, it is concluded
that the ESM is appropriate for the evaluation of resistance of steel-sheathed shear walls in
Eurocodes if some specifications are given regarding wall overstrength.

REFERENCES

AISI. 2015. S400-15 North American Standard for Seismic Design of Cold formed Steel Structural Sys-
tems. American Iron and Steel Institute (AISI). Washington. USA.
AISI. 2016. S100-16 North American Specification for Seismic Design of Cold-Formed Steel Structural
Members. American Iron and Steel Institute (AISI). Washington. USA.
Balh N. 2010. Development of seismic design provisions for steel sheathed shear walls. Master thesis.
Department of Civil Engineering & Applied Mechanics. McGill University.
Balh N. & Rogers C.A. 2010. Development of seismic design provisions for steel sheathed shear walls,
Research Report. Department of Civil Engineering & Applied Mechanics. McGill University.

227
CEN. 2004. EN 1998-1 Eurocode 8: Design of Structures for earthquake resistance-Part 1: General rules,
seismic actions and rules for buildings. Brussels: European Committee for Standardization.
CEN. 2006. EN 1993- 1-3Eurocode 3: Design of steel structures-Part 1-3: General rules-Supplementary
rules for cold-formed members and sheeting. Brussels: European Committee for Standardization.
Da Breo J. 2012. Impact of Gravity Loads on the Lateral Performance of Cold-Formed Steel Frame/
Steel Sheathed Shear Walls, PhD Thesis. Department of Civil Engineering & Applied Mechanics.
McGill University.
Da Breo J. & Rogers C.A. 2012. Steel sheathed shear walls subjected to combined lateral and gravity
loads, Research Report. Department of Civil Engineering & Applied Mechanics. McGill University.
Kechidi, S. Bourahla, N. & Castro, J.M. 2017. Seismic design procedure for cold-formed steel sheathed
shear wall frames: Proposal and evaluation. Journal of Constructional Steel Research. 128: 219–232.
Landolfo, R. 2019. Lightweight steel framed systems in seismic areas: Current achievements and future
challenges”, Thin-Walled Structures 140: 114–131.
Ong-Tone C. & Rogers C.A. 2009. Tests and evaluation of cold-formed steel frame/steel sheathed shear
walls, Research Report. Department of Civil Engineering & Applied Mechanics. McGill University.
Montreal. Canada.
Serrette R. L. 1997. Additional shear wall values for light weight steel framing. Rep. No. LGSRG-1-97.
Santa Clara University. Santa Clara. California.
Yanagi N. & Yu, C. 2014. Effective Strip Method for the Design of Cold-Formed Steel Framed Shear
Wall with Steel Sheet Sheathing. ASCE Journal of Structural Engineering, 140 (4).
Yu C. 2007. Steel Sheet Sheathing Options for Cold-Formed Steel Framed Shear Wall Assemblies Pro-
viding Shear Resistance. Report No. UNT-G76234, Department of Engineering Technology. Univer-
sity of North Texas.
Yu C, Chen Y. 2009. Steel sheet sheathing options for cold-formed steel framed shear wall assemblies
providing shear resistance – phase 2. Report no. UNT-G70752. Department of Engineering Technol-
ogy. University of North Texas.
Yu C. 2010. Shear resistance of cold-formed steel framed shear walls with 0.686-mm, 0.762-mm, and
0.838-mm steel sheet sheathing. Engineering Structures, 32 (6): 1522–1529.
Yu C. & Chen Y. 2011. Detailing recommendations for 1.83 m wide cold-formed steel shear walls with
steel sheathing. Journal of Constructional Steel Research, 67: 93–101.

228
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Behaviour of a concrete slab in compression in composite


steel-concrete frame joints

P. Červenka & J. Dolejš


Faculty of Civil Engineering, Czech Technical University in Prague, The Czech Republic

ABSTRACT: The objective of this research is a concrete slab in compression in composite steel-
concrete joint. en 1998-1 prescribes a simplified formula and it describes two mechanisms. The
standard procedure for calculating the effective width of the slab is inaccurate, as it is stated
repeatedly in a number of publications. Mechanism 1 acts in the concrete slab with a concentrated
compressive force on the column flange. in the standard procedure for calculating the bearing cap-
acity of this mechanism, the positive effect of the state of stress in the concrete slab in the joint
region is not taken into account. the goal of this research is to derive more accurate formula for
calculation of effective width of fibre-reinforced concrete slab in compression in the nodal zone.
New aspects have been taken into account: confinement effect of the mechanism 1 and more pre-
cise inclinations of concrete struts. An experiments, a validation of a numerical models and
a parametric study of the composite joints with slab made of common concrete and fibre-
reinforced concrete are main steps of this research.

1 INTRODUCTION

The concrete slab in compression is a component of steel-concrete composite joint. This compo-
nent occurs mainly in the following situations: 1) an asymmetric loading of the joint by hogging
moments; 2) an exceptional loss of the inner column; 3) horizontal loading by seismic action.
Current code formula for a bearing capacity and stiffness calculation is inaccurate for this
component of steel concrete composite joint with a common concrete slab. A failure of this
component leads to the loss of the bearing capacity and the stiffness of the whole joint.
A fibre-reinforced concrete slab is a modern construction solution with many advantages,
but until now no research has dealt with the bearing capacity and the stiffness of fibre-
reinforced concrete slab in compression in a nodal zone of steel-concrete composite frame.

2 STATE OF THE ART

An Appendix C of EN 1998-1 prescribes a formula for calculating of the effective width beff of the
common concrete slab:
beff ¼ 0; 7hc þ bc ð1Þ

where hc and bc are cross-sectional dimensions of the column depicted on Figure 1. This formula is
based on two mechanisms, which are defined in Figure 1. Mechanism 1 represents a direct com-
pression force between the concrete slab and the column flange. Mechanism 2 is determined by
struts and tie model.
This simplified formula is inaccurate. In the standard procedure for calculating of the bearing
capacity of the mechanism 1, the positive effect of the state of stress in the concrete slab in the
joint region is not taken into account. This positive state of stress leads to consideration of higher
compression strength of the concrete slab in this region (Plumier, 2001, Bennacer, 2015, du Plessis,
1973, Kato, 1985). The length of the partial effective width created by mechanism 2 depends on
the inclination of the concrete struts. For the joint subjected to a sagging moment there is an

229
Figure 1. Transfer of forces – F1 is developed by the mechanism 1, F2 is developed by the mechanism 2;
hc and bc are cross-sectional dimensions of the column.

inclination 45° taken into account in the described formula given in EN 1998-1. Some authors
reported a different inclination (Bennacer, 2015, Somja, 2017, Civjan, 2001).

3 EXPERIMENTAL STUDY

Authors prepared an experiment on a full-size construction of inner steel-concrete composite


frame joint (Červenka, 2018). The aims of the experiment is to measure normal compression stress
distribution in concrete slab in the nodal zone and to measure the inclination of minimal principle
stress of mechanism 2. The joint will be subjected to hogging moment in one connection and to
sagging moment in other connection until a collapse of the joint. Basic test configuration is given
in Figure 2: specimen of composite steel-concrete joint (1), test frame girder (2), a member inserted
between test frame girders (3), hydraulic jacks applying compressive force (4) and tensile force (5),
column of main test frame (6), beam of minor test frame (7), position of lateral restrains of steel-
concrete cantilever (8), pin connection (9).
A steel-concrete composite joint design used in the experiment should guarantee: 1) a sagging
moment resistance of the joint will be smaller than a hogging moment resistance; 2) a critical
component of the joint will be represented by concrete slab in compression. For these purposes,
the bearing capacity of other components are designed higher than the bearing capacity of the
examined component concrete slab in compression. The design of steel-concrete composite joint
was done by the component method. A scheme of designed joint is shown in Figure 3.

Figure 2. Test configuration (Červenka, 2018).

230
Figure 3. Scheme of designed joint (Červenka, 2018).

A solid slab was designed with a thickness 100 mm and a concrete class C25/30 reinforced
by steel fibres (the amount of the fibres is 80 kg/m3). The bearing capacity of the concrete slab
in compression was calculated by simplified formula given in Appendix C of EN 1998-1. The
stiffness of this component was determined as the concrete in compression by EN 1993-1-8.
According to this preliminary calculation, the concrete slab should collapse when an elastic
distribution of a forces is still present in the joint. An actual strength of the concrete slab will
be probably higher than a nominal strength, hence components of the steel part of the joint
are here designed to be ductile to enable a plastic distribution of the forces.
In order to establish the concrete slab in compression as the critical component of the joint,
a column web panel is equipped with supplementary web plates on both sides. As a consequence,
a bearing capacity of the component web panel in shear increased. The column web panel in shear
is usually the critical component of the joints subjected to significantly unequal moments.
According to the design by the component method carried out with the nominal
strength of members, hydraulic jacks forces for collapse of the joint should be ±80 kN.
A maximum vertical deflection caused by hydraulic jacks should be 35 mm for the sag-
ging moment and 44 mm for the hogging moment at the collapse of the composite
joint. This calculation is taking into account:
1) a secant stiffness of steel-concrete composite joint; 2) an initial stiffness of the joints of
the test frame; 3) an equivalent cross-section properties of steel-concrete composite beams; 4)
an influence of deformation of the whole test frame.

4 INSTRUMENTATION FOR THE MEASUREMENT

Strain gauges are located on the concrete slab in the sagging moment area and their location
are described in Figure 4, where: 1) red marked strain gauges are located on the upper side of
slab; 2) blue marked strain gauges are located on both upper and bottom sides of the slab.
A horizontal lines of the strain gauges are marked by numbers 1 to 7. The strain gauges in the
line 1, 3 and 5 will record transverse tension forces caused by the longitudinal forces in the
concrete slab, by mechanism 1 and mechanism 2. The strain gauge rosettes located in lines 6
and 7 will measure the minimal principal stress caused by mechanism 2 and its inclination.
Strain gauges located in line 6 will also measure the distribution of the normal compression
stress caused by mechanism 1. On the bottom side of the slab there are located „blue col-
oured” strain gauges (Figure 4), because the bearing capacity of the mechanism 1 on the
bottom side of the concrete slab should be influenced by positive effect of a vertical restrain
ensured by transverse reinforcement (Plumier, 2001).
Together with the strain gauge measurement, the distribution of stress on concrete slab will
be measured by a digital image correlation method (DIC).

231
Figure 4. Location of strain gauges on the concrete slab.

A location of strain gauges on the concrete slab reinforcement is depicted in Figure 5. Strain
gauges on the transverse reinforcement will measure transverse tension forces caused by the mech-
anism 1 and mechanism 2 in the sagging moment area. Based on a preliminary analysis in the
program Scia Engineer, strain gauges on the longitudinal reinforcement were located in the hog-
ging moment area.
A location of inclinometers is depicted in the Figure 6. It was chosen with a respect to recom-
mendations given by publication (Demonceau, 2008). A location of displacement sensors in
a nodal zone of the composite joint is also depicted in the Figure 6. Displacement sensors along
with inclinometers will be used to gain a relationships between the moment and a rotations of con-
nections, whole joint and column web panel (Jaspart, 1992, Bursi, 2004). The relationships
between the moment and the rotation obtained by inclinometers and displacement sensors will be
compared according to recommendation given by publication (Jaspart, 1992). On both connec-
tions there are three displacement sensors to get the connection rotations. For the evaluation of
column web panel rotation there are five displacement sensors connected to an independent struc-
ture. Further two sensors are located on the top and bottom of the column to get a horizontal
movement on its supports (these sensors are also connected to an independent structure). Other
two displacement sensors are placed in the ends of steel-concrete composite cantilevers to get a slip
between the steel beams and the concrete slab. Hydraulic jacks will also record their own vertical
movements.
A location of strain gauges on steel part of joint is depicted in Figure 7. It was chosen with
respect to results of an analysis carried out by the program Idea Statica. The strain gauges are
placed on following components: column flange in bending, endplate in bending, beam web in ten-
sion, beam flange in compression and the strain gauge rosette is situated in the column web panel
in shear.
A monotonic test will be carried out by hydraulic jacks controlled by deformation. An unload-
ing phase will be realized in the elastic stage to get the initial stiffness of the composite joint.
Photos of the preparation of the experiment are in Figures 8 and 9.

Figure 5. Location of strain gauges on the concrete slab reinforcement.

232
Figure 6. Location of displacement sensors (left) and inclinometers (right).

Figure 7. Location of strain gauges on the steel part of joint.

Figure 8. Slab formwork (left), slab reinforcement (right).

Figure 9. Examined real-size steel-concrete composite structure.

233
5 CONCLUSION

The subject of this paper is a preparation of the experimental test on steel-concrete composite
joint. It is expected that the experiment will be performed in March 2019. Results of the experi-
mental test will be used for a validation of a numerical tool and a parametric study will be carried
out with the isolated composite steel-concrete frame joint.

ACKNOWLEDGEMENT

Current research is supported by grant TAČR - TJ01000045 “Advanced procedures of steel


and composite structure connections design and production”.

REFERENCES

Bennacer, M.A., A. Beroual, A. Kriker & Demonceau, J.-F. 2015. Analytical model for composite joints
under sagging moment. Engineering Structures 101: 399–411.
Bursi,O., Caramelli, S., Fabbrocino, G., Molina, J., Salvatore, W., Taucer, F. & Zandonini, R. 2004. 3D
full-scale seismic testing of a steel–concrete composite building at ELSA. Report No. EUR 21299 EN.
Joint Research Center, Institute for the Protection and the Security of the Citizen, European Labora-
tory for Structural Assessment (ELSA). Ispra.
Civjan, S.A., Engelhardt, M.D. & Gross, J.L. 2001. Slab Effects in SMRF Retrofit Connection Test.
Journal of Structural Engineering 127(3).
Červenka, P. Zkouška styčníku spřaženého ocelobetonového rámu. 2018. In: Studnička, J. and J. Fíla, eds.
Sborník semináře doktorandů katedry ocelových a dřevěných konstrukcí 13.2. a 20.9. 2018. Seminář dok-
torandů katedry ocelových a dřevěných konstrukcí, Praha, 2018-09-20. Praha: ČVUT, Fakulta stavební,
Katedra ocelových a dřevěných konstrukcí, pp. 23–26.
Demonceau, J.-F. 2008. Steel and composite frames: sway response under conventional loading and
development of membrane effects in beams further to an exceptional action. Doctor of philosophy
thesis, Civil and Environmental Engineering, University of Liège.
du Plessis, D.P. & Daniels, J.H. 1973. Strength of Composite Beam-to-Column Connections. Fritz
Engineering Lab. Lehigh University, Bethlehem, Pennsylvania.
Jaspart, J.P. & Marquoi, R. 1992. Investigation by testing of the structural response of semi-rigid joints.
Testing of metals for structures; edited by F.M. Mazzolani.
Kato, B. & Tagawa, Y. 1985. Strength of composite beams under seismic loading. Composite and Mixed
Construction, ASCE.
Plumier, A. & Doneux, C. 2001. Seismic behaviour and design of composite steel concrete structures. LNEC
Edition.
Somja, H. & Marie, F. 2017. Concrete Structures Reinforced by Steel Profiles. SMARTCOCO Project.
RFCS GRANT AGREEMENT N° RFSR-CT-2012-00031. Deliverable D8.3 Strut-and-tie models for
shear keys. Contribution to the calibration report. INSA Rennes, Internal Report.
EN 1998-1: Eurocode 8: Design of structures for earthquake resistance - Part 1: General rules, seismic
actions and rules for buildings. European committee for standardization, 2004.
EN 1993-1-8: Eurocode 3: Design of steel structures - Part 1-8: Design of joints. European committee for
standardization, 2005.

234
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

On the development of IoT platforms for the detection of damage


in steel railway bridges

R. Chacón
Universitat Politècnica de Catalunya, Barcelona, Spain

A. Rodriguez
Geocisa, Geotécnia y Cimientos, Madrid, Spain

P. Sierra & X. Martínez


Universitat Politècnica de Catalunya, Barcelona, Spain
Centro Internacional de Métodos Numéricos en la Ingeniería (CIMNE), Spain

S. Oller
Centro Internacional de Métodos Numéricos en la Ingeniería (CIMNE), Spain

ABSTRACT: This paper presents the ongoing research related to the development of IoT
platforms for the detection of damage in steel railway bridges. The development encompasses
several fields from sensors, to damage detection, to data transmission, and to the infrastruc-
ture management system. The paper introduces the concept of regular availability of up-to-
date data sets through an Internet-of-Things (IoT) approach that integrates information in
cloud-based BIM platforms. Initial case studies have been developed in routinely performed
load tests. Furthermore, laboratory experiences aimed at calibrating numerical methods under
a controlled environment are being developed. Subsequently, complete systems encompassing
different parts will be implemented in the form of real-case studies. The steps that have been
performed so far include i) measurement of key characterizing data from these structures
using different platforms ii) data transmission using different ranges and platforms iii)
damage detection using advanced FE-models and iv) understanding the data management
platforms and the corresponding interoperability. The ongoing research project includes
experiences in a joint effort between academia and industrial partners. The assessed structures
include (but are not limited to) either new or existing steel railway bridges.

1 INTRODUCTION

Due to numerous different reasons, steel railway bridges are subject to progressive deterior-
ation in performance throughout their service life. Varying degrees of damage may occur and
propagate in time due to degradation of stiffness in structural members, in connections, in
supports or at material levels. The potential reasons for the initial development of damage are
countless. Overloading, catastrophic accidental loading, material degradation (scouring,
fatigue and/or corrosion) or collisions may trigger a progressive deterioration of the overall
structure or of particular structural elements.
One great challenge that is posed in bridge engineering is related to infrastructure manage-
ment and the analysis of long-term performance of assets (Frangopol 2011). Bridge manage-
ment performed by the incumbent agencies requires encompassing a wide variety of information
from different data sources (Sánchez-Silva et al. 2016). This information is typically related to
the initial design stages of the project, to fabrication and erection, to the as-built information of
the structure as well as to the continuous inspections and monitoring of these assets. Data

235
sources vary considerably from one asset to another. In bridges built decades ago, information
is fragmented and is seldom available digitally. In recently built projects, information is increas-
ingly centralized and updated in various interoperable BIM (Building Information Modelling)
forms (Jeong et al. 2017). For the former case, bridge management is performed in most cases
by manually transferring data to electronic formats and treating the aggregate information
together. For the latter, modern data-hubs provide inclusive environment for data-transfer; for
this purpose, a certain level of data-standardization is needed.
With the advent of BIM methodologies in the construction sector, information management
is tending towards standard procedures. Several BIM standard models are available nowadays
such as the Industry Foundation Classes (buildindSMART 2018), which represents
a platform-neutral set of rules for interoperability or the Linked Building Data, another open
initiative that adds the benefits of the Semantic Web and the Web of Data frameworks (w3c
2019). Wide varieties of construction-related applications in which BIM and other platforms
are linked have been presented in recent years. For instance, academic and industrial initia-
tives for the integration of BIM and geographic information systems in supply chain manage-
ment (Irizarry et al, 2013) or urban facility management (Mignard & Nicole 2014).
Furthermore, the integration of information obtained with laser scanning techniques into
BIM platforms has also been a matter of development (Bosché et al. 2015, Brilakis et al.,
2010). The integration of sensors, with the aim of enriching the life cycle, operational and
maintenance capabilities of BIM platforms is one area in which data-exchange and system
integration have been discussed. Data flows from sensors to BIM platforms (Kensek 2014)
and machine learning tools integrated in BIM platforms using multisensory arrays of informa-
tion (Park et al. 2017) suggest a potential use of sensors in the years to come. The pervasive
emergence of sensing technologies for structural health monitoring (SHM) and the ubiquitous
digitalization in engineering (“Industry 4.0”) pose increasing demands on information model-
ing concepts in civil engineering as well. From other fields like machine health monitoring
(Zhao et al. 2019), it is foreseen that applications of massive use of ubiquitous sensors coupled
with machine learning techniques may also permeate the construction sector.
SHM has provided crucial information about the performance of bridges in last decades. It
has been demonstrated in a vast array of examples in which sensors are deployed in short-,
medium- or long-term. It is recognized that SHM can be a valuable tool in understanding the
operational condition and health of assets and therefore, in the deployment of damage detec-
tion techniques. Damage detection in civil infrastructure has been active field of research in
several engineering disciplines. A wide variety of damage detection techniques can be found
but one may classify these techniques in two categories (Gonzalez & Karoumi 2015): Model-
based and Model-free. For the former, accurate FE models are usually required. A physical
interpretation of damage may be performed by using signals and the corresponding long-term
structural behavior. For the latter, the need of FE simulations vanishes but when detecting
damage, the physical interpretation of results vanishes as well. For either case, algorithms
must be trained based on the gathered data. One technique that has been explored in last dec-
ades is based upon the vibrational conditions of the structure. The use of the natural fre-
quency parameter as a sensitive indicator of structural integrity has been a matter of scientific
debate (Salawu 1997, Fan & Qiao 2011, Doebling et al. 1998, Alves et al. 2015). Measurement
of acceleration in real structures, and in particular, railway steel bridges may be nowadays
performed reliably with different techniques.
Notwithstanding, fully integrated bridge management systems encompassing all potential
monitoring aspects are available for a limited (though numerous) amount of case studies.
A systematic use of such tools is not typical for the vast majority of bridges around the world.
To cite a few, some examples of bridge management systems encompassing design, construction,
maintenance, exploitation and life-cycle analysis have been developed in the United States,
Taiwan, China and Sweden. Computational platforms integrating several aspects in manage-
ment of highway bridges implemented in real case studies have been presented (Okasha & Fran-
gopol 2011). This tool is aimed at providing a decision support for a given bridge by performing
an integrated process of interactive computations. Jeong et al. 2017 developed a bridge informa-
tion modelling system for supporting bridge monitoring and management applications; in this

236
case, the OpenBrim data model (a standard framework that encompasses sensor-data and struc-
tural analysis) was systematically used. European projects aimed at infusing IoT technologies
into bridge management have been also developed (IoTBridge, 2016). In China, following simi-
lar data-integration techniques, examples of automation of safety management during bridge
construction have been presented (Jia-Rui et al. 2019). Another example of tailor-made bridge
monitoring system for the case of prestressed composite box-girders with corrugated steel webs
was presented in Taiwan (Sung et al 2016). A profuse set of sensors and statistical treatment of
results shows the potential and needs of data-treatment in real time.
In this paper, attempts for encompassing data-gathering, data-transmission, real-time data-
analysis and the corresponding integration into existing platforms are discussed. Research and
development is performed in a joint effort between academia and industrial partners. A set of
numerical methods for damage detection in steel bridges are systematically tested with labora-
tory experiences. Several data-gathering and data-transmission techniques are explored as well
as different standard protocols as potential management frameworks. Preliminary results are
analyzed for the sake of designing full deployments in real case studies.

2 DESCRIPTION OF THE SYSTEM

The system encompasses several steps such as i) data-gathering via sensors and microcontrol-
lers, ii) data-transmission via different protocols, iii) damage detection by means of data-
processing and iv) interoperable management in the form of data-hubs. The first part of the
project is focused on develop points i) and iii). Data-transmission (ii) as well as data-
management (iv) belong to the second stage of the project. Figure 1 shows schematics of all
the involved parts. These parts are meant to be interconnected in IoT frameworks in which
data may flow seamlessly. The damage-detection strategy that is chosen at this stage of the
project is a model-, vibration-based identification of damage. In this paper, only information
related to i) and iii) is dealt with.
At this first part, the methodology used to validate the different developments that will be
conducted encompasses two different approaches:
• A set of experiments performed at the Laboratory of Scale Reduced Models at the Depart-
ment of Civil and Environmental Engineering, Barcelona, Spain. Small-scale tests have
been useful for prototyping.

Figure 1. Overall schematics of the system.

237
A set of load tests in full-scale bridges located at various places in Spain. Full-scale tests
have been useful for setting and further consideration of realistic site conditions.
In the following, a brief description of these studies is presented for each case:

3 DATA-GATHERING

For both laboratory and in situ tests, different sensors (accelerometers) and data-acquisition
systems have been employed. Equipment differ primarily in cost and versatility. Firstly, open-
source electronic platforms have been used primarily for prototyping solutions using both
analog (ADXL335) and digital (ADXL345) accelerometers. These prototyping platforms
allow gathering raw data that may be transmitted in several formats to other devices. Further-
more, highly optimized electronic devices developed at Tokyo Denki University (sensors pro-
vided with processing capacity) have been also tested. Both autonomy and robustness of these
devices are considerably high as well as accuracy and precision for their reduced size. Finally,
high-precision professional multi-purpose HBM data-acquisition equipment with wired accel-
erometers have also been used. Accuracy and precision in these cases is obtained at
a relatively high cost, less autonomy and heavier weight. Figure 2 displays three kind of tech-
nologies employed in this study.
The observations performed so far allow pointing out the following remarks:
• Open-source electronics boards such Arduino ® DUE and Arduino ® MKR1000 have
been useful for prototyping several different solutions in which acceleration needs to be
recorded. The open-source integrated development environments (IDE) found in such
boards set an ideal platform for testing different codification and further transmission
of signals. Both analog and digital devices can be plugged to these boards. For analog
devices, 12-bit Analog-to-digital converters with 50 Hz sampling rates have been suc-
cessfully tested with no data loss when acquiring signals. For digital devices, both SPI
and I2C protocols available in the boards have been successfully tested. Slightly lower
sampling rates have been observed in such cases (without data-loss). One of the main
benefits of these platforms is the possibility of developing tailor-made Java and/or C++
codes that can be uploaded in the microcontroller by the user. These prototyping
boards may be functional in their generic form but ideally, prototyped solutions need to
be optimized further when needed. Sensors and microcontrollers present a low energy
consumption. Boards are compatible with a vast array of data-transmission systems
that are also under study in part 2. One of the disadvantages of these boards is that all
features need to be set by the user (including low-pass filters or any other post-
processing technique).

Figure 2. Three types of accelerometers.

238
• Fully integrated and optimized high precision, 3-axis accelerometers provide a versatile
solution for measuring vibrations. These devices provide interesting levels of autonomy,
resolution (0.00005 m/s2) and selectable sampling rates ranging from 200 Hz to 10 Hz with
no data-loss. These devices are set on the desired surface with batteries or other power
supply. Data is stored in a local memory and is accessible after the measurement finishes.
• Professional equipment has also been used and analyzed. These devices provide high-
precision sensors, high resolution data acquisition systems and user-friendly platforms as
graphical user interfaces. Occasionally, these technologies are considerably more expensive
than the aforementioned alternatives. In addition, computers need to be plugged to these
devices during measurements, which subtracts versatility to the system. Piezoelectric uni-
axial accelerometers were connected to a HBM® MGC-plus data acquisition system.

4 DAMAGE DETECTION

In the following is presented the damage-detection method that has been developed for the
project. The formulated algorithm is primarily based on the analysis of the dynamic perform-
ance of the structure. Damage detection, localization and quantification are sought in the
form of an optimization problem using the natural frequencies and the dynamic response of
the structure as objective functions. Several authors combine modal parameters with opti-
mization algorithms (Ren & De Roeck, 2002) and (Hao & Xia, 2002).
The numerical method is defined as model-based. The method consists of comparing the
dynamic response of several structures with a given damage pattern, with the dynamic
response of the reference damaged structure. An optimization algorithm proposes the damage
patterns in order to find which one minimizes the difference in the response obtained with the
reference and the calculated structure. The general scheme of the algorithm can be summar-
ized in the flowchart shown in Figure 3.
There is a vast array of optimization algorithms that can be used in for the proposed pro-
cedure. As the structural problems that have to be solved contain a large number of variables,
the best choice may not be clearly defined. Heuristic or soft optimization algorithms are con-
sidered useful for solving these problems (Wei & Pizhong, 2011). Genetic algorithms, as
RMOP (Espinoza & Coma, 2012) and derivative-free algorithm, as MODIR (Campana et al,
2016) are also being used and tested. The damage scenarios proposed by the optimization
algorithm are based on a set of parameters to be specified in the numerical FE model. These

Figure 3. Damage-detection algorithm flowchart.

239
parameters must be defined previously and must represent the different damage cases that the
real structure may suffer. The uniqueness of the solution, and therefore the success of the opti-
mization procedure, must be guaranteed by means of soft objective functions.
A structural finite element software, known as Plastic Crack Dynamic (PLCd, 2019), is used
to obtain the dynamic and modal response of the structure. The objective functions relate the
results obtained in the structural analysis with those measured in the real structure. One or
more functions may be minimized in the optimization process until the defined verification
criteria is satisfied. If it is not satisfactory, the algorithm will propose another set of param-
eters, based on its own selection criteria. An example of an objective function (Eq. 1) may be
the relative difference between the first N natural frequencies of the model analysis (fm;i ) and
the first N natural frequencies measured in the reference damaged structure (fref ;i ). In the
objective functions different modal parameters such as frequency, modal shapes, modal curva-
tures, maximum acceleration, deflections may be combined.

N f

P 
FOBJ ¼  m;i  fref ;i  ð1Þ
 f 
i¼1 ref ;i

5 PRELIMINARY RESULTS

5.1 Laboratory tests


Laboratory tests were conceived and implemented with a twofold objective: i) testing different
measuring platforms and ii) generating scale-reduced specimens under undamaged and dam-
aged structural conditions for further studies. The evaluation of the prototyping platforms as
well as of other used devices have provided an overall understanding of the capabilities of the
studied technologies. Sampling rates, ranges, implementation of low-pass filters, resolution,
precision and limitations have been systematically identified. The requirements that are set in
these experimental vibration-based studies were intended to match those routinely encoun-
tered in steel railway bridges. Furthermore, the deployment of an experimental program on
steel undamaged beams that were subsequently damaged provided a controlled environment
in which the damage detection algorithms were systematically tested. Figure 4 displays a test
of a beam loaded dynamically by using a Brüel and Kjaer modal exciter. Two accelerometers
were located at both ends and signals were systematically measured for a wide range of fre-
quencies (frequency sweep). As an example, Figure 5 shows measurements taken with one of

Figure 4. Laboratory tests for the analysis of equipment.

240
Figure 5. Comparisons laboratory tests for the analysis of equipment.

the analyzed instruments: a digital accelerometer. With this device, the measuring scale can be
set by the user (from ±2G to ±16G). Advantages and disadvantages were pinpointed when
studying its potential application in the measurement of vibrations in steel bridges

5.2 Damage prediction


The optimization procedure described in section 4 has been tested for a cantilever beam.
Figure 6 shows the beam simulated using PLCD in which sectors (1 to 10 in the plot) can be
modified by introducing a pre-defined damage. The response of the reference beam is obtained
numerically, and corresponds to having a damage that reduces the material stiffness to
a 0.05% in sectors 2 and 3.
The design variables assigned to the damage algorithm were the section that could be dam-
aged, and the level of damage of that section. The level of damage is considered by reducing
the material stiffness between a 0% (no damage) and a 99.95% (complete damage). After sev-
eral iterations, the optimization code provided a damage configuration that matches quite
accurately the reference beam considered. The optimization algorithm that provided a better
result was MODIR using the sum of the first 3 modal frequencies as objective function.
Figure 6 shows the damaged beam found, red for damaged, grey for undamaged zones.

5.3 Load tests


Real load tests have been another source of experimentation. Realistic site conditions provide
the necessary perspective when prototyping ad-hoc solutions. During the development of the
joint effort between industrial and academic partners, several dynamic load tests have been
used as prototyping platforms as well. Figure 7 displays a one-railway steel bridge assembled
with a truss and a concrete slab in northern Spain. Different measuring technologies have
been implemented in this example. In this case, several features such as autonomy, robustness,
ease-of-implementation as well as transmission requirements have been scrutinized. In add-
ition, preliminary results in which one may see that versatile platforms provide comparable

Figure 6. Model-based damage detection of scale-reduced specimens using PLCD (left). Damaged
structure found by MODIR optimization algorithm (right).

241
Figure 7. Real load tests for the analysis of equipment.

results to other optimized solutions are presented. Recognizably, the amount of data that may
be gathered and sent is considerably higher when versatile, low-cost yet accurate solutions are
designed. The ongoing development of research is related to the appraisal of the influence of
such data quantity in the performance of damage-detection algorithms.

6 CONCLUSIONS

Preliminary results related to the ongoing research presented herein show how IoT solutions
for the acquisition of acceleration data may be performed using versatile solutions. The subse-
quent usage of vibration-based damage detection is being tested in the form of controlled
laboratory tests. The next stage of the project is related to the application of real case studies
in which data-acquisition, data-transmission, data-processing and data management are
encompassed altogether.

ACKNOWLEDGEMENTS

Financial support provided by the Spanish “Centro de Desarrollo Tecnológico Industrial” CDTI.
Iberoeka project “E-testing” as well as by Bec.Ar. Secretaria de Políticas Universitarias. Minis-
terio de Educación, Cultura, Ciencia y Tecnología. República Argentina for the third author.

REFERENCES

Alves, V., Meixedo, A., Ribeiro, D., Calçada, R. & Cury, A. 2015. Evaluation of the performance of
different damage indicators in railway bridges. Procedia Engineering. 114: 746–753.
Bosché, F., Ahmed, M., Turkan, Y., Haas, C. & Haas, R. 2015. The value of integrating Scan-to-BIM
and Scan-vs-BIM techniques for construction monitoring using laser scanning and BIM: The case of
cylindrical MEP components. Automation in Construction. 49: 201–213.
Brilakis, I., Lourakis, M., Sacks, R., Savarese, S., Christodoulou, C, Teizer, J. & Makhmalbaf, A. 2010.
Toward automated generation of parametric BIMs based on hybrid video and laser scanning data.
Advanced Engineering Informatics. 24 (4):456–465. buildingSMART. 2019. http://www.buildingsmart-
tech.org/.
Campana, E. F., Diez, M. & Liuzzi, G. 2018. A multi-objective direct algorithm for ship hull
optimization. Computational Optimization and Applications. 71: 53:72.
Chen, S. & Shirolé, A. 2006. Integration of information and automation technologies in bridge engineering
and management: extending the state of the art. Transportation Research Record. 1976(1): 3:12.
Doebling, S., Farrar, C. & Prime, M. 1998. A summary review of vibration-based damage identification
methods. The Shock and Vibration Digest. 30: 91–105.
Espinoza, H. & Coma, M. 2012. RMOP-GUI Manual.
Fan, W. & Qiao, P. 2011. Vibration-based damage identification methods. A review and comparative
study. International Journal of Structural Health Monitoring. 10(1): 83–111.

242
Frangopol, D. 2011. A life-cycle performance, management and optimization of structural systems under
uncertainty: accomplishments and challenges. Structure and Infrastructure Engineering 7(6): 389–413.
Gonzalez, I. & Karoumi, R. 2015. BWIM aided damage detection in bridges using machine learning.
Journal of Civil Structural Health Monitoring. 5: 715–725.
Hao, H. & Xia, Y. 2002. Vibration-based damage detection of structures by genetic algorithm. Journal of
Computing in Civil Engineering. 16: 222–229.
IoTBridge. 2016. http://www.iotbridge.eu/.
Irizarry, J., Karan, E. & Jalaei, F. 2013. Integrating BIM and GIS to improve the visual monitoring of
construction supply chain management. Automation in Constuction. 31:241–254.
Jeong, S., Hou, R., Lynch, J., Sohn, H. & Law, K. 2017. An information modeling framework for bridge
monitoring. Advances in Engineering Software 114: 11–31.
Kensek, K. 2014. Integration of Environmental Sensors with BIM: case studies using Arduino, Dynamo,
and the Revit API. Informes de la Construcción. 66: 536.
Mignard, C. & Nicolle, C. 2014. Merging BIM and GIS using ontologies application to urban facility
management in ACTIVe3D. Computers in Industry. 65(9): 1276–1290.
Okasha, N. & Frangopol, D. 2011. Computational platform for the integrated life-cycle management of
highway bridges. Engineering Structures 33: 2145–2153.
Park, J., Chen, J. & Cho, Y. 2017. Self-corrective knowledge-based hybrid tracking system using BIM
and multimodal sensors. Advanced Engineering Informatics. 32: 126–138.
Plastic Crack Dynamic. 2019. PLCD. http://www.cimne.com/PLCd.
Ren, W.X. & De Roeck, G. 2002. Structural damage identification using modal data. I: Simulation
verification. Journal of Structural Engineering-ASCE, 128, 87–95.
Salawu, O. 1997. Detection of structural damage though changes in frequency: a review. Engineering
Structures. 19(9): 718–723.
Sánchez-Silva, M, Frangopol, D., Padgett, J., Soliman, M. 2016. Maintenance and Operation of Infra-
structure Systems: Review. Journal of Structural Engineering. 142(9)
Sung, Y., Lin, T., Chiu, Y., Chang, K., Chen, K. & Chang, C. 2016. A bridge safety monitoring system
for prestressed composite box-girder bridges with corrugated steel webs based on in-situ loading
experiments and a long-term monitoring database. Engineering Structures 126: 571–585.
Web of Data. 2019. https://www.w3.org/2013/data/.
Wei, F. & Pizhong, Q. 2011. Vibration-based damage identification methods: a review and comparative
study. Structural Health Monitoring. 10: 83–111.
Zhao, R., Yan, R., Chen, Z., Mao, K., Wang, P., Gao, R. 2019. Deep learning and its applications to
machine health monitoring. Mechanical Systems and Signal Processing. 115: 213–237.

243
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Experimental and numerical studies on shear behaviour of stainless


steel plate girders with inclined stiffeners

X.W. Chen, H.X. Yuan & X.X. Du


School of Civil Engineering, Wuhan University, Wuhan, PR China

E. Real
Department of Civil and Environmental Engineering, Universitat Politècnica de Catalunya, Barcelona, Spain

ABSTRACT: Experimental and numerical studies on shear buckling behaviour of stainless


steel plate girders with inclined stiffeners have been carried out in this study. Material proper-
ties of the two adopted stainless steel alloys, including austenitic grade EN 1.4301 and duplex
grade EN 1.4462 were obtained, and initial geometric imperfections of the cross-sections were
determined. Two stainless steel plate girders with inclined stiffeners were tested under simply
supported conditions. The critical buckling strengths and ultimate shear resistances were all
obtained and further compared to those of previously tested plate girders with transverse stiff-
eners only and both transverse and diagonal stiffeners. Elaborated finite element (FE) models
developed by means of the FE software package ABAQUS were validated against the
obtained experimental results. The validated FE models was subsequently used to evaluate the
influence of the stiffener inclination on the load-carrying capacity of plate girders.

1 INTRODUCTION

The introduction of web stiffeners into slender plate girders, such as transverse and longitu-
dinal stiffeners, has often been regarded as economical in material and fabrication. The stiff-
ened web panels exhibit much higher critical buckling and postbuckling strength than
unstiffened ones, resulting in significant increase in resistance to shear and bending. More-
over, considerable research has been performed on evaluating the effects of inclined or diag-
onal stiffeners. Lotz (1936) performed two tests on plate girders employing inclined stiffeners
and indicated that a considerable increase in ultimate resistance could be achieved, when the
inclination of the stiffeners coincided with the direction of the compressive stresses in the web
panel. Jensen & Antoni (1941) conducted 13 steel plate girder testes and concluded that plate
girders with both transverse and diagonal stiffeners, compared to those with transverse or
inclined stiffeners only, enhanced the exploitation of the load-carrying capacity. Yonezawa
et al. (1978) presented comprehensive studies on diagonally stiffened plate girders and devel-
oped calculation methods for the shear buckling coefficient of diagonally stiffened web panels
and the ultimate shear resistance. Guarnieri (1985) carried out an extensive experimental
study on 98 cantilever plate girders with inclined stiffeners, considering four different angles
of inclination of the stiffeners and three material grades. More recently, Azmi et al. (2017)
reported experimental tests on steel plate girders with perforated webs and inclined stiffeners.
Stainless steel has been increasingly used in structural applications owing to the prominent
corrosion resistance and aesthetic appeal. Meanwhile, particular attention should be paid to
developing separate design formulae for stainless steel structures in view of the nonlinear
material behaviour that differ significantly from carbon steel structures. Shear buckling
behaviour of stainless steel plate girders employing only transverse stiffeners (Olsson 2001,
Real et al. 2007, Estrada et al. 2007, Saliba et al. 2013, Chen et al. 2018) or combining trans-
verse with longitudinal (Estrada et al. 2008) or diagonal stiffeners (Yuan et al. 2019) has been
extensively studied, and efficient design approaches were also presented. However, limited

244
attention has been given to stainless steel plate girders with inclined stiffeners, and few related
experimental studies has been publicly reported yet. Hence, the focus of this study is the inves-
tigation of shear buckling behaviour of stainless steel plate girders with inclined stiffeners.
Experimental tests on two stainless steel plate girders with inclined stiffeners were carried
out under simply supported conditions. The obtained test results were carefully analysed and
further compared to those of previously tested stainless steel plate girders with transverse stiff-
eners only or with diagonal stiffeners between transverse stiffeners. Elaborated FE models,
developed by mean of ABAQUS, were also validated against the test results, which were fur-
ther employed to conduct a parametric study to examine the influence of the stiffener inclin-
ation on the ultimate shear resistance of plate girders.

2 EXPERIMENTAL PROGRAMME

2.1 Material properties


Tensile coupon tests were conducted to determine the material properties of two adopted
stainless steel grades – austenitic grade EN 1.4301 and duplex grade EN 1.4462. The test
results were reported in detail by Chen et al. (2018) and are also listed in Table 1, where the
presented strain hardening exponents n and m for the compound Ramberg-Osgood model
(EN 1993-1-4+A1 2015) were obtained by fitting the test data points in this study.

2.2 Test specimens


Two plate girders with inclined stiffeners were fabricated by welding hot-rolled stainless steel
plates of the two material grades. The average measured geometric dimensions are tabulated
in Table 2, in which w0 is the initial local imperfection amplitude, and other symbols are
defined with reference to Figure 1. The inclined stiffeners were carefully designed according to
the provisions in EN 1993-1-4+A1 (2015), so as to achieve rigid supports for the web panels.
The test specimens were labelled according to the related material grade and stiffener config-
uration, in which the symbol “IS” referred to plate girders with inclined stiffeners.
Prior to testing, the initial local geometric imperfections of the cross-sections were carefully
determined. The experimental setup involved a laser displacement sensor (LDS) recording
imperfections at six representative cross-sections parallel to the inclined stiffeners for each

Table 1. Average measured material properties from tensile coupon tests.

t E0 σ0.01 σ0.2 σ1.0 σu εu εf Exponents


Grade (mm) (MPa) (MPa) (MPa) (MPa) (MPa) (%) (%) n m

4.11 189,000 192.3 279.7 316.8 756.3 - 69.5 11.6 1.8


1.4301 7.85 180,700 191.4 291.7 338.9 706.0 - 62.9 9.4 2.4
11.85 182,800 184.7 280.4 319.1 719.6 - 57.7 9.0 2.0
3.90 204,800 345.3 539.6 604.1 761.4 26.1 40.2 8.9 3.6
1.4462 7.72 188,700 296.5 551.4 614.5 738.4 19.3 33.0 6.7 3.8
12.59 184,000 227.8 464.6 552.8 705.3 23.3 37.4 4.8 3.9

Table 2. Measured geometric dimensions for the test specimens.


Material
Plate girder
grade L a Lx e hw b bs tw tf ts tis w0

V-304-IS 1.4301 1696.3 747.9 374.0 100.3 498.8 150.0 60.0 4.11 11.85 4.11 7.85 3.57
V-2205-IS 1.4462 1698.0 748.4 374.2 100.6 499.3 150.0 60.0 3.90 12.59 3.90 7.72 2.11

All dimensions are in mm.

245
Figure 1. Geometry of the test specimens.

specimen, as shown in Figure 2. The maximum recorded value among these six cross-sections
was taken as the local imperfection amplitude w0 for the specimen. Figure 3 presents the meas-
ured imperfection distributions for specimen V-304-IS, whilst the recorded values of the
imperfection amplitude w0 for the two specimens are tabulated in Table 2.

2.3 Plate girder tests


The employed test setup for plate girder tests is shown in Figure 4. Each plate girder subjected to
a concentrated load at mid-span was simply supported by two end bearings – served as roller
support and pin support respectively. The clamping frame was adopted at each end, where the
lateral displacement and rotation about the longitudinal axis were restrained. Moreover, two lat-
eral supports were also introduced to prevent the specimen from lateral torsional buckling.
The instrumentation diagram for each test is shown in Figure 5, including a total of twenty
triaxial strain gauges, fourteen uniaxial strain gauges and seven laser displacement sensors.
Three triaxial strain gauges were attached along the centreline of each side of a web panel and

Figure 2. Location of six measured cross-sections.

Figure 3. Measured imperfection distributions for specimen V-304-IS.

246
Figure 4. Test setup.

Figure 5. Instrumentation diagram for plate girder tests.

another two were located on the upper corner. Two uniaxial strain gauges were placed at the
centre of intermediate triangular web panel below the loading point and another twelve uni-
axial strain gauges were affixed at both the upper flange and transverse stiffeners. Three LDSs
were installed to record the vertical displacements at mid-span and two end supports, whilst
four symmetrically located LDSs were used to monitor the out-of-plane lateral deformations
of each web panel.
The plate girder tests were carried out using a 5000 kN hydraulic testing machine. Each spe-
cimen was initially loaded at a rate of 0.2 kN/s until the buckle in web panels became notice-
able, and then the specimen was tested to failure under displacement control at a constant rate
of approximately 0.01 mm/s. The structural response of each plate girder with inclined stiff-
eners was recorded continuously, which were later compared to the results of previously tested
plate girders with transverse stiffeners only (labelled as “TS”) and diagonally stiffened plate
girders (labelled as “DS”) (Chen et al. 2018, Yuan et al. 2019).

3 DISCUSSION OF THE EXPERIMENTAL RESULTS

3.1 Critical buckling strength


The critical shear buckling strengths of web panels were determined experimentally by means
of the strain reversal method (Vann & Sehested 1973). Four principal surface strains, involv-
ing principal tensile strain (PT) and principal compressive strain (PC), were calculated from
the two paired triaxial strain gauges of each web panel. The reversal point at which the princi-
pal compressive strain on the convex side of the buckle begins to decrease was taken as the
critical shear buckling strength. Figure 6 presents the shear resistance versus principal surface
strain curves for specimen V-2205-IS, and the obtained critical shear buckling strengths of all
tested specimens including the previously tested plate girders are tabulated in Table 3. It can
be seen that by placing the intermediate stiffeners inclined to the flange, skew web panels were
formed and hence a considerable increase in critical shear buckling strength was achieved -

247
Figure 6. Shear resistance versus principal surface strain curves for specimen V-2205-IS.

Table 3. Summary of the plate girder test results.


Critical shear buckling strength Ultimate shear resistance
Vu,FE (kN) Vu,FE/Vu,TEST
Plate girder Vcr,Test (kN) Vu,Test (kN)

V-304-IS 171 395.2 388.3 0.98


V-304-TS 131 243.2 240.8 0.99
V-304-DS 278 328.7 328.8 1.00
V-2205-IS 163 648.7 641.1 0.99
V-2205-TS 127 385.9 387.0 1.00
V-2205-DS 230 558.4 560.6 1.00

31% for V-304-IS and 28% for V-2205-IS compared to the plate girders with transverse stiff-
eners only. By using the Rayleigh-Ritz method, theoretical analysis conducted by Xiang et al.
(1995) revealed that a skew plate exhibited higher critical shear buckling strength than
a similar rectangular plate, which was due to that applied shear force tended to reduce the
skewness of the plate. Moreover, highest critical shear buckling strengths were obtained from
diagonally stiffened web panels for both material grades, which can be attributed to the fact
that the presence of diagonal stiffeners would subdivide the web panel into triangular subpa-
nels and provide additional out-of-plane restraint for the web panel, resulting in substantial
increase in the critical buckling strength.

3.2 Ultimate shear resistance


The recorded shear resistance versus mid-span vertical displacement curves and failure modes
from all tested specimens are shown in Figure 7, and the obtained ultimate shear resistances
are listed in Table 3. It is evident that the ultimate shear resistances of plate girders with vari-
ous stiffener configurations are considerably higher than the corresponding critical buckling
strengths, exhibiting significant postbuckling strength.

248
Figure 7. Shear resistance versus mid-span vertical displacement curves for all tested specimens,
(a) Austenitic grade EN 1.4301, (b) Duplex grade EN 1.4462.

Well-defined tension field developed in web panels were observed for plate girders
with transverse stiffeners only and plate girders with inclined stiffeners, reflected by long
flat plateaus corresponded to the full development of plastic hinges formed in the flanges
and transverse stiffeners. Moreover, it can be clearly observed that due to the presence
of inclined stiffeners, the tension field was developed on the short diagonal direction of
the skew web panel and a wider tension field was achieved, resulting in considerable
strength improvement reached up to 63% and 68% for V-304-IS and V-2205-IS com-
pared to the plate girders with transverse stiffeners only, respectively. Meanwhile, it is
expected that the stiffener inclination can be significant for the buckling behaviour of
web panels with inclined stiffeners, and hence, numerical study will be further conducted
to examine the influence of the stiffener inclination on the ultimate resistance of plate
girders, considering different web aspect ratios.
For diagonally stiffened plate girders, the development of the full tension field was
delayed until the diagonal stiffeners failed in buckling and no evidence of plastic hinges
was obtained, accompanied by an increase in flexural stiffness. The ultimate resistance
also increased considerably to the extent of 35% for V-304-DS and 45% for V-2205-DS,
due to the introduction of the diagonal stiffeners. Although it can be noticed from the
test results that the highest load-carrying capacity was achieved by the introduction of
inclined stiffeners in this study, a previous study (Yuan et al. 2019) has concluded that
increasing the stiffness of the diagonal stiffeners would significantly raise the ultimate
resistances since the diagonal stiffeners directly undertake a certain amount of shear
force in addition to providing restraint for the web panel. Therefore, a comparison
between the numerical results of these two plate girders will also be conducted to evalu-
ate their strength potentiation.

4 NUMERICAL MODELLING

4.1 Validation of numerical models


The numerical models of plate girders with inclined stiffeners were developed by means of the
FE software package ABAQUS. The four-node shell element S4R, widely used in modelling
slender stainless steel cross-sections, was employed herein. The measured geometric and
material properties were carefully incorporated into the FE models for the validation of
numerical models, while in subsequent parametric analysis, the two-stage Ramberg-Osgood
expression recommended in EN 1993-1-4+A1 (2015) was employed to represent the stress-
strain relationship of stainless steel alloys using the material parameters listed in Table 1. The
initial local geometric imperfections with the shape of the lowest relevant elastic buckling

249
mode and measured amplitudes of imperfections were introduced into the FE models by
means of the *IMPERFECTION command. The concentrated load at mid-span, simply sup-
ported conditions and lateral supports were carefully prescribed in the FE models. The non-
linear analysis was performed using the modified Riks method, which was found to be
capable of solving buckling problems considering both geometric and material nonlinearities.
The validation of FE models was performed by comparing the experimental and numerical
shear resistance versus vertical displacement curves, deformed shapes and principal membrane
strains. The comparison between test and FE results of specimen V-304-IS is shown in Figure 8,
and the numerically predicted ultimate shear resistance Vu,FE is summarised in Table 3. It is
shown that close agreement between experimental and numerical shear response has been
achieved.

4.2 Parametric analysis


Based upon the validated FE models, parametric analysis was performed herein to examine
the sensitivity of the stiffener inclination and the web aspect ratio to the shear resistance of
plate girders with inclined stiffeners. Austenitic stainless steel grade EN 1.4301 was considered
and the local imperfection amplitude of b/200, as recommended in EN 1993-1-5, was
employed. The same cross-section (b = 150 mm, tf = 12 mm, hw = 500 mm, tw = 4 mm) was
chosen for all FE models, and the width bs of inclined stiffeners was set as equal to 60 mm.
A total of three different aspect ratios (a/hw=1.5, 2.0, 2.5) and fourteen stiffener inclinations θ
varied from 38° to 90° were included, wherein the models with stiffener inclination θ equal to
90° corresponded to plate girders with transverse stiffeners only.
Figure 9 shows the ultimate shear resistances of plate girders with inclined stiffeners of vary-
ing inclinations compared to that of plate girders with transverse stiffeners only. It can be seen
that the shear resistance increases with smaller values of the stiffener inclination θ, and this
effect becomes more prominent for plate girders with lower web aspect ratios. The peak
values (denoted by VIS,Max) in shear resistance improvement are achieved with the stiffener
inclination θ equal to 48° for all three different web aspect ratios, beyond which the strength-
ening effect is undermined due to the web crippling occurred in the intermediate triangular
web panel below the loading point. Meanwhile, the resistance enhancement is less pronounced
for plate girders with smaller values of Lx/a, corresponding to less inclined web panels, and
the knee regions of the VIS - Lx/a curves for three web aspect ratios are found to be close to Lx
/a = 0.2. It is therefore concluded that substantial increases in shear resistance can be achieved
by placing the intermediate stiffeners inclined to the flanges with Lx/a > 0.2 and θ > 48°. Fur-
thermore, the comparison of the numerical results obtained in this study with those from the

Figure 8. Comparison between test and FE results of specimen V-304-IS (a) Shear resistance versus ver-
tical displacement curves, (b) Shear resistance versus principal membrane strain curves.

250
Figure 9. Influence of the stiffener inclination on ultimate shear resistance (a) VIS versus θ curves,
(b) VIS versus Lx/a curves.

Figure 10. Influence of the diagonal stiffener thickness on ultimate shear resistance.

previous study on diagonally stiffened plate girders (Yuan et al. 2019) is presented in Figure 10,
in which VDS is the ultimate resistance of diagonally stiffened plate girders with the same cross-
section considered in this study and the width-to-thickness ratio (bds/tds) of diagonal stiffeners
was taken as 7.5. It is revealed that the diagonally stiffened plate girders exhibited higher ultim-
ate capacities once sufficient stiffness of diagonal stiffeners was guaranteed (e.g. the thickness of
diagonal stiffeners tds > 6.16 mm for web aspect ratio a/hw = 1.5). Reduced requirement for the
stiffness of diagonal stiffeners can be used for web panels with higher aspect ratios, which was
attributed to the fact that the strengthening effect due to the presence of diagonal stiffeners was
more remarkable for plate girders with higher aspect ratios, while this effect was less pro-
nounced in case of inclined stiffeners.

5 CONCLUSIONS

Shear behaviour of stainless steel plate girders with inclined stiffeners has been investigated
both experimentally and numerially in this study. Two plate girders fabricated from hot-rolled
stainless steel plates were tested under simply supported conditions. Initial geometric imperfec-
tions of the cross-sections were measured prior to testing. By using the attached strain gauges
and the installed laser displacement transducers, the critical buckling strengths and ultimate
resistances were all determined. The experimental results were carefully analysed and further

251
compared to those of previously tested plate girders with transverse stiffeners only and both
transverse and diagonal stiffeners. It was revealed that by placing the intermediate stiffeners
inclined to the flanges, a significant increase in shear resistance was achieved, which was even
higher than that of diagonally stiffened plate girder. Elaborated FE models were developed
using ABAQUS and were validated against the obtained test results, based on which parametric
analysis was performed to examine the influence of the stiffener inclination and the web aspect
ratio. It was concluded that a substantial increase in shear resistance could be achieved by pla-
cing the inclined stiffeners at an appropriate position (Lx/a > 0.2 and θ > 48°). Moreover, the
diagonally stiffened plate girders exhibit most favorable structural performance with highest
ultimate shear resistance once sufficient stiffness of diagonal stiffeners was achieved.

ACKNOWLEDGEMENTS

The financial supports from the Natural Science Foundation of Hubei Province (Grant no.
2018CFB441), National Natural Science Foundation of China (Grant no. 51508424), the Fun-
damental Research Funds for the Central Universities (Grant no. 2042017gf0047) are greatly
acknowledged by the authors. The first author also appreciates the financial support from the
China Scholarship Council for sponsoring the study in the UPC.

REFERENCES

Azmi, M.R. Yatim, M.Y.M. Esa, A. & Wan Badaruzzaman, W.H. 2017. Experimental studies on perfor-
ated plate girders with inclined stiffeners. Thin-Walled Structures 117: 247–256.
Chen, X.W. Yuan, H.X. Du, X.X. Zhao, Y. Ye, J. & Yang, L. 2018. Shear buckling behaviour of welded
stainless steel plate girders with transverse stiffeners. Thin-Walled Structures 122: 529–544.
Estrada, I. Real, E. & Mirambell, E. 2007. General behaviour and effect of rigid and non-rigid end post
in stainless steel plate girders loaded in shear. Part I: Experimental study, Journal of Constructional
Steel Research 63 (7): 970–984.
Estrada, I. Real, E. & Mirambell, E. 2008. Shear resistance in stainless steel plate girders with transverse
and longitudinal stiffening, Journal of Constructional Steel Research 64 (11): 1239–1254.
EN 1993-1-5. 2006. Eurocode 3: Design of steel structures–Part 1.5: Plated structural elements. CEN.
Brussels.
EN 1993-1-4+A1. 2015. Eurocode 3: Design of steel structures–Part 1.4: General rules–supplementary
rules for stainless steel. CEN. Brussels.
Guarnieri, G. 1985. Collapse of plate girders with inclined stiffeners. Journal of Structural Engineering,
111: 379–399.
Jensen, C.D. & Antoni, C.M. 1941. Welded girders with inclined stiffeners. Welding Journal 20: 170–182.
Lotz, W.F. 1936. Welded girders with inclined stiffeners. B.S. Thesis. Lehigh University. UAS.
Olsson, A. 2001. Stainless steel plasticity-material modeling and structural application. PhD Thesis. Lulea
University of Technology. Sweden.
Real, E. Mirambell, E. & Estrada, I. 2007. Shear response of stainless steel plate girders. Engineering
Structures 29 (7): 1626–1640.
Saliba, N. & Gardner, L. 2013. Experimental study of shear response of lean duplex stainless steel plate
girders, Engineering Structures 46: 375–391.
Vann, W.P. & Sehested, J. Experimental techniques for plate buckling, in: Proceedings of the 2nd spe-
cialty conference on cold-formed steel structures, University of Missouri-Rolla, USA, October, 1973
Xiang, Y., Wang, C. M., & Kitipornchal, S. 1995. Buckling of skew Mindlin plates subjected to in-plane
shear loadings. International journal of mechanical sciences 37(10): 1089–1101.
Yonezawa, H. Miakami, I. Dogaki, M. & Uno, H. 1978. Shear strength of plate girders with diagonally
stiffened webs, Transactions of the Japan Society of Civil Engineers 10: 17–27 (in Japanese).
Yuan, H.X. Chen, X.W. Theofanous, M.; Wu, Y.W. Cao, T.Y. & Du, X.X. 2019. Shear behaviour and
design of diagonally stiffened stainless steel plate girders. Journal of Constructional Steel Research 153:
588–602.

252
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Ultimate strength analysis of steel-concrete cross-sections


at elevated temperatures

C.G. Chiorean, M. Selariu & S.M. Buru


Faculty of Civil Engineering, Technical University of Cluj-Napoca, Cluj-Napoca, Romania

ABSTRACT: This paper presents a new computational method for ultimate strength analysis
of composite steel-concrete cross-sections with arbitrary shape subjected to elevated temperatures.
The analysis method is carried out in two main steps: (i) thermal analysis, used to evaluate the
temperature distribution throughout the cross-section at a specific time and (ii) mechanical ana-
lysis, in which the interaction diagrams are determined. The main feature of the proposed
approach consists in controlling the inelastic response at the strains level enforcing in the same
time the elasto-plastic equilibrium for a prescribed axial force and bending moments ratio. The
ultimate (maximum) strength capacity is formulated as a problem of unconstrained mathematical
optimization by applying the method of Lagrange multipliers. The optimized function is defined
as a total internal bending moment and two constrains are defined by enforcing the constant axial
force and bending moment ratio. Hence, the ultimate bending moment capacities are directly
obtained by solving just three coupled nonlinear equations. The developed procedure has been
used to predict the bending moment capacity diagrams of several cross-sections under fire actions
and the comparisons made prove the effectiveness and the reliability of the proposed method of
analysis.

1 INTRODUCTION

Ultimate strength capacity of a cross-section subjected to bi-axial bending moments and axial
force is defined as maximum values of the bending moments under constant axial force associ-
ated to a specific deformed state when further increase in deformation results in reduced
response. On the other hand, the nominal strength capacity of the cross-section is reached
when in the most compressed concrete fibre or tensioned steel the predefined nominal ultimate
values of the strains are attained. In general, conventional (nominal) strength capacity occurs
prior to maximum (ultimate) strength capacity when considering material standard laws with-
out strain-softening branches. However is important to note that when strain-softening effect
exhibited either by the concrete in compression or the steel susceptible to local buckling the
ultimate strength capacity of the composite cross-sections occurs prior to conventional failure.
During the last decades several methods have been presented for the nominal (conventional)
strength analysis of various concrete and composite steel–concrete sections under uniaxial and
biaxial moments and axial loads subjected to elevated temperatures (e.g. Caldas et al. 2010, El-
Fitiany and Youssef 2014). These approaches are based upon the solution of the nonlinear equi-
librium equations consisting of an iterative sequence of linear predictions and nonlinear correc-
tions to obtain either the strain distribution or the location and inclination of the neutral axis
which determines the ultimate load, where one or two components of the section forces can
remain constant. In general, these methods does not generates genuinely plane interaction curves,
give fast solutions but no evidence of numerical robustness have been proved yet and also are
sensitive to the origin of the loading axes and some problems in convergence may arise, particu-
larly when the initial or starting values of the variables are not selected properly and under large
axial forces in conjunction with strain-softening effect. Some numerical procedures on the ultimate
(maximum) strength analysis of cross-sections are available in literature (e.g. Kodur & Dwaikat,

253
2008; Law & Gillie, 2010). For instance in (Kodur & Dwaikat, 2008) ultimate strength capacity
evaluation, involves the construction of full moment-curvature diagrams for a specified axial load
and neutral axis orientation and detecting in this way when the maximum response is reached (i.e.
descending branch is initiated). Such an indirect approach exhibits relatively well numerical stabil-
ity but the main drawback of this approach is given by the fact that, for arbitrary shape cross-
section, the resulted bending moments associated to the prescribed total curvature do not laying
in the same plane. Furthermore, because for each point of the failure curve (surface) is necessary
to construct the full moment-curvature diagram this method is less effective from the computa-
tional point of view. A direct analysis approach for calculating the biaxial bending moment-axial
force ultimate strength capacity for arbitrary cross-section through the use of the cross-sectional
tangent stiffness has been developed in (Law & Gillie, 2010). Since this approach does not enforce
equilibrium for bending moments about each principal axis, spurious bending moments may
occur and hence some inaccuracies may be introduced in representing in the same plane the fail-
ure diagrams of the cross-sections with arbitrary shape.
A new direct numerical procedure is proposed herein in order to determine the ultimate
strength capacity of composite steel-concrete cross-sections with arbitrary shape at elevated
temperatures. The ultimate (maximum) strength capacity is formulated as a problem of
unconstrained mathematical optimization by applying the method of Lagrange multipliers for
finding the local maxima of a function subjected to equality constraints. The optimized func-
tion is defined as a total internal bending moment and two constrains are defined by imposing
the constant axial force and bending moment ratio. Both vertical and horizontal interaction
diagrams are evaluated taking into account the material property degradation due to the tem-
perature increase and the additional thermal strains under the assumption that spalling of con-
crete does not occur. A computer program was developed, aimed at obtaining the ultimate
strength capacity of composite cross-sections under combined biaxial bending and axial load.
In order to illustrate the proposed method and its accuracy and efficiency, this program was
used to study several representative examples, which have been studied previously by other
researchers using independent fibre element solutions. The examples and the comparisons
made prove the effectiveness, time saving and reliability of the proposed method of analysis.

2 MATHEMATICAL FORMULATION

2.1 Basic assumptions and constitutive material models


Consider the cross-section subjected to the action of the external bending moments about both
global axes and axial force as shown in Figure 1. The cross-section may assume any shape with
multiple polygonal or circular openings. It is assumed that plane section remains plane after
deformation. This implies a full interaction between the steel and concrete components of
a composite concrete-steel cross section. Shear and torsional interaction effects are not

Figure 1. Model of arbitrary composite cross-section subjected to fire analysis.

254
accounted for, neither in the concrete nor in the steel constitutive models. Thus, at the ambient
temperatures, the resultant strain distribution corresponding to the curvatures about each global
axes Φ={ϕz, ϕy} and the axial strain u can be expressed at a generic point, in concrete matrix
fiber and structural steel or ordinary reinforced bars, of coordinates ðz; yÞ in a linear form as:
 
ε u; z ; y ¼ u þ y z þ z y ð1Þ

In fire conditions, the implicit nonlinear constitutive models both for steel and concrete are
considered, without an explicit formulation of the creep and transient strains (Štefan et al.,
2019). As a consequence of the above assumption, for a given fiber (steel or concrete), total
strain is assumed to follow an additive decomposition of mechanical and thermal strain as:
 
εtot u; z ; y ; T ¼ u þ y z þ z y ¼ εmech ðT Þ þ εth ðT Þ ð2Þ

where εmech represents the mechanical strain and εth represents the thermal strain a function of
the temperature T at the specified fiber. Thermal transient and creep strains are implicitly
included in nonlinear constitutive models adopted for steel and concrete. From Eq. (2) the
mechanical strain can be expressed as:
εmech ¼ εtot  εth ð3Þ

and the implicit nonlinear material model for concrete and steel at high temperatures can be
defined in the form:
σ ¼ σðεmech ; T Þ ð4Þ

In this paper, moisture effects and all the material properties, both thermal and mechanical,
are assumed according to European Standards (EN 1992-1-2, EN 1993-1-2, EN 1994-1-2). In
the thermal analysis reinforcement is neglected, with the temperatures of the bars considered
equal with the surrounding concrete temperature. Residual stresses for encased steel profile
may be incorporated in the analysis as described in (Chiorean, 2017). The standard fire curve
is used in this paper to obtain the gases temperature:

T∞ ¼ 20 þ 345  logð8  t þ 1Þ ð5Þ

where T∞ is the gas temperature (°C) and t is the time of exposure to fire in minutes.

2.2 Thermal analysis


For a given cross-section, thermal analysis is performed by solving the 2-D heat conduction
equation:

∂T
kr2 T ¼ ρc ð6Þ
∂t

where k is the thermal conductivity, c is the specific heat and ρ is the mass density of material,
t represents time, and T represents temperature. The combined effect of convective and radia-
tive boundary conditions is considered as:

krT  n ¼ hðT  T∞ Þ ð7Þ

where T∞ is the ambient (gas) temperature, h is the temperature-dependent combined convective-


radiative heat transfer coefficient and n is the vector normal to the boundary. The initial conditions
consist of the temperature of every point throughout the cross-section at the beginning of the

255
analysis T(y, z, t0) = T0 (y, z) where the temperature T0 is specified. In the FEM context in which
the above mentioned transient heat transfer problem is solved, the equilibrium equation is described
as:
 
∂T
C þ KfTg ¼ R ð8Þ
∂t

where C, K and R below are the capacitance matrix (thermal capacity), the thermal conductiv-
ity matrix and the nodal heat flux vector, respectively:
ð ð ð ð
C ¼ ρcNT NdO; K ¼ BT DBdO þ h NT Nd; R ¼ hT∞ NT d ð9Þ
O O  

where B is the matrix containing the derivatives of the interpolation functions, N is the matrix
containing the interpolation shape functions, and D is the thermal conductivity matrix. To
obtain the solution of Eq. (8) an explicit numerical time integration strategy based on the finite
difference method is adopted (Barros et al., 2018). Since the Eq. (8) is strongly nonlinear due to
the dependence of the material properties on temperature some incremental or incremental-
iterative strategies have to be considered in solving the nonlinear system of equations given by
Eq. (8). In the proposed approach a simple incremental procedure has been applied and imple-
mented in the developed computer program. In this respect two types of 3 and 6-noded two-
dimensional triangular finite elements, with three integration points, were implemented in the
computer program to solve the heat transfer of steel-concrete cross-sections.

2.3 Formulation of the proposed method for ultimate strength analysis of cross-sections
The analysis method is carried out in two main steps: (i) thermal analysis, used to evaluate the
temperature distribution throughout the cross-section at a specific time and (ii) mechanical ana-
lysis, in which the interaction diagrams are determined. The main feature of the proposed
approach consists in controlling the inelastic response at the strains level enforcing in the same
time the elasto-plastic equilibrium for a prescribed axial force and bending moments ratio. The
failure diagrams correspond to a specific deformed state when further increase in deformation
results in reduced response. A new numerical procedure is proposed in order to determine the
ultimate strength capacity of composite steel-concrete cross-sections with arbitrary shape. The
ultimate (maximum) strength capacity is formulated as a problem of mathematical optimization
by applying the method of Lagrange multipliers for finding the local maxima of a function sub-
jected to equality constraints. The optimized function is defined as a total internal bending
moment and two constrains are defined by imposing the constant axial force and bending
moment ratio. The total internal bending moment can be expressed as:
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Mtot ¼ 2
Mz;int þ My;int 2 ð10Þ

where:
ðð
  
Mz;int ¼ σ εmech u; z ; y ; T ydA
O
ðð ð11Þ
  
My;int ¼ σ εmech u; z ; y ; T zdA
O

and where the surface integral is extended over steel and concrete areas (Ω). In the following
we will consider that the bending moment ratio is kept constant and the internal axial force
(Nint) has a predefined value (N0), such that:

256
My;int ¼ tgðαÞMz;int
ðð
     ð12Þ
Nint u; z ; y ¼ σ εmech u; z ; y ; T dA ¼ N0
O

The reinforcement bars contribution is added to Eqs. (11-12) above and its temperature is
assumed to be equal to the temperature in the corresponding position of the concrete mesh. The
problem of evaluation of the maximum value of the total bending moment for a given value of
the axial force (N0) and a given bending moment ratio (tg(α)) can be considered as an optimiza-
tion problem. Considering the axial strain (u) and curvatures about each principal bending
moment axis (ϕz, ϕy) as the main unknowns and the constraints defined by enforcing the con-
stant axial force and bending moment ratio, the optimization problem can be defined as:
   
Maximize f u; z ; y ¼ Mtot u; z ; y
(    
g1 u; z ; y ≡Nint u; z ; y  N0 ¼ 0 ð13Þ
Subject to      
g2 u; z ; y ≡My;int u; z ; y  tgðαÞMz;int u; z ; y ¼ 0

Introducing the new variables (λ1, λ2) called the Lagrangian multipliers, in the following we
will study the maximum value of the Lagrange function which incorporates both the original
objective function and constraint equations:
   
L u; z ; y ; λ1 ; λ2 ¼ f u; z ; y þ λ1 g1 ðu; z ; y Þ þ λ2 g2 ðu; z ; y Þ
or
   qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ð14Þ
L u; z ; y ; λ1 ; λ2 ¼ Mz;int u; z ; y 1 þ tg2 ðαÞ þ λ1 Nint ðu; z ; y Þ  N0

   
þ λ2 My;int u; z ; y  tgðαÞMz;int u; z ; y

In order to solve the above unconstrained optimization problem the following nonlinear
system of equations has to be derived:
8
  <   P 2  
rf u; z ; y þ λk rgk u; z ; y ¼ 0
ru;z ;y ;λ1 ;λ2 L u; z ; y ; λ1 ; λ2 ¼ 0 , ð15Þ
:   
k¼1 
g1 u; z ; y ¼ g2 u; z ; y ¼ 0

which can be further explicitly expressed as:


8 ÐÐ   pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ÐÐ  
>
> Et u; z ; y ydA 1 þ tg2 ðαÞ þ λ1 Et u; z ; y dA þ λ2
>
>
>
> O  O 
>
> ÐÐ   ÐÐ  
>
>  ;   ð αÞ  ;  ¼0
>
> E t u; z y zdA tg E t u; z y ydA
>
> Ω pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi Ω
>
> ÐÐ   ÐÐ  
>
> Et u; z ; y y2 dA 1 þ tg2 ðαÞ þ λ1 Et u; z ; y ydA þ λ2
>
>
>
> O  O 
>
> ÐÐ   ÐÐ  
>
< Et u; z ; y zydA  tgðαÞ Et u; z ; y y2 dA ¼ 0
ÐÐ   Ωpffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ÐÐ  Ω  ð16Þ
>
> Et u; z ; y zydA 1 þ tg2 ðαÞ þ λ1 Et u; z ; y zdA þ λ2
>
>
>
> O  O 
>
>
>
> ÐÐ   2 ÐÐ  
>
> Et u; z ; y z dA  tgðαÞ Et u; z ; y zydA ¼ 0
>
>
>
>ÐÐ   Ω Ω
>
>
> σ εmech u; z ; y dA  N0 ¼ 0
>
>
> ÐÐ 
O   ÐÐ ÐÐ   
>
>
>
: σ εmech u; z ; y xdA  tg ðαÞ σ εmech u; z ; y ydA ¼ 0
O O O

257
where
  
  ∂σ εmech u; z ; y ; T
Et u; z ; y ; T ¼   ð17Þ
∂εmech u; z ; y ; T

represents the tangent Young modulus and is evaluated for each fibre, concrete and struc-
tural steel, as function of the fibre strains and the assumed nonlinear stress-strain constitu-
tive equations at a specified temperature T. Introducing the following notations:
ðð ðð
       
EAt u; z ; y ; T ¼ Et u; z ; y ; T dA; ESy u; z ; y ; T ¼ Et u; z ; y ; T zdA;
O O
ðð ðð
       
ESz u; z ; y ; T ¼ Et u; z ; y ; T ydA; EIz u; z ; y ; T ¼ Et u; z ; y ; T y2 dA; ð18Þ
O O
ðð ðð
       
ESzy u; z ; y ; T ¼ Et u; z ; y ; T zydA; EIy u; z ; y ; T ¼ Et u; z ; y ; T z2 dA;
O O

and eliminating from the first two equations above the dummy variables λ the above
system of equations reduces at the following three coupled nonlinear equations with
unknowns (u, ϕz, ϕy):
8    2
>
>  y  EAt EIy  ESz  EAt ESzy  ESz ESy ¼0
2 2
> EA EI ES
>
< ÐÐ 
t z
 
σ εmech u; z ; y ; T dA  N0 ¼ 0 ð19Þ
>
> O 
ÐÐ   ÐÐ   
>
> ð Þ
: σ εmech u; z ; y ; T xdA  tg α σ εmech u; z ; y ; T ydA ¼ 0
O O

The above nonlinear system of equations can be solved numerically using, for instance, the
iterative Newton method or bracketing methods (Chiorean, 2019) and taking into account the
fact that the stresses are implicit functions of curvatures through the mechanical strain distri-
bution given by Eq. (3). In this way, for given bending moments ratio and given value of the
axial force we can obtain directly the strain state throughout the cross-section defined by the
parameters (u*, ϕ*z, ϕ*y) solution of the system (19) corresponding to a maximum value of the
total bending moment. Then the ultimate bending moments can be simply evaluated as:
ðð  
Mz;int ¼ σ εmech u ; z ; y ydA; My;int ¼ tgðαÞMz;int ð20Þ
O

This procedure can be applied for evaluation of the interaction diagram and moment capaci-
ties contours. For instance for determination of the moment capacity contour (My-Mz) this
procedure is applied repetitively for different values of the angle α and same value for the
axial force, whereas for determination of the interaction diagram (N-M) the procedure is
applied repetitively for different values of the axial force N0 and keeping constant the bending
moment ratio. Further details regarding the numerical implementation and the global root
bracketing approach implemented by combining the Newton with Brent-Decker method for
solving the nonlinear system (19) is given in (Chiorean, 2019).

3 COMPUTATIONAL EXAMPLE

The composite steel-concrete cross-section depicted in Figure 2a (Chen et al., 2001, Caldas et.al.,
2010) consists of the concrete matrix, fifteen reinforcement bars of diameter 18 mm, a structural
steel element and a circular opening. Characteristic strengths for concrete, structural steel and

258
Figure 2. (a) Composite steel-concrete cross-section (dimensions in mm); (b) Finite element mesh.

reinforcement bars, at ambient temperature are: σc = 20 MPa,, fys = 323MPa, fyr = 400MPa,
respectively. The Young modulus for all steel sections was 200 GPa. The section is subjected to
the standard fire (Eq. (5)). The cross-section has been subdivided into a 2018 total number of
three-noded finite elements (Figure 2b), and the thermal properties for concrete are taken from
Eurocode 2 for siliceous aggregates with 0% moisture. The tensile strength of concrete is neglected
in the proposed analysis as in the comparatives one presented in literature (Chen et al., 2001,
Caldas et.al., 2010).
Using the proposed method developed here, the ultimate strength capacities of the cross-
section is evaluated in terms of the bending-moment capacity diagrams for a compressive axial
force of N = 4120 kN. The capacity diagrams are reported at the plastic centroid (Chen et al.,
2001) of the cross-section computed for ambient temperature. The temperature fields for a 30,
90, 180 and 300 minutes exposure are depicted in Figure 3. Caldas et al. studied this cross-
section and nominal bending moment capacities have been determined associated to the ultimate
compressive strain adopted conservatively as the value corresponding to the peak stress.

Figure 3. Temperature distribution for different exposure times.

259
Figure 4. Moment capacity contours for compressive axial force N = 4120 kN.

The comparative curves of the bending moment capacities diagrams for a compressive value
of axial force equal with 4120 kN for ambient temperature and for several exposure times are
depicted in Figure 4. As expected, higher exposure times leads to contraction of the moment
capacity diagrams, this means lower strength capacity of cross-section. As it can be seen fairly
good agreement exist between the results obtained from the proposed approach and those
selected from literature. However, some differences can be observed between the predicted
results using the proposed approach and those given in (Caldas et al., 2010). The procedure
developed in Caldas et al., 2010 determine the nominal strength capacity assuming lower limit
strain for concrete (associated to the peak stress) whereas the proposed approach determine the
ultimate strength capacity associated to the maximum total bending moment of the cross-section.
Moreover, it is important to note that the methods used here for comparisons do not generate
genuinely plane moment capacity curves, as is the main feature of the proposed approach.

4 CONCLUSIONS

A new computational method for the ultimate strength analysis of composite steel-concrete cross-
sections with arbitrary shape subjected to elevated temperatures has been developed. The ultimate
strength capacity is formulated as a problem of unconstrained mathematical optimization of the
total internal bending moment and two constrains are defined by enforcing the constant axial
force and bending moment ratio. Comparing the proposed method with the existing ones several
important features may be highlighted: (i) the proposed approach is able to compute the ultimate
strength capacity of cross-sections directly, with a high degree of accuracy, without the need to
evaluate the entire moment-curvature diagram, by means of solving just three coupled nonlinear
equations enforcing in the same time equilibrium for bending moments about each principal axis
resulting genuinely plane interaction diagrams; (ii) the numerical procedure adopted for solving
the nonlinear system of equations involves a direct and explicit tangent stiffness strategy thus
resulting in a high rate of convergence; (iii) from the extensive numerical experiments, some of
them presented in this work, has been found that the convergence stability is not sensitive to the
initial/starting values of the iterative process, to how the origin of the reference loading axes is
chosen or to the strain softening exhibited by the concrete in compression. The developed proced-
ure has been used to determine the bending moment capacity diagrams of composite-steel concrete
cross-section with arbitrary shape and different fire exposure regimes. The comparisons made
prove the effectiveness and reliability of the proposed method of analysis. Further studies are
envisaged for extending the proposed approach for nominal strength capacity evaluation and also
studies concerning the numerical stability and computational efficiency of this approach are
underway and will be presented in a future work.

260
REFERENCES

Barros, R.C., Pires, D., Silveira, R.A.M., Lemes, I.J.M., Rocha, P.A.S. Advanced inelastic analysis of
steel structures at elevated temperatures by SCM/RPHM coupling, Journal of Constructional Steel
Research, 145, 368–385, 2018.
Caldas, R.B., Sousa, J.B.M., Fakury, R.K. Interaction diagrams for reinforced concrete sections sub-
jected to fire, Engineering Structures, 32, 2832–2838, 2010.
Chiorean, C.G. A computer method for moment-curvature analysis of composite steel-concrete
cross-sections of arbitrary shape. Engineering Structures and Technologies, 9(1), 25–40, 2017.
Chiorean, C.G., Ultimate and nominal strength capacity evaluation of composite sections with arbitrary
shapes at elevated temperatures, Submitted for publication, 2019.
Chen, S.F., Teng, J.G., and Chan, S.L. Design of biaxially loaded short composite columns of arbitrary
section, Journal of Structural Engineering, 127, 840–846, 2001.
El-Fitiany, S.F., Youssef, M.A. Interaction diagrams for fire-exposed reinforced concrete sections, Engin-
eering Structures, 70, 246–259, 2014.
EN 1992-1-2. Eurocode 2: Design of concrete structures-Part 1-2: General rules-Structural fire design.
CEN; 2004.
EN 1993-1-2. Eurocode 3: Design of steel structures-Part 1-2: General rules-Structural fire design. CEN;
2005.
EN 1994-1-2. Eurocode 4: Design of composite steel and concrete structures-Part 1-2: General
rules-Structural fire design. CEN; 2005.
Kodur, V.K.R., Dwaikat, M. A numerical model for predicting the fire resistance of reinforced concrete
beams, Cement & Concrete Composites, 30, 431–443, 2008.
Law A. and Gillie M. Interaction diagrams for ambient and heated concrete sections, Engineering Struc-
tures, 32, 1641–1649, 2010.
Štefan, R., Sura, J., Prochazka, J., Kohoukova, A., Wald, F. Numerical investigation of slender
reinforced concrete and steel-concrete composite columns at normal and high temperatures using sec-
tional analysis and moment-curvature approach, Engineering Structures, 190, 285–305, 2019.

261
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Seismic design of two-storey X-bracings

S. Costanzo, M. D’Aniello, G. Di Lorenzo, A. De Martino & R. Landolfo


Department of Structures for Engineering and Architecture, University of Naples Federico II

ABSTRACT: Two-storey X-bracings are becoming very popular in European practice,


owing the possibility to reduce the bending demand typically observed on the brace-
intercepted beam in V and inverted-V configurations. However, they are not properly
addressed within the current Eurocode 8, and specific design provisions are missing. In this
paper the main critical issues affecting the seismic behaviour of two-story X concentrically
braced frames are discussed. With this regard, seismic design criteria provided by both US
and European codes for concentrically are summarized and critically discussed with specific
reference to two-story X configuration.

1 INTRODUCTION

Concentrically braced frames are commonly used lateral resisting systems in seismic design of
steel buildings at any seismicity. With reference to such structural typology, seismic codes pro-
vide capacity design criteria devoted restraining plastic deformation into the diagonal mem-
bers, while the remaining elements, namely columns, beams and connections, elastically
behave.
Several Authors (Elghazouli, 2010, 2017; Broderick et al. 2008; Brandonisio et al. 2012, Bosco
et al, 2014, 2016a,b,c, 2017a,b; Costanzo & Landolfo, 2017; Costanzo et al., 2015, 2018, 2019;
D’Aniello et al. 2015; Marino, 2013; Fiorino et al. 2016, 2017; Piluso et al. 2019a,b, Wang et al.
2018) critically discussed seismic criteria codified both in European and American framework.
Remarks from numerous researches focused on the brace-intercepted beam in V and inverted
V configuration. Indeed, after brace under compression buckles, an unbalanced vertical force due
to the different forces transferred by tension and compression braces is applied on the brace-
intercepted beam, which is subjected to large bending moment. In such condition, the formation
of a plastic hinge at mid-span of the beams should be avoided, otherwise it would result in a drop
of storey lateral resistance with consequent inelastic drift concentration at the storey with yielded
beams and significant deterioration of the overall response.
With the attempt to reduce large flexural demand typically observed in V and inverted
V configuration, growing attention has been put on two-storey X bracings, obtained by using
V and inverted V bracings in two consecutive stories (see Figure 1) (Shen et al 2014, 2015).
Such configuration allows reducing the beam depth as respect to V and chevron traditional
bracings, because the unbalanced forces applied on the brace-intercepted beam coming from
above and below storeys are in opposite direction and they would theoretically result in mod-
erate bending moment at the beam mid-length.
Moreover, arranging bracings as shown in Figure 1 allows avoiding geometrical and
technological difficulties commonly occurring in the design of traditional X braced frames.
Indeed, for typical values of interstorey height and span length used in steel structural build-
ings, to equip the frame with diagonal members in X-configurations leads improper brace-to-
beam angles (namely smaller than 30°), resulting in brace-to-beam/column connections
impractical from constructional and technological points of view (e.g. the dimensions of the
gusset plates can occupy up to 1/3 of the span). Conversely, the two-Storey X configuration,

262
Figure 1. Two-storey X CBF.

similarly to the traditional V one, allows slope of diagonal members in the range [30°-60°],
namely suitable for correct design of gusset plates at brace-to-frame connection location.
Even though many practice engineers are often opting for two-storey bracings, this config-
uration is not specifically addressed within EN 1998-1, which provides no specific rules. At
the contrary, two-storey X concentrically braced frames are specifically mentioned by AISC
Design Manuals.
In the following Sections design criteria provided by both European and US codes
for concentrically braced frames in high ductility class are critically discussed with spe-
cific reference to the two-storey X configuration. For the sake of comparison, the
European notation is used hereinafter to define the elements resistances also for the
US code.

2 DESIGN RULES ACCORDING TO AISC 341-16

In United States of America (USA), the AISC 341 provisions and the applicable building
code, typically ASCE 7, give the rules for the analysis and design of seismic resisting
structures. In this framework, two-storey X special concentrically braced frames (i.e. high
ductile bracings, acronym X-SCBF) are conceived as made by two adjacent V and
inverted V bracings and the relevant design provisions are extended to the two-storey
X configurations, as described hereinafter.
According to AISC 341 the required strength of dissipative diagonals (whatever configur-
ation of bracings is considered) is calculated by performing linear-elastic global analysis on
a tension-compression (T-C) structural scheme (Figure 2b) in which both tension and

Figure 2. T-O (a) and T-C (b) structural schemes.

263
compression diagonals are specifically accounted for and the bracings should be designed to
meet the following condition:

χNpl;br  NEd;br ð1Þ

where χ is the buckling reduction factor and NEd,br is the required strength.
To avoid significant degradation of brace response in the post-buckling range and to prevent
buckling under service condition and fracture phenomena, both global and local slenderness of
diagonal members should be verified: AISC 341 provides specific limitation on the brace geo-
metrical slenderness KL/r (being K the effective length factor; L the unsupported length and r is
the radius of gyration) which should be less than 200, resulting in less stringent requirement
than the European one (see Section 3), for typical material properties. As numerous Authors
(Tang & Goel, 1989; Elchalakani et al, 2003; Lee & Bruneau, 2005) highlighted, beside the
global slenderness, the local slenderness (i.e. the width-to-thickness ratio) significantly affects the
brace hysteretic response, particularly in case of stocky diagonal that are more prone to fracture
due to local buckling phenomena. Differently from the European approach (see Section 3),
AISC 341 mandates specific limitations for dissipative elements depending on the ductility class.
Concerning the design of non-dissipative members, capacity design criteria are deemed to
be satisfied by calculating the relevant required strength considering the most unfavourable
condition between:
i) the seismic-induced effects evaluated by mean of an elastic analysis with both braces in ten-
sion and compression, magnified by using the system overstrength factor Ωo = 2;
ii) the required strength evaluated by performing a plastic mechanism analysis: a free-body
force representative of the nonlinear behaviour is considered, in which all braces in tension
are assumed to attain forces corresponding to their expected tensile strength, and all braces
in compression are assumed to exhibit their expected post-buckling strength, evaluated as
the 30% of their relevant compression capacity.
In detail the is assumed:

NT ¼ γov  Npl;br ð2Þ

NC ¼ 0:3  χ  γov  Npl;br ð3Þ

According to the design examples given in AISC design manual, this approach is applied to
the two-story X bracings case considering that global plastic mechanism compliant to the first
mode of vibration shape occurs. Therefore, the diagonals at two consecutive stories (equipped
with V and inverted V bracings) are assumed reaching the same level of inelastic deformation
at the same time, thus resulting in very poor unbalanced force on the brace-intercepted beam.

3 DESIGN RULES ACCORDING TO EN 1998-1

Differently from US code, in EN 1998-1 the seismic design provisions for concentrically
braced frames are referred to two different subcategories: (i) traditional cross bracings
(XCBF) and V and inverted V bracings, which even differs for the force reduction factor,
smaller for the latter configuration, that is expected to provide limited ductility.
As previously mentioned, two-storey X bracings are not properly addressed in the frame-
work of EN 1998-1 and specific provisions are missing.
According EN 1998-1 the diagonal members in traditional one-storey X configuration are
designed according to a simplified design provisions: the code allows using a tension only
(T-O) model in which the solely the tension bracings resist the seismic action, while the contri-
bution of diagonal under compression is neglected. Thereby, the diagonals are designed to
guarantee the following requirement is satisfied:

264
Npl;br  NEd;br ð4Þ

where Npl,br,Rd is the plastic strength of brace cross-section and NEd,br is the design action cal-
culated by using T-O model (see Figure 2a).
On the contrary, EC8 stipulates to design the braces of inverted-V CBFs to resist the forces
calculated by using a tension-compression (T-C) model (Figure 2b) in which the contribution
of compression diagonals is specifically accounted for. Thereby, the braces are designed such
that:

χNpl;br  NEd;br ð5Þ

where χ is the buckling reduction factor and NEd,br is the required strength.
In addition, for both V and X configurations, to assure an uniform distribution of plastic
deformation along the building height and at avoiding detrimental soft-storey mechanisms,
EC8 mandates that the overstrength ratio Ωi = Npl,br,i/NEd,br,i should vary within the range Ω
to 1.25Ω.
Differently from AISC, EN 1998-1 limit the slenderness ratio λ which should fall in the
range [1.3, 2] for X bracings and should be minor than 2 for the V and inverted
V configuration. The introduction of lower bound limit for the cross bracings is due to the
simplified tension-only diagonal model assumed for structural analysis. Indeed, the presence
of the compression diagonal is neglected, and the lower bound slenderness limit is introduced
to avoid overloading of the column.
EN 1998-1 does not provide width-to-thickness ratio limitations specifically conceived for
the dissipative members, expected to undergo severe plastic deformation and the non-seismic
rules given by EN 1993-1-1 are extended to the seismic case, with reference to the ductility
class considered. This approach result is very less stringent limitation as respect to the US one.
According to EN 1998-1 the required strength of non-dissipative members is obtained by
performing a linear-elastic analysis and by magnifying the seismic-induced effect by the over-
strength factor Ω = min(Ωi) where Ωi = Npl,br,i/NEd,br,i.
Plastic mechanism analysis is required solely for the brace-intercepted beam in V and inverted
V configuration. In such cases the tension braces are assumed attaining their plastic strength
(Npl,br), while the compression ones transfer their post-buckling capacity calculated as:

NC ¼ 0:3  Npl;br ð6Þ

Differently from AISC341, the European code mandate assuming the 30% of the plastic resist-
ance of the member, rather than its compression capacity, thus disregarding the diagonal slen-
derness; it should be noted that Equation 6 leads to force distribution that is not consistent
for very slender brace (close to the upper bound limit 2) for which the buckling strength tends
to the 20% of the plastic one (the buckling reduction factor is around 0.2), thus smaller than
the post-buckling residual capacity recommended by EN 1998-1.

4 CRITICAL DISCUSSION

As discussed in previous Sections, two-storey X-bracings are very attractive for practice engineers;
however, this configuration is not properly addressed within Eurocode 8, while AISC341 design
manual explicitly consider the two-storey X configuration, extending the detailing rules provided
for V and inverted V bracings.
This approach is implicitly based on the assumptions that global plastic mechanism compliant
to the first mode of vibration shape develops and the braces at two adjacent stories transfer the
same forces at the same time. Since the forces at the stories above and below the beam under
consideration are applied in opposite direction, under this assumption the beam is subjected to

265
very poor bending demand. However, it should be noted that the storey drift ratios of two con-
secutive storeys are generally different from each other and different possible loading patterns
applied at the brace-intercepted beams need to be considered (Shen et al., 2014, 2015). Indeed, if
soft-storey mechanism occurs, bracings above/below the beam behave elastically and the unbal-
anced force applied at the brace-intercepted section could be significantly larger.
Both US and European codes allow performing a linear elastic analysis of the structure to
calculate the required strengths of the diagonal members.
Moreover, EN 1998-1 allows using simplified T-O model to calculate the internal force for
XCBF. This simplified approach is based on the assumption that negligible contribution in
terms of strength and stiffness is provided for by the compression diagonals after buckling
occurs (Costanzo & Landolfo 2017, Costanzo et al., 2019). However, this ideal model is
a reasonable approximation only for the post-buckling situation, while in the first stages of
seismic event both diagonals are active and axial force is transferred at the column and brace
connection by the compression diagonal. Even for stocky diagonals the T-O model does not
accurately represent the structural behaviour and the compressed diagonals could transmit to
the non-dissipative zones significant effects.
As observed by numerous Authors, the homogeneity condition of the overstrength ratios is not
adequate to assure uniform distribution of plasticity and to avoid soft-storey mechanisms. More-
over, the interrelation and juxtaposition of the requirement on the Ω variation and normalized
slenderness limitations significantly influence the sizing process of diagonal members. Indeed, in
typical configurations, the diagonals at the roof storey are characterized by the largest over-
strength ratio, due to the need of contemporarily meeting the slenderness upper bound limit. As
a consequence, the designer is forced to oversize the bracings at lower storeys in order to satisfy
the Ω homogeneity condition, resulting in poor plastic engagement of dissipative zones and unsat-
isfactorily energy dissipation capacity (Elghazouli, 2010, 2017; Broderick et al. 2008; Brandonisio
et al. 2012, Bosco et al, 2014, 2016a, 2016b, 2017a, 2017b; Costanzo & Landolfo, 2017; Costanzo
et al., 2015, 2018, 2019, D’Aniello et al, 2015; Marino, 2013).
Different approaches can be recognized between US and European codes to evaluate the
required strength of non-dissipative members. The response of concentrically braced frames is
basically ruled by the behaviour of the diagonal members, which exhibit large plastic engagement
after the buckling of braces and the nonlinear response of the system significantly differs from the
elastic behaviour; thereby, to perform plastic mechanism analysis, namely assuming a free-body
force distribution representative of post-buckling behaviour could be more conservative and
effective.

5 CONCLUSION

The paper discusses on the seismic design of two-storey X concentrically braced frames.
The design criteria provided by both European and US codes for concentrically braced
frames in high ductility class have been critically discussed with specific reference to the two-
storey X configuration, in order to identify the most effective design rules for seismic design of
this type of systems. The discussion inferred the following remarks:
– Two-storey X bracings are very attractive for practice engineers due to the possibility to
reduce the high bending demand typically observed on the brace-intercepted beam in trad-
itional V and inverted V configuration.
– EN 1998-1 does not specifically address concentric bracings in two-storey configuration,
while AISC extend the seismic rules provided for V and inverted V configuration. The
latter approach assumes that global plastic mechanism compliant to the first mode of vibra-
tion shape develops. However, even different scenarios could likely occur and should be
considered. Indeed, if soft storey mechanism develops, the force transfer mechanism tends
to the traditional V response and high bending demand can occur on the beam.
– The simplified T-O model allowed by EN 1998-1 for traditional one-storey X bracings is
often not adequate to represent the structural behavior.

266
– The homogeneity condition of the overstrength ratios is not effective to assure uniform dis-
tribution of plasticity and to avoid soft-storey mechanisms and it even entails significant
efforts in design process of bracings.

REFERENCES

American Institute of Steel Construction, Inc. (AISC). (2016). Seismic Provisions for Structural Steel
Buildings. ANSI/AISC Standard 341-16. AISC, Chicago, Illinois.
ASCE 7–10: minimum design loads for buildings and other structures. Reston, VA: American Society of
Civil Engineers; 2010.
Bosco M., Brandonisio G., Marino, E.M. Mele, E. De Luca, A. 2017. Ω* method: An alternative to
Eurocode 8 procedure for seismic design of X-CBFs. Journal of Constructional Steel Research 134,
135–147.
Bosco, M. Ferrara, E. Ghersi, A. Marino, E.M. Rossi, P.P. 2016b. Improvement of the model proposed
by Menegotto and Pinto for steel. Engineering Structures 124: 442–456.
Bosco, M. Ferrara, G.A.F. Ghersi, A. Marino, E.M. Rossi, P.P. 2016b. Application of nonlinear static
method with corrective eccentricities to steel multi-storey braced buildings. Geotechnical, Geological
and Earthquake Engineering 40: 193–203, 2016b.
Bosco, M. Ghersi, A. Marino, E.M. Rossi, P.P. 2016a. A capacity design procedure for columns of steel
structures with diagonals braces. Open Construction and Building Technology Journal 8: 196–207.
Bosco, M. Ghersi, A. Marino, E.M. Rossi, P.P. 2016c. Generalized corrective eccentricities for nonlinear
static analysis of buildings with framed or braced structure. Bulletin of Earthquake Engineering.
Brandonisio, G. Toreno, M. Grande, E. Mele, E. De Luca, A. 2012. Seismic design of concentric braced
frames. Journal of Constructional Steel Research 78: 22–37.
Broderick, B.M. Elghazouli, A.Y. Goggins, J. M. 2008. Earthquake testing and response analysis of
concentrically-braced-sub-frames. Journal of Constructional Steel Research 64:9, 997–1007.
Costanzo, S. & Landolfo, R. 2017. Concentrically braced frames: European vs North American seismic
design provisions. The Open Civil Engineering Journal 11: (Suppl 1: M11) 453–463. DOI: 0.2174/
1874149501711010453.
Costanzo, S. D’Aniello, M. Landolfo, R. 2019. Proposal of design rules for ductile X-CBFS in the frame-
work of EUROCODE 8, Earthquake Engineering and Structural Dynamics 48(1): 124–151. DOI:
https://doi.org/10.1002/eqe.3128.
Costanzo, S. D’Aniello, M. Landolfo, R. De Martino, A. 2018. Critical discussion on seismic design cri-
teria for cross concentrically braced frames. Ingegneria Sismica: International Journal of Earthquake
Engineering 35(2): 23–36.
D’Aniello, M. Costanzo, S. Landolfo, R. 2015. The influence of beam stiffness on seismic response of
chevron concentric bracings. Journal of Constructional Steel Research: 112, 305–324, DOI: https://doi.
org/10.1016/j.jcsr.2015.05.021
Elchalakani M. Zhao X., Grzebieta R. 2003. Tests of Cold-Formed Circular Tubular Braces under
Cyclic Axial Loading. Journal of Structural Engineering 129: No. 4.
Elghazouli, A.Y. 2017. Seismic design of steel structures to Eurocode 8. Structural Engineer 85 (12):
26–31.
Elghazouli, A.Y. 2010. Assessment of European seismic design procedures for steel framed structures.
Bulletin of Earthquake Engineering: 8, 65–89.
EN 1993:1-1, Eurocode 3: Design of steel structures – Part 1-1: General rules and rules for buildings.
CEN; 2005.
EN 1998-1-1. Eurocode 8: Design of structures for earthquake resistance – Part 1: General rules, seismic
actions and rules for buildings. CEN; 2005.
Fiorino, L. Macillo, V. Landolfo, R. 2017. Shake table tests of a full-scale two-story
sheathing-braced cold-formed steel building. Engineering Structures 151: 633–647. doi: 10.1016/j.
engstruct.2017.08.056.
Fiorino, L. Terracciano, M.T. Landolfo, R. 2016. Experimental investigation of seismic behavior of low
dissipative CFS strap-braced stud walls. Journal of Constructional Steel Research 127: 92–107.
doi:10.1016/j.jcsr.2016.07.027.
Lee K. & Bruneau M. 2005. Energy Dissipation of Compression Members in Concentrically Braced
Frames: Review of Experimental Data. Journal of Structural Engineering: 5524.
Marino, E.M. 2013. A unified approach for the design of high ductility steel frames with con-centric
braces in the framework of Eurocode 8. Earthquake Engng Struct. Dyn: 43, 97–118.

267
Piluso, V Pisapia, A. Castaldo, P. Nastri, E. 2019b. Probabilistic Theory of Plastic Mechanism Control
for Steel Moment Resisting Frames. Structural Safety 76: 95–107.
Piluso, V. Montuori, R. Nastri, E. Paciello, A. 2019a. Seismic response of MRF-CBF dual systems
equipped with low damage friction connections. Journal of Constructional Steel Research: 154,
263–277.
Shen, J. Wen, R. Akbas, B. 2015. Mechanisms in Two-story X-braced Frames, J. of Constr. Steel Res.
106: 258–277.
Shen, J. Wen, R. Akbas, B. Doran, B. Uckan, E. 2014. Seismic demand on brace-intersected beams in
two-story X-braced frames, Eng. Struct. 76: 295–312.
Tang, X. & Goel, S.C. 1989 Brace Fractures and Analysis of Phase I Structure, Journal of Structural
Engineering, ASCE Vol. 115, No. 8, August, pp. 1960–1976, Reston, VA.
Wang, Y., Nastri, E., Tirca, L., Montuori, R., Piluso, V. 2018. Comparative response of earthquake
resistant CBF buildings designed according to Canadian and European code provisions. Key Engineer-
ing Materials 763: 1155–1163.

268
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Experimental tests on bolted end-plate connections using thermal


insulation layer attached to steel structures

M. Couchaux & A. Alhasawi


Laboratoire de Génie Civil et Génie Mécanique, INSA, Rennes, France

A. Ben Larbi
CTICM, Saint-Aubin, France

ABSTRACT: This paper deals with the mechanical performances of thermal break for exter-
nal steel structures (balconies, passageways) attached to a steel facade with external thermal
insulation. The proposed solution is composed of a PVC or plywood layers inserted between an
end-plate connection and a steel column. Static and cyclic tests are performed on a cantilever
beam connected to a steel column by a thermal break in order to investigate the effect of the
thermal insulation layer on the rotational stiffness, bending moment resistances and failure
modes. These tests highlight on one hand the influence of the thermal insulation and on another
hand, the bolt configuration (extended and flush end-plate, stiffeners, RHS or I profile).

1 INTRODUCTION

Due to the risk of global warming, there is a necessity to decrease the CO2 emissions by redu-
cing energy consumption. With 40% of total energy consumption in Europe, and about 36%
of CO2, the building is a key sector in the fight against climate change. Although improving
thermal performances of the envelope components seems to be easy to reach, the problem of
singular points and interfaces between materials need to be solved. In most cases, the singular
points necessitate to be analysed considering the interaction between various normative
requirements. The case of thermal breaks that must meet thermal, mechanical and often
acoustic performance requirements is a good illustration.
Many researches have been conducted in recent years to develop solutions to reduce thermal
bridging in concrete buildings (Hardock and Roppel (2013), Keo et al (2018)). However, most of
these researches concerne linear thermal bridges, mainly at the junctions between the elements of
the building envelope. Few works have been devoted to point thermal bridges due to the imple-
mentation of systems and structures on a façade of a building (solar panels, balconies, passage-
ways) or roof (solar panels, HVAC systems). Recent works have been carried out on the
mechanical behaviour of connections in steel structures using “insulation” intermediate layers as
thermal breaks. Nasdala et al (2007) and Sulcova et al (2007) proposed to adapt the component
method of Eurocode 3 in order to take into account the influence of deformable insulation plates.
Cleary et al (2016) conducted tests on steel connections using filler plates made of Fiber-
Reinforced Resin as thermal barrier. Simple connections and moment connections tested under
service loading showed that the thermal filler plates increase the rotation of moment connections
and the vertical displacement of simple connections. Peterman et al (2016) performed tests on con-
nections insulated by Fiber-reinforced polymer fillers to characterize their long term behaviour
and to evaluate the modification of the shear resistances of bolts in presence of fillers. In fact, the
presence of fillers introduces a bending moment in the bolts in addition to shear forces and reduce
the resistance to this latter force. Ben Ben Larbi et al (2017) proposed a solution of thermal break
with end-plate connection for external steel structures (balconies, passageways) attached on
a concrete façade with external thermal insulation. This solution consists in the implementation of
a rigid PVC layer and an acoustic insulation between the end-plate of the beam and the support

269
(floor slab, concrete wall). Hamel and White (2016) performed thermal and mechanical tests on
thermal break composed of extended end-plates resting on neoprene and connected to a steel sup-
port. The use of neoprene as thermal break clearly decrease the rotational stiffness of the connec-
tion and favour the failure of bolts in tension, as a result decrease the ultimate resistance. The
crushing of the neoprene in the compressive area was very important and that confirmed the
results obtained by Ben Larbi et al (2017) when they tested an elastomer associated with a PVC.
Thus, the use of thermal insulation stiffer than elastomer such as PVC and plywood is necessary.
This paper presents an experimental study of thermal breaks composed of PVC or plywood
layers placed between the bolted end-plate of a steel beam and a support composed of a steel
column. The objective of the tests, the experimental test set-up and the mechanical characteris-
tics of materials are presented in section 2. The results obtained during monotonic tests per-
formed on a cantilever beam subjected to a major bending moment and a minor shear force
are depicted in section 3. The effect of bolt configuration, thermal insulation on the rotational
stiffness and failure mode are highlighted. The image correlation is also used to evaluate the
deformed shape of the end-plate during the loading and estimate the area of the compressive
zone. Finally cyclic tests have been performed on these configurations (see section 4) to esti-
mate the effect of this new design on the seismic behaviour of the structures.

2 TEST PROGRAM

2.1 Presentation of tests


The aim of the tests was to study the effect of bolt configuration and the type of thermal insulation
layer on the mechanical behaviour of thermal break (rotational stiffness, resistances and failure
mode) subjected to a major bending moment and a minor shear force. A total of 25 experimental
tests have been conducted on a wide range of thermal breaks in the laboratory of INSA Rennes.
A series of 16 connections have been subjected to monotonic loadings whilst the remaining connec-
tions were tested under cyclic loading. A total of six geometries of steel connections have been
studied (from PA-1 to PA-6) that contain end-plate welded to IPE 200 (PA-1, PA-2 and PA-5) or
tubes (RHS for PA-3 and PA-4, SHS for PA-6) permitting to evaluate the effect of the beam-shape
on the connection’s behaviour. The details of the connections are depicted in Figure 1 and Table 1.
Specimens PA-1, RTPA-1-P/B and PA-3, RTPA-3-P/B are extended end-plate of 15 mm
thickness connected to IPE and RHS respectively. To improve their mechanical characteristics
some specimens have been reinforced by stiffeners in tensile and compressive area. The afore-
mentioned specimens, for both IPE and RHS, correspond to PA-4, RTPA-4-P/B and RTPA-
5-P respectively.

Figure 1. Geometry of thermal break tested.

270
Table 1. Geometrical properties of specimen tested.
ta hp e1 Thermal
Connection Beam (mm) (mm) (mm) insulation Stiffener Test

PA-1 IPE200 - 340 35 - No Monotonic/cyclic


RTPA-1-P IPE200 20 340 35 PVC No Monotonic/cyclic
RTPA-1-B IPE200 30 340 35 Wood No Monotonic/cyclic
PA-2 IPE200 - 220 50 - No Monotonic/cyclic
RTPA-2-P IPE200 20 220 50 PVC No Monotonic/cyclic
RTPA-2-P1 IPE200 10 220 50 PVC No Monotonic/cyclic
RTPA-2-B IPE200 30 220 50 Wood No Monotonic/cyclic
PA-3 RHS 200 × 120 × 6 - 340 35 - No Monotonic
RTPA-3-P RHS 200 × 120 × 6 20 340 35 PVC No Monotonic
RTPA-3-B RHS 200 × 120 × 6 30 340 35 Wood No Monotonic
PA-4 RHS 200 × 120 × 6 - 340 35 - Yes Monotonic
RTPA-4-P RHS 200 × 120 × 6 20 340 35 PVC Yes Monotonic/cyclic
RTPA-4-B RHS 200 × 120 × 6 30 340 35 Wood Yes Monotonic
RTPA-5-P IPE200 20 340 35 PVC Yes Monotonic/cyclic
PA-6 SHS 80 × 5 - 180 35 - No Monotonic
RTPA-6-P1 SHS 80 × 5 10 180 35 PVC No Monotonic

In addition, specimens PA-2, RTPA-2-P/P1/B are made of flush end-plate to investigate the
impact of insulations on these short connections. Furthermore, thermal break could be
employed to install elements on the building roof using small dimensions of SHS; specimens
PA-6, RTPA-6-P1 are suitable solutions of thermal breaks for these elements. The thermal
insulation layers are composed of structural pine plywood or PVC plates. All these connec-
tions are used with PVC layer of 10 and 20 mm thickness and plywood of 30 mm. The thick-
ness of plywood is increased due to its high thermal conductivity comparatively with PVC.

2.2 Test set-up and measurements


The test set-up is composed of a cantilever beam connected to a strong column designed to
remain elastic during all the campaign test (see Figure 2). The column is braced by a diagonal
made of an HEA 200. Lateral supports are placed at mid-length of the beam in order to
avoid/delay the lateral torsional buckling of the steel profile. The specimen is loaded by one
load-jack with a capacity of 1000 kN in both direction positioned at 1100 mm from the con-
nection. The specimens can be replaced after each test.
Displacement transducers (LVDT) are positioned to measure the displacements at different
locations of the test specimen as shown in Figure 2-a. The horizontal displacement transducers

Figure 2. Test set-up and measurements.

271
Table 2. Mechanical characteristics of steel.
Yield strength Tensile strength Elongation Necking
Element 2 2
N/mm N/mm % %
End-plate 10 247 329 26,0 74,4
End-plate 15 331 452 18,3 69,8
IPE200-flange 300 427 20,3 68,2
IPE200-web 307 433 21,7 70,6
RHS200-flange 474 568 14,3 63,4
RHS200-web 486 565 14,7 61,3
SHS 80 514 550 21,3 64,9
Bolt M16 1025 1117 4,4 76,3
Bolt M12 813 923 - 61,6

Ut1 and Ub1 provide a direct estimation of the rotation of the connection. Inclinometers were
also positioned on the beam near the end-plate and on the column in order to evaluate the
rotation of the connection. Bolts have been instrumented with axial strain gauges BTM-6C in
order to measure the axial strain and thus by calibration, the tensile force in the bolts.
Digital image correlation was used to evaluate the displacements of the end-plates, beam
and column in order to determine the rotation of the connection as the deformation of the
end-plate relatively to the column (see Figure 2-b). This application is based on the compari-
son between two images taken at different stages whilst the connection is deforming. One of
these images is referred as reference image and the other as a deformed image.

2.3 Material characteristics


Tensile tests were carried-out on steel coupons taken from beams, end-plates, and bolts. The
average mechanical characteristics are listed in Table 2. Cylinders of PVC and plywood have
been tested in compression (see Figure 3). For the two materials, the behaviour is elastic for
stress lower than 6-7 N/mm2 and become elasto-plastic for higher stress values with
a hardening and important permanent strains. The yield strength of PVC and plywood are
respectively equal to 4,4 and 6 N/mm2 and the elasticity module to 200 and 302 N/mm2.

3 MONOTONIC TESTS

The moment-rotation curves of connections obtained for different specimens are presented in
Figure 4. The rotation of the connection, θc,U, is obtained from horizontal LVDT, Ut1 and
Ub1. Furthermore, from the aforementioned curves, the following parameters are extracted
and provided in Table 3:

Figure 3. Compression tests on thermal insulation layer.

272
Figure 4. Moment-rotation curves during monotonic tests.

• the secant rotational stiffness Sj evaluated for 2/3Mj,pl,


• the plastic bending resistance Mj,pl calculated according to ECCS requirements,
• the ultimate bending resistance Mj,u corresponding to the maximum bending moment,
• elements that yield/crack before failure and the failure mode.
Most of specimens, except for thermal breaks attached to RHS without stiffeners, failed by bolt
ruptures in tension. For specimens RTPA-3-P/B, PA-6 and RTPA-6-P1 failure is due to weld
tearing at the corner. Excluding the specimens PA-3, the additional thermal insulation do not

Table 3. Bending resistances and failure modes for monotonic tests.


Sj Mj,pl Mj,u Failure mode Yielding/cracking
Specimen kNm/rad kNm kNm - -
PA-1 9880 44,6 73,8 Bolt Bolt, flange
RTPA-1-P 8106 32,6 66,6 Bolt Bolt, flange, PVC
RTPA-1-B 8624 34,8 71,2 Bolt Bolt, flange, Wood
PA-2 10195 40,0 60,7 Bolt Bolt, flange
RTPA-2-P 2880 12,2 50,4 Bolt Bolt, PVC
RTPA-2-P1 3196 29,2 50,7 Bolt Bolt, flange, PVC
RTPA-2-B 4266 15,4 48,1 Bolt Bolt, wood
PA-3 16860 32,8 67,6 Bolt Bolt, flange
RTPA-3-P 6540 38,2 66,4 Weld Bolt, weld, flange, PVC
RTPA-3-B 6364 41,0 58,5 Weld Bolt, weld, flange, wood
PA-4 19000 57,0 88,8 Bolt Bolt, flange, weld
RTPA-4-P 12914 42,7 85,0 Bolt Bolt, flange, PVC
RTPA-4-B 11973 44,5 85,0 Bolt Bolt, flange, wood
RTPA-5-P 13343 42,1 79,5 Bolt Bolt, flange, PVC, weld
PA-6 1425 7,5 14,6 Weld(1) Bolt, weld, flange
RTPA-6-P1 705 8,0 15,1 Weld(1) Bolt, weld, flange, PVC

Note 1: The test was stopped after major vertical displacements

273
influence the failure mode. For specimens PA-3, the single fillet weld that attach the end-plate
to the tube is bent until tearing. The additional insulating layer increases this phenomena.
The introduction of thermal insulation layer composed of PVC (10 or 20 mm thickness) or ply-
wood (30 mm thickness) decrease the ultimate resistance. For extended end-plate with stiffeners,
this decrease is limited, less than 5%. For non-stiffened end-plates, with I or RHS profile, this
decrease do not exceed 13% in the worst case. For flush end-plate connections, the reduction of
resistance goes to 15-20%. The geometry of the end-plate has a clear impact on the resistance of
thermal break particularly concerning the size of the potential compression area that can develop.
The thermal insulation being flexible compared to steel, an important compression area is needed
to equilibrate the tensile force due to the bending moment. In presence of flush end-plate, this
compression area is rather limited and the lever arm decreases also. The stiffeners permit to
favour the distribution of contact pressures between the flange and the thermal insulation.
The consequence of the geometry of the end-plate are also important for the rotational stiff-
ness that is divided by 2 when using thermal insulation layer with extended end-plate and by 3
for flush end-plate connections. This decrease is due to the crushing of the thermal insulation
that is more substantial in case the compression area is reduced.
The stiffeners improve the mechanical characteristics (rotational stiffness, plastic and ultim-
ate bending moments) of the thermal break as an extended end-plate with stiffeners and ther-
mal insulation exhibit the same mechanical characteristics as an extended end-plate without
stiffeners and thermal insulation.
Image correlation have been used to evaluate the relative displacement between the column
flange and the end-plate (see Figure 6) in order to estimate the crushing of the thermal insulation
and thus the size of the compression area. The deformed shape of specimens with and without
thermal insulation layer are quite different, particularly the size of the compressive area. For spe-
cimen PA-1, the bottom extended end-plate is in contact (see Figure 6-a) and uplift start around
the bottom flange of the beam. For specimens with thermal insulation, RTPA-1-P and RTPA-
1-B, the compression area extends over 30% of the beam depth. For flush end-plate connections
(see Figure 6-b), the results are completely different. For specimens with thermal insulating layer,
the compression develops over half the beam depth. For plywood and PVC of 20 mm thickness,
the deformed shape is almost linear even in the tensile area.

Figure 5. Specimens at failure.

274
Figure 6. Relative displacements of end-plate and column flange at failure.

Figure 7. Moment-rotation curves and failure of specimen RTPA-4-P.

4 CYCLIC TESTS

The cyclic tests were performed considering ECCS requirements, based on the incremental dis-
placements previously determined with monotonic tests. The failure modes and elements that
yield/crack during cyclic tests are listed in Table 4, the maximum bending moments obtained
during monotonic and cyclic tests, noted Mj,u,mon and Mj,u,cyclic respectively, are also added.

Table 4. Bending resistances and failure modes for cyclic tests.


Specimen Mj,u,mon Mj,u,cyclic Failure mode Yielding/cracking
kNm kNm - -
PA-1 73,8 74,1 Welds Bolts, flange, welds
RTPA-1-P 66,6 60,5 Bolts, welds Bolts, flange, welds, PVC
RTPA-1-B 71,2 62,7 Bolts, welds Bolts, flange, welds, wood
PA-2 60,7 55,9 Bolt Bolt flange
RTPA-2-P 50,4 50,1 Bolts Bolt, PVC
RTPA-2-P1 50,7 53,2 Bolts Bolt, PVC, flange
RTPA-2-B 48,1 50,9 Bolts Bolt, wood
RTPA-4-P 85 88,0 Bolts Bolt, flange, PVC, weld
RTPA-5-P 79,5 90,6 Bolts Bolt, flange, PVC, weld

275
Except for specimens PA-1 and RTPA-1-P/B, the failure modes are similar to that observed
during static tests and corresponds to bolt ruptures in tension. The resistance decreases in pres-
ence of extended end-plate un-stiffened but it is not necessarily the case with flush end-plate con-
nections and particularly with extended end-plate stiffened. For the latter, prying effect
disappear as the thermal insulation thickness reduce during the previous loading.

5 CONCLUSIONS

This paper give the results of an experimental campaign performed on thermal break of steel
beam attached to a steel support. This thermal break solution consists in inserting a PVC or ply-
wood layer between the end-plate of a steel beam and the steel support (column, beam, sleeve. . .).
Cantilever beams attached to steel column with thermal insulation layer were tested under mono-
tonic and cyclic loadings and the main parameters studied were the material of thermal insulation,
the bolt configuration (flush end-plate, extended end-plate with and without stiffeners), the type
of beam (I profile, RHS or SHS). The addition of thermal insulation composed of PVC of 20 mm
or plywood of 30 mm divide by two the rotational stiffness of the extended end-plate compara-
tively to a conventional connection. For flush end-plate, this rotational stiffness is divided by 3
except when using a PVC layer of 10 mm thickness. The addition of thermal insulation decrease
the ultimate bending due to the modification of the size of the compression area. This decrease is
limited by 13% with extended end-plate and drop to 20 % with flush end-plate. The related failure
modes were not modified by the addition of thermal insulation layer except for end-plate of RHS
where weld tearing develops after major yielding of end-plates in bending. An extended end-plate
with thermal insulation and stiffeners develops mechanical characteristics similar to that of an
extended end-plate without thermal insulation and stiffeners.

REFERENCES

Ben Larbi, A. Couchaux, M. & Bouchaïr, A. 2017. Thermal and mechanical analysis of thermal break
with end-plate for attached steel structures, Engineering Structures 131: 362–379.
CECM, Recommended testing procedures for assessing the behavior of structural elements under cyclic
loads, European Convention for Constructional Steelwork, Technical Committee 1, TWG 13 – Seismic
Design 45, 1986.
Cleary, D.B. Riddell, W.T. Camishion, N. Downey, P. Marko, S. Neville, G. Oostdyk, M. & Panaro, T.
Steel connections with fiber-reinforced resin thermal barrier filler plates under service loading, Journal
of Structural Engineering, 142, Issue 11. 2016
EN 1993-1-8, Eurocode 3: Design of steel structures. Part 1-8: Design of joints, 2005.
Hardock, D. & Roppel P. 2013. Thermal Breaks and Energy Performance in High-rise Concrete
Balconies, CTBUH Journal, Issue IV: 32–37.
Hamel, S. & White, S. 2016. Thermo-mechanical modelling and testing of thermal breaks in structural
steel point transmittances, Anchorage, AK.
Keo, P. Le Gac, B. Somja, H. & Palas, F. 2018. Experimental study of the behavior of a steel concrete
hybrid thermal break system under vertical actions, High tech concrete: where technology and engineer-
ing meet. Springer: 2573–2580.
Keo, P. Le Gac, B. Somja, H. & Palas, F. 2018. Low-cycle fatigue life of a thermal break system under
climatic actions, Engineering Structures, 168: 525–543.
Nasdala, L. Hohn, B. & Rühl, R. 2007. Design of end-plate connections with elastomeric intermediate
layer, Journal of Constructional Steel Research 63: 494–504.
Peterman, K. D. Kordas, J. A. Moradei, J. Coleman, K. D’Aloisio, J. A. Webster, M. D. & Hajjar, J. F.
2016. Thermal Break Strategies for Cladding Systems in Building Structures: Report to the Charles
Pankow Foundation, Boston, MA.
Peterman, K.D. Moradei, J. D’Aloisio, J. Webster, M. & Hajjar, J.F. 2019. The behaviour of double lap splice
bolted connections with fiber-reinforced polymer fills, Workshop Connection VIII, Boston, May 2016.
Sulcova, Z. Sokol, Z. & Wald, F. 2007. Structural connections with thermal separation, CESB 07 Prague
Conference: 672–677.

276
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Assessment of Eurocode fire design rules for structural members


made of high strength steels

Carlos Couto & Paulo Vila Real


RISCO - Department of civil engineering, University of Aveiro, Portugal

ABSTRACT: This paper investigates the accuracy of simple design rules of Eurocode 3 Part
1-2 (EN 1993-1-2) and FIDESC4 for columns and beams made of high strength steel (HSS).
The comparison is made against numerical results calculated from GMNIA using the present
material model and reduction factors from EN 1993-1-2, which was defined for mild strength
steel (MSS). Globally, satisfactory results are obtained with a reliability level close to what
would be obtained for MSS, if one considers that no specific changes are introduced, despite
the behaviour of HSS being essentially different from that of MSS. Nevertheless, for columns,
the flexural buckling curve of EN 1993-1-2 seems slightly unconservative, a trend also
observed for MSS, and for beams the results show that the predicted resistance from the
simple design rules may be too much on the safe sided and, for that reason, will lead to uneco-
nomical solutions, thus it should be further improved in the future.

1 INTRODUCTION

The demand for structural members made of high strength steel (HSS) is increasing throughout
the world. To reflect this, the future revision of Part 1-1 of the Eurocode 3 (CEN, 2018), devoted
to the design of steel structures, includes rules for steel with a yield strength up to 700 MPa and
this decision must be accompanied by rules for the case of fire to avoid a gap in the design codes.
For the time being, within the scope of CEN, a new version of Eurocode 3 Part 1-2 (EN
1993-1-2) is being prepared to replace the present one (CEN, 2005). Thus, the present paper
aims at contributing to the definition of the current reliability level of existing design rules and
their applicability to members made of HSS.
A review about previous research on HSS can be found in Ban and Shi (2018) and, more
particularly, about the mechanical properties of HSS for the case of fire, in Maraveas et al.
(2017). As noticed on these references, until now, the focus has been mainly in the character-
ization of the mechanical properties of HSS at elevated temperatures or in post-fire condi-
tions, and research on the structural behaviour of members made of HSS are still scarce.
Previous studies on columns (Chen and Young, 2008; Fang and Chan, 2018; Winful et al.,
2018), concluded that the EN 1993-1-2 design equations led to conservative results when com-
pared to numerical results, but different reduction factors for the mechanical properties and
a Ramberg-Osgood constitutive model were used in the FE-simulations, still no comparisons were
provided using the EN 1993-1-2 non-linear constitutive model. Moreover, it was concluded that
for slender section (Class 4) columns results were too much on the safe side. To overcome this,
Fang and Chan (2018) have analysed the use of a stress corresponding to a higher strain of 0.5%
and 1% instead of the 0.2% proof strength as design yield strength for steel concluding that it led
to better capacity prediction. The same conclusions were reached by Couto et al. (2014, 2015)
who observed that using the 0.2% proof strength led to lower capacity prediction for Class 4
cross-sections, and therefore developed and proposed proper effective width expressions together
with the use of the stress at 2% total strain, as done for other section classes, for improved results.
For beams, studies for the case of fire are scarce. Varol and Cashell (2017) have performed
a numerical study to assess the accuracy of Eurocode buckling curves for beams, using EN

277
1993-1-2 constitutive model in FE-simulations, and concluded that said buckling curves were
not satisfactory in their format for high strength steel beams, both at room and elevated tem-
perature and more severe buckling curves were proposed by the authors. The numerical inves-
tigation from Shakil et al. (2018) shown that the HSS beams have greater strength reserve
compared to mild strength counterparts, before runaway failure due to plastic mechanism,
with a similar observation for a single storey frame. Thus, higher critical temperatures were
reached when compared to those obtained for a similar frame made of mild strength steel.

2 HIGH STRENGTH STEELS AT ELEVATED TEMPERATURES

2.1 Reduction in mechanical properties


The supplementary rules for the extension of EN 1993 up to steel grades S700 (CEN, 2009)
states that EN 1993-1-2 standard is applicable to steels with grades greater than S460 up to
S700 without further additional rules. On this, previous research on HSSs at elevated tempera-
ture, or post-fire conditions, generally have shown that the mechanical properties differ from
that of conventional mild steels (Ban and Shi, 2018). In terms of the reduction in strength and
stiffness with temperature, the EN 1993-1-2 equations are generally conservative when com-
pared to said experimental results, as shown in Figure 1. The results on this figure might be
found elsewhere, see Ban and Shi (2018) and Maraveas et al. (2017).
Regarding the constitutive model, as abovementioned, some authors rely on a Ramberg-
Osgood model to define the stress-strain relationship of HSS at elevated temperatures as
better agreement to the experimental results seems to be obtained when compared to the
material model defined in EN 1993-1-2 (see Figure 3). However, in this paper, the EN 1993-
1-2 model is considered, but the effect of other constitutive models in the accuracy of simple
design models ought to be study in future.

Figure 1. Comparison between the reduction factors for steel mechanical properties and equations from
EN 1993-1-2.

278
2.2 Material imperfections (residual stresses)
The distribution of residual stresses in HSS members is generally similar to that of mild steel
counterparts (Ban and Shi, 2018), however, different patterns are found in the literature
(Couto and Vila Real, 2019) resulting from test measurements, in particular, for welded sec-
tions, it may lead to some differences in capacity prediction by the numerical models. Never-
theless, a common aspect observed by different studies (Ban et al., 2012, 2013; Ban and Shi,
2018) is that the ratios of magnitudes between the tensile and compressive residual stresses to
the yield strength is lower than that of mild steels, for instance, in the studies of Ban et al.
(2013) a maximum 460 MPa (tensile at weld and flame cut edges) and 288 MPa (compressive)
residual stress was considered for the S960 steel which has an yield strength of 960 MPa. For
the fire situation, Wang and Qin (2016) have measured the residual stresses in welded thin-
walled H-section before and after heating and cooling the specimens and demonstrated that
the magnitude of residual stresses decreased under such circumstances. However, in this
regard, Franssen (1996a) argues that residual stresses are an initial stresses and must be mod-
elled as such and therefore are not dependent on the heating conditions, and history, reached
during a fire.
Following this latter approach, for hot-rolled cases studied in this paper, the maximum differ-
ence in the ultimate strength obtained in the numerical simulations, between considering or not
the residual stresses, was about 6%, while, for the welded cases, this difference increased to 20%.

3 NUMERICAL MODEL

The software packages SAFIR (Franssen and Gernay, 2017) and ANSYS (2018) were used to
obtain the numerical results with the Finite Element Method by performing non-linear analysis to
assess the ultimate strength of the members at elevated temperatures. After a sensitive analysis
a mesh with maximum area of 0.001 m2 for each shell element was chosen for the models. Fork-
supports were used at both ends of the structural member and the loading was applied using
nodal forces. Additionally, end-plates were included with a thickness of 10 times the web thickness
to ensure correct load distribution. To allow the comparison with simple design rules, the tem-
perature was considered uniform in the member. The numerical model is shown in Figure 2.
Residual stresses and geometrical imperfections were included in the numerical models. For the
residual stresses, they were defined as initial state conditions consisting on the patterns given in
Figure 4. For the geometrical imperfections, both local (or plate) and global (or member) imper-
fection’s shape were considered affine to those given by a Linear Buckling Analysis (LBA).
A global imperfection amplitude of 80% of L/750 according to EN 1090-2:2011 (CEN, 2011) and
following the recommendations of EN1993-1-5 Annex C was considered. The plate imperfections
were defined according to EN 1090-2:2011 (CEN, 2011), as used by Couto and Vila Real (2019).
The steel S690 was considered in this study, with the constitutive law defined according to
EN 1993-1-2, as shown in Figure 3. At ambient temperature, the steel has a yield strength of
690 MPa, a Young modulus of 210 GPa and a Poisson ratio of 0.3.

Figure 2. Detail of the numerical model.

279
Figure 3. EN 1993-1-2 constitutive law.

Figure 4. Pattern of the residual stresses considered in this study.

4 CASES STUDIED

Eight commercial cross-sections were considered as shown in Table 1. The classification of


flange and web, for the case of fire and S690 steel, is also given for loading in compression or
major-axis bending. In comparison to normal temperature classification (according to
EN1993-1-1), also given in the table, one notices the more stringent classification limits at ele-
vated temperatures where, for example, the HE 220 B profile is Class 2 at normal temperature
and Class 3 for the case of fire.
Also, eight welded cross-sections were considered, in this study, with a geometry and classi-
fication given in Table 2.
In welded sections, plate imperfections were always considered. In commercial sections,
only cases with Class 4 sections were analysed with plate imperfections included.

5 BUCKLING OF COMPRESSED MEMBERS

Figures 5–7 show the results for columns with hot-rolled profiles HE 160 B and IPE 220 and
welded cross-section I600 × 8 + 180 × 15. Both strong- and weak-axis buckling are plotted. In
each figures, the cross-sectional resistance was calculated according to EN 1993-1-2 (CEN,
2005) for Class 1 to 3 profiles and according to FIDESC4 proposal for Class 4 profiles (see
Tables 1 and 2). The latter corresponds to the equations proposed by Couto et al. (2015)
which are likely to be adopted in the future code revision.
Accordingly, it is observed that the buckling curve defined in EN 1993-1-2 might slightly over-
predict the capacity of columns made of HSS for a certain range of relative slenderness. However,

280
Table 1. Commercial cross-sections and classification according to EN 1993-1-2* in S690 steel.
Class in compression Class in Major-axis bending

Commercial designation Flange Web Flange Web

HE 160 B 2 (1) 1 (1) 2 (1) 1 (1)


HE 220 B 3 (2) 1 (1) 3 (2) 1 (1)
HE 300 B 3 (3) 3 (1) 3 (3) 1 (1)
HE 280 A 4 (4) 4 (3) 4 (4) 1 (1)
IPE 100 1 (1) 2 (1) 1 (1) 1 (1)
IPE 220 1 (1) 4 (4) 1 (1) 1 (1)
IPE 240 1 (1) 4 (4) 1 (1) 1 (1)
IPE 400 2 (1) 4 (4) 2 (1) 2 (1)

* section class at normal temperature according to EN 1993-1-1 is given within brackets.

Table 2. Welded cross-sections and classification according to EN 1993-1-2* in S690 steel.


Class in compression Class in Major-axis bending

Designation (I hw × tw + b × tf)** Flange Web Flange Web

I600 × 6 + 180 × 10 4 (4) 4 (4) 4 (4) 4 (4)


I600 × 8 + 180 × 15 3 (2) 4 (4) 3 (2) 4 (4)
I600 × 10 + 300 × 35 1 (1) 4 (4) 1 (1) 3 (3)
I600 × 30 + 300 × 35 1 (1) 3 (3) 1 (1) 1 (1)
I450 × 5 + 150 × 15 2 (1) 4 (4) 2 (1) 4 (4)
I450 × 6 + 150 × 10 4 (3) 4 (4) 4 (3) 4 (4)
I450 × 10 + 150 × 30 1 (1) 4 (4) 1 (1) 3 (2)
I450 × 25 + 180 × 30 1 (1) 2 (1) 1 (1) 1 (1)

* section class at normal temperature according to EN 1993-1-1 is given within brackets.


** hw and tw are the web height and thickness; b and tf are the flange width and thickness.

Figure 5. Results for HE 160 B (Class 2) profile for buckling in strong-axis (left) and weak-axis (right).

such behaviour has been observed for mild steel columns by other authors, as well, thus the same
reliability level was considered to have been attained. In fact, the capacity of columns experimen-
tally tested in fire is known to be higher than that obtained numerically, as observed by Franssen
et al. (1995,1996b) during an European research project (Schleich et al., 1998).

281
Figure 6. Results for IPE 220 (Class 4) profile for buckling in strong-axis (left) and weak-axis (right).

Figure 7. Results for the welded profile I600×8+180×15 (Class 4) for buckling in strong-axis (left) and
weak-axis (right).

6 LATERAL-TORSIONAL BUCKLING

Figures 8–10 show the results for laterally unrestrained beams with different cross-sections
where lateral-torsional buckling is the main failure mode. As for the previous section, in each
figures, the cross-sectional resistance was calculated according to EN 1993-1-2 (CEN, 2005)
for Class 1 to 3 profiles and according to FIDESC4 proposal for Class 4 profiles (see Tables 1
and 2). The buckling curves defined as FIDESC4 are based on the original proposal of Couto
et al. (2016) which was developed within the European Research Project FIDESC4.
Results show that the simple design methods lead to conservative predictions of the lateral-
torsional buckling resistance of the analysed members. It was observed that, for profiles with
a height-to-width ratio of h/b > 2, results might be too conservative, however neither EN
1993-1-2 nor FIDESC4 methods make a distinction based on this ratio. These cases should be
further investigated in the future.

7 OVERALL EVALUATION

Figure 11 compares the reduction factor for the instability phenomena for the cases analysed
in this paper, either flexural buckling (1391 total cases) or lateral-torsional buckling (791 total
cases). Value of χMethod plotted in the ordinate refers to the simple design value calculated

282
Figure 8. Beam results for IPE 220 (left) and HE 160 B (right), Class 1 and 2 profiles, respectively.

Figure 9. Beam results for a welded Class 3 profile I450 × 10 + 150 × 30 (left) and hot-rolled Class 4 pro-
file HE 280 A (right).

Figure 10. Beam results for two welded Class 4 profiles I600 × 8 + 180 × 15 (left) and I600 × 6 + 180×10
(right).

with EN 1993-1-2 or FIDESC4 proposal for Class 1-3 and Class 4 members respectively,
while χGMNIA plotted in the abscissa refers to the numerical result, given as the ratio between
the ultimate strength (axial force or bending moment) and the EN 1993-1-2 or FIDESC4
cross-sectional resistance.

283
Figure 11. Comparison of χMethod vs. χGMNIA for flexural buckling (left) and lateral-torsional buckling
(right).

The ratio χMethod/χGMNIA yields a mean value of 0.99 for flexural buckling and 0.87 for lat-
eral-torsional buckling with a respective standard deviation of 0.077 and 0.104. For the flex-
ural buckling cases, 94% of all predictions are within ± 15% of the numerical results, while for
lateral-torsional buckling, only 64% of the cases are within these boundaries, being, in this
case, most of the results on the safe side.

8 CONCLUSIONS

In this paper the EN 1993-1-2 and FIDESC4 provisions for columns and beams made of HSS
were evaluated in terms of their ability to predict the ultimate capacity when compared to
numerical simulations.
Although some question remain open, for instance, the adequacy of present reduction fac-
tors for the mechanical properties, the constitutive law of HSS in fire, the influence of residual
stresses in the buckling resistance of members made of HSS, the consistency of the present
rules for hybrid members constructed from MSS and HSS plates, globally, the results have
shown good agreement between the capacity predicted by the simple design rules when com-
pared to results from numerical simulations.
For the columns, both the weak- and strong-axis failure modes could be better covered by
a slightly more severe buckling curve. In particular, for the relative slenderness ranges between
0.75 and 1.25 the capacity was overpredicted when compared to the numerical results, how-
ever, the same reliability level as for MSS columns was obtained. This is generally considered
as acceptable because the capacity of columns experimentally tested at elevated temperatures
is known to be higher than that obtained numerically.
For the beams, it was observed that the design rules might underestimate the ultimate cap-
acity and, thus, lead to uneconomical solutions. Therefore, it is recommended that future
studies on this subject be conducted to improve such situation.

ACKNOWLEDGMENTS

This research work was funded by the Portuguese Government through the FCT (Fundação
para a Ciência e a Tecnologia) under the Post-doc grant SFRH/BPD/114816/2016 awarded to
the first author.

284
REFERENCES

ANSYS® (2018) ‘Academic Research Mechanical, Release 18.2ʹ.


Ban, H. et al. (2012) ‘Overall buckling behavior of 460 MPa high strength steel columns: Experimental
investigation and design method’, Journal of Constructional Steel Research.
Ban, H. et al. (2013) ‘Experimental investigation of the overall buckling behaviour of 960 MPa high
strength steel columns’, Journal of Constructional Steel Research.
Ban, H. and Shi, G. (2018) ‘A review of research on high-strength steel structures’, Proceedings of the
Institution of Civil Engineers - Structures and Buildings.
CEN (2005) ‘EN 1993-1-2, Eurocode 3: Design of steel structures - Part 1-2ʹ.
CEN (2009) ‘EN 1993-1-12, Eurocode 3 - Design of steel structures - Part 1-12ʹ.
CEN (2011) ‘EN 1090-2 : 2008 + A1 Execution of steel structures and aluminium structures - Part 2 :
Technical requirements for steel structures’.
CEN (2018) ‘prEN 1993-1-1, Eurocode 3 — Design of steel structures — Part 1-1ʹ.
Chen, J. and Young, B. (2008) ‘Design of high strength steel columns at elevated temperatures’, Journal
of Constructional Steel Research.
Couto, C., Vila Real P., Lopes N., Zhao B. (2014) ‘Effective width method to account for the local buck-
ling of steel thin plates at elevated temperatures’, Thin-Walled Structures.
Couto, C., Vila Real P., Lopes N., Zhao B. (2015) ‘Resistance of steel cross-sections with local buckling
at elevated temperatures’, Journal of Constructional Steel Research.
Couto, C., Vila Real P., Lopes N., Zhao B. (2016) ‘Numerical investigation on the lateral torsional buck-
ling of beams with slender cross-sections at elevated temperatures’, Engineering Structures.
Couto, C. and Vila Real, P. (2019) ‘Numerical investigation on the influence of imperfections in the local
buckling of thin-walled I-shaped sections’, Thin-Walled Structures.
Fang, H. and Chan, T.-M. (2018) ‘Axial compressive strength of welded S460 steel columns at elevated
temperatures’, Thin-Walled Structures.
Franssen, J.-M., Schleich, J.-B. and Cajot, L.-G. (1995) ‘A simple model for the fire resistance of axially-
loaded members according to Eurocode 3ʹ, Journal of Constructional Steel Research.
Franssen, J.-M. (1996a) ‘Residual stresses in steel profiles submitted to the fire: an analogy’, in Proc. 3rd
CIB/W14 FSF workshop on modelling. TNO Building and construction research.
Franssen, J.-M. et al. (1996b) ‘A simple model for the fire resistance of axially loaded members—com-
parison with experimental results’, Journal of Constructional Steel Research.
Franssen, J. and Gernay, T. (2017) ‘Modeling structures in fire with SAFIR ® : Theoretical background
and capabilities’, Journal of Structural Fire Engineering.
Neuenschwander, M., Scandella C., Knobloch M., Fontana M. (2017) ‘Modeling elevated-temperature
mechanical behavior of high and ultra-high strength steels in structural fire design’, Materials &
Design.
Maraveas, C., Fasoulakis, Z. C. and Tsavdaridis, K. D. (2017) ‘Mechanical properties of High and Very
High Steel at elevated temperatures and after cooling down’, Fire Science Reviews.
Schleich, J.-B. et al. (1998) Buckling curves of hot rolled H steel sections submitted to fire, C.E.C. Agree-
ment No 7210-SA/316/515/618/931.
Shakil, S., Lu, W. and Puttonen, J. (2018) ‘Response of high-strength steel beam and single-storey frame
in fire: Numerical simulation’, Journal of Constructional Steel Research.
Varol, H. and Cashell, K. A. (2017) ‘Numerical modelling of high strength steel beams at elevated tem-
perature’, Fire Safety Journal.
Wang, W. and Qin, S. (2016) ‘Experimental investigation of residual stresses in thin-walled welded H-sec-
tions after fire exposure’, Thin-Walled Structures.
Winful, D., Cashell K.A., Afshan S., Barnes A.M., Pargeter R.J. (2018) ‘Behaviour of high strength steel
columns under fire conditions’, Journal of Constructional Steel Research.

285
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Numerical investigation on thin-walled CFS columns in fire

Hélder D. Craveiro, José Henriques, Aldina Santiago & Luís Laím


ISISE – Institute for Sustainability and Innovation in Structural Engineering, University of Coimbra,
Portugal

José Henriques
CERG - Construction Engineering Research Group, Faculty of Engineering Technology,
University of Hasselt, Belgium

ABSTRACT: The structural performance of cold-formed steel compressed elements sub-


jected to fire is investigated using finite element analysis. The numerical study is based on a set
of experimental tests on cold-formed steel columns subjected to fire with restrained thermal
elongation. On the fire resistance tests, different levels of restraint and load were considered.
This allowed studying the influence of the generated internal forces during heating in the over-
all structural fire behavior of the cold-formed steel columns. Specifically, in this study, two
cross-section shapes are investigated, namely C-channels and Σ-channels (sigma). Based on
experimental results the numerical models were developed and calibrated using all mechanical
and thermal properties determined experimentally at both ambient and elevated temperatures.
To further investigate the influence of some key parameters, such as initial applied load level,
level of restraint to thermal elongation and slenderness, a parametric study is presented. The
level of restraint to thermal elongation will significantly influence the critical temperature of
the cold-formed steel columns.

1 INTRODUCTION

The market share of light steel framing has increased in the past few decades and is now-
adays a widely used solution for residential and mid-rise buildings. Its increasing popular-
ity in the building construction industry is due to several advantages over other more
traditional materials, such as lightness, high strength-t-weight ratio and ease of fabrication
and erection (Craveiro, HD, et al. (2014)). However, thin-walled sections also present
some disadvantages related to the high susceptibility to local sectional buckling phenom-
ena and low fire resistance. It is well known that the fire resistance of unprotected cold-
formed steel structures is less than 30 minutes and that fire protection is required to fulfil
the fire resistance requirements (Craveiro, HD (2015).
The great majority of investigations on cold-formed steel elements subjected to fire were
conducted using the steady state method (Shanshan Cheng, et al. (2015)) of mainly
focused on the study of individual buckling modes at elevated temperatures, without con-
sidering possible interactions between global, local and distortional buckling (Ranawaka,
Mahendran (2009 and 2010)).
Currently, the design methods available in the EN 1993-1-2 (2005) for hot-rolled steel
member are also applicable to CFS members with class 4 cross-sections, establishing the same
reduction factors for the yield strength and limiting the critical temperature to 350ºC. Based
on extensive research it was found that the reduction factors for class 4 sections are not suit-
able for cold-formed steels (Craveiro, HD, et al. (2016)). Some studies have shown that ambi-
ent temperature design methods can be used but adequate mechanical properties at an
elevated temperature should be considered (Shanmuganathan Gunalan, et al. (2015)).

286
The influence of the surrounding structure on the overall behaviour of cold-formed steel
structural elements has been neglected, and only for hot-rolled steel members, this type of
investigations have been undertaken regularly. Consequently, it is assumed that no internal
forces are generated due to restraint to thermal elongation during the fire exposure. However,
the forces generated during heating due to restraint to thermal elongation can play a relevant
role in the fire resistance of a cold-formed steel column, causing premature collapse (Craveiro,
HD. (2015), Ali, F. (2001), Franssen, J.M. (2000)).
In this paper, an investigation on the behavior of individual cold-formed steel lipped chan-
nels and Sigma channels subjected to fire with restrained thermal elongation is presented. The
experimental investigation is briefly presented as well as the numerical models developed. The
experimental results were used for calibration purposes. After calibration, the numerical
models were used to conduct a parametric study, in order to assess the influence of several
parameters on the overall behavior of cold-formed steel columns in fire with restrained ther-
mal elongation. The obtained experimental and numerical results were compared with design
predictions according to the EN 1993-1-2 (2005).

2 EXPERIMENTAL TESTS

2.1 Tested column


In this investigation, two cross-section shapes were investigated, namely lipped channels and
Sigma channels. In Figure 1 a schematic representation of the tested cross-sections is pre-
sented. The lipped channel profiles were fabricated with S280GD+Z steel and the Sigma pro-
files were fabricated with S320GD+Z steel. The length of the profiles was 2950 mm.

2.2 Test set-up


The experimental set-up comprised a two-dimensional (2D) reaction steel frame ((1) in Figure 2 a)
and b)) and a three dimensional (3D) restraining steel frame adaptable for different levels of stiff-
ness ((2) in Figure 2 a) and b)) in order to simulate the axial and rotational restraint imposed by
the surrounding structure to a CFS column in fire. This restraining system intended to reproduce
the actual boundary conditions of a CFS column when inserted in a real building.

Figure 1. Cross-section details. a) Lipped channel. b) Sigma channel.

287
Figure 2. Schematic representation of the test set-up for fire resistance tests with restraint to thermal
elongation.

The connections between the peripheral columns and the top beams of the restraining frame
were made with threaded rods. A hydraulic jack, placed in the 2D reaction frame, was used
to apply the serviceability load ((3) in Figure 2 a) and b)). The thermal action was applied
by a vertical modular electric furnace ((4) in Figure 2 b)) programmed to reproduce the
standard fire curve ISO 834 (1999). When the two parts of the furnace are closed a chamber
(2.5 m × 1.5 m × 1.5 m) was created around the test column.
To measure the restraining forces generated on the testing column during the heating pro-
cess a special device was built ((5) in Figure 2 a)), consisting of a hollow steel cylinder where
a stiff steel cylinder Teflon (PTFE) lined slides through it. On the top of the stiff steel cylin-
der, a 500 kN load cell was placed and compressed against the top end plate of the hollow
steel cylinder.

2.3 Test procedure


In each experimental test, the service load (30 and 50% Nb,Rd) was applied to the CFS columns
using the hydraulic jack fixed to the reaction frame and with the nuts of the threaded rods loos-
ened. This means that the orthogonal top beams of the restraining frame had free vertical move-
ment, guaranteeing that the compressive load was totally transferred to the CFS column. The
applied service load was controlled by a load cell placed between the top beams of the restrain-
ing frame and the hydraulic jack. Reaching the serviceability load, the nuts of the threaded rods
of the peripheral columns to top beams connections were tightened and from that moment the
restraining frame started to impose axial and rotational restraint to the CFS column being
tested in fire. During the entire test, the initial applied load was kept constant.
The thermal action was applied by a vertical modular electric furnace ((4) in Figure 2 a) and b))
programmed to reproduce the standard fire curve ISO 834. With heating the CFS column started
to expand due to thermal elongation of steel and since the column was restrained additional axial
forces were generated. The generated restraining forces were measured by a load cell. At the same
time and with temperature increase the mechanical properties of the steel degraded. After reaching
a maximum the column cannot withstand additional axial forces and the generated restraining
forces started to decrease, reaching once again the initially applied serviceability load (failure
criteria).

288
3 FINITE ELEMENT MODELLING

3.1 Numerical model


The experimental tests were reproduced using the finite element method and finite element
software Abaqus (2017). The objective was to develop a finite element model (FEM) able to
provide accurate strength estimations for cold-formed steel columns in fire. The modelling
was divided into three stages, firstly a heat transfer analysis was undertaken to calibrate
temperature evolution in the cross-section, secondly, an eigenvalue buckling analysis was
undertaken and then a non-linear static analysis to estimate the load bearing capacity and
failure modes under fire situation. The buckling modes of interest are used to input the geo-
metric imperfections in the nonlinear analysis. The magnitude for global imperfections was
L/1000, for distortional was the thickness of the profile (t) and for local imperfections h⁄200
(Duarte Oliveira (2016)). For the heat transfer analysis, two types of surfaces were defined,
namely “film condition” and “radiation to ambient” corresponding respectively to heat
transfer by convection and radiation.
The finite element type selected for this investigation was the S4R which is a four-node gen-
eral-purpose shell, quadrilateral and stress/displacement shell element with reduced integra-
tion, a large strain formulation, hourglass control and a first-order interpolation. The mesh
size used in this investigation was 10 mm.
The material model was created based on material testing also conducted in the scope of
this investigation (Craveiro, H.D., et al. (2016)). In Figure 3 the true stress and logarithmic
plastic strain curves used as input for the S280GD+Z steel are depicted.
To reproduce the behavior of CFS columns under simulated fire conditions with restraint
to thermal elongation the surrounding structures used in the experimental tests were replaced
by linear springs (3 and 13 kN/mm) connected to the centroid of the column to be simulated.
As previously mentioned, and just like in the experimental tests the serviceability loading was
applied on the columns before the thermal action. Then the determined temperature profiles
were used in the heating phase. The serviceability loading was kept constant throughout the
simulation. The additional forces are generated due to the action of the axial spring imposing
axial restraint to thermal elongation.

3.2 Calibration of the numerical models


The validation of the finite element model developed to reproduce the behavior of
CFS columns under fire condition with restraint to thermal elongation consisted of
comparing the evolution of the non-dimensional ratio between the generated restrain-
ing forces during fire tests with the initial applied service load (P/P0) as a function of

Figure 3. True stress and logarithmic plastic strain curves as a function of temperature used as input in
the finite element model.

289
the mean temperature of the CFS column ( θc ). In Figures 4 and 5 the comparison
between experimental and FEA results is presented in order to show the accuracy of
the finite element model for both tested cross-sections.

Figure 4. Comparison between experimental and numerical results for the lipped channel column with
pinned and semi-rigid boundary conditions and with different levels of axial restraint (Helder Craveiro, 2015).

Figure 5. Comparison between experimental and numerical results for the Sigma column in fire
(Mariana Ferreira, 2018).

290
Observing the obtained results in Figures 4 and 5 the influence of axial restraint to thermal
elongation is clear. Increasing the level of restraint led to an increase in the generated axial
forces during the fire resistance test. Analyzing the results for the lipped channel columns,
considering pin-ended support conditions and the lowest level of stiffness of the surrounding
structure to the CFS column (3 kN/mm of axial restraint), the Pmax =P0 ratio is 1.987 times the
value of the initial applied load for 50% load level. Analyzing the results obtained for the high-
est level of stiffness of the surrounding structure to the CFS column (13 kN/mm of axial
restraint) it was found out that the mean Pmax =P0 ratio is 2.262 times the initial load for the
50% load level which is significantly higher than those obtained for the lowest axial restraint
level. For the semi-rigid end-support condition the Pmax =P0 ratio was 2.473 and 3.702, respect-
ively for the lowest and highest axial restraint levels. Regarding critical temperatures, it as
observed that for the pinned support condition the critical temperatures were very similar for
both axial restraint levels, however for the semi-rigid support condition the critical temperat-
ures reduced from 453ºC to 415ºC, increasing the level of restraint to thermal elongation.
For the Sigma cold-formed steel column, considering an axial restraint of 30 kN/mm the
Pmax =P0 is 1.54 times the initial applied load.
Generally, it can be stated that the developed finite element model is able to accurately
reproduce the behavior of CFS columns under simulated fire conditions with restraint to ther-
mal elongation.

4 PARAMETRIC STUDY DISCUSSION

Adopting the developed and calibrated numerical models a parametric study was undertaken to
further understand the behavior of cold-formed steel columns in fire with restrained thermal
elongation. Several parameters were assessed such as thicknesses (1.5, 2.5 mm), web heights
(150 and 250 mm), lengths (1000, 1500, 2000, 2500, 3000 mm), initial load level (serviceability
load; 30, 50, 70% Nb;Rd ), boundary conditions and also different axial restraint ratios, relating the
axial stiffness of the surrounding structure to the column with the actual axial stiffness of the
cold-formed steel column (ka =ka;c ). This particular parameter is of outmost importance on the
overall behavior of cold-formed steel columns in fire since it will dictate the magnitude of the gen-
erated axial forces due to restraint to thermal elongation. In Figure 6 the influence of the axial
restraint ratio is depicted for a lipped channel profile (C 150  43  15  1:5 mm) with 3 m,
assuming semi-rigid end support conditions and a serviceability load of 50% Nb;Rd .
Observing the obtained results, the level of restraint to thermal elongation plays a very rele-
vant role in the overall behaviour of cold-formed steel columns in fire. Clearly, a critical tem-
perature reduction is observed when the level of restraint increases, however, the critical

Figure 6. P/P0 ratio vs θcr for different axial restraint ratio, for a 50% load level considering a semi-rigid
end support condition.

291
Figure 7. P/P0 ratio vs critical temperature as a function of the initial load level, assuming a axial
restraint level of 30 kN/mm.

temperature tends to a limit value, independent of the level of axial restraint to thermal elong-
ation. For instance, for a ratio of 5% and above the critical temperature reduction is negligible.
For the Sigma shaped profile similar type of investigation is still being undertaken. Specific-
ally, in terms of the influence of the initial load level applied to the column (Figure 7). As
expected for lower initial load levels higher P=P0 ratios are found for the same level of axial
restraint to thermal elongation. Moreover, increasing the initial load level led to a significant
decrease in the observed critical temperature.

5 COMPARISON WITH DESIGN PREDICTIONS

The suitability of the current design methods was assessed by comparing the FEA results with the
design predictions according to the EN 1993-1-2. The assessment was undertaken for both cross-
section shapes, considering the pinned support condition and different levels of restraint. In
Figure 8 (Laim, L. et al. (2018)) and 9 the obtained results for both cross-sections are depicted.
The obtained results have shown that the current design methodologies are not suitable for
cold-formed steel compression elements subjected to fire with restrained thermal elongation.
Particularly, it can be seen in Figure 8 that the code predictions may be slightly unsafe (for pin-
ended columns with no axial restraints) and over non-conservative (for pin-ended columns with
a 0.20 axial restraint ratio) whatever the load level and the column slenderness. For example,
when the methods established in EN 1993-1.2:2005 are used for CFS C columns under a 50%
load level the estimated critical temperatures may have an average non-conservative error equal
to 8% (with a maximum error equal to 14%) for non-restrained columns and an average non-
conservative error equal to 80% for columns with a 0.20 axial restraint ratio. Last but not least,
such code predictions fit better for the Sigma channels, with the estimated critical temperatures
being on the safe side for the unrestrained members (Figure 9b).

6 CONCLUSION

The main conclusion is related to the fact that the currently available design guidelines
in the EN 1993-1-2 are not suitable to properly design cold-formed steel compression
elements in a fire situation, especially considering the influence of restraint to thermal
elongation, which may significantly reduce the critical temperature and consequently
lead to premature collapse.

292
Figure 8. Comparison between design predictions and FEA results for the lipped channel columns
assuming different load levels (Laim, L. et al. (2018)).

Figure 9. Comparison between design predictions and FEA results for the Sigma shaped cold-formed
steel columns, for the 50% load level and different levels of axial restraint.

This is ongoing research and the authors intend to conduct extensive parametric studies on
this subject, aiming to develop new/improved calculation methods for the design of cold-
formed steel columns in fire with restrained thermal elongation.

293
ACKNOWLEDGEMENTS

The authors gratefully acknowledge to the Portuguese Foundation for Science and Technol-
ogy for its support under the framework of the research projects PTDC/ECM/116859/2010
(FireColdFSteel), POCI-01-0145-FEDER-031435 (FireImprov_ColdFSteel) and POCI-01-
0145-FEDER-031858 (INNOCFSCONC).

REFERENCES

Duarte Nuno Ferreira de Oliveira (2016). Numerical Analysis of Compressed Cold-Formed Steel Elem-
ents subjected to Ambient and High Temperatures, MSc Thesis in Civil Engineering, Faculty of Sci-
ences and Technology, University of Coimbra, Portugal (in Portuguese).
EN 1993-1-2 (2005), Design of steel structures. General rules. Structural fire design. European committee
for standardization, Brussels, Belgium.
Faris Ali, David O’Connor (2001), Structural performance of rotationally restrained steel columns in
fire. Fire Safety Journal, 36, pp. 679–691.
Ferreira, Mariana Afonso Correia de Queirós (2018), Parâmetros que afetam o comportamento ao fogo
de colunas com perfis em sigma de aço enformados a frio, MSc Theis in Civil Engineering, Depart-
ment of Civil Engineering, University of Coimbra, Portugal.
Helder D. Craveiro, João P.C. Rodrigues, Luís Laím (2014), Cold-formed steel columns made with open
cross-sections subjected to fire, Thin-Walled Structures, 85, pp. 1–14.
Helder D. Craveiro, João P. Rodrigues, Aldina Santiago, Luís Laím (2016), Review of the high tempera-
ture mechanical and thermal properties of the steels used in cold-formed steel structures – The case of
the S280GD+Z steel, Thin-Walled Structures, Vol. 98, Part A, pp. 154–168.
Helder David da Silva Craveiro (2016), Fire resistance of cold-formed steel columns. Thesis for the
degree of Doctor of Philosophy (PhD) in Fire Safety Engineering, Department of Civil Engineering,
Faculty of Sciences and Technology, University of Coimbra, Portugal.
Jean-Marc Franssen (2000), Failure temperature of a system comprising a restrained column submitted
to fire, Fire Safety Journal, 34, pp. 191–207.
Laím, L., Rodrigues, J.P.C., Gardner, L. (2018), Numerical parametric study of cold-formed steel
C-shaped columns exposed to fire, In: Proceedings of the 10th International Conference on Structures
in Fire (SiF 2018), Belfast, UK, pp. 695–702.
Ranawaka, T., Mahendran, M. (2009), Distortional buckling tests of cold-formed steel compression
members at elevated temperatures, Journal of Constructional Steel Research, Vol. 65, pp. 249–259.
Ranawaka, T., Mahendran, M. (2010), Numerical modelling of light gauge cold-formed steel compres-
sion members subjected to distortional buckling at elevated temperatures, Thin-Walled Structures,
Vol. 48, pp. 334–344.
Shanshan Cheng, Long-Yuan Li, Boksun Kim (2015), Buckling analysis of partially protected
cold-formed steel channel section columns at elevated temperatures, Fire Safety Journal, 72, pp. 7–15.
Shanmuganathan Gunalan, Yasintha B. Heva, Mahen Mahendran (2015), Local buckling studies of
cold-formed steel compression members at elevated temperatures, Journal of Constructional Steel
Research, 108, pp. 31–45.

294
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

FEM analysis of the buckling behavior of thin-walled


CFS columns. Part I—Channel (C) and Double Channel (I)
cross-sections

Hélder D. Craveiro, L.S. da Silva & J.P. Martins


ISISE – Institute for Sustainability and Innovation in Structural Engineering,
University of Coimbra, Portugal

José Henriques
CERG - Construction Engineering Research Group, Faculty of Engineering Technology,
University of Hasselt, Belgium

ABSTRACT: One of the main advantages of thin-walled CFS profiles is the “freedom” the
designer has to conceive a cross-section shape tailored for the structural application. Conse-
quently, different cross-sections shapes have been developed and may be found in the market. Fur-
thermore, the combination of thin-walled CFS profiles, in the form of built-up cross-section
members, further increases the range of available options. On the other hand, this diversity of solu-
tions brings a challenge to the designer, as for some uncommon cross-section shapes, the mechan-
ical behavior is not entirely known and therefore not covered in the design codes. Hence, in this
paper, the buckling behavior of thin-walled CFS columns is investigated by means of FEM. The
numerical models were developed and calibrated using experimental results on lipped channels
and open built-up I cross-sections. Upon validation, a parametric analysis is performed, to further
assess the influence of different parameters, such as steel thickness, material grade, cross-section
dimensions, boundary conditions and connection spacing, in the case of the built-up cross-section.

1 INTRODUCTION

The demand for structures using cold-formed steel (CFS) sections has increased significantly
since it has been recognized as a valuable and competitive structural solution. One of the main
advantages of thin-walled CFS profiles is the “freedom” the designer has to conceive a cross-
section shape tailored for the structural application. Consequently, different cross-sections
shapes have been developed and may be found in the market. Furthermore, the combination of
thin-walled CFS profiles, in the form of built-up cross-section members, further increases the
range of available options. On the other hand, this diversity of solutions brings a challenge to
the designer, as for some uncommon cross-section shapes, the mechanical behavior is not
entirely known and therefore not covered in the design codes. Extensive studies can be found in
the literature regarding plain and lipped channels with a single edge or double edge stiffeners
(Young, Rasmusses (1998), Popovich, Hancock, Rasmussen (1999), Yan, Young (2002)).
This particular subject of built-up elements has been widely studied in the past few years
through experimental testing and finite element modelling. At an experimental level several studies
presented compressive test results on different types of cross-sections, namely open built-up an
closed built-up (Craveiro, et al (2016b), Fratamico, et al (2018), Roy, et al. (2018), Jia-Hui Zhang,
et al. (2015)). The main objectives were the assessment of the load bearing capacity of built-up
elements, evaluating the governing buckling modes and its interactions, gathering results for direct
comparison with currently available design guidelines (EN 1993-1-3 (2006), AISI S100 (2012)) and
for calibration of finite element models. It was found that generally the design predictions accord-
ing to the Direct Strength Method (DSM) (Schafer,Pekoz, (1998)) may be conservative for

295
columns failing due to global buckling, whereas for the EN 1993-1-3 some differences were found
for open and closed sections (Kherbouche,Megnounif, (2019), Craveiro, et al (2016a, 2016b)).
Regarding the design of built-up members, the EN 1993-1-3 (2006) simply states that the
buckling resistance of a closed built-up member should be determined using the buckling
curve b in association with the basic yield strength fyb , and buckling curve c in association
with the average yield strength, fya , provided that Aeff ¼ Ag .
Additional experimental and numerical investigations are needed to further understand and
characterize the behavior of built-up elements under compression, especially due to the high
versatility of these solutions, aiming to develop accurate and reliable design methods able to
tackle different built-up configurations. In this paper, a set of experimental results on lipped
channels and open built-up I sections is presented and used as a base for the calibration of the
numerical models. Both pinned and fixed support conditions were investigated. Based on the
calibrated numerical models a small parametric study was undertaken and the obtained results
were then compared with the design predictions according to the EN 1993-1-3.

2 EXPERIMENTAL TESTS

The experimental test set-up comprised a steel reaction frame (1) and a reinforced concrete
footing where the hydraulic jack, controlled by a servo-hydraulic unit, was positioned ((2) and
(3)). The boundary conditions were materialized by a device that was able to simulate both
pinned and fixed supports (3). A load cell was positioned on top to measure the applied load-
ing (4). In each experimental test, lateral and axial displacements were monitored and
recorded as well as strains at mid-height of the columns in several points in the cross-section
(Figure 1).
In Figure 1. a1) the cross-sections tested are presented and in Figure 1. b1) the instrumenta-
tion used in the cross-section is depicted. The profiles were fastened using self-drilling screws
along the length of the column. The spacing was L/4. The test programme comprised 12 com-
pression tests. Both pinned and fixed boundary conditions were tested to estimate the upper
and lower bounds to the strength of the tested columns. In Table 1 the test programme is
detailed.

Figure 1. Cross-sections tested and schematic representation of the test set-up. a1) C. a2) I. b1) and b2)
Positioning of strain gauges. c) Experimental test set-up. (Craveiro, HD. (2016a)).

296
Table 1. Test programme.

Test Reference λ Loading End-support Repetitions


C_PP_i 2.11 3
Pinned
I_PP_i 1.62 Until failure 3
C_FF_i 1.21 3
I_FF_i 0.78 Fixed 3

3 FINITE ELEMENT MODELLING

The experimental tests were reproduced using the finite element method and finite element
software Abaqus (2017). The objective was to develop a finite element model (FEM) able to
provide accurate strength estimations for cold-formed steel columns. The geometry of the
modelled elements was based on the middle line dimensions of the cross-section of the individ-
ual profiles used in the experimental tests. The modelling was divided in two stages, firstly an
eigenvalue buckling analysis was undertaken and then a non-linear static analysis to estimate
the load bearing capacity and failure modes. The buckling modes of interest are used to input
the geometric imperfections in the nonlinear analysis.
The finite element type selected for this investigation was the S4R which is a four node gen-
eral-purpose shell, quadrilateral and stress/displacement shell element with reduced integra-
tion, a large strain formulation, hourglass control and a first order interpolation. The mesh
size used in this investigation was 10 mm.
The material model was created based on material testing also conducted in the scope of
this investigation (Craveiro, et al. (2016)). Tensile coupon tests were conducted, and the
obtained properties were as follows: Modulus of elasticity of 204 GPa and yield strength of
307 MPa. Considering Ramberg-Osgood model, stress-strain equations were developed. The
equation used was as follows (Eq. 1):
 n
f f
ε¼ þ 0:002  ð1Þ
E fy

The guidelines presented in the MMPDS-01 were considered for the determination of the
Ramberg-Osgood coefficient (Eq. 2 and 3):
 
fu
εus ¼ 100 εr  ð2Þ
E
 
nθ ¼ lnðεus =0:2Þ=ln fu =fy ð3Þ

Where ε is the strain, εr is the strain at rupture, εus is the plastic strain at maximum tension load, f
is the stress, E is the modulus of elasticity, fu is the ultimate strength and fy is the yield strength.
The estimated n coefficient was 14.602. The obtained stress-strain relationship was then converted
to true stress and logarithmic plastic strain using the following equations (Eq. (4) and (5):

σtrue ¼ σnom ð1 þ εnom Þ ð4Þ


σtrue
εpl
ln ¼ lnð1 þ εnom Þ  ð5Þ
E

In Figure 2 the logarithmic plastic strain curve for the S280GD+Z steel, used as input in the
numerical model is depicted.
As for the experimental tests, both pinned and fixed boundary conditions were considered.
Based on experimental observations it was clear that the buckling lengths for the fixed col-
umns were bigger than the theoretical value of 0.5L. Hence, an ideal fixed boundary condition

297
Figure 2. Logarithmic plastic strain curve obtained for the S280GD+Z steel.

was not achieved in the experimental testing. Taking this into consideration, and for calibra-
tion purposes, a hinge was modelled with a high rotational stiffness of 80% of 4EI=L.
For the open built-up columns, the self-drilling screws were modelled using connector elem-
ents in Abaqus. The connector element used was the CONN3D2 with 2 nodes and 6 DOF per
node. To model the interaction between the surfaces of the lipped channels, surface-to-surface
interaction was considered assuming a tangential friction coefficient of 0.2 for the contact
behavior in the tangential direction and a hard contact for the contact behavior in normal
direction between the web surfaces of the profiles.
In the numerical models, loading was applied by imposing an axial displacement, just like in
the experimental tests. The displacement was imposed to a node positioned in the centroid of
the cross-section and connected to the perimeter of the cross-section by a coupling constraint.
Based on the eigenvalue analysis the contours for the imperfections were established. Despite
the first buckling mode may be the critical one, interaction between different buckling modes
was observed, hence in the nonlinear analysis a combination of buckling modes was con-
sidered. The magnitude for global imperfections was L/1000, for distortional was the thickness
of the profile (t) and for local imperfections h=200.

4 RESULTS AND DISCUSSION

4.1 Calibration of the numerical models


For calibration purposes, the experimental and numerical results were compared through
axial compression load (P) versus axial shortening (dv) curves for both cross-section
shapes investigated and for both boundary conditions tested. In Figures 3 and 4
a comparison between experimental and numerical results is established. From the com-
parison, it can be stated that the numerical results are in very good agreement with the
experimental ones.
In Tables 2 and 3 the obtained results in the experimental tests are compared with the ones
obtained in the numerical models and with design predictions. Once again, the good agree-
ment between experimental and numerical results can be observed in Table 2. The design pre-
dictions were determined assuming the support systems were able to provide significant
restraint against torsion and warping, hence considering a value of 0.7 for the determination
of the buckling length for torsional or torsional-flexural buckling. Comparing both experi-
mental and numerical results with design predictions according to the EN 1993-1-3 it seems
that the design buckling load is too conservative for fixed lipped channel columns and slightly
unconservative for pinned open built-up I columns.
Using the calibrated numerical models, it is possible to conduct additional parametric stud-
ies assessing more in detail de influence of several parameters in the overall behavior of lipped
channel and open built-up I cold-formed steel columns (Craveiro, (2016)).

298
Figure 3. Comparison between FEA and experimental results. a) C_PP. b) I_PP.

Figure 4. Comparison between FEA and experimental results. a) C_PP. b) I_PP.

Table 2. Experimental and numerical results obtained for the lipped channel columns.
Experimental Numerical Nb,Rd
Test Pmax [kN] PFEA [kN] [kN] PFEA/Pmax Pmax/Nb,Rd PFEA/Nb,Rd

C_PP-1 24.53 1.10 0.91


C_PP-2 29.14 26.98 27.1 0.93 1.08 0.99
C_PP-3 27.04 1.00 1.00
μ 26.90 26.98 27.1 1.01 0.99 0.99
C_FF-1 65.62 0.98 0.98
C_FF-2 64.35 64.14 67.1 1.00 0.96 0.96
C_FF-3 69.22 0.93 1.03
μ 66.40 64.14 67.1 0.97 0.99 0.96

Table 3. Experimental and numerical results obtained for the open built-up I columns.
Experimental Numerical Nb,Rd PFEA Pmax PFEA
Test
Pmax [kN] PFEA [kN] [kN] /Pmax /Nb,Rd /Nb,Rd

I_PP-1 74.69 1.08 0.87


I_PP-2 69.06 80.61 85.51 1.17 0.81 0.94
I_PP-3 83.18 0.97 0.97
μ 75.64 80.61 85.51 1.07 0.88 0.94
I_FF-1 198.75 0.99 0.95
I_FF-2 182.07 197.7 208.9 1.09 0.87 0.95
I_FF-3 180.11 1.10 0.86
μ 186.98 197.7 208.9 1.06 0.90 0.95

4.2 Parametric study


Using the developed and calibrated numerical model an extensive parametric study is ongoing
to get a deep understanding of the influence of several parameters in the overall behavior of
compressed cold-formed steel elements. In this investigation single sections, open built-up and
closed built-up cross-section shapes are being assessed. For this parametric study three

299
thicknesses (1.5, 2.5 and 3.5 mm), three lengths (1000, 2000 and 3000 mm) and two boundary
conditions (pinned and fixed) were assessed. For the built-up cross-section, the adopted spa-
cing for the fasteners was L/4. In the future, this parameter will be investigated in detail.
The most relevant results are detailed in Tables 4–7, summarizing the obtained results for
the tested lengths, thicknesses, cross-section shapes and boundary conditions.
Summarizing the obtained results, the strength curves according to the EN 1993-1-3 and
EN 1993-1-5 are plotted in Figure 5 (flexural buckling – curve b).
For the tested situations, the predominant failure mode was flexural buckling about the
minor axis, however distortional buckling and interaction between flexural and distortional
buckling was also prevalent in several cases. For instance, for fixed columns, the governing
failure mode was the interaction between flexural buckling and distortional buckling for both

Table 4. Experimental and numerical results obtained for the pinned lipped channel columns.
Nb,Rd PFEA Nc,Rd PFEA/Nc,Rd PFEA/Nb,Rd
Test λ [kN] [kN] [kN] [_] [_]

C_PP_1000_1.5 0.59 57.4 57.51 68.30 0.84 1.002


C_PP_1000_2.5 0.68 112.71 113.79 142.03 0.80 1.010
C_PP_1000_3.5 0.75 168.26 197.30 222.54 0.89 1.173
C_PP_2000_1.5 1.19 32.92 28.98 68.3 0.42 0.880
C_PP_2000_2.5 1.37 56.23 55.95 142.03 0.39 0.995
C_PP_2000_3.5 1.5 76.22 98.64 222.54 0.44 1.294
C_PP_3000_1.5 1.76 17.86 15.44 68.30 0.23 0.865
C_PP_3000_2.5 2.01 29.37 28.73 142.03 0.20 0.978
C_PP_3000_3.5 2.22 38.52 45.19 222.54 0.20 1.173

Table 5. Experimental and numerical results obtained for the fixed lipped channel columns.
Nb,Rd PFEA/Nc,Rd PFEA/Nb,Rd
λ PFEA [kN] Nc,Rd [kN]
Test [kN] [_] [_]

C_FF_1000_1.5 0.41 63.02 69.2 71.01 0.975 1.098


C_FF_1000_2.5 0.46 127.9 141.9 142.0 0.999 1.109
C_FF_1000_3.5 0.49 200.3 190.0 222.9 0.852 0.949
C_FF_2000_1.5 0.81 51.1 55.4 71.01 0.780 1.084
C_FF_2000_2.5 0.88 97.7 101.4 145.1 0.699 1.038
C_FF_2000_3.5 0.89 150.1 179.9 222.9 0.807 1.198
C_FF_3000_1.5 1.18 34.7 46.6 71.01 0.656 1.343
C_FF_3000_2.5 1.23 67.5 84.7 145.1 0.584 1.255
C_FF_3000_3.5 1.18 110.8 154.2 222.9 0.692 1.392

Table 6. Experimental and numerical results obtained for the pinned open built-up I columns.
Nb,Rd PFEA Nc,Rd PFEA/Nc,Rd PFEA/Nb,Rd
Test λ [kN] [kN] [kN] [_] [_]

I_PP_1000_1.5 0.47 127.3 138.5 142.0 0.975 1.088


I_PP_1000_2.5 0.53 252.5 249.4 290.3 0.859 0.988
I_PP_1000_3.5 0.57 384.7 374.2 451.7 0.828 0.973
I_PP_2000_1.5 0.94 89.8 115.9 142.0 0.816 1.291
I_PP_2000_2.5 1.06 161.5 158.8 290.3 0.547 0.983
I_PP_2000_3.5 1.14 230.9 282.1 451.7 0.625 1.222
I_PP_3000_1.5 1.42 52.77 55.4 142.0 0.390 1.050
I_PP_3000_2.5 1.59 89.77 88.7 290.3 0.306 0.988
I_PP_3000_3.5 1.71 124.0 140.8 451.7 0.312 1.135

300
Table 7. Experimental and numerical results obtained for the fixed open built-up columns.
Nb,Rd PFEA Nc,Rd PFEA/Nc,Rd PFEA/Nb,Rd
Test λ [kN] [kN] [kN] [_] [_]

I_FF_1000_1.5 0.25 139.4 143.0 142.0 1.007 1.026


I_FF_1000_2.5 0.28 282.2 283.8 290.2 0.978 1.006
I_FF_1000_3.5 0.29 437.3 447.1 451.7 0.990 1.022
I_FF_2000_1.5 0.49 125.9 134.8 142.0 0.949 1.071
I_FF_2000_2.5 0.53 252.5 240.2 290.2 0.828 0.951
I_FF_2000_3.5 0.57 348.7 358.6 451.7 0.794 1.028
I_FF_3000_1.5 0.71 109.9 121.3 142.0 0.854 1.104
I_FF_3000_2.5 0.79 210.7 227.1 290.3 0.782 1.078
I_FF_3000_3.5 0.85 311.3 335.0 451.7 0.742 1.076

Figure 5. Strength curves for lipped and open built-up I columns for both end support conditions.

lipped channel columns. For the higher thickness tested of 3.5 mm it seems that the design
predictions may be too conservative, especially for fixed columns.
Comparing the obtained results with design predictions it was found that design methods
are able to estimate adequate accuracy the buckling resistance of lipped channel and open
built-up I cold-formed steel columns for both end-support conditions investigated.

5 CONCLUSIONS

This paper presents an experimental and numerical analysis using the finite element
method on the buckling behavior of single lipped channel and open built-up I cold-
formed steel columns under compression. The assumptions considered in the numerical
modelling led to a good agreement between experimental and numerical results. The
parametric study will provide lower and top boundaries for the resistance of this type of
cold-formed steel columns.
Comparing the obtained results in the parametric studies with the design predictions accord-
ing to the EN 1993-1-3, it can generally be stated that a good agreement was observed, except in
some cases with 3.5 mm thickness and for the columns with lipped channel cross-section.
This is an ongoing investigation and the authors envisage that in the near future additional
parametric studies will be undertaken, further improving the available models and expanding
the current range of dimensions and parameters. For instance, the assessment of the influence

301
of spacing of fasteners used to fabricate built-up sections will be one of the key topics to be
addressed.

ACKNOWLEDGEMENTS

The authors gratefully acknowledge to the Portuguese Foundation for Science and Technol-
ogy for its support under the framework of the research projects PTDC/ECM/116859/2010
(FireColdFSteel), and POCI-01-0145-FEDER-031858 (INNOCFSCONC).

REFERENCES

AISI. North American Specification for the design of cold-formed steel structural members. North
American Cold-Formed Steel Specification. Washington, DC: American Iron and Steel Institute, AISI
S100-12; 2012.
B. Young, J. R. Rasmussen, Tests of fixed-ended plain channel columns, Journal of Structural Engineer-
ing, 124 (2) (1998), 131–139.
B. Young, J. Chen, Design of cold-formed steel built-up closed sections with intermediate stiffeners, Jour-
nal of Structural Engineering, 134 (5), (2008), 727–737.
Craveiro, Hélder David, Fire resistance of cold-formed steel columns. Coimbra, 2016a. PhD Thesis.
Craveiro, H.D., Rodrigues, João P., Laím, L., Buckling resistance of axially loaded cold-formed steel
columns, Thin-Walled Structures, 106 (2016b) 358–375.
Craveiro, H.D., Rodrigues, J.P. Santiago, A., Laim, L., Review of the high temperature mechanical and
thermal properties of the steels used in cold-formed steel structures – The case of the S280GD-Z steel,
Thin-Walled Structures, Vol. 98, part A, 2016, 154–168.
Craveiro, H., Ribeiro, J., Breda, R., Structural analysis of C and I shaped cold-formed steel columns,
Proceedings of Eurosteel 2017, 2017, pp. 1726–1735.
D. Popovich, G.J. Hancock, K.J.R. Rasmussen, Axial compression tests on cold-formed angles loaded
parallel with a leg, Journal of Structural Engineering, 127 (6) (1999), 600–607.
David C. Fratamico, Shahabeddin Torabian, Xi Zhao, Kim J. Rasmussen, Benjamin Schafer, Experi-
ments on the global buckling and collapse of built-up cold-formed steel columns, Journal of Construc-
tional Steel Research, Vol. 144, 2018, 65–80.
EN 1993-1-3:2006, Eurocode 3: Design of steel structures, Part 1-3: General Rules, Supplementary Rules
for Cold-Formed Members and Sheeting, European Committee for Standardization, Brussels, Bel-
gium, p.125.
J. Yan, B. Young, Column tests of cold-formed steel channels with complex stiffeners, Journal of Struc-
tural Engineering, 128 (6) (2002) 737–745.
Jia-Hui Zhang, Ben Young, Numerical investigation and design of cold-formed steel built-up open sec-
tion columns with longitudinal stiffeners, Thin-Walled Structures, Volume 89, 2015, 178–191.
Kherbouche, S., Megnounif, A., Numerical study and design of thin walled cold formed steel built-up
open and closed section columns, Engineering Structures, Vol. 179, 2019, 670-682.
Krishanu Roy, Tina C.H. Ting, Hieng H. Lau, James B.P. Lim, Effect of thickness on the behaviour of
axially loaded back-to-back cold-formed steel built-up channel section – Experimental and numerical
investigation, Structures, Vol. 16, 2018, 327–346.
Li Yuanqui, Li Yinglei, S. Wang, Z. Shen, Ultimate load-carrying capacity of cold-formed thin-walled
columns with built-up box and I section under axial compression. Thin-Walled Structures, 79 (2014)
2020–217.
Schafer B.W., Peköz T. Direct strength prediction of cold-formed steel members using numerical elastic
buckling solutions. In: Proceedings of the 14th international specialty conference on cold-formed steel
structures. St. Louis, MO, USA, University of Missouri-Rolla; 1998. p. 69–76.
Simulia 2017. Abaqus FEA (version 6.17) Simulia Dassault Systémes.
T.A. Stone, R.A. Laboube, Behavior of cold-formed steel built-up I-sections, Thin-Walled Structures, 43
(12) (2005) 1805–1817.

302
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

FEM analysis of the buckling behaviour of thin-walled CFS


columns. Part II—monosymmetric (R) and double symmetric
built-up box cross-sections

H.D. Craveiro
ISISE, University of Coimbra, Portugal

J. Henriques
CERG, Faculty of Engineering Technology, University Hasselt, Belgium

L.S. Silva
ISISE, University of Coimbra, Portugal

J.P. Martins
ISISE, University of Coimbra, Portugal

ABSTRACT: One of the main advantages of thin-walled CFS profiles is the “freedom” the
designer has to conceive a cross-section shape tailored for the structural application. Conse-
quently, different cross-sections shapes have been developed and may be found in the market.
Furthermore, the combination of thin-walled CFS profiles, in the form of built-up cross-section
members, further increases the range of available options. On the other hand, this diversity of
solutions brings a challenge to the designer, as for some uncommon cross-section shapes, the
mechanical behavior is not entirely known and therefore not covered in the design codes.
Hence, in the present paper, the second part of a numerical study on the subject is presented.
Columns consisting of the monosymmetric built-up box cross-section (R-section) and doubly
symmetric built-up box cross-section (2R-section), comprising lipped and plain channels are
investigated by means of FEM. The numerical models were developed and calibrated using
experimental results on thin-walled CFS columns using these cross-section shapes. Upon valid-
ation, a parametric analysis is performed, to further assess the influence of different parameters,
such as steel thickness, column length, boundary conditions and connection spacing.

1 INTRODUCTION

The demand for structures using cold-formed steel (CFS) sections has increased significantly since
it has been recognized as a valuable and competitive structural solution. One of the main advan-
tages of thin-walled CFS profiles is the “freedom” the designer has to conceive a cross-section
shape tailored for the structural application. Consequently, different cross-sections shapes have
been developed and may be found in the market. Furthermore, the combination of thin-walled
CFS profiles, in the form of built-up cross-section members, further increases the range of avail-
able options. On the other hand, this diversity of solutions brings a challenge to the designer, as
for some uncommon cross-section shapes, the mechanical behavior is not entirely known and
therefore not covered in the design codes. Adopting built-up sections will provide additional load
bearing capacity and higher torsional stiffness, and the combination of several individual profiles
allows to prescribe symmetric cross-sections eliminating eccentricities between shear and gravity
centers, enhancing the stability of the cold-formed steel elements. Moreover, the use of built-up
sections may present relevant economic advantages since the field of applicability of individual
cold-formed steel elements can be further expanded without changing the manufacture procedure.

303
This particular subject of built-up elements has been studied in the past few years through
experimental testing and finite element modelling. At an experimental level several studies pre-
sented compressive test results on different types of cross-sections, namely open built-up an
closed built-up (Craveiro, H.D., et al (2016), Fratamico, D.C. et al (2018), Krishanu Roy,
et al. (2018), Jia-Hui Zhang, et al. (2015), Li Yuanqui, et al. (2014), Reyes Wilson (2011), Ben
Young, J. Chen (2008)). The main objectives were the assessment of the load bearing capacity
of built-up elements, evaluating the governing buckling modes and its interactions, gathering
results for direct comparison with currently available design guidelines (EN 1993-1-3 (2006),
AISI S100 (2012)) and for calibration of finite element models.
Regarding the design of built-up members, the EN 1993-1-3 (2006) simply states that the
buckling resistance of a closed built-up member should be determined using the buckling
curve b in association with the basic yield strength f_yb, and buckling curve c in association
with the average yield strength, f_ya, provided that A_eff = A_g.
Additional experimental and numerical investigations are needed to further understand and
characterize the behavior of built-up elements under compression, especially due to the high
versatility of these solutions, aiming to develop accurate and reliable design methods able to
tackle different built-up configurations.
This paper presents the second part of a numerical study on thin-walled CFS columns.
More precisely, a set of experimental results on columns consisting of the monosymmetric
built-up box cross-section (R-section) and doubly symmetric built-up box cross-section
(2R-section), comprising lipped and plain channels, are used to validate and calibrate the
numerical modelling. Both pinned and fixed support conditions were investigated. Then,
a small parametric study was undertaken and the obtained results were compared with the
design predictions according to the EN 1993-1-3.

2 EXPERIMENTAL TESTS

The experimental tests on columns was part of a larger experimental campaign. Detail on the
test set-up can be obtained in the companion paper which presents the first part of this numer-
ical study (Craveiro, H.D., et al (2019)). In the present, columns consisting of the monosym-
metric built-up box cross-section (R-section) and doubly symmetric built-up box cross-section
(2R-section) are addressed. In Figure 1 are illustrated the shape of the built-up cross-sections
and the connection points between the two profiles within the cross-section. For the R-section,
connection is achieved between the flanges of the lipped channel and channel profiles. In the
2R-section, two lipped channel profiles are first positioned back-to-back and connected with
two fasteners within the web. Then, to perform the double hollow cross-section (2R-section) is
achieved connecting channel profiles as described before to realize the R-section. The connec-
tion between the two profiles was accomplished with self-drilling screws.
The test programme comprised 12 compression tests. Both pinned (PP) and fixed (FF)
boundary conditions were tested to estimate the upper and lower bounds to the strength of
the tested columns. In Table 1 the test program is detailed.

Figure 1. Shape of the investigated cross-sections. a) R-section and b) 2R-section.

304
Table 1. Test programme.
Test Reference λ Loading End-support Repetitions

R_PP_i 1.66 Pinned 3


2R_PP_i 1.03 Until 3
R_FF_i 0.77 failure Fixed 3
2R_FF_i 0.48 3

3 FINITE ELEMENT MODELLING

The experimental tests were reproduced using the finite element method and finite element
software Abaqus (2017). The objective was to develop a finite element model (FEM) able to
provide accurate strength estimations for cold-formed steel columns. The geometry of the
modelled elements was based on the middle line dimensions of the cross-section of the individ-
ual profiles used in the experimental tests. The modelling was divided in two stages, firstly an
eigenvalue buckling analysis was undertaken and then a non-linear static analysis to estimate
the load bearing capacity and failure modes. The buckling modes of interest are used to input
the geometric imperfections in the nonlinear analysis.
The finite element type selected for this investigation was the S4R which is a four-node gen-
eral-purpose shell, quadrilateral and stress/displacement shell element with reduced integra-
tion, a large strain formulation, hourglass control and a first order interpolation. The mesh
size used in this investigation was 10 mm.
Detail on the following aspects can be found in the companion paper presenting the first
part of this numerical study (Craveiro, H.D., et al (2019)): i) material model; ii) connection
between CFS profiles and interactions (contact); iii) loading strategy; iv) types of analysis; v)
geometric imperfections.

4 RESULTS AND DISCUSSION

4.1 Validation of the numerical model


The validation of the numerical model was accomplished through the comparison of the axial
compression load versus axial deformation curves between FEM model and experimental
tests. In Figure 2 and Figure 3 is shown the comparison for both profile configurations
(R and 2R) and for both boundary conditions. A good agreement between experiments and
numerical is evident, in particular up to maximum load capacity of the columns. The approxi-
mation in the post-buckling range depends on the configuration. The accuracy in this region

Figure 2. Comparison of axial compression versus axial deformation curves between FEM model and
experimental tests for R specimens. a) Pinned columns. b) Fixed columns.

305
Figure 3. Comparison of axial compression versus axial deformation curves between FEM model and
experimental tests for 2R specimens. a) Pinned columns. b) Fixed columns.

can be improved with a sensitivity study on the distortional and local modes to consider as
initial geometric imperfections. This is currently under development.
In Table 2 and Table 3 the obtained results in the experimental tests are compared with the
ones obtained in the numerical models and with design predictions according to the EN 1993-1-3.
The design predictions were determined assuming the support systems were able to provide sig-
nificant restraint against torsion and warping, hence considering a value of 0.7 for the determin-
ation of the buckling length for torsional or torsional-flexural buckling. For comparison
purposes, ratios between the results obtained with the different approaches are included. It is pos-
sible to observe the following:
• FEA/EXP: the numerical exceeds the mean of the resistance obtained within the experimen-
tal results but the maximum obtained in the experimental tests is close or even exceeds the

Table 2. Experimental and numerical results obtained for the R-section columns.
Experimental Numerical Nb,Rd
Test Pmax [kN] PFEA [kN] [kN] PFEA/Pmax Pmax/Nb,Rd PFEA/Nb,Rd

R_PP-1 66.26 1.13 0.87


R_PP-2 70.44 74.88 76.55 1.06 0.92 0.98
R_PP-3 73.38 1.02 0.96
μ 70.02 74.88 76.55 1.07 0.91 0.98
R_FF-1 148.03 0.90 0.76
R_FF-2 160.79 133.58 195.52 0.83 0.82 0.68
R_FF-3 138.22 0.97 0.71
μ 149.01 133.58 195.52 0.90 0.76 0.68

Table 3. Experimental and numerical results obtained for the 2R-section columns.
Experimental Numerical Nb,Rd
Test Pmax [kN] PFEA [kN] [kN] PFEA/Pmax Pmax/Nb,Rd PFEA/Nb,Rd

2R_PP-1 268.79 0.99 0.88


2R_PP-2 257.71 265.60 305.57 1.03 0.84 0.87
2R_PP-3 235.15 1.13 0.77
μ 253.88 265.60 305.57 1.05 0.83 0.87
2R_FF-1 388.67 1.03 0.88
2R_FF-2 377.88 398.78 471.07 1.06 0.85 0.85
2R_FF-3 356.59 1.12 0.80
μ 374.38 398.78 471.04 1.07 0.84 0.85

306
numerical result. Real boundary conditions and initial perfections may be the reason for the
variations, nevertheless, the estimation of the numerical presents a good approximation.
• EXP/EC3: In all cases the analytical estimations exceed the experimental results. Two main
reasons are found for these deviations; first, the boundary conditions, in particular for the fully-
fixed case; second, the imperfection factor (buckling curve) indicated by the EN1993-1-3 may
not be indicated for built-up cross-sections form hollow sections. According to this code, curve
b (α = 0.34) should be used for all built-up cross-section members. In the EN 1993-1-1 (2005),
for hollow sections made of cold formed steel, the buckling curve to be use id curve c (α = 0.49).

4.2 Parametric study


In order to extend the limitations of the experimental campaign, namely in what concerns the vari-
ation properties of the test specimens, a parametric study is currently under development. The pre-
liminary results of this parametric study are here discussed. In Table 4 are summarize the varied
parameters and their range of variation. At this stage, no combined variation of the problem vari-
ables has been considered. The varied parameters are the column length (L), the plate gauge thick-
ness of the steel profiles (t) and the number of connections along the length (s). The reference for
these variables is the following: L = 300 mm, t = 2.5 mm and s = L/4. When a parameter is under
variation, the refence value of the other parameters was used. The boundary conditions consider
ideal conditions, pinned (PP) and fully fixed (FF). For all other properties of columns models, the
values used in the validation process, resulting from the test specimens, were considered, as for
example the material grade. The results are discussed here below for each of the varied parameter
• Column length (L)
In Table 5 are detailed the results of the variation of the column length. The table includes the
numerical and analytical predictions. For the latter, the ultimate load capacity (Nb,Rd) and the

Table 4. Summary of the parametric variations.


Varied parameter Range of variation Name of the model

R_PP_Li; R_FF_Li;
Column length L= 3000 mm; L = 2000 mm; L = 1000 mm
2R_PP_Li; 2R_FF_Li
R_PP_ti; R_FF_ti;
Plate gauge thickness t = 3.5 mm; t = 2.5 mm; t = 1.5 mm
2R_PP_ti;2R_FF_ti
R_PP_si; R_FF_si;
Connection spacing s1 = L/4; s2 = L/8
2R_PP_si; 2R_FF_si

Table 5. Analytical and numerical results obtained for variation of the column length(L).
Nb,Rd PFEA Nc,Rd
Test λ [kN] [kN] [kN] PFEA/Nc,Rd PFEA/Nb,Rd

R_PP_1000 0.54 228.23 243.48 0.92 1.07


R_PP_2000 1.09 143.53 136.84 0.52 0.95
R_PP_3000 1.63 78.55 75.64 0.29 0.96
R_FF_1000 0.27 257.10 280.61 263.77 1.06 1.09
R_FF_2000 0.54 228.23 245.32 0.93 1.07
R_FF_3000 0.81 188.95 201.26 0.76 1.07
2R_PP_1000 0.34 501.39 557.06 1.06 1.11
2R_PP_2000 0.67 421.17 419.51 0.80 1.00
2R_PP_3000 1.01 311.44 268.57 0.51 0.86
2R_FF_1000 0.17 527.54 559.01 527.54 1.06 1.06
2R_FF_2000 0.34 501.39 546.48 1.04 1.09
2R_FF_3000 0.51 465.22 443.25 0.84 0.95

307
cross-section capacity (Nc,Rd) were computed. In Figure 4 the numerical results are compared
with the strength curves. For both cross-sections shape, and majority of the simulated cases,
the governing failure mode was the flexural buckling about the minor axis. The short columns,
the cross-section capacity was attained.
• Plate gauge thickness (t)
In Table 6 are detailed the results of the variation of the plate gauge thickness. The table
includes the numerical and analytical predictions. In Figure 5 the numerical results are com-
pared with the strength curves. For all cases and for both cross-sections, the governing failure
mode is the flexural buckling about the minor axis.
• Connection spacing (s)
In Table 7 are detailed the results of the variation of the plate gauge thickness. The table
includes the numerical and analytical predictions. In Figure 6 the numerical results are com-
pared with the strength curves. The results indicate that the reduction on the connector spa-
cing are reflected as an increase on the resistance. The increase is more significant with the
decrease of the column slenderness.

Figure 4. Strength curves for variation of the column length (L).

Table 6. Analytical and numerical results obtained for variation of the plate gauge thickness(t).
Nb,Rd PFEA Nc,Rd
Test λ [kN] [kN] [kN] PFEA/Nc,Rd PFEA/Nb,Rd

R_PP_1.5 1.40 44.20 37.99 116.33 0.33 0.86


R_PP_2.5 1.63 78.55 75.64 263.77 0.29 0.96
R_PP_3.5 1.73 113.59 114.29 432.46 0.26 1.01
R_FF_1.5 0.70 91.23 96.75 116.33 0.83 1.06
R_FF_2.5 0.81 188.95 201.26 263.77 0.76 1.07
R_FF_3.5 0.88 291.82 311.64 432.46 0.72 1.07
2R_PP_1.5 0.88 157.07 133.06 232.67 0.57 0.85
2R_PP_2.5 1.01 311.44 268.57 527.54 0.51 0.86
2R_PP_3.5 1.08 472.62 430.50 864.91 0.50 0.91
2R_FF_1.5 0.44 211.78 214.36 232.67 0.922 1.01
2R_FF_2.5 0.51 465.22 462.27 527.54 0.88 0.99
2R_FF_3.5 0.54 864.91 684.65 864.91 0.78 0.78

308
Figure 5. Strength curves for variation of the plate gauge thickness (t).

Table 7. Analytical and numerical results obtained for variation of the connection spacing (s).
Nb,Rd PFEA Nc,Rd
Test λ [kN] [kN] [kN] PFEA/Nc,Rd PFEA/Nb,Rd

R_PP_L/4 75.64 0.29 0.96


1.63 78.55
R_PP_L/8 79.15 0.30 1.01
263.77
R_FF_L/4 201.26 0.76 1.07
0.81 188.95
R_FF_L/8 227.24 0.86 1.20
2R_PP_L/4 268.57 0.51 0.86
1.01 311.44
2R_PP_L/8 300.0 0.57 0.96
2R_FF_L/4 443.28 527.54 0.84 0.95
2R_FF_L/8 0.51 465.22 50.28 0.96 1.09

Figure 6. Strength curves for variation of the connection spacing (s).

5 CONCLUSIONS

This paper presented a numerical study under development on columns consisting of the mono-
symmetric built-up box cross-section (R-section) and doubly symmetric built-up box cross-section
(2R-section). The validation of the numerical model was accomplished against an experimental

309
campaign on columns using the same type of cross-sections. The comparison of results, namely
the ultimate load capacity, demonstrated the good approximation between both approaches. Fur-
thermore, the prediction of the ultimate load capacity according to EN 1993-1-3 was included.
The results of the latter show an overestimation the experimental results. Two main reasons are
identified for the deviations: i) the ideal support conditions assumed in the analytical calculations;
ii) possible lack of suitability of the buckling curve (curve b), indicated in the EN 1993-1-3 for
built-up cross-section members, for the hollow sections investigated within the present paper. In
the EN 1993-1-1 (2005), for cold formed hollow cross-sections, the appropriate buckling curve is
curve c. This is issue is currently under investigation.
Then, using the validated numerical model, the results of a parametric study, that is under
execution, were presented and discussed. The following properties of the studied column were
varied: length (L), plate gauge thickness (t) and spacing between the connection to form the
built-up cross-section (s) along the length of the column. The main conclusions are:
• The predominant failure mode, for the generality of the studied cases, is flexural buckling
around the weak axis.
• The investigated members consist of hollow cross-sections and therefore when the slender-
ness is significantly reduced, the capacity of the cross-sections is attained and no predomin-
ant local or distortional modes are observed.
• The reduction of the spacing of the connection within the length of the column results in an
increase of resistance, the more compact is the member the greater is the increase. This
observation indicates the presence of distortional and local modes even if the governing
failure mode is flexural buckling.

ACKNOWLEDGEMENTS

The authors gratefully acknowledge to the Portuguese Foundation for Science and Technol-
ogy for its support under the framework of the research projects PTDC/ECM/116859/2010
(FireColdFSteel), POCI-01-0145-FEDER-031435 (FireImprov_ColdFSteel) and POCI-01-
0145-FEDER-031858 (INNOCFSCONC).

REFERENCES

AISI. North American Specification for the design of cold-formed steel structural members. North
American Cold-Formed Steel Specification. Washington, DC: American Iron and Steel Institute, AISI
S100-12; 2012.
B. Young, J. Chen, Design of cold-formed steel built-up closed sections with intermediate stiffeners,
J. Struct. Eng., 134 (5) (2008), pp. 727–737
B. Young, J. Chen, Design of cold-formed steel built-up closed sections with intermediate stiffeners, Jour-
nal of Structural Engineering, 134 (5), (2008), 727–737.
Craveiro, H.D., Rodrigues, João P., Laím, L., Buckling resistance of axially loaded cold-formed steel
columns, Thin-Walled Structures, 106 (2016) 358–375
Craveiro, H.D., Rodrigues, J.P. Santiago, A., Laim, L., Review of the high temperature mechanical and
thermal properties of the steels used in cold-formed steel structures – The case of the S280GD-Z steel,
Thin-Walled Structures, Vol. 98, part A, 2016, 154–168.
Craveiro, H.D., Henriques, J., Simões da Silva, L., Martins, J. P., FEM analysis of the buckling behav-
iour of thin-walled CFS columns. Part I- Channel (C) and Double Channel (I) cross-sections, SDSS –
The Internaltional Colloquium on Stability and Ductility of Steel Structures, 11-13 September 2019,
Prague, Czech Republic. (Submitted)
David C. Fratamico, Shahabeddin Torabian, Xi Zhao, Kim J. Rasmussen, Benjamin Schafer, Experi-
ments on the global buckling and collapse of built-up cold-formed steel columns, Journal of Construc-
tional Steel Research, Vol. 144, 2018, 65–80.
EN 1993-1-3:2006, Eurocode 3: Design of steel structures, Part 1-3: General Rules, Supplementary Rules
for Cold-Formed Members and Sheeting, European Committee for Standardization, Brussels, Bel-
gium, p.125.

310
EN 1993-1-1:2005, Eurocode 3: Design of steel structures, Part 1-1: General Rules and rules for build-
ings, European Committee for Standardization, Brussels, Belgium, p. 91.
Jia-Hui Zhang, Ben Young, Numerical investigation and design of cold-formed steel built-up open sec-
tion columns with longitudinal stiffeners, Thin-Walled Structures, Volume 89, 2015, 178–191.
Kherbouche, S., Megnounif, A., Numerical study and design of thin walled cold formed steel built-up
open and closed section columns, Engineering Structures, Vol. 179, 2019, 670–682.
Krishanu Roy, Tina C.H. Ting, Hieng H. Lau, James B.P. Lim, Effect of thickness on the behaviour of
axially loaded back-to-back cold-formed steel built-up channel section – Experimental and numerical
investigation, Structures, Vol. 16, 2018, 327–346.
Li Yuanqui, Li Yinglei, S. Wang, Z. Shen, Ultimate load-carrying capacity of cold-formed thin-walled
columns with built-up box and I section under axial compression, Thin-Walled Struct., 79 (2014),
pp. 202–217
Li Yuanqui, Li Yinglei, S. Wang, Z. Shen, Ultimate load-carrying capacity of cold-formed thin-walled
columns with built-up box and I section under axial compression. Thin-Walled Structures, 79 (2014)
2022–217.
Reyes Wilson, Guzmán Andrés, Evaluation of the slenderness ratio in built-up cold-formed box sections,
J. Constr. Steel Res., 67 (2011), pp. 929–935
Schafer B.W., Peköz T. Direct strength prediction of cold-formed steel members using numerical elastic
buckling solutions. In: Proceedings of the 14th international specialty conference on cold-formed steel
structures. St. Louis, MO, USA, University of Missouri-Rolla; 1998. p. 69–76
Simulia 2017. Abaqus FEA (version 6.17) Simulia Dassault Systémes
T.A. Stone, R.A. Laboube, Behavior of cold-formed steel built-up I-sections, Thin-Walled Structures, 43
(12) (2005) 1805–1817.

311
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Comparison of two different innovative solutions for IPE beam


to CHS column connections

R. Das
Construction Engineering Research Group, Hasselt University, Hasselt, Belgium

A. Kanyilmaz
Department of Architecture, Built Environment and Construction Engineering, Politecnico di Milano, Italy

H. Degee
Construction Engineering Research Group, Hasselt University, Hasselt, Belgium

ABSTRACT: Local stiffeners and gusset plates used in conventional open-to-Circular Hollow
Section (CHS) column connections result in high welding quantities, whereas direct welding
exposes the CHS to severe local distortions. Such disadvantages hinder the use of open-to-CHS
column connections. To minimize these drawbacks and improve the structural behaviour, this
study proposes two moment resisting connection configurations developed within the European
research project “LASTEICON”. It deals specifically with I-beam resp. individual plates passing
through the CHS column. The “passing-through” system is made possible by resorting to an easy
fabrication process thanks to efficient laser cut & weld technology and avoids the localized dam-
ages, premature flange failures and excessive use of stiffening plates required by the conventional
I-beam-to-CHS column connection configurations. The new configurations are investigated for
two different load cases with experimentally calibrated nonlinear finite element models and are
also compared among each other to differentiate the force-transfer mechanism involved.

1 INTRODUCTION

Conventional open-to-CHS column connections use stiffeners and gusset plates to transfer the
loads efficiently between the members. Resulting extensive welding quantities, premature flange
fracture and local distortions on the CHS issues unavoidable complexities and economic losses
(Alostaz & Schneider 1996, Fukumoto & Morita 2005). Despite these shortcomings, CHS offer
massive potentials such as excellent resistance in all directions and provides economic advantages
in fire protection along with an aesthetic appeal (Wardenier et al. 2008, Rondal et. al 1992). To
further their cause, CIDECT has provided significant understanding regarding stability, fire pro-
tection, composite construction, and the static and fatigue behaviour of several types of CHS con-
nections through distinguished design guides (Kurobane et al. 2004, Dutta et al. 1998). However,
to further improve the open-to-CHS connections, several different “passing-through” connections
were proposed through the “LASTEICON” research project (Castiglioni et al. 2016) such as: two
different types of nominally pinned connections, namely C1 and C2; as well as two different
moment resisting connections, namely C3 and C4. This present study focuses on the aforemen-
tioned moment resisting connections: (1) the C3 connection (Figure 1a), constituting of main
beams connected to an I-beam passing through the CHS column via Laser Cut slots and (2) the
C4 connection (Figure 1b), constituting of main beams connected via three individual plates (two
transverse flange plates and a longitudinal web plate) passing through the CHS column. The
force transfer mechanism was identified to be somewhat different for these LASTEICON connec-
tions compared to the conventional ones. As the Laser Cutting Technology (LCT) offers several
advantages compared to the mechanical cutting procedure (Kanyilmaz & Castiglioni 2018) and

312
Figure 1. All parametric dimensions for (a) LASTEICON C3 configuration (b) LASTEICON C4
configuration.

a swift fabrication process with precise tolerance, it encourages the application of these “passing-
through” connections. This article presents the characterization of these two “passing-through”
open-to-CHS column connections and compares them to highlight the advantages and disadvan-
tages through experimentally calibrated numerical models.

2 MODELLING APPROACH AND EXPERIMENTAL CALIBRATION

The LASTEICON configurations were analyzed through nonlinear static analyses by means of
3D solid models implemented in DIANA 10.2 (2018). The laser cut slots on the CHS column were
taken into account to place the through members. To avoid any secondary connection failure and
focus on the “passing-through” zone, the slots in these numerical models were made with zero
tolerance, thus perfectly connecting the CHS column with the through members assuming
a perfectly welded connection. To avoid complicated numerical models, save more computational
time and avoid other possible failure mechanisms, the through members were directly connected
to the main beams as a perfectly welded connection rather than modelling the bolts explicitly. The
models were designed to be symmetric around all axes. Two different load cases - a gravitational
loading LC1 (Figure 2a) and an opposite bending loading LC2 (Figure 2b) - were considered for
the numerical studies. Both vertical loads were incremented simultaneously in each step.
A preliminary experimental campaign was carried out by INSA, Rennes (Couchaux et al. 2019)
to validate the numerical models for both load cases. Two additional solid circular plates were
connected to each extremity of the CHS column (see Figure 1) which were finally pinned by rollers
following the boundary conditions. Bracings were placed to limit the lateral torsional buckling of

313
Figure 2. Schematic diagram showing the application of (a) LC1-Monotonic gravitational (unidirec-
tional) loading and (b) LC2-Monotonic opposite bending loading on C4 connection.

the beam. Four stiffening plates were welded to each beam flange in the connection zones to limit
local buckling. These additional members were also considered in the numerical models (see
Figure 1) to get a reliable replica of the experimental specimens and thus provide an appropriate
validation. Loads were applied at the furthest extremity of the main beams. Inclinometers and
LVDTs were placed in relevant positions to measure corresponding displacements.
Four specimens, one for each configuration (C3 and C4) and each load case (LC1 and
LC2), were investigated with the aforementioned test set-up. Relevant geometrical specifica-
tions regarding the different components are provided in Table 1. Experimental results were
compared with the numerical results in terms of force-displacement curves and observed fail-
ure modes. Material properties for the numerical models were adopted according to the stress-
strain relationship obtained from the experimental coupon tests. Decent agreement can be
seen between the experimental and numerical results in terms of initial stiffness and ultimate
resistance of the connections, as shown in Figure 3. Failure modes obtained from the numer-
ical models also agreed with those of the experimental prototypes (Figure 4).

3 COMPARISON BETWEEN LASTEICON C3 AND C4 CONFIGURATION

Three numerical prototypes were considered for each configuration and under each load case LC1
and LC2, as shown in Table 2, for the numerical parametric study. The overall length of the beam

Table 1. Geometric specifications of specimens tested in the preliminary experimental campaign.


“Main” Beam Through member specifications CHS specifications

Lb h bf hw tw tf dc tc Lc
Specimen & load case IPE (mm) (mm) (mm) (mm) (mm) (mm) (mm) (mm) (mm)

C3-LC1 IPE 400 5000.0 400 180 373 08.6 13.5 355.6 08.8 2340.0
C3-LC2 IPE 400 3400.0 400 180 373 08.6 13.5 355.6 10.0 2340.0
C4-LC1 IPE 400 5000.0 424 180 320 08.0 10.0 355.6 08.8 2340.0
C4-LC2 IPE 400 3400.0 424 180 320 10.0 12.0 355.6 12.5 2340.0

Figure 3. Comparison of load-displ curves between numerical and experimental results.

314
Figure 4. (a) C3-LC1 failure due to through beam flange plasticity, (b) C3-LC2 failure due to beam-to-
CHS column connection tearing, (c) C4-LC1 failure due to flange plate buckling in compression inside
the CHS column and (d) C4-LC2 failure due to beam-to-CHS column connection tearing.

(Lb) and CHS column (Lc), as indicated in Figure 1, was taken as 5000 mm and 2340 mm respect-
ively. A CHS column with a diameter (dc) of 355.6 mm and thickness (tc) of 10 mm was chosen for
all the numerical prototypes aforementioned in Table 2. Material is now considered as elastic-
perfectly plastic: overstrengthening due to strain hardening is neglected to further be reliably com-
pared with an analytical procedure to calculate the design resistance of the respective connections.
The ultimate resistances calculated for all these 12 case studies following a newly developed design
approach (Das et al. 2019a, Das et al. 2019b) are also listed in Table 2 along with the predicted
failure mode. The width and thickness of the through flange plates in C4 were kept equal to that
of the through I-beam in the C3 connection (Figure 1a and Table 2). The geometry of the CHS
column as well as the main beams were also kept identical (Figure 1b and Table 2) to avoid any
other influence than the joint configuration and the resulting load transfer mechanism. Figure 5
illustrates the identified force transfer mechanisms for both connections under LC1 and LC2. The
vertical force-displacement curves under LC1 and LC2 are plotted in Figure 6a and Figure 6b
respectively.

3.1 Comparison under monotonic gravitational loading, LC1


Under LC1, although the through I-beam section was varied in C3.1 and C3.2 and different steel
grades were considered in C3.3, the flanges of the through I-beams started to yield just outside
the CHS column wall prior to all other components of the connection configuration in all three
cases. This occurred due to a rigid-body behaviour of the main joint panel. Compressive stresses
generated by the moments at each side of the connection panel were not affecting the through

315
Table 2. Analytical values corresponding to the ultimate resistances for C3 and C4 configurations
under LC1 and LC2.
“Through” Beam (C3) or Plates (C4)

LASTEICON Joint Failure


“Main” db Ultimate
Config. Mode
Beam fyb, fyc fy,in (or h) bf tf hw tw Strength
(IPE) (MPa) (MPa) (mm) (mm) (mm) (mm) (mm) (kN)

C3_LC1
C3.1 IPE 400 355.0 355.0 400.0 180.0 13.5 373.0 8.6 199.8 Beam flexure
C3.2 IPE 500 355.0 355.0 500.0 200.0 16.0 468.0 10.2 335.4 Beam flexure
C3.3 IPE 400 275.0 440.0 400.0 180.0 13.5 373.0 8.6 247.6 Beam flexure
C4_LC1
Plate
C4.1 IPE 400 355.0 355.0 427.0 180.0 13.5 320.0 8.6 160.4
Buckling
Plate
C4.2 IPE 500 355.0 355.0 532.0 200.0 16.0 320.0 10.2 258.8
Buckling
C4.3 IPE 400 275.0 440.0 427.0 180.0 13.5 320.0 8.6 169.5 Beam flexure
C3_LC2
Joint Panel
C3.1 IPE 400 355.0 355.0 400.0 180.0 13.5 373.0 8.6 119.0
Shear
Joint Panel
C3.2 IPE 500 355.0 355.0 500.0 200.0 16.0 468.0 10.2 179.3
Shear
CHS
C3.3 IPE 400 275.0 440.0 400.0 180.0 13.5 373.0 8.6 118.0
Punching
C4_LC2
Joint Panel
C4.1 IPE 400 355.0 355.0 427.0 180.0 13.5 320.0 8.6 104.0
Shear
Joint Panel
C4.2 IPE 500 355.0 355.0 532.0 200.0 16.0 320.0 10.2 155.1
Shear
Joint Panel
C4.3 IPE 400 275.0 440.0 427.0 180.0 13.5 320.0 8.6 100.1
Shear

I-beam flanges in C3 thanks to the anchorage provided by the continuous web. So, the fracture is
triggered solely by the bending of the through I-beam flanges just outside the CHS column sur-
face, as evidenced by the von mises equivalent stresses and strains shown in Figure 7. Therefore,
in all the C3 connection models under LC1, analytical calculations considering only the plastic
flexural resistance of the through I-beam at the face of the CHS column, was suitable to predict
the ultimate resistance. However, the through flanges also eventually reached the yield stress, thus
confirming the effective transmission of the bending moments.
For the C4 connections under LC1, two different types of failure were observed from the
numerical analyses: (1) through flange plate buckling under compression (Figure 8a, 8b, 8d
and 8e) and (2) flexural failure of the main I-beams with eventual plastic flange buckling near the
main beam-to-through flange plate connection zone (the furthest connection line shared between
the through bottom flange plate and the main beam flange) as shown in Figure 8c and 8f. The
first type of failure matches the one obtained from C4-LC1 tests, as mentioned in section 2. These
two failure modes can be easily predicted by the following force-transfer mechanism:
The through flange plate primarily resists the compressive stresses developed by the vertical
loads applied at the extremities of the main beams. Failure Mode (1) occurs when the through
flange plate fails to provide sufficient resistance to the joint panel, and therefore buckles inside the
CHS column. Contrary to the C3 connection, as the through web plate is not directly connected to
the through flange plates in the C4 connection, it cannot provide anchorage to resist the through
flange plate buckling. After this, the forces are redistributed to the through web plate via the CHS
column. When the web plate loses its capacity, it buckles under compressive stresses due to bend-
ing. Finally, the ultimate failure occurs due to compressive crushing of the CHS column wall.

316
Figure 5. Free body diagram of forces at joint panel for (a) C3 under gravitational loading (LC1) (b)
C3 under opposite bending loading (LC2) (c) C4 under gravitational loading (LC1) (d) C4 under opposite
bending loading (LC2).

Figure 6. Vertical force-displacement curve comparisons for C3 and C4 under (a) LC1 and (b) LC2.

Figure 7. Von mises equivalent stresses obtained at failure under LC1 for (a) C3.1, (b) C3.2, (c) C3.3
and von mises equivalent strains obtained at failure under LC1 for (d) C3.1, (e) C3.2, (f) C3.3.

317
Figure 8. Von mises equivalent stresses obtained at failure under LC1 for (a) C4.1, (b) C4.2, (c) C4.3
and von mises equivalent strains obtained at failure under LC1 for (d) C4.1, (e) C4.2, (f) C4.3.

However, Failure Mode (2) occurs if the through flange plates are strong enough to provide
the necessary resistance. In this case, the main beams eventually fail due to flexural plasticity
just outside the main beam-to-through flange plate connection zone. The force-transfer mech-
anism for C3 and C4 under LC1 can be visualized from Figure 5a and 5c respectively.
Therefore, the lack of connection between the through plates inside the CHS is seen as the
main governing factor for the behaviour of the connection C4 in comparison with the C3
type. While in the C3 connection the web of the through I-beam is stiffening the flanges
against local elastic buckling under compressive stresses, the through web plate is unable to
provide the same in the C4 connection thus making it vulnerable towards a premature flange
plate buckling inside the CHS. This phenomenon therefore reduces the ultimate resistance of
the C4 connection compared to the C3 one. As shown in Table 2 and Figure 6a, C3.1 and
C3.2 have greater joint strengths than C4.1 and C4.2 respectively.
So, as an additional investigation point, an additional case was analysed with an increased
yield limit in to avoid flange buckling in configuration C4 (model C4.3 where the through plates
were made of a stronger steel grade S450, fy,in = 440MPa and the main beams as well as the CHS
column of a weaker steel grade S275, fyb = fyc = 275MPa). As predicted by the analytical calcula-
tions shown in Table 2 for C4.3, the numerical prototype was observed to avoid the through
flange plate buckling and therefore to fail due to flexural plasticity in the main beams (Figure 8c).
For a relevant comparison, a C3.3 connection was also modelled numerically. The through
I-beam was modelled with S450 ( fy,in = 440MPa) steel and all the other components were mod-
elled with S275 ( fyb = fyc = 275MPa) steel. As the failure occurred similarly just outside the CHS
column, the ultimate resistance was determined by calculating the flexural resistance of the beam
corresponding to S450 steel. The C3.3 was again observed to provide a larger resistance than
C4.3 in terms of vertical forces acting at the furthest extremities of the main beams. Therefore, C3
could be established as a stronger LASTEICON configuration than a C4 connection with equally
thick through flange and web plates as well as the CHS column. Although the C4 connection
might prove to be more economical for the constructional industry than the C3, if both of these
LASTEICON “passing-through” connections are compared from a practical design perspective,
C3 might appear to be the preferred one as it provides comparatively more resistance and can
also avoid a plate buckling, thus escaping a brittle failure as well as the substantial amount of
analytical calculations required to calculate it.

3.2 Comparison under monotonic opposite bending loading, LC2


A totally different force transfer mechanism is activated for the models under LC2, as shown in
Figure 5b for C3 and Figure 5d for C4 connections. Yielding occurred simultaneously in both the
through I-beam web (or through web plate) as well as the CHS column surface as illustrated in
Figure 9. The antisymmetrical vertical loads applied at the furthest extremities of the main beams
created moments in opposite directions at each side of the joint panel. These were resisted by
a combination of longitudinal shear resistance provided by the through I-beam web (resp. web
plate) and the transverse tensile/compressive resistance of the CHS chord face. As a result, the
web yielded in the “passing-through” zone with consecutive yielding of the CHS chord face

318
Figure 9. Von mises equivalent stresses obtained at failure under LC2 for (a) C3.1 and (b) C4.1
connection.

Figure 10. Shear stress (kN/mm2) distribution in (a) the through I-beam in C3.1 and (b) the through
web plate in C4.1 at failure under LC2.

surrounding the flange connection zone in tension for both configurations. In all cases, C3 was
observed to provide a greater resistance compared to a similar C4 connection, as shown in Figure
6b. Although the CHS chord provided equal resistance (analytically as well as numerically) for
each corresponding pair of C3 and C4 configurations (i.e. C3.1 and C4.1, C3.2 and C4.2 and
C3.3 and C4.3), a substantial difference was noticed in the shear behaviour of the through I-beam
web in C3 and the through web plate in C4 respectively. This is related to a different shear stress
distribution in the given elements. While the shear stresses in the through I-beam web (C3) were
observed to develop according to a uniform rectangular distribution throughout the whole section
in both vertical and longitudinal directions (flanges anchor the web thanks to a continued link) as
shown in Figures 5b and 10a, it failed to do so along the vertical direction of the through web
plate in the C4 connections, as illustrated in Figures 5d and 10b. The shear stress distribution
along the vertical direction of the through web plate was noticed to be rather parabolic instead of
a uniform rectangular one as illustrated in Figure 5d. This is also combined with significant flex-
ural stresses developing at the four corners of the through web plate, therefore further limiting the
development of the uniform shear stress distribution along its vertical axis. So, a parabolic shear
stress distribution was considered along the vertical direction of the through web plate for the C4
τv;avg
¼ 1:5ypffiffi3. As a consequence,
f
connections, where the maximum allowable shear stress is τv;max ¼ 1:5
according to the Cauchy Reciprocal Theorem, to maintain equality in shear stresses, this phenom-
enon limits the development of the shear stresses in the longitudinal direction of the through web
plate. Although the shear stresses develops in a rectangular manner along the longitudinal direc-
tion of the through web plate (τh,avg in Figure 5d) – their values are reduced to
τv;avg
instead of pyffiffi3.
f
τh;avg ¼ τv;max ¼ 1:5
Additionally, the length of the through web plate is slightly reduced than that of the through
I-beam web due to its eccentric position (as denoted by “e” in Figure 1b). These two differences
therefore reduced the resistance offered by a C4 connection compared to a similar C3 connection.

4 CONCLUSIONS

This paper compares two innovative open-to-CHS connections, namely the LASTEICON C3 and
C4 connection, which consist of main beams connected to each side of the CHS column by means

319
of a through I-beam or two transverse and one longitudinal plate passing through the CHS
column via Laser Cut slots. The force-transfer is identified through multiple numerical models ini-
tially calibrated with experimental prototypes. For a monotonic gravitational loading, the C4 joint
showed crucial vulnerability towards a through flange plate buckling whereas, in the C3 connec-
tion, such a failure was avoided thanks to the continuous anchorage provided by the directly con-
nected I-beam web. As a result, failure in the C3 connections occurred due to flexural plasticity in
the through I-beam flanges just outside the CHS. For the opposite bending loading, although the
force transfer mechanism and the failure modes were found to be similar for the connections, cer-
tain differences were noticed from a resistance standpoint. While, the through I-beam web in C3
allowed the shear stresses to be distributed in a uniform rectangular manner thanks to the anchor-
ing provided by the I-beam flanges, the through web plate in C4 failed to do so, thus reducing the
resistance of the C4 connection compared to C3. Under both load cases C3 provided a greater
resistance than C4 with approximately same amount of steel and avoided complicated analytical
calculations which might further support their preference from a designer’s perspective. However,
the C4 connection can offer economic advantage as well as easy accommodation to design a four-
way “passing-through” joint where a through beam is inserted through the CHS column in one
direction and the plates can take care of the other (Das et al. 2019b). To further encourage such
LASTEICON “passing-through” connections, a standardized design procedure will be docu-
mented in the near future with detailed experimental and numerical verification.

REFERENCES

Alostaz Y.M. & Schneider S.P. 1996. Connections to concrete-filled steel tubes. A Report on Research
Sponsored by the National Science Foundation. NSF CMS 93-00682.
Castiglioni C.A., Kanyilmaz A., Salvatore W., Morelli F., Piscini A., Hjiaj M., Couchaux M., Calado L.,
Proença J.M., Hoffmeister B., Korndörfer J., Bigelow H., Degee H., Das R., Raso S., Valli A.,
Brugnolli M., Galazzi A., Hojda R. 2016. EU-RFCS Project LASTEICON (Laser Technology for
Innovative Joints in Steel Construction). www.LASTEICON.eu.
Couchaux M., Vyhlas V. & Hjiaj M. 2019. RFCS Project LASTEICON Test Report: Tests on Joints C3
and C4. INSA, Rennes, March 2019.
Das R., Castiglioni C., Couchaux M., Hoffmeister B. & Degee H. 2019a. Design and analysis of laser-cut
based moment resisting passing-through I-beam-to-CHS column joints. Journal of Constructional
Steel Research (Submitted, under review).
Das R., Kanyilmaz A., Couchaux M., Hoffmeister B. & Degee H. 2019b. Characterization of moment
resisting IPE beam-to-CHS column connections by passing-through plates, Engineering Structures
(Submitted, under review).
DIANA User’s Manual, DIANA Release 10.2, May, 2018.
Dutta D., Wardenier J., Yeomans N., Sakae K., Bucak O. & Packer J.A. 1998. Design guide for fabrica-
tion, assembly and erection of hollow section structures. CIDECT Design Guide 7.
Fukumoto T. & Morita K. 2005. Elastoplastic behavior of panel zone in steel beam-to-concrete filled
steel tube column moment connections. Journal of Structural Engineering 131: 1841–1853.
Kanyilmaz A. & Castiglioni C.A. 2018. Fabrication of laser cut I-beam-to-CHS-column steel joints with
minimized welding. Journal of Constructional Steel Research 146: 16–32.
Kurobane Y., Packer J.A., Wardenier J. & Yeomans N. 2004. Design Guide for Structural Hollow Sec-
tion Column Connections. CIDECT Design Guide 9.
Rondal J., Würker K.G., Dutta D., Wardenier J., & Yeomans N. 1992. Structural stability of hollow
sections. CIDECT Design Guide 2.
Wardenier J., Kurobane Y., Packer J.A., van der Vegte G.J. & Zhao X.-L. 2008. Design guide for circu-
lar hollow section (CHS) joints under predominantly static loading. CIDECT Design Guide 1.

320
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Performances of moment resisting frames with slender steel and


composite sections in low and moderate seismic areas

H. Degée
Hasselt University, Belgium

Y. Duchêne
Design office Greisch, Belgium

B. Hoffmeister
RWTH Aachen, Belgium

ABSTRACT: Specific investigations have been carried out regarding typical beam profiles
commonly used for steel and composite frames. In a first stage, experimental tests on class-3 and
class-4 built up steel profiles and composite beam-to-column nodes were performed. The meas-
urement results were evaluated with regard to the development of the hysteretic behavior with
particular emphasis on the cyclic degradation. These test results have been used as reference for
the calibration and validation of numerical models aiming at extending the scope of the experi-
mental outcomes through appropriate parametric variations regarding the behavior of nodal con-
nections as well as towards the global analysis and behavior of structures made of class 3 and 4
profiles. Based on the outcomes of these investigations, practical design recommendations are
finally derived for moment resisting frames located in low and moderate seismicity regions.

1 INTRODUCTION

According to current version of Eurocode 8 only cross-sectional classes 1 or 2 are permitted


for steel or composite structures when a considerable behaviour factor (≥1.5 or 2.0) is intended
to be taken into account. Within moment resisting frames almost all members are affected by
this limitation:
– Single-bay – single-storey frames: Such structures are part of a highly competitive market
requiring a strong optimisation in terms of material requirement and ease of construction.
Light weight steel frames are an optimum solution so far, with the exception of seismic
regions where the above mentioned restrictions strongly limit their applicability despite
their low masses. On the other hand such structures may develop plastic hinges alternatively
in the columns or beams and thus offering the possibility to adjust the static stiffness and
resistance e.g. by means of haunches. In order to remain competitive however both columns
and beams need to be kept slender yielding cross-sectional classes 3 or 4. Such cross-
sections however can develop significant cyclic bending capacity which, although at a lower
level than the elastic one, can be sufficient in order to resist seismic actions.
– Multi-bay – multi-storey composite buildings: Due to the high position of neutral axis of
composite cross-sections in negative bending the web of steel profiles is very often to be
classified as class 3 or 4 section, although almost all rolled steel profiles may be classified
as class 1 or 2 cross-sections. On the other hand in such kind of buildings under seismic
actions there are always both positive and negative moments developing at the corners of
the frame and thus providing a significant resistance and dissipation capability when the
entire system is considered.

321
The above mentioned aspects are of particular importance in countries with moderate seis-
micity regions, where a special adoption of the otherwise common types of structures presents
a great obstacle for the small and medium size companies active in this market. Hence an
extension of the applicability of the afore mentioned structural types to moderate seismic
regions would largely enhance their competitiveness. This leads to the following objectives:
– Extension of the applicability of class 3 or 4 steel cross-sections within single-bay –
single-storey frames to moderate seismic actions. The objective will be achieved by
means of characterization of class 4 cross section in terms of rotational behaviour and
energy dissipation; comparison to class 1 and 2 section with equal static design resist-
ance in terms of cyclic performance considering the influence of local buckling. The
tests will include constant depth and haunched girders with plastic hinges in the beams
or columns. The following numerical studies will comprise the verification of the local
behaviour as well as the performance of complete frames under seismic actions.
– Extension of the applicability of class 3 (web) composite sections to moderate seismic actions.
The objective will be achieved by means of characterization of class 3 composite cross section
in terms of rotational behaviour and energy dissipation; comparison to class 1 or 2 sections in
terms of cyclic resistance considering the influence of local buckling, determination of the
importance of partial class 3 or 4 presence in the cross section (web). The studies will com-
prise experimental investigations as well as numerical simulations aiming at the verification of
the local behaviour of frame corners and in particular at the determination of the interaction
between the positive and negative moment regions in a multi-bay – multi-storey structure.
– Identification and definition of b/t values rather than cross-sectional classes to be limited
depending on the targeted energy dissipation and the corresponding deformations by
means of numerical analyses.

2 EXPERIMENTAL INVESTIGATIONS

2.1 Tests on steel portal frames


The test program considered typical welded profiles typically used for light weight frames of
small industrial or storage halls. In total 6 cyclic tests on frame corners with welded and bolted
connections as well as haunched and constant depth girders were carried out. The test specimens
were designed such that for constant depth girders the plastic hinges developed in the beams
whereas for the haunched girders the plastic hinges developed in the top of the column.
The tests were performed according to the ECCS testing procedure with increasing ampli-
tudes of deformation cycles and were executed until collapse (e.g. rupture due to low cycle
fatigue or extensive local buckling). The measurements were evaluated with regard to the
development of the hysteretic behaviour with particular emphasis on the degradation.
Different examples for moment resisting frames (MRF) have been compiled from literature as
well as from direct dialogue with manufacturers to mirror the state-of-the-art. Following these
examples the geometries, plate thicknesses and joint details have been assigned. Furthermore it
has been decided to test only frame corners as subsystems of a complete MRF and due to the test
setup the length of the beams and columns were limited. All specimens were made as class 3 cross
sections and in order to reach plastic hinges in the beams and columns the panel zones were
reinforced, whereby one additional test on a frame corner without a reinforcement and one add-
itional test on a frame corner made of class 1 cross sections with comparable monotonic resistance
and stiffness were carried out. The 6 different specimens that was decided on were:
– S1 - Welded frame corner with constant girder
– S2 - Bolted frame corner with constant girder
– S3 - Welded frame corner with haunched girder
– S4 - Bolted frame corner with haunched girder
– S5 - Welded frame corner with constant girder - class 1
– S6 - Welded frame corner with constant girder - unreinforced panel zone

322
Figure 1. Overview of the specimens – type of steel corner frames.

Figure 1 shows the above mentioned specimens or types of frame corners respectively.
The boundary conditions of the test setup were designed to represent the cut-out from
a complete MRF; the boundary conditions have been particularly focused on a best possible
representation of the real load-deformation behaviour of a complete MRF. The test setup is
shown schematically in Figure 2.
The detailed results are available in Degée et al. (2018). The main observations can be sum-
marized as follows:
– All test specimens failed after a large number of plastic cycles with deep plastic deform-
ations and formation of local buckles in the beam or column.
– There was visible direct influence of the flange slenderness on the resistance and cyclic
performance observed.
– The bolted connection of specimen S2 led to a strong local loading of the column flanges.
The plastic mechanism was partly similar to the plastic behaviour of T-stub connections.
– Specimens with haunched beam (welded and bolted) led to the formation of plastic
hinges in the columns. The performance was comparable to specimens with plastic mech-
anisms in the beam.
– In test specimen with not reinforced web panel (S6) the final collapse involved plastifica-
tion in the panel zone and cracks in the welding zones including the web panel. The
achieved number of plastic cycles, however, was very high.
– Expect specimen S2, the other specimens failed – after the repeated formation of plastic
local buckles – due to formation of cracks in the areas of high cyclic strains. The cracks
formed in the base material as well as in the HAZ of the welds.

2.2 Tests on composite portal frames


The test program considered typical steel composite beam profiles commonly used for multi-bay -
multi-storey composite frames. In total three cyclic test on frame corners with welded connections
and constant depth girder were carried out. The test specimens were designed such that the plastic
hinges developed in the beams.
The tests were performed according to the ECCS testing procedure with increasing amplitudes
of deformation cycles and were executed until collapse. The measurements were evaluated with
regard to the development of the hysteretic behaviour with particular emphasis on the degradation.

Figure 2. Test set-up with specimen S2.

323
Figure 3. Test set-up for steel-concrete specimens.

The large scale test specimens represent a section of an exterior corner of a composite frame.
The main dimensions in length of column and beam were equal for all specimens and were
chosen so, that the resulting moment distribution is comparable to that of a complete MRF
loaded by vertical and horizontal loads (hinges were set at points of zero moment). All columns
consisted of a HEB 360, Class 1 profile. The girder was joined to the column with double side
full penetration butt welds. The effective width of the concrete slab was determined according
to EN 1998-1. Specimen S01 and S02 were classified as Class 3 Beams under negative bending.
The only difference between these specimens was an additional vertical stiffener in the beam.
The third test consisted of a class 1 Beam, with comparable monotonic bending resistance and
stiffness. The reinforcement arrangements were the same for each specimen, as well as the
arrangement of the headed shear studs. The 3 different specimens that was decided on were:
– Specimen 1 (S01) IPE 450 – Class 3 Beam
– Specimen 2 (S02) IPE 450 – Class 3 Beam, reinforced
– Specimen 3 (S03) HEA 360 – Class 1 Beam
A general overview of the test set-up is given in Figure 3.
The detailed results are available in Degée et al. (2018). The main observations can be sum-
marized as follows:
– All test specimen provided a full composite action for positive and negative bending;
– The achieved maximum resistance in positive and negative direction was in the same range;
– In all specimen plastification in the steel beam was observed only in the lower flange;
– No significant crushing of concrete (slab in compression) was observed;
– Clearly visible cracking of concrete in tension was observed, leading to partial disintegra-
tion of concrete parts; the contribution of the concrete and reinforcement in tension,
however, remained present almost until the end of the tests;
– Plastic buckles in the lower flange were only slightly developed;
– The failure of all three specimens developed – after several plastic cycles – rather fast due
to the formation of a severe crack in the lower flange of the steel beam.
– The evaluated relative resistance and relative absorbed energy were very similar for all
three specimens, the slender cross-sections performed slightly better.
– The absolute absorbed energy showed very similar development, with a slight advantage
for the compact section.

3 NUMERICAL INVESTIGATIONS

The test results presented above have been used as references for the calibration and validation
of numerical models aiming at extending the scope of the experimental outcomes through

324
appropriate parametric variations regarding the behaviour of nodal connections as well as
towards the global analysis and behaviour of structures made of class 3 and 4 profiles.
Two types of numerical models have been calibrated. First, detailed 3D FE models of the
nodal zone have been used. These models have been implemented in ABAQUS resorting to
8-node reduced-integration solid elements (type C3D8R) considering geometric and material
non-linearities. A non-linear kinematic hardening according to Lemaître and Chaboche
(1990) is used to take into account the cyclic degradation. More details on the parameters and
calibration can be found in Degée et al. (2018). These models have essentially been used to
extend the scope of the experimental conclusions summarized in the section 2.
These detailed local models have then been used to calibrate equivalent spring models of
the nodes and/or of the cross-sections implemented in the context of a global model resorting
to one-dimensional plane beam elements (using the FE package FINELG), leading to conclu-
sions on the global performances of the plane frame structures.

3.1 Investigations at node level


The numerical simulations were performed in order to determine the performance of slender sec-
tions in the plastic range and to compare them to the performance of similar compact sections.
A total of 45 parametric variations of the steel test specimens and of 4 parametric variations of
the composite specimens have been carried out.
Regarding the steel nodes, the extension of the tests by numerical simulations comprised the
following parameters:
– Flange thickness (class 1. . . 4) with constant depth and width of profiles,
– Web thickness (4 . . . 12 mm, the web remained class 1) with constant depth and width of
profiles,
– Profile depth with constant width and plate thickness, web (6 mm) class 1 . . . 4
Regarding the composite nodes, the extension of the tests by numerical simulations com-
prised the following parameters:
– Variation of the longitudinal reinforcement in the slab (decreasing from 26 cm² to 6,5 cm²)
– Variation of the concrete strength (decreasing from 54 N/mm² to 28 N/mm²)
A detailed assessment of the results is available in Degée et al. (2018). The main observa-
tions can be summarized as follows:
Variation of beam flange thickness from class 1 to 4:
– The absolute amplitudes until reaching the maximum moment failure decreased signifi-
cantly with increased slenderness;
– The number of cycles until reaching the theoretic failure due to accumulation of plastic
strains decreased from 22 to 17;
– The degradation after exceeding the maximum moment resistance was more severe for
compact flanges, as observed from the hysteresis;
– The relative resistance (compared to yield initiation, evaluated according to [2]) decreased
slightly with for higher slenderness. The strength degradation gradient for compact sec-
tions was significantly higher than for slender sections;
– The relative absorbed energy was the highest for compact sections; however, the decrease
of the absorbed energy for higher slenderness was moderate;
– The absolute absorbed energy decreased with higher slenderness – this is obvious, as
these values consider also the absolute resistance. Generally, until a partial ductility value
of 4 to 6, for each cross-section the total absorbed energy increased constantly.
Variation of the web slenderness from class 1 to 4 (constant depth of the beam):
– A clear change in the shape of the plastic zone was observed. With increasing slenderness
of the web shear buckling became governing the ultimate resistance. Although this failure

325
mode prevents the achievement of plastic moment in the beam cross-section, the cyclic
performance offered a high amount of dissipation capability.
– The numerically achieved ultimate resistance of the frame eaves connections exceeded the
nominal, characteristic resistance corresponding to the ultimate shear force. This indicates
that the web was able to develop significant tension field action, similar to shear links;
– The relative resistance of all members achieved a similar level; the evolution of the hyster-
esis shows an increasing asymmetry with increasing web slenderness.
– The relative absorbed energy was also similar in all cases, with a rather clear decrease
after achieving the maximum resistance
– The absolute dissipated energy decreases with increasing web slenderness, according to
the decreasing ultimate resistance. Until the partial ductility value of 4 a rather stable
accumulation of the total absorbed energy can be observed;
– The expected number of cycles until crack-formation was nearly constant (22 . . . 25) with-
out a clear tendency regarding the web slenderness.
Variation of the beam depth (web slenderness from class 1 to 4):
– Similar to the above parameter study, the increasing slenderness of the web leads to
a change of the failure mode from flexural bending to plastic shear buckling of the web.
– Observations b) to d) and f) of the previous study apply;
– The absolute absorbed energy was very similar for all members until the partial ductility
value of 4.
Variation of reinforcement:
– The achieved relative resistance was the same for all members;
– The achieved relative absorbed energy increased with lower reinforcement ratio. This indi-
cates, that with less reinforcement an earlier plastification of the steel profile occurred;
– The absolute absorbed energy was nearly identical for all members;
– The expected number of cycles was nearly constant.
Variation of concrete strength:
– The variation of concrete strength did not lead to significant changes in the performance
(relative resistance, absorbed energy).
– As expected, the ultimate absolute resistance of the member with the lowest concrete
strength reduced slightly in the positive direction (concrete in compression);
– For negative moment direction no influence of the (tensile) concrete strength was observed;
– The expected number of cycles remained nearly constant;
Haunched steel profile (compared to specimen S01):
– The plastic zone moved (as expected) from the area of beam-to-column welds to the
beginning of the haunch;
– The relative resistance remained nearly unchanged;
– The relative and absolute absorbed energy increased slightly;
– The expected number of cycles until failure did not change significantly; however, the
maximum strains are observed in the base material of the lower flange. In this region the
disadvantageous influence of the welds shall be minimised, despite the presence of the
welded haunch.

3.2 Investigations at frame level


A number of realistic frames is defined, whose design is governed by seismic actions (moderate
seismic zone) for different number of storey (ns = 1 to 6/nb = 1) “case studies A” or bay (nb = 1
to 3/ns = 3) “case studies B”.
For these case studies, beams are generally designed by stiffness conditions (minimum iner-
tia needed to respect limitations of interstorey drift) and columns are contrarily designed by

326
resistance conditions and ductility conditions (WBSC checking: minimum resistance of col-
umns in comparison with beams “Wy,pl,b”/ULS checking: resistance with the increase of loads
by overstrength factor “Ωb”).
Based on a final design following strictly Eurocodes 8 rules for DCM, required geometrical
characteristics for beams “Wmax.y” and “Imin.y” (flexural modulus and second moment of
area) were deduced for each case study for several size of column. For each coupe of condi-
tions, several possible cross-sectional shapes have been evaluated, with profiles being classified
in different cross-section classes.
The actual performances of these frames are then evaluated by mean of Non-linear Time-
history analyses. The main observations of these analyses are as follows:
Dissipative zones: They are located at beam to column joints for negative bending moment,
at ≈ 1/4 or ≈ 3/4 span for positive bending moments and at the bottom of column. Four dissi-
pative zones are observed by floor but only two are activated at the same time (one at the
node and one in the span). So 26 different zones (6 × 4 + 2) could be potentially used for dissi-
pate energy in the reference case study with 6 storeys. The most activated dissipative zones are
located on floors 2 to 4.
Recorded moment/rotation curves: Despite the difference of cross-section classes, shapes of
curves are quite similar. It can be noted that:
– Rotation capacity in moment/rotation (M/ф) curves is smaller for cross section class 1
(0.059 rad/beam 2.10) than for cross section class 3 (0.068 rad/beam 2.2), if inertia and
bending modulus of beams are equal.
– Maximum bending moment in moment/rotation (M/ф) curves is greater for cross section
class 1 (overstrength: ≈ 153% / beam 2.10) than for cross section class 3 (overstrength: ≈ 133%
and 135% / beams 2.2 and 2.1).
Rotation demand: maximum rotation capacities is never reached. The rotations
observed are well below the maximum rotations capacities of main beams (18% to 35%
is actually exploited), as illustrated in Figure 4. Due to the rather high effect of the
gravity loads on the beams, the rotation demand is essentially concentrated in the dissi-
pative zone at beam end, while the ‘in span’ dissipative zone is almost not leaving the
elastic range.

Figure 4. Rotation demand in corner springs (case-study with 6 levels).

327
4 CONCLUSIONS

Weak Beam Strong Column criteria seem essential in the design procedure, because forces in
the columns are larger with step by step dynamic analysis. Therefore, dynamic instabilities in
columns could be a weak point in the structure and should be prevented. Moreover, a high
overstrength is observed in moment/rotation (M/ф) curves due to the strain hardening of
beams. This point is related to the assumption that beam to column joints are resistant and
ductile enough to avoid brittle failures in joints. “Weak Beam Strong Column” criterion leads
to create more dissipative zones in beams.
For case studies including slender sections (cross section class 3), structures designed based
on q = 4 proved to exhibit an appropriate behaviour (i.e. sufficient ductility and rotation cap-
acity, reduction of seismic effects thanks to additional spring flexibility):
– No local/global mechanisms;
– No excessive rotations (failures in beams);
– No excessive interstorey drifts;
– No instabilities for the columns in frame plane.
Based on the incremental dynamic analyses, an estimation of the actual available behaviour
factor for the designed frames has been achieved as ranging from 4,4 to 4,9. These values of
behaviour factors (q > 4) are mainly explained by the presence of many dissipative zones on
case studies (up to 26 dissipative zones per frame), by a high overstrength in beams (133% to
153% rate of design bending moment to effective resistance) and by a relatively low demand
on supply ratio in terms of rotation capacities in beams (demand of 18% to 35% rate of max-
imum rotation capacities), even for beams cross sections class 3.
As general outcome of the research project, these observations have been formulated as
practical design rules that could be implemented in future revisions of the seismic design codes.

ACKNOWLEDGEMENT

This research has been carried out with the support of the Research Fund for Coal and Steel
(RFCS) of the European Commission under the grant agreement RFSR-CT-2013-00022.

REFERENCES

Bisch P, Carvalho E, Degee H, Fajfar P, Fardis M, Franchin P, Kreslin M, Pecker A, Pinto P, Plumier
A, Somja H, Tsionis G. (2012). Eurocode 8: Seismic Design of Buildings worked examples, JRC tech-
nical report. Publications Office of the European Union/Joint Research Centre, Luxembourg.
Degée, H. et al. “Meakado - Design of steel and composite structures with limited ductility requirements
for optimized performances in moderate earthquake areas” – Final report. RFCS publications (2018).
ECCS. Recommended Testing Procedure for Assessing the Behaviour of Structural Steel Elements under
Cyclic Loading, ECCS - Technical Committee 1 - Structural Safety and Loadings, Technical Working
Group 1.3 - Seismic Design, Brussels, Belgium (1986).
Eurocode 8: Design of structures for earthquake resistance - Part 1: General rules, seismic actions and
rules for buildings; EN 1998-1:2004 + AC:2009.
Elgahzouli, A. Assessment of European seismic design procedures for steel framed structures, Bulletin of
Earthquake Engineering, 8, N° 1 (2010), 65–89.
Kato B. Rotation Capacity of H-Section Members as Determined by Local Buckling, Journal of Con-
struction Steel Research 13 (1989), 95–109.
Lemaître, J. and Chaboche J-L. Mechanics of Solid Materials, Cambridge University Press (1990).
Sedlacek, G. and Feldmann, M. Background Document 5.09 for chapter 5 of Eurocode 3, Part 1.1, The
b/t-ratios controlling the applicability of analysis models in Eurocode 3, Part 1.1. (1995).

328
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Laser technology for innovative connections in steel construction:


An overview of the project LASTEICON

H. Degée
Hasselt University, Belgium

A. Kanyilmaz & C. Castiglioni


Fincon, Italy

L. Calado
IST Lisbon, Portugal

M. Couchaux
INSA Rennes, France

B. Hoffmeister
RWTH Aachen, Germany

F. Morelli
University of Pisa, Italy

ABSTRACT: Circular hollow sections present several advantages, such as uniform behavior in
all directions or possibility to obtain composite behavior by concrete infill, and hence high strength,
stability and good fire resistance. Moreover, their aesthetic appeal has a great potential for decision
makers (architects, building owners…). However, their adoption in the practice often suffers from
the complexity and high cost of their joint detailing. The project Lasteicon aims at proposing
a novel and simple jointing solution with reduced complexity and fabrication costs, to be adopted
for connections between CHS-columns and I-beams or for truss structures with CHS chords. The
proposed solution relies on a versatile laser cutting and welding technology. The present contribu-
tion gives a broad overview of different aspects of the project: assessment of fabrication procedures
and tolerances, experimental testing, numerical investigation and comparison with traditional
design and fabrication methods in terms of structural performances in normal and cyclic loading.

1 INTRODUCTION

Modern laser tube cutting machines are capable of making any type of cut on steel tubes, in
entirely automatic, programmed cycles. CAD programming and the laser beam eliminate the
traditional fixed costs determined by punches, clamps, tools, templates and dies. The laser cut-
ting technology (LCT) has so far generated significant benefits in both processing and cost
terms, for a wide range of applications. ADIGESYS, partner of the project reported in this
paper and provider of laser cutting machines, reports that their customers obtained production
improvements in the range from 70 to 80% through LCT, with respect to conventional pro-
cesses. This was mainly thanks to more stringent machining tolerances resulting in improved
quality of joints, fittings, easier fastenings, process efficiency and vast flexibility (see also Aloke
et al., 1997 and Harnicarova et al, 2010).
Circular hollow sections (CHS) have excellent properties with high compression, tension and
bending resistance in all directions, thanks to their inherent shape and geometrical properties.
Structures produced using tubular sections have lighter overall weights in the order of 40% and

329
Figure 1. Traditional joint (a) and Lasteicon solution with I-beam passing though CHS elements (b).

they require smaller volume of fire protection material than their equivalent H section. Indeed
CHS columns can be designed to have a fire resistance up to 120 minutes, without using fire pro-
tection. Moreover, tubular profiles also have reduced transport costs. They are also architecturally
very appealing offering more space, and freedom to the designers. However, these profiles have
complex detailing, fabrication and erection requirements for their connections, always requiring
stiffeners and gusset plates (see e.g. Schneider et al, 1998, or Elremaily et al, 2001). From an exten-
sive literature study including the largest steel users in the world construction market such as
Europe, to be found extensively in Kanyilmaz et al (2018), United States and Japan, it is seen that
most common types of I-beams-to-CHS columns connections are directly welded, or diaphragm
connections. Both solutions include vast amount of welded local stiffeners or gusset plates, which
makes the design and construction of these joints complicated and expensive, slow down the
design and construction process and spoil the aesthetics of the design.
The solution investigated within th project LASTEICON proposes using LCT in the fabri-
cation of I-beam-to-CHS-column joints. This can drastically reduce fabrication costs, as well
as meeting the structural requirements, expanding also the freedom of architects and engineers
when developing new projects. Figure 1 shows a prototype of laser-cut joint (b) compared
with a typical diaphragm joint (a).
The project LASTEICON, entering its last financed year at the moment of writing the pre-
sent contribution, has as primary objective to overcome the complexity of steel joints between
CHS-columns and I beams, and let engineers and architects exploit the outstanding structural
and architectural properties of hollow section profiles. This is achieved making use of laser cut-
ting technology (LCT) in the shop fabrication of steel joints. Laser cutting allows reducing
shop-fabrication time, obtaining better precision with higher quality in joint fabrication, and
improving the workplace safety with less manual work and more computer-programmed auto-
mation. A preliminary estimate made by industrial partners involved in the research proposal
shows that, applying LCT, fabrication costs of steel structures can be reduced by 8-9%. Further-
more, shop fabrication will become more environmental friendly by releasing much less slag.
Therefore, thanks to LCT, design and fabrication of steel joints will be much simpler, faster,
more economic, and steel joints will become more aesthetically appealing and eco-friendly.

2 FABRICATION PROCEDURE AND TOLERANCES

For the development LASTEICON joints, the first step was to investigate the fabrication details
including the tolerances required for the slots, laser cutting parameters, and welding aspects (see
Kanyilmaz et al, 2017 and 2018). Then based on the observations from this investigation, final
design of joint specimens has been made. The whole fabrication process applied to the joints
assembled with different column and beam sizes and welding types has been described,

330
quantifying the time and resources spent during the process. Furthermore, a detailed description
of laser cutting procedure has been provided, to show its potential in the steel construction sector.
In case of fabrication with LCT, the selected cutting tolerance values can minimize the slot
sizes to optimize the welding quantity and take full advantage of the LCT accuracy, at the same
time facilitating the fabrication. The LCT allows cutting a “slot” in a steel profile, with a very
high precision (with an accuracy of 10 μm). Laser cutting machines are also able to notice the
differences between the expected profile and the real profile, automatically calculating the neces-
sary compensations. The process can hold quite close tolerances, often within 25 μm. The “slot”
in the CHS-column cannot be cut “a priori”, “sized” on the “nominal geometry” of the profile,
because the profile is affected by geometric imperfections (geometric tolerances), that are regu-
lated by EN 10034:1995, which provides the allowable size scatters for each nominal dimension
of the profiles (e.g. flange thickness, flange width, depth, web thickness, out-of-square, web off-
centre). This tolerance issue can be dealt in two ways. The “slot” can be oversized with respect to
the “nominal” size of the I-beam, to facilitate assembly operation or the “slot” can be “sized” on
the “actual geometry” of the I-beam profile by “measuring” the “actual geometry” of the I-beam
to be inserted through the CHS-column (manual measurement or laser scanning). In Kanyilmaz
et al (2018), the results of an experimental study analyzing 5 types of beam profiles ordered from
different steel suppliers have been presented, investigating these two options with the aim of opti-
mizing the cut length, welding quantity, and easiness of assembly. Furthermore, laser cutting
parameters and welding issues have been examined and discussed.
These fabrication studies are illustrated in Figure 2.
Adopting LCT, special connections can be easily designed to join different types of struc-
tural elements, with pleasant aesthetic results as well as improving global structural integrity,
without the need of costly prototypes being used for evaluation purposes. It ensures better uni-
formity of parts, less machine tooling time and most cost-effective management as well as
excellent performance in terms of cutting quality. Laser cutting can be up to thirty times faster
than standard methods. As a first step to a better quantified approach, a reference multi-storey
building already designed and constructed with CHS columns and I beams according to Euro-
pean Standards is taken as a reference to evaluate the added value of the new jointing proced-
ure. This leads to the set of CHS to I beams connections investigated in the next sections.

Figure 2. Pictures from the fabrication studies.

331
3 EXPERIMENTAL TESTING

In a first experimental stage, 4 different types of connections are investigated. Two of them can
be considered as nominally pinned (Configurations C1 and C2 – see Figure 3.a and 3.b) and two
can be considered as moment-resisting (Configurations C3 and C4 – see Figure 4.a and 4.b). For
each configuration, different CHS diameter and thickness are investigated, as well as different
types of welding (full penetration or fillet welds). Each connection is evaluated for two types of
loading: (i) symmetrical loading, corresponding to a global vertical loading and (ii) unsymmetrical
loading corresponding to a global horizontal loading. For the unsymmetrical loading, tests are
carried out monotonically and cyclically.

Figure 3. Configuration C1 resp. C2 – nominally pinned connection with passing-through (a) IPE pro-
file (b) web-plate.

Figure 4. Configuration C3 resp. C4 – moment-resisting connection with passing-through (a) IPE pro-
file (b) flange- and web-plates.

332
As a matter of illustration, Figure 5 shows the two typical failure mode of these connec-
tions, i.e. failure or the IPE profile out of the CHS and failure of the weld zone at the interface
CHS-flange of the IPE.
The main observations obtained from the experimental testing can be summarized as follows:
• For simple joint configurations (C1 and C2), the failure is controlled by a combination of
buckling of the crossing element in bending (plate or profile), connection plate outside the
CHS and local crushing of the tube wall. Design rules matching these observations are
derived and available in Kayilmaz et al. 2019;
• For moment resisting configurations with passing-through beam, symmetrically loaded,
failure is controlled by plastic local buckling of the beam flange, with no significant influ-
ence of the weld type and tube thickness with a part of the ductility activated by yielding of
the inside part of the crossing I-profile;
• For moment resisting configurations with passing-through plates, symmetrically loaded,
the failure is controlled by the local buckling of the compression flange plate and/or web
plate, coupled with local crushing of the tube;
• For moment resisting configurations with passing-through beam or plates, unsymmetrically
loaded, failure is triggered by weld tearing in the HAZ or tube wall crushing. The resistance
is therefore significantly influenced by the thickness of the tube;
• A comparison between a conventional configuration with beams welded to the tube with
no crossing elements and a Lasteicon configuration shows that the performance of the Las-
teicon system with a CHS thickness of 10 mm matches the performances of a conventional
configuration with 22 or 16 mm depending on the loading direction (symmetrical resp.
unsymmetrical, as illustrated in Figure 6.
Following this first set of tests on relatively simple 2-way connections, 3 additional configur-
ations are also investigated within the project LASTEICON, namely:
– Four-way joints, i.e. one CHS column crossed by IPE beams in the two perpendicular dir-
ections (See Figure 7)
– 2 way joints in a steel-concrete configuration, i.e. node with steel-concrete composite
beams, with or without concrete infill in the CHS column
– Trusses (See Figure 8).
Due to the limited length of this paper, no details are however given herein on these add-
itional configurations. More can be found in Kanyilmaz et al. (2019).

Figure 5. Typical failure modes (a) IPE profile outside the CHS column (b) Weld zone.

333
Figure 6. Comparison of the performances of a conventional to a Lasteicon moment-resisting connec-
tion (force-displacement curves).

Figure 7. Four-way configurations.

Figure 8. Truss configurations.

334
4 ANALYTICAL AND NUMERICAL INVESTIGATIONS

In view of enlarging the scope of the experimental results and to validate practical design pro-
cedures, all the tested configurations are also investigated by numerical modelling, as illus-
trated in Figure 9 for two-ways configurations C3 as well as truss configurations.
After a due calibration, the following parameters are investigated:
• Size of the crossing profile or plates
• Thickness of the tube
• Diameter of the tube
• Material properties
• Moment to shear ratio of the loading applied to the connection
Next to these numerical models, analytical models suitable for the practical design of the
Lasteicon solutions are also proposed based on simple mechanical load transfer models, as
illustrated for instance in Figure 10 for a moment resisting configuration with crossing plates.
Detailed results on moment-resisting configurations are presented in a companion paper
also submitted to SDSS conference (Das et al, 2019). As a matter of illustration, Figure 11
shows the influence of a variation of the CHS column thickness for C4 configuration.

5 CONCLUSIONS AND PERSPECTIVES

The results achieved so far within the project Lasteicon in terms of feasibility and perform-
ances of the LCT for innovative CHS columns to IPE beams are found very promising. They
cover a wide range of applications, from simple two-way joints for steel structures to moment
resisting four-ways joints in steel or steel-concrete context, as well as truss configurations.
Based on a wide experimental and numerical database, reliable design procedures are final-
ized and the economical and life-cycle performances is derived as well. All the outcomes and
their resulting practical implementation are planned to be presented during a specific work-
shop to be organized in the autumn of 2019.

Figure 9. Numerical models (C3 and truss).

Figure 10. Simplified load transfer model (a) symmetric loading (b) unsymmetric loading.

335
Figure 11. Influence of the CHS column thickness (configuration C4).

ACKNOWLEDGEMENTS

The project LASTEICON is carried out with the support of the Research Fund for Coals and
Steel (RFCS) of the European Commission, under grant agreement 709807.

REFERENCES

A. Elremaily, Atorod Azizinamini, Design provisions for connections between steel beams and concrete
filled tube columns, Journal of Constructional Steel Research 57 (2001) 971–995.
Castiglioni, C.A., Kanyilmaz, A et al. Laser Technology for Innovative Connections in Steel Construc-
tion. Research report, 2019.
Das R., Kanyilmaz A, Degée H. Comparison of two different innovative solutions for IPE beam to CHS
column connections. Proceedings of the SDSS 2019 conference, 2019.
Harničárová et.al., Comparison of different material cutting technologies in terms of their impact on the
cutting quality of structural steel, Tehnički vjesnik 17, 3(2010), 371–376.
Kanyilmaz A., Castiglioni C.A., Raso, S., Valli A., Brugnolli M., Galazzi A., Hojda R., Circular hollow
section joint fabrication using laser cutting technology: Tolerance assessment, Taylor & Francis (Conf.
proc.), Tubular Structures XVI – Heidarpour & Zhao (Eds), Melbourne, Australia, 4–6 December 2017,
p. 631–637, ISBN: 978-0-8153-8134-1, 2018.
Kanyilmaz A., Castiglioni C.A., Brambilla G., Gjoka K., Galazzi A., Raso S., Valli A.,
Brugnolli M., Hojda R., Experimental assessment of tolerances for the fabrication of laser-cut
steel joints, Ernst & Sohn (Conf. proc.) EUROSTEEL 2017, September 13-15, 2017, Copen-
hagen, Denmark, Ernst Sohn Verlag fr Architektur und technische Wissenschaften GmbH
Co. KG, Berlin, CE/ papers, 2017.
R. Aloke et al., A model for prediction of dimensional tolerances of laser cut holes in mild steel thin
plates, Int. J. Math. Tools Manufact. Vol. 37, No. 8, pp. 1069–1078, 1997.
Schneider and Alostaz, Experimental behavior of connections to concrete filled steel tubes, JCSR 45 (3):
1998, pp 321–352.

336
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Analytical assessment of CFS wall-panels sheathed with


MgO board

Abhinav Dewangan & Govardhan Bhatt


National Institute of Technology, Raipur, India

Chanchal Sonkar
CSIR-Central Building Research Institute, Roorkee, India

ABSTRACT: Currently, Cold-formed steel (CFS) framed structures are being utilized at a large
scale for residential as well as commercial purposes. CFS wall panels (CFSWP) are conventionally
sheathed using Gypsum boards (GB) and Oriented Strand boards (OSB). Use of MgO (Magne-
sium-Oxide) boards have also been started recently for sheathing. Few studies have been per-
formed using MgO boards sheathing. This paper presents an analytical study on CFSWP with
MgO board sheathing and effect of parametric variation on strength of wall-panels’. In this study
a total of 49 specimen has been evaluated analytically. The analytical study presented is done in 3
steps (1) Calculating the fastener-sheathing stiffness parameters (2) Calculating the load factors
for the considered section through finite strip method (CUFSM tool: Constrained and Uncon-
strained Finite strip method; version 4.05) (3) Calculating nominal load (strength) by the Direct
Strength Method (DSM) of AISI S-100 has been utilized. In the end the available strength of the
CFSWP has been evaluated by load and resistance factor design method (LRFD) method. The
aspect ratio of the wall-panels analyzed for this study is ranging from 1:1 to 1:2 with a frame
height of 2400 mm. In this study, it has been found that the strength of wall panel with stud spa-
cing 600 mm, screw spacing 150 mm and aspect ratio of 1:1 is having 45% greater strength than
bare frame wall panel of same specification. The paper also compares a strength variation of wall-
panel with MgO Board Sheathing with respect to CFSWP with GB and OSB sheathing.

Keywords: Cold-Formed Steel Wall-Panels (CFSWP); Magnesium Oxide (MgO) Board;


Direct Strength Method (DSM); CUFSM.

1 INTRODUCTION

In the past two decades there has been a lot of theoretical as well as practical research has been
taking place in the field of the CFS framed structures has been done including testing the axial
bearing capacity of CFSWP and the performance of the CFSWP when the shear force is acting
over it. The basic concept behind the testing of CFSWP is to ascertain its behavior and capacity
to bear the desired load. In the recent trends of low-cost housing, mass-housing, rapid industrial
building the utilization of the CFS based technology has been observed a giant leap in the field.
The significance of utilizing the CFS technology in the current scenario is its eco-friendly nature
and high strength to weight ratio as compared to other means of construction and development
of building structures. In terms of load bearing, sustainability, recyclable nature, rapid construc-
tion and development along with the high safety (Fiorino et al. 2014) like features it has upper
hand over other technologies available in civil engineering field.
In the earlier attempt by Ye et al. 2016 in the area of the CFS based column sections for the
best combination of screw spacing and stud spacing CFS wall-panels with Oriented Strand
Boards (OSB), Gypsum Boards (GB), and Fiber Cement Boards (FCB) sheathing has been

337
utilized here as the benchmark for pre-validation and comparison of the analytical results calcu-
lated at par in the end. The c-section with the edge stiffeners plays vital role in improving the
strength (Zhang et al. 2007) whether they are inclined stiffeners or perpendicular. The edge stiff-
eners increase the stability of the column stud which in return improvises the buckling capacity of
the section. There have been studies like Young & Rasmussen (1998) to check the column load
capacity with the boundary condition of pinned-pinned and fixed-fixed end supports.
The major contribution in assessing the CFS based structure capacity on the basis of its
strength in the analytical format is driven by performing the elastic buckling analysis (Li &
Schafer (2010)) and calculating the nominal load carrying capacity. Here in this study to
assess the desired CFSWP with sheathing applied on the both sides, Direct Strength method
(DSM) has been utilized by: (1) evaluating the stiffness acting on CFSWP bare frame because
of sheathing, (2) evaluating the load factors for the considered section through finite strip
method (CUFSM tool: Constrained and Unconstrained Finite strip method; version 4.05) (3)
Calculating nominal load (strength) using DSM procedure. Later on, to get the actual avail-
able axial strength of CFSWP employing the load and resistance factor design (LRFD)
method. The DSM practice has been recommended by AISI S100 for accessing the considered
CFSWP. The sheathing boards utilized here to brace the (Vieira & Schafer (2013)) wall-panels
using the screws with variation of connection density are Magnesium-Oxide (MgO) boards
(Sinh building solutions (2017)), Heavy-duty Oriented Strand Boards (HDOSB) (Thomas
(2003)) and Gypsum boards (GB) (Rahmanian (2011)). Generally, it has been observed in the
available literatures the boards used for sheathing purpose are OSB, GB and fiber cement
boards (FCB) among which the OSB and GB were found to provide least strength improve-
ment as compared to FCB boards. Whereas very few studies have been done to evaluate the
impact on the strength and stability of CFS wall-panels sheathed with MgO and HDOSB
sheathing boards.

2 ANALYTICAL MODEL FOR THE CFS WALL-PANEL (CFSWP)

Analytical modelling of the CFSWP with sheathing is worked out sequentially as per the Vieira
& Schafer (2013). The sequential analytical modelling comprises of estimating the quantity of
bracings supplied by the sheathing in terms of stiffness. Theoretically sheathing connected with
the stud imposes three types stiffness in x-direction, y-direction and in rotational z-direction as
per the sign convention considered for the study (Figure 2). These stiffnesses are introduced to
the model using the springs having lateral translational stiffness (kx ), vertical translational stiff-
ness (ky), and z-direction rotational stiffness (k ). The plane of action for these respective con-
sidered stiffnesses are: in plane translation, out of plane translation, and in-plane rotation.
These springs restrains section against the buckling modes associated with weak-axis of global
mode of buckling. The spring restraints are being provided on both flange of the section con-
sidered. The stiffness of these springs depends upon the physical properties of the sheathing
material considered, density of fasteners in vicinity of the stud to connect stud & sheathing
board, CFS material along with its cross-section configuration and screw geometric properties.
Lateral Translational Stiffness ðkx Þ (Vieira & Schafer (2013))

kx ¼ 1=ð1=kxl þ 1=kxd Þ ð1Þ

3Ed 4 t3 π
kxl ¼ ð2Þ
4t2board ð9d 4 π þ 16tboard t3 Þ
π2 Gtboard df wtf
kxd ¼ ð3Þ
L2
Out-of-Plane Translational Stiffness (ky ) (Vieira & Schafer (2013))
ðEIÞW π4 df
ky ¼ ð4Þ
L4

338
Rotational Stiffness (k ) (Vieira & Schafer (2013))
 
k ¼ 1= 1=kc þ 1=kw ð5Þ

kc ¼ 0:00035Et2 þ 75 ð6Þ

ðEIÞw
kw ¼ ð7Þ
df

Where,
kxl ¼ local translational stiffness (Green et al. 1947; Winter 1960)
kxd ¼ diaphragm translational stiffness
E ¼ Young’s modulus of the steel stud,
d ¼ fastener diameter,
t ¼ flange thickness
tboard ¼ board or sheathing thickness,
G ¼ Shear modulus of the sheathing, ¼ fastener tributary width, df ¼ distance between fas-
teners, and L= sheathing height
ðEI ÞW ¼ sheathing rigidity

2.1 Section configuration


The wall-panel sections considered in this study ranges from 1:1 to 1:2 for length-to- height
ration. The height considered here for the study is 2400 mm, which is kept constant for all the
sections. The length of the section considered ranges from 1200 mm to 2400 mm. The aspect
ratio [R] for the frame considered varies from 1:1 to 1:2. The evaluation of the wall-panel here
is based on the spacing of stud along the length, spacing of the screw in the field space, and
effect of sheathing material configuration. There is one single isolated stud has also been con-
sidered in this study for the strength increase comparison purpose. The cross-section configur-
ation could be read from Figure 1. The section configuration illustrative image could be
observed in Figure 2. And the details of the configuration are given in Table 1.

2.2 Material configuration


In the wall-panel section both the stud and tracks are made up-of galvanized steel thin sheets of
yield strength 347Mpa and ultimate tensile strength of 400Mpa as found through the tensile
coupon test. The tensile coupon test here has been performed in accordance with the IS 1608:2005
(Indian Standard). The material used here for the sheathing purpose are MgO board, HDOSB
board, and Gypsum boards whose physical properties as assumed as per the available data of pre-
vious studies (Rahmanian (2011) and Thomas (2003)). and technical sheets (Sinh building solu-
tions (2017)). The respective engineering properties of materials are discussed in Table 2.

2.3 Wall-panel fastener configuration


The fastener considered for connecting the stud and track is the dry wall screws diameter 4.0 mm
screw having the length of 24 mm whereas the stud and the sheathing is connected using the self-
drilling bulge head screws of diameter 4.0 mm and 32 mm long as shown in the Figure 3.

2.4 Elastic stability of the analytical model


The main idea behind testing the stability of the CFS based structure is the incremental variation
in strength (Ye et al. (2016)) of the wall-panel for in-plane loading condition. To test the stability
of the section considered here CUFSM (Confined-Unconfined Finite Strip Method) software tool
of version 4.05 has been employed. The stability of the wall-system has been checked by

339
Figure 1. CFS element cross-section configuration. (a) Stud Cross-section, (b) Track Cross-section.

Table 1. Section configuration details.


Stud 300 mm 400 mm 500 mm 600 mm
spacing

Field screw
spacing
Specimen
(CFS frame 1.2 × 2.4 sq.m 1.6 × 2.4 sq.m 2.0 × 2.4 sq.m 2.4 × 2.4 sq.m
with) frame [R = 1:2] frame [R = 1:1.5] frame [R = 1:1.2] frame [R = 1:1]

MgO
150 mm M1A M2A M3A M4A
Sheathing
200 mm M1B M2B M3B M4B
300 mm M1C M2C M3C M4C
400 mm M1D M2D M3D M4D

HDOSB
150 mm O1A O2A O3A O4A
Sheathing
200 mm O1B O2B O3B O4B
300 mm O1C O2C O3C O4C
400 mm O1D O2D O3D O4D

GB
150 mm G1A G2A G3A G4A
Sheathing
200 mm G1B G2B G3B G4B
300 mm G1C G2C G3C G4C
400 mm G1D G2D G3D G4D

*note: R stands for aspect ratio of frame

performing the elastic buckling analysis here. The naive idea behind proposing the strength deter-
mination through this method is considering the unconfined finite strip method results for cor-
rected buckling loads. Here the initial local-global buckling load interaction has been considered
there after by employing the existing design expression for axial strength determination either by
utilizing DSM as per AISI S-100.

2.5 Direct strength method


Consecutively axial load carrying capacity has been calculated using Direct Strength Method
as per AISI S-100, according to which the elastic buckling loads; Local Buckling (Pcrl ), Distor-
tional Buckling (Pcrd ) and Global Buckling (Pcre ) will be obtained as per the member strength
determination procedure given by Vieira & Schafer (2013) using the elastic buckling load

340
Figure 2. Typical frame-arrangement. (a) Analytical model in CUFSm. (b) Stud-track arrangement. (c)
CFS frame arrangement with sheathing.

Figure 3. Screw used for connecting CFS frame members. (a) 4 mm diameter, length 24 mm (Stud-
track connecting screw). (b) 4 mm diameter, length 32mm (Sheathing-CFS member connecting screw).

Table 2. Material configuration.


Section Element Material Thickness (mm) Engineering properties

Stud and Track CFS 1.0 E = 203 GPa, U = 0.3


Fy = 347 MPa, Fu = 400 Mpa
Sheathing MgO board 10.0 E = 4500 MPa, U = 0.2
HDOSB board 10.0 E = 4190 MPa, U = 0.21
Gypsum board 10.0 E = 2500 MPa, U = 0.2
*note: E stands for Elasticity Modulus, U stands for Poisson’s ratio

factors and the squash load, Py obtained from CUFSM analysis for predicting the nominal
strength of the wall-panel by DSM as per section A1.2.1 of AISI S-100. The nominal strength
of the wall-panel will be determined by equation number (14), according to which the least
buckling strength among the global strength (Pne ), local global interaction strength (Pnl ), and
distortional strength (Pnd ) which will be determined by equation number 4 to 6. Then the
available strength of the section will be calculated by LRFD method of AISI S-100 section
C4.1. The resistance factor in LRFD method is 0.85.

341
Local Buckling, (Pcrl )

Pcrl ¼ ðL:F:Þlocal  Py ð8Þ

Distortional Buckling, (Pcrd )

Pcrd ¼ ðL:F:Þdistortional  Py ð9Þ

Global Buckling, (Pcre )

Pcre ¼ ðL:F:Þglobal  Py ð10Þ

The details of the parameters such as required for calculating the various stiffness’s has been dis-
cussed in equation number (1) to (7). In the local buckling case the sheathing does not have any
effect on local buckling. The sheathing restrains the flange, but local buckling is largely driven by
the web Theoretically, kx and k (if located at the exact mid-width of the flange) have no influ-
ence on local buckling, only ky . For local buckling predictions it is recommended to ignore the
sheathing. A conventional finite strip signature curve result completed on the bare stud (or similar
shell finite element model) is adequate for finding the local elastic buckling load, In the Distor-
tional buckling case the sheathing provides beneficial rotational restraint against distortional
buckling, and k; should be included when determining the elastic distortional buckling load
(Pcrd ). For studs with deep webs (and narrow flanges) the additional restraint supplied by kx may
be influential. In the global buckling case Sheathing greatly influences the global buckling load
(Pcre ) For determining Pcre , inclusion of all available fastener-sheathing springs (kx ; ky ; &k; ) is
recommended, but kx is critical as it provides the primary fastener-sheathing restraint for both
weak-axis flexure and torsion (when present on both flanges). The last step is to utilize existing
design expressions mentioned in equation number (8) - (10) to convert the elastic buckling loads
(and knowing the squash load, Py ) to predict the nominal strength Pn .
Global Strength, Pne

 sffiffiffiffiffiffiffiffi
0:658 λ2c Py for λc  1:5 Py
Pne ¼ and λc ¼ ð11Þ
0:877Pcre for λc 41:5 Pcre

Local-global interaction, Pnl


8 sffiffiffiffiffiffiffiffi
< P for λc  0:766
 ne 0:4  0:4 Pne
Pnl ¼ and λl ¼ ð12Þ
: 1  0:15 Pne
Pcrl Pcrl
Pne Pne for λc 40:766 Pcrl

Distortional Buckling, Pnd


8 sffiffiffiffiffiffiffiffiffi
< Py for λc  0:561
 0:4  0:6 Py
Pnd ¼ and λd ¼ ð13Þ
: 1  0:25
Pcrd
Py
Pcrd
Py Py for λc 40:561 Pcrd

Finally, the nominal capacity is the minimum (note that Pnl is strictly less than or equal to Pne ,
and that Pnd is just an intermediary calculation) and there after evaluating the theoretical
available strength by the product of nominal strength and resistance factor of LRFD method.

Pn ¼ minðPne ; Pnl ; Pnd Þ ð14Þ

342
3 VALIDATION OF THE MODEL CONSIDERED

The very first step before proceeding for determining the theoretical strength of the section the
validation of the model considered has been carried with reference to the data made available
by Ye et al. 2016. These data comprise of Experimental strength of CFS-wall-panel with pre-
diction of strength by AISI S-100 based DSM procedure along with the FEM prediction.

4 RESULT AND DISCUSSIONS

The increment in the resulting strength if the stud spacing is decreased and increase in the
screw connection density has been observed here in this study. The important outcome of this
study is the trend of the results observed according to which the less is the elastic nature in
terms of elastic modulus of the sheathing material the lesser will be its strength for same thick-
ness and other properties of the specimen. The stiffness of the springs assigned in the model
also tend to vary with the elastic properties of the steel used for the stud and track. The avail-
able strength for the bare frame model for single stud was found to be 22.27 kN.
Based on the utility demand the configuration of wall panel will be decided. The load bearing
demand from the wall-panel is calculated based on the dead load, live load, superimposed dead
load and seismic load or wind load. The optimized combination of all these loads will set
a demand benchmark. Here in the results it has been observed that increase in strength with
respect to bare frame for 300mm stud spacing is 49%, 400mm is 47.5%, 500mm is 46% and
600mm is 45%. Here the basis of idealization of single stud bare frame to the respective bare
frame of the specimen is the number of studs present and interelement distance between them.

Table 3. Validation details.


Ye et al. 2016 Evaluation

Strength %vari-
by Analyt- %variation with %variation with ation with
ical model experimental Analytical FEM pre-
Specimen considered Experimental Analytical FEM solution prediction diction

W8 28.80 24.30 29.65 26.86 18.5 2.86 7.22


W11 28.58 29.10 29.65 32.25 1.78 3.60 11.37

Table 4. Axial strength for various specimen.


Strength Speci- Strength Speci- Strength
Specimen (kN/m) men (kN/m) men (kN/m)

M1A 110.157 O1A 110.136 G1A 109.931


M1B 109.137 O1B 109.574 G1B 108.817
M1C 108.695 O1C 108.684 G1C 108.529
M1D 107.907 O1D 107.883 G1D 107.670
M2A 82.662 O2A 82.656 G2A 82.527
M2B 82.230 O2B 82.212 G2B 82.110
M2C 81.549 O2C 81.533 G2C 81.387
M2D 80.962 O2D 80.940 G2D 80.802
M3A 66.153 O3A 67.838 G3A 66.058
M3B 65.801 O3B 65.791 G3B 65.716
M3C 65.254 O3C 65.246 G3C 65.145
M3D 64.775 O3D 64.761 G3D 64.659
M4A 55.140 O4A 55.131 G4A 55.063
M4B 54.842 O4B 54.836 G4B 54.774
M4C 54.391 O4C 55.233 G4C 54.300
M4D 53.983 O4D 53.975 G4D 53.584

343
5 CONCLUSIONS

The elastic modulus of sheathing material plays important role is bracing the CFS wall
panel frame and increasing its strength considerably. The variation of the engineering
properties of the CFS material also affects the strength of the CFSWP. Sheathing on
both sides of the wall-panel increases the stability as compared to case of uneven sheath-
ing applied on the both sides. The increase in the screw density in the vicinity gives the
increment in the strength of the frame.
The field screw spacing is the strength governing quantity in the case of axial in-plane load-
ing whereas the Edge screws are major strength governing parameter for lateral in-plane load-
ing cases. The heavy-duty OSB sheaths are quite comparable with MgO board sheathed
CFSWP as observed from the trend of results observed. The HDOSB boards are found to be
stronger than MgO board when the stud spacing was 500mm and screw spacing was 150 mm.
This case gives the result out of general assumption that with increase in elastic modulus the
strength increases.

REFERENCES

AISI S-100. 2016. North American Specification for Cold-Formed Steel Structural Members.
Building Solutions, Sinh. 2017. SINH™ Magnesium Oxide Boards. Sinh-Brochure-English-2017.
Fiorino, L., Iuorio, O. and Landolfo, R. 2014. Designing CFS structures: The new school bfs in naples.
Thin-Walled Structures, 78, pp.37–47.
Green, G.G., Winter, G., and Cuykendall, T.R. 1947. Light gage steel columns in wall-braced panels.
Rep. 35, Pt. 2, Engineering Experiment Station, Cornell Univ., Ithaca, NY.
IS 1608 (Indian Standard). 2005. Metallic Materials - Tensile Testing at Ambient Temperature.
Li, Z. and Schafer, B.W. 2010. Buckling analysis of cold-formed steel members with general boundary
conditions using CUFSM conventional and constrained finite strip methods.
Rahmanian, I. 2011. Thermal and mechanical properties of gypsum boards and their influences on fire
resistance of gypsum board-based systems (Doctoral dissertation, The University of Manchester
(United Kingdom)).
Schafer, B.W. 2008. The direct strength method of cold-formed steel member design. Journal of construc-
tional steel research, 64(7–8), pp.766–778.
Schafer, B.W. 2012. CUFSM 4.05–finite strip buckling analysis of thin-walled members. Baltimore,
USA: Department of Civil Engineering, Johns Hopkins University.
Schafer, B.W. 2013. Sheathing Braced Design of Wall Studs.
Schafer, B.W., Vieira Jr, L.C., Sangree, R.H. and Guan, Y. 2010. Rotational restraint and distortional
buckling in cold-formed steel framing systems. Revista Sul-americana de Engenharia Estrutural, 7 (1).
Thomas, W.H. 2003. Poisson’s ratios of an oriented strand board. Wood science and technology, 37
(3–4), pp.259–268.
Tian, Y.S., Wang, J., Lu, T.J. and Barlow, C.Y. 2004. An experimental study on the axial behaviour of
cold-formed steel wall studs and panels. Thin-walled structures, 42(4), pp.557–573.
Vieira, L.C.M., Jr. 2011. Behavior and design of sheathed cold-formed steel stud walls under compres-
sion. Ph.D. thesis, Johns Hopkins Univ., Baltimore.
Vieira Jr, L.C. and Schafer, B.W. 2012. Lateral stiffness and strength of sheathing braced cold-formed
steel stud walls. Engineering Structures, 37, pp.205–213.
Vieira Jr, L.C. and Schafer, B.W. 2013. Behavior and design of sheathed cold-formed steel stud walls
under compression. Journal of Structural Engineering, 139 (5), pp.772–786.
Winter, G. 1960. Lateral bracing of beams and columns. J. Struct. Div. 102(1),77–92.
Young, B. and Rasmussen, K.J., 1998. Design of lipped channel columns. Journal of Structural Engineer-
ing, 124(2), pp.140–148.
Ye, J., Feng, R., Chen, W. and Liu, W. 2016. Behavior of cold-formed steel wall stud with sheathing
subjected to compression. Journal of Constructional Steel Research, 116, pp.79–91.
Zhang, Y., Wang, C. and Zhang, Z. 2007. Tests and finite element analysis of pin-ended channel columns
with inclined simple edge stiffeners. Journal of Constructional Steel Research, 63(3), pp.383–395.

344
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Direct Strength Method (DSM) design of simply supported


short-to-intermediate hot-rolled steel equal-leg angle columns

P.B. Dinis & D. Camotim


CERIS, DECivil, Instituto Superior Técnico, Universidade de Lisboa, Portugal

ABSTRACT: This work reports results of an ongoing investigation on the DSM design of
hot-rolled steel equal-leg angle columns with pinned (spherically hinged) supports and short-to-
intermediate lengths, i.e., buckling in flexural-torsional modes. It extends the scope of similar
studies involving cold-formed steel angle columns with similar characteristics (but higher leg
width-to-thickness ratios). After collecting the experimental and numerical failure loads available
in the literature, concerning columns with several geometries, the paper addresses the mechanical
reasoning behind a DSM-based design approach proposed in the context of cold-formed steel
short-to-intermediate angle columns, which must be modified to handle spherically hinged sup-
ports. Then, the failure loads are used to assess the quality of their estimates by this design
approach, namely by determining LRFD resistance factors – it is shown that it can be success-
fully applied to predict hot-rolled steel simply supported angle column failure loads, yielding
LRFD resistance factors above 0.90.

1 INTRODUCTION

The structural behaviour and design of thin-walled equal-leg angle columns has attracted the
attention of several researchers in the past, namely Kitipornchai & Lee (1986), Adluri & Madu-
gula (1996), Popovic et al. (1999), Young (2004), Rasmussen (2006) and, more recently, Shi et al.
(2011), Dinis et al. (2012), Silvestre et al. (2013) and Mesacasa Jr. et al. (2014). As far as the
design is concerned, a fair amount of work has been devoted to develop rules and procedures to
predict the ultimate strength of short-to-intermediate cold-formed steel columns with quite slen-
der legs (width-to-thickness ratios b/t above 20), adopting mostly local buckling concepts. Note
that these columns buckle in (major-axis) flexural-torsional modes, associated with an almost
horizontal “critical load plateau” of the “signature curve” fcr vs. L, where fcr and L are the
column critical buckling stress and length (logarithmic scale) – Figure 1(a) shows typical signature
curves of fixed-ended (F), cylindrically-hinged (PC – major-axis bending rotations prevented) and
spherically-hinged (PS) columns, all with prevented secondary warping and torsional rotations,
while Figure 1(b) provides the mechanical characterisation of the flexural-torsional critical buck-
ling (and failure) modes: combination of a tiny (often virtually imperceptible to the naked eye,
but not negligible) amounts of major-axis flexure with highly dominant amounts of torsion.
Numerical simulations reported by Dinis et al. (2012) and concerning F, PC and PS columns
shed new light on key mechanical aspects of the structural behaviour of such members, namely
the strong influence of length-dependent interaction between major-axis flexural-torsional and
minor-axis flexural buckling (two global instabilities) – this interaction is particularly severe in
PC and PS columns. It was also shown that a single design approach cannot handle adequately
short-to-intermediate equal-leg angle columns with the above three end support conditions.
The above findings led Dinis & Camotim (2015) to propose a novel, rational and unified
design approach applicable to both F and PC columns, which is based on the Direct Strength
Method (DSM – e.g., Camotim et al. 2016) and combines (i) numerically-based genuine flex-
ural-torsional strength/design curves, specifically developed for this purpose, with (ii) the cur-
rently codified DSM global design curve.

345
Figure 1. (a) Typical “signature curves” fcr vs. L of F, PC and PS columns, and (b) mechanical charac-
terisation of the short-to-intermediate column flexural-torsional critical buckling (and failure) modes.

This design approach was shown to be both accurate and reliable – indeed, the prediction
quality of the available experimental and numerical failure loads allows for the use of the
Load and Resistance Factor Design (LRFD) resistance factor ϕc = 0.85, recommended for
compression member design by the current North American Specification for cold-formed
steel structural elements (AISI 2016). This design approach was subsequently slightly
improved by Landesmann et al. (2016) and has been recently cast in a form ready for codifica-
tion by Dinis & Camotim (2018).
Recently, Dinis et al. (2016) and Dinis & Camotim (2016) extended the above research to hot-
rolled steel (HRS) F and PC columns, which differ from their cold-formed counterparts in the
fact that they (i) exhibit stockier legs and (ii) are affected by residual stresses. These studies pro-
vided promising indications that the DSM-based design approach developed in the context of
cold-formed steel angle columns can also be applied to hot-rolled ones – however, experimental
validation is still lacking.
This work continues the research effort described in the previous paragraphs and reports the
latest results concerning extending the scope of the investigation to HRS columns with spheric-
ally hinged supports (PS columns) – note that this simply supported end conditions (pinned
with respect to major and minor-axis bending, and fixed with respect to torsion) (i) corresponds
to having rigid plates attached to the column end cross-sections and resting on half-spheres, to
prevent torsional rotations, and (ii) alter visibly the column flexural-torsion behaviour, i.e., the
mechanics underlying the existing DSM-based design approaches. Therefore, the same concepts
and procedures of the methodology proposed by Dinis & Camotim (2018) are now adopted to
develop and assess the merits of a new DSM-based design approach for PS columns – in par-
ticular, the search for design curves simpler than those recently reported by the authors (Dinis
& Camotim 2017) is addressed. The merits of the proposed DSM-based design approach are
assessed through the comparison with experimental and numerical failure loads – while the
latter are obtained by means of a parametric study involving columns with several leg width-to-
thickness ratios and yield stresses, the former are taken from the literature (a fair number of PS
column experimental failures loads are available, which is not true for their F and PC counter-
parts). It is shown that the proposed design approach leads to efficient (safe and reliable) failure
load predictions for the short-to-intermediate PS columns considered in this work, which exhibit
a wide slenderness range – i.e., there is also a DSM-based design approach able to handle col-
umns with these support conditions.

2 FAILURE LOAD DATA CONCERNING HOT-ROLLED STEEL PS COLUMNS

The experimental failure loads considered in this work concern (i) 4 column tested by Waka-
bayashi & Nonaka (1965), with b/t = 12.9, (ii) 4 column tested by Kennedy & Murty (1972),
with 18.5≥ b/t ≥12.8, (iii) 7 column tested by Kitipornchai & Lee (1986), with 15.9≥ b/t ≥13.3, (iv)
3 column tested by Adluri & Madugula (1996), with 15.8≥ b/t ≥12.9, (v) 9 column tested by Fan
(2009), with 16.0≥ b/t ≥12.5, and (vi) 57 column tested by Ban et al. (2013), with 16.0≥ b/t ≥14.0 –
a total of 84 experimental failure loads, a number deemed acceptable to assess the merits of the

346
proposed DSM-based design approach for PS angle columns. The specimen cross-section
dimensions b and t, lengths L, yield stresses fy and ultimate strengths fu have been recently
reported by the authors (Dinis & Camotim 2017).
The numerical failure loads considered in this work were obtained by means of Abaqus shell
finite element analyses (SFEA), adopting a model previously used by the authors that (i) is based
on an elastic-perfectly plastic steel material behaviour (Prandtl-Reuss’s model: von Mises yield
criterion and associated flow rule), (ii) discretises the columns into fine meshes of 4-node isopara-
metric elements (length-to-width ratio roughly equal to 1 – the rounded corners were disre-
garded), (iii) simulates the support conditions by attaching rigid plates to the column end
sections, which are pinned with respect to major and minor-axis flexure, and fixed with respect
to torsion. The parametric study carried out involved steel (E = 210 GPa, ν = 0.3) columns with
20≥ b/t ≥11, exhibiting lengths and yield stresses selected to ensure (i) buckling in flexural-
torsional (predominantly torsional) modes, and (ii) covering a wide slenderness range – the yield
stresses adopted were 150, 300, 500 and 700 MPa. A total of 144 short-to-intermediate columns
were analysed, all containing initial geometrical imperfections combining (i) a critical flexural-
torsional component, with amplitude equal to 10% of the wall thickness t, and (ii) a non-critical
minor-axis flexural component, with amplitude equal to L/1000 (leg tips under compression –
most detrimental situation), value commonly adopted in HRS columns (e.g., Može et al. 2014)
and in line with the measurements reported for cold-formed steel columns by Popovic et al.
(1999). Moreover, all the columns analysed contained a 3-point residual stress distribution – note
that this residual stress distribution is almost universally adopted in the numerical simulation of
hot-rolled steel columns. The column cross-section dimensions, lengths and ultimate strengths
were also reported by Dinis & Camotim (2017).

3 DIRECT STRENGTH METHOD DESIGN CONSIDERATIONS

The use of the Direct Strength Method (DSM) to design cold-formed steel angle columns has
attracted the attention of a few researchers in the past: (i) Young (2004), for F columns, (ii)
Rasmussen (2006), for PC columns, and Silvestre et al. (2013), for both F and PC columns,
put forward DSM-based design approaches based on the combined use of global and local
strength curves to obtain ultimate strength estimates (fnle). Even if these design approaches
were found to predict the available failure load data quite adequately, the fact that they are
mostly empirical (the length-dependence of the column post-buckling and failure behaviours
is never explicitly taken into account) led Dinis & Camotim (2015) to propose a more rational
DSM-based design approach for F and PC columns with slender legs (b/t >20), which was
subsequently slightly modified/simplified by Landesmann et al. (2016) and, finally, cast in
a form ready for codification by Dinis & Camotim (2018). It was found that (i) the length-
dependent interaction between major-axis flexural-torsional and minor-axis flexural buckling
(two global instabilities) heavily influences the column strength, and that (ii) the effective cen-
troid shift effects strongly affect the PC column failure loads (not the F ones). The length-
dependence of the column post-buckling behaviour can be linked to the characteristics of the
critical (flexural-torsional) buckling mode, namely the amount of major-axis flexure, accur-
ately quantifiable by means of a parameter Δf, defined as

fbt  fcrft
Δf ¼  100 ð1Þ
fbt

where fbt and fcrft are the pure torsional and major-axis flexural-torsional (critical) buckling
stresses – such buckling stresses can be straightforwardly determined by means of analytical
expressions derived by Dinis & Camotim (2015). Moreover, it was shown to be necessary to
develop (i) a set of genuine flexural-torsional strength curves, intended to play the role of the
local strength curve in the traditional DSM design against column local-global interactive fail-
ures, and (ii) curves providing a parameter able to capture the ultimate strength erosion

347
stemming from the effective centroid shift effects. Performing these tasks led to a DSM-based
design approach (fnfte) for F and PC short-to-intermediate angle columns cast in the form
8
>  1
>
> pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
>
< β  f ne if λfte  0:5 þ
0:25  b 2a
fnfte ¼  a  a 1 ð2Þ
>
>  pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
>
> fcrft fcrft
: β  fne 1b if λfte 4 0:5 þ 0:25  b 2a
fne fne
8 8
< 0:19 Δf þ 0:4 if
> Δf 53 < 0:014 Δf þ 0:15
> if Δf  7
a¼ b¼ ð3Þ
>
: >
:
0:97 if Δf ≥ 3 0:248 if Δf 47

where (i) parameter b should not be confused with the angle leg width, (ii) the values a = 0.4 and
b = 0.15 were adopted for Δf = 0, which amounts to saying that Equation. (2) coincides with the
currently codified DSM local-global interactive strength curve for the shorter columns (fbt/fcrft
close to 1.0), and (iii) the slenderness λfte = (fne/fcrft)0.5 is based on the column nominal strength
against minor-axis flexural collapses (fne), obtained from the codified DSM global design curve
8 
>
>
λ2c
if λc  1:5 sffiffiffiffiffiffiffi
< f y 0:658 ! fy
fne ¼ 0:877 with λc ¼ ð4Þ
>f
> if λc 41:5 fcre
: y
λ2c

where fcre is the column minor-axis flexural buckling stress.


The coefficient β, which accounts for the effective centroid shift effects and follows an idea
originally put forward by Rasmussen (2006), is obtained from
8
< 1 for F columns
β¼ 0:68 ð5Þ
:  1 for PC columns
ðλfte  cÞd

c ¼ 0:2 Δf þ 0:55 d ¼ 0:08 Δf þ 0:72 ð6Þ

It is still worth noting that the expressions providing parameters a, b, c and d (Equations. (3)
and (6)) were obtained through “trial-and-error curve-fitting procedures”, based on the F and
PC column numerical failure load data reported by Dinis & Camotim (2015), and have been
very recently simplified for codification purposes (Dinis & Camotim 2018). Moreover, recent
investigations by Dinis et al. (2016) and Dinis & Camotim (2016) indicated that the above DSM
design approach, developed for F and PC cold-formed steel columns, can be readily applied to
HRS columns with the same end support conditions (F or PC) – however, experimental valid-
ation is necessary before codification can be considered.

3.1 DSM design approach for PS columns


Very recently, Dinis & Camotim (2017) showed that the above DSM design approach need to
be modified to handle PS steel columns (either cold-formed or hot-rolled) – this is due to the
change in major-axis flexure end support conditions (from fixed to pinned), which has
a pronounced influence in both the column flexural-torsional strength and effective centroid
shift effects – in particular, it changes significantly the function Δf (L). Therefore, it was neces-
sary to use the concepts and procedures employed earlier, in the context of F and PC columns,
to develop a design approach able to handle PS cold-formed and hot-rolled steel columns.
The main features of this design approach are the following:

348
(i) A fully numerical approach is again adopted to determine new length-dependent (i1) flexural-
torsional strength curves fnft and (i1) curves providing the reduction coefficient β.
(ii) Both curves are qualitatively similar to those developed in the context of F and PC columns:
“Winter-type” curves that depend on the length via parameter Δf – for Δf = 0, the new fnft
curve still coincides with the currently codified DSM local curve. Thus, the modifications,
with respect to the proposal of Dinis & Camotim (2018), consist of new expressions for
functions a (Δf), b (Δf), c (Δf), d (Δf).
(iii) The set of genuine flexural-torsional strength curves (fnft) is developed in the context of
columns with fully prevented minor-axis bending displacements – i.e., “forced” to fail in
flexural-torsional modes. These strength curves make it possible to capture the progressive
column post-critical strength drop as its length increases along the fcr (L) curve “horizontal
plateau”. The expressions for the functions a (Δf) and b (Δf), proposed by Dinis & Camotim
(2017), read
8
>
< 0:001 Δf þ 0:014 Δf  0:007 Δf þ 0:4
3 2
if Δf  1
a¼ 0:001 Δf þ 0:04 Δf þ 0:365
2
if 15Δf 510
>
: 0:865 if Δf ≥ 10
8
< 0:001 Δ3f þ 0:001 Δ2f  0:011 Δf þ 0:15 if Δf  1
b¼ 0:005 Δf þ 0:134 if 15Δf 510 ð7Þ
:
0:184 if Δf ≥ 10

(iv) The procedure adopted to search for an expression providing the function β (Δf) is based
on an “elastic reduction factor” concept and also reflects the change in column post-
buckling behaviour along the fcr (L) curve plateau. This procedure assumes that the differ-
ence between the applied loads (P) obtained for identical PCm (cylindrically pinned with
respect to minor-axis flexure) and PS columns (PPCm and PPS) stems exclusively from the
effective centroid shift effects – then, the PPS/PPCm ratio provides a good approximation
for β (i.e., it is assumed that β ≈ PPS/PPCm). The expressions for the functions c (Δf) and d
(Δf) proposed by Dinis & Camotim (2017), which were obtained through another “trial-
and-error curve-fitting procedure”, read
8
>
< 300 Δ3f þ 110 Δ2f  12:8 Δf þ 1 if Δf  0:2
c ¼ 0:001 Δ3f þ 0:01 Δ2f  0:058 Δf þ 0:451 if 0:25Δf 5 8
>
: 0:115 if Δf ≥ 8
8
>
< 290 Δf  98 Δf þ 10:8 Δf þ 0:25
3 2
if Δf  0:2
d¼ 0:001 Δf þ 0:03 Δf þ 0:804
2
if 0:25Δf 5 9:5 ð8Þ
>
: 0:999 if Δf ≥ 9:5

The DSM design approach (fnfte) proposed by Dinis & Camotim (2017) leads to quite accurate
predictions of the experimental and numerical ultimate strengths: indeed, the failure-to-
predicted load ratio fu/fnfte averages, standard deviations, maximum/minimum values are 1.16/
0.19/1.78/0.88 (experimental data) and 1.01/0.08/1.17/0.90 (numerical data) – the higher scatter
of the experimental failure load predictions is easily explained by the very high sensitivity of
the column failure load ultimate to the minor-axis flexural initial geometrical imperfection
sign (curvature towards leg tips or corner) and amplitude.
Since it is felt that the expressions providing parameters a, b, c and d, which are given in
Equations (7)-(8), are a bit too complicated, it was decided to make an attempt to simplify
them, without sacrificing the quality of the failure load predictions – next, linear approxima-
tion are sought for these expressions.

349
3.1.1 Modification of the flexural-torsional strength curves – parameters a and b
The search for simpler expressions for parameters a and b is based on the numerical failure
load data reported by Dinis & Camotim (2017) concerning HRS PS columns with fully
restrained minor-axis bending displacements, i.e., “forced” to fail in a mode combining major-
axis flexure and torsion – as before, a = 0.4 and b = 0.15 are adopted for Δf = 0. The proposed
expressions constitute bi-linear approximations of Equations. (7) and their coefficients were
selected, through a “trial and error” procedure, to ensure flexural-torsional ultimate strength
predictions as accurate as possible. The best approximation was found to be
8 8
< 0:04 Δf þ 0:4 if Δf 510 < 0:01 Δf þ 0:15 if Δf  4
a¼ b¼ ð9Þ
: 0:80 if Δf ≥10 : 0:19 if Δf 44

and Figure 2(a) compares the functions a (Δf) and b (Δf) provided by Equations. (7) and (9).
Moreover, Figure 2(b) compares the flexural-torsional strength curves obtained with Equa-
tions. (9) and (7) for Δf = 0.5; 9.8, as well as the numerical failure loads that they must predict.
This last figure clearly shows that the fnft values provide fairly accurate underestimations of
the numerical failure loads. It is worth noting that Equations. (7) and (9) are valid only for Δf
values up to 11, which stems from the fact that Dinis & Camotim (2017) showed that, regard-
less of the cross-section geometry, Δf never reaches 10.9, i.e., the flexural-torsional “horizontal
plateau” never extends that far – within this Δf range the above a (Δf) and b (Δf) bi-linear
approximations are perfectly adequate.

3.1.2 Modification of the reduction coefficient – parameters c and d


The search for simpler expression for parameters c and d is also based on the elastic post-buckling
results reported by Dinis & Camotim (2017) and concerning identical PCm and PS columns. As
before, the new expressions constitute linear approximations of Equations. (8) and their coeffi-
cients were selected, once more, by means of a “trial and error” procedure aimed at capturing, as
accurately as possible, the numerically determined strength erosion caused by the effective cen-
troid shift effects. Moreover, it was decided to cease enforcing that c = 1.0 and d = 0.25 for Δf = 0,
a condition previously adopted for the sake of making Equation (8) “compatible” with the expres-
sion proposed by Rasmussen (2006) for the very short columns (Δf ≈ 0). Note that adopting the
above condition was not a very rational choice, since it led to functions c (Δf) and d (Δf)
exhibiting very high gradients in the close vicinity of Δf = 0 (see Figure 3), a feature not
stemming from the numerical results. The best linear approximations were found to be
8 8
< 0:02 Δf þ 0:4 if Δf 510 < 0:04 Δf þ 0:8 if Δf  5
c¼ d¼ ð10Þ
: 0:20 if Δf ≥ 10 : 1:0 if Δf 45

and Figure 3(a) compares the functions provided by Equations (8) and (10). Moreover, Figure 3
(b) compares, for PS columns with Δf = 0.5; 5.0; 9.8, the numerical β values reported by Dinis &
Camotim (2017) with the β (λft) curves obtained with Equations (10). These figures clearly show

Figure 2. (a) Plots of parameters a and b against Δf, and (b) plots fu/fy vs. λft and fnft strength curves for PS
columns with Δf = 0.5; 9.8 – note that the minor-axis bending displacements are fully prevented in both cases.

350
Figure 3. (a) Plots of parameters c and d against Δf and (b) β (λft) curves of PS columns with Δf = 0.5, 5.0, 9.8.

that the new β (λft) curves follow quite well the trends of the numerical result obtained, even if
there are some perceptible differences – recall that Equations (8)–(10) only need to be used for Δf
values up to around 11 (Dinis & Camotim 2017).

3.1.3 Assessment of the proposed DSM-based design approach for HRS PS columns
In order to assess the performance/merits of the above DSM-based design approach in esti-
mating PS column failure loads, the failure-to-predicted strength ratios fu/fnfte concerning the
whole set of failure loads (experimental and numerical) gathered in Section 2, are considered
next. Moreover, the prediction quality is also assessed through the determination of the Load
and Resistance Factor Design (LRFD) resistance factor associated with the estimates of the
proposed/modified DSM design approach. In particular, it is intended to check whether
a factor equal or higher than ϕc=0.90 is obtained – this is the value recommended, for com-
pression members, by the current North American Specification (AISC 2016). Since this speci-
fication does not include any procedure to calculate ϕ, it was decided to use the formula given
in Section K2.1.1 of Chapter K of the North American Specification for cold-formed steel
structural members (AISI 2016), which reads
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi  
β VM 2 þ V2 þ C V2 þ V2
P P 1 m
 ¼ C ðMm Fm Pm Þ e 0 F Q
with CP ¼ 1þ ð11Þ
n m2

where (i) Cϕ is a calibration coefficient, (ii) β0 is the target reliability index, (iii) Mm and Fm are the
mean values of the material and fabrication factors, respectively, (iv) VM, VF and VQ are the coef-
ficients of variation of the material factor, fabrication factor and load effect, respectively, (v) CP
is a correction factor depending on the numbers of tests (n) and degrees of freedom (m = n–1),
and (vi) Pm and VP are the mean and coefficient of variation of the “exact”-to-predicted ultimate
strength ratios fu/fnfte. The material and fabrication factors adopted in the North American speci-
fications are mostly due to the work of Galambos and co-workers – they are Mm = 1.10 and VM
= 0.10 (Galambos & Ravindra 1978), Fm = 1.00 and VF = 0.05 (Ravindra & Galambos 1978),
respectively. On the other hand, the target reliability index, calibration coefficient and coefficients
of variation of the load effect differ slightly in the cold-formed and hot-rolled steel specifications:
(i) β0 = 2.6, CP = 1.0, Cϕ = 1.48 and VQ = 0.19 (based on the load combination 1.2D + 1.6L with
a live-dead load ratio L/D = 3.0) (Ellingwood & Galambos 1982, AISC 2000), and (ii) β0 = 2.5, Cϕ
= 1.52 and VQ = 0.21 (1.2D + 1.6L with a live-dead load ratio L/D = 5.0) (AISI 2016) – see
Meimand & Schafer (2014).
In order to assess the impact of replacing functions a (Δf), b(Δf), c (Δf) and d (Δf), Figures 4(a)-
(b) plot the PS column fu/fnfte ratios against λfte and include the corresponding averages, standard
deviations and maximum/minimum values for the whole sets of PS column experimental and
numerical failure loads considered in this work (see Section 2). The comparison between the stat-
istical indicators in Figures 4(a)-(b) and those reported by Dinis & Camotim (2017) shows that
the new expressions have virtually no impact on the proposed DSM-based design approach statis-
tical indicators. Indeed, the averages, standard deviations, maximum values and minimum values
of the experimental/numerical fu/fnfte ratios vary only by –0.01/0.00, –0.04/–0.02, –0.19/–0.03,
+0.02/0.00, respectively – the most meaningful variations concern the maximum values (they

351
Figure 4. Proposed DSM-based design approach: plots fu/fnfte vs. λfte concerning the PS column (a)
experimental and (b) numerical failure loads.

decrease). It is still worth noting that several experimental failure loads are considerably underesti-
mated, which is due to the fact that the design approach was developed on the basis of numerical
failure loads of columns containing the most detrimental initial imperfections – given the failure
load high sensitivity to the initial imperfection sign and amplitude, tested specimens containing
less detrimental initial imperfections are likely to exhibit visibly larger failure loads.
The ϕ values stemming from the proposed DSM-based procedure, with the new a, b, c,
d expressions, are never below the value prescribed by AISC (2016) for compression members.
Indeed, n, Pm, VP and ϕ read 84/1.16/0.13/0.97 (experimental data), 144/1.01/0.06/0.91 (numer-
ical data) and 228/1.06/0.12/0.91 (experimental and numerical data) – note that the ϕ values
obtained with the AISI (2016) expression are practically identical: ϕ=0.98 (experimental), 0.91
(numerical), 0.92 (experimental and numerical). Finally, it should still be noted that all available
PS column experimental failure loads concern a quite narrow and fairly low slenderness
range (1.09 ≥ λfte ≥ 0.50), thus evidencing the need to acquire experimental results for
columns with moderate and high slenderness before proper validation can be achieved.

4 CONCLUSION

This paper reported the results of an investigation on the strength and DSM design of hot-
rolled steel angle columns (stocky legs – b/t ≤ 20) with spherically-hinged support conditions (PS
columns) and short-to-intermediate lengths (i.e., buckling in flexural-torsional modes). Initially,
a fairly large set of column failure loads was assembled, comprising (i) experimental data, col-
lected from the literature and concerning 84 angle column tests, and (ii) numerical data,
obtained from Abaqus shell finite element analyses and involving 144 columns with various
cross-section dimensions, lengths and yield stresses. Then, a recently developed rational DSM-
based design approach for cold-formed steel fixed-ended and pin-ended (cylindrical or spherical
hinges) angle columns and their applicability to hot-rolled steel PS angle columns were briefly
reviewed – particular attention was paid to a modification/simplification of the expressions pro-
viding the design curves previously developed by the authors in (Dinis & Camotim 2017). The
proposed DSM-based design approach was shown to yield safe and reliable failure load esti-
mates for hot-rolled steel PS angle columns buckling in flexural-torsional modes – note that the
failure load predictions lead to resistance factors above the value recommended by AISC (2016)
(ϕc=0.90). Nevertheless, further experimental (mostly for slender columns) and numerical valid-
ation is indispensable before the above design approach can reach the codification stage.

REFERENCES

Adluri, S.M.R., Madugula, M.K.S. (1996). Flexural buckling of steel angles: Experimental investigation,
Journal of Structural Engineering (ASCE), 122(3), 309–317.
AISC (American Institute of Steel Construction) (2000). Load and Resistance Factor Design Specification
for Structural Steel Buildings and Associated Commentary, Chicago, Illinois.

352
AISC (American Institute of Steel Construction) (2016). Specification for Structural Steel Buildings and
Associated Commentary (ANSI/AISI-360-16), Chicago, Illinois.
AISI (American Iron and Steel Institute) (2016). North American Specification (NAS) for the Design of
Cold-Formed Steel Structural Members and Associated Commentary (AISI-S100-16), Washington DC.
Ban, H.Y., Shi, G., Shi, Y.J., Wang, Y.Q. (2013). Column buckling tests of 420 MPa high strength steel
single equal angles, International Journal of Structural Stability and Dynamics, 13(2), 1–23.
Camotim, D., Dinis, P.B., Martins, A.D. (2016). Direct strength method (DSM) – a general approach for
the design of cold-formed steel structures, Recent Trends in Cold-Formed Steel Construction, C. Yu
(ed.), Woodhead Publishing (Series in Civil and Structural Engineering), Amsterdam, 69–105.
Dinis, P.B., Camotim, D. (2015). A novel DSM-based approach for the rational design of fixed-ended and
pin-ended short-to-intermediate thin-walled angle columns, Thin-Walled Structures, 87(February),
158–182.
Dinis, P.B., Camotim, D. (2016). Behavior and design of hot-rolled steel pin-ended short-to-intermediate
angle columns, USB Key Drive Proceedings of Seventh International Conference on Coupled Instabilities
in Metal Structures (CIMS 2016 – Baltimore, 7-8/11), Paper 23.
Dinis, P.B., Camotim, D. (2017). Spherically-hinged short-to-intermediate angle columns: stability,
non-linear behavior and DSM design, Website Proceedings of Structural Stability Research Council
(SSRC) Annual Stability Conference (San Antonio, 21-24/3).
Dinis, P.B., Camotim, D. (2018). Proposal to improve the DSM design of cold-formed steel angle columns:
need, background, quality assessment and illustration, Journal of Structural Engineering (ASCE), in press.
Dinis, P.B., Camotim, D., Silvestre, N. (2012). On the mechanics of angle column instability, Thin-
Walled Structures, 52(March), 80–89.
Dinis, P.B., Camotim, D., Preto, V. (2016). Behaviour and design of short-to-intermediate hot-rolled
steel angle columns, Proceedings of International Colloquim on Stability and Ductility of Steel Stuctures
(SDSS 2016 – Timisoara, 30/5 to 1/6), D. Dubina, V. Ungureanu (eds.), Wiley/Ernst & Sohn (Mem
Martins), 477–484.
Ellingwood, B., Galambos, T.V. (1982). Probability-based criteria for structural design, Structural
Safety, 1(1), 15–26.
Fan, J.K. (2009). Theoretical and Experimental Study on Q460 Single Equal-Leg Angles under Axial Com-
pression, Master Thesis, Xian University of Architecture and Technology, Xian, China.
Galambos, T.V., Ravindra, M.K. (1978). Properties of steel for use in LRFD, Journal of the Structural
Division (ASCE), 104(9), 1459–1468.
Kennedy, J.B., Murty, M.K.S. (1972). Buckling of steel angle and tee struts, Journal of the Structural
Division (ASCE), 98(11), 2507–2521.
Kitipornchai, S., Lee, H.W. (1986). Inelastic experiments on angle and tee struts, Journal of Construc-
tional Steel Research, 6(3), 219–236.
Landesmann, A., Camotim, D., Dinis, P.B., Cruz, R. (2016). Short-to-intermediate slender pin-ended
cold-formed steel equal-leg angle columns: experimental investigation, numerical simulations and
DSM Design, Engineering Structures, 132(1 February), 471–493.
Meimand, V.Z., Schafer, B.W. (2014). Impact of load combinations on structural reliability determined
from testing cold-formed steel components, Structural Safety, 48(May), 25–32.
Mesacasa E. Jr., Dinis, P.B., Camotim, D., Malite, M. (2014). Mode-interaction in thin-walled equal-leg
angle columns, Thin-Walled Structures, 81(August), 138–149.
Može, P., Cajot, L.-C., Sinur, F., Rejec, K., Beg, D. (2014). Residual stress distribution of large steel
equal leg angles, Engineering Structures, 71(July), 35–47.
Popovic, D., Hancock, G.J., Rasmussen, K.J.R. (1999). Axial compression tests of cold-formed angles,
Journal of Structural Engineering (ASCE), 125(5), 515–523.
Rasmussen, K.J.R. (2006). Design of slender angle section beam-columns by the direct strength method,
Journal of Structural Engineering (ASCE), 132(2), 204–211.
Ravindra, M.K., Galambos, T.V. (1978). Load and Resistance Factor Design for steel, Journal of the
Structural Division (ASCE), 104(9), 1337–1353.
Shi, G., Liu, Z., Ban, H.Y., Zhang, Y., Shi, Y.J., Wang, Y.Q. (2011). Tests and finite element analysis on
the local buckling of 420 MPa steel equal angle columns under axial compression, Steel and Composite
Structures, 12(1), 31–51.
Silvestre, N., Dinis, P.B., Camotim, D. (2013). Developments on the design of cold-formed steel angles,
Journal of Structural Engineering (ASCE), 139(5), 680–694.
Wakabayashi, M., Nonaka, T. (1965). On the buckling strength of angles in transmission towers, Bulletin
of the Disaster Prevention Research Institute, 15(2), 1–18.
Young, B. (2004). Tests and design of fixed-ended cold-formed steel plain angle columns, Journal of
Structural Engineering (ASCE), 130(12), 1931-1940.

353
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Stability of ring stiffened steel liners under external pressure—


comparison of the existing design concept with 3D-FEM analysis

A. Ecker & H. Unterweger


Institute of Steel Structures, TU Graz, Austria

ABSTRACT: External pressure is often the decisive load case for the empty hydraulic pres-
sure tunnel for calculation of the wall thickness of the steel liner. Due to large diameters and
thin wall thicknesses the steel liner is susceptible to shell buckling under external pressure.
Therefore, the steel liner is often stiffened with rings with a rectangular cross section. In the
Sixties of the last century comprehensive research activities started, including tests, and analyt-
ical formulae were developed for the design. Basically, these formulae are used nowadays for
the dimensioning of the ring stiffened steel liner. This paper deals with the nowadays used
analytical design concept for ring stiffened steel liners and the comparison with
a comprehensive 3D-FEM analysis. This design concept is very complex and not easy to
handle. Two different verifications for buckling of the ring stiffener and for buckling of the
cylindrical shell between the rings are necessary. The combined buckling behavior of the ring
and the cylindrical shell between the rings can only be performed with a 3D-FEM analysis.
This paper sums up the analytical design concept and the results of a comprehensive compari-
son with the 3D-FEM analysis are shown, based on steel liner geometries used in practice
with varying the distance of stiffeners eRing. Due to the lack of accuracy of the analytical
design concept a short overview of a potential new design concept, using the results of the
numerical analysis, is shown.

1 INTRODUCTION

Steel liners for pressure tunnels of hydroelectric power plants are susceptible to shell buckling due
to external pressure (see Figure 1). In case of revisions of the steel liner (also called pipe in this
paper) or the turbines, when the pipe isn´t filled with water, the external pressure is often the
decisive load case. External pressure is caused by infiltration water or by cement injection pressure.
Cement injection is used during erection to fill the gap between the steel liner and the concrete
filling.
Due to large diameters of the pressure tunnels and the nowadays used high strength steels
(up to steel grade S890; this leads to very thin wall thicknesses and high pipe slenderness, up
to R/ts = 170) the steel liner is often stiffened with ring stiffeners (see Figure 1), when external
pressure occurs. The common used ring stiffeners have a rectangular cross section (height hv
and thickness tv) and have equidistant distances eRing between the rings. They are welded over
the hole circumference with full penetration welds onto the pipe.
In different tests (Ullman 1964) and cases of damage (Berti et al. 1998) only single lobe
buckling of steel liners with and without ring stiffeners occur (Figure 1). Usual multi lobe
buckling occurs in cylindrical shells under external pressure, but due to the surrounding rock
in form of a nearly rigid support in radial direction the single lobe buckling mode leads to the
minimum critical external pressure.
The buckling of a ring stiffened steel liner under external pressure is a very complex process. It
is a combined buckling of the ring stiffener itself and the pipe between the rings. As a result, the
accurate external pressure can only be determined by FEM-analysis. Nevertheless, an analytical
model, used in practice, exists to determine the critical external pressure of the ring stiffened steel

354
Figure 1. Geometry of the ring stiffened steel liner – single lobe buckling.

liner with the assumptions of Amstutz (1970) and Feder (1971). The typical buckling shape of the
analytical solution shows Figure 1. The analytical solution for the critical external pressure for the
steel liner without ring stiffeners from Amstutz (1970) is also the basic solution for ring stiffened
steel liners. The well-known formulae of Jacobsen (1974) is also appropriate for determination of
the critical external pressure of a steel liner without ring stiffeners. The critical pressure in these
solutions can´t be explicitly solved and multiple equations are given for iteration. Taras et al.
(2007) prepared the solution of Amstutz (1970) in such a way, that it is possible to get the elastic
and inelastic (assuming local plasticity at the critical cross section) buckling capacity of the steel
liner. All solutions have in common that the surrounding rock is considered as a rigid radial sup-
port, that only allows deformations to the inside of the pipe.
This paper gives an overview of the analytical solutions for the critical external pressure for
ring stiffened steel liners and the performed FEM-analyses for many different examples used
in practice. Also a comparison of the results of the analytical solution for the critical external
pressure with the FEM-analysis is shown. Last but not least a short overview of a possible
new design concept is shown.

2 ANALYTICAL SOLUTION AND FEM-ANALYSIS

2.1 Analytical solution for the ring stiffened steel liner


The analytical solution for the critical external pressure is based on two different design
checks, shown in Figure 2. The minimum of these two critical pressures is decisive for the
design. The first design check (N1) is based on the critical pressure of the ring pcr,Ring.
For the verification N1 the solutions of Amstutz (1970) and Jacobsen (1974) are used. The
same solutions as for an unstiffened steel liner are used, with the difference, that the cross sec-
tion of the ring and the pipe with the effective width bm has to be used (see Figure 2 and

Figure 2. Design due to external pressure p0; Two different design checks (N1 and N2) and assumptions
for calculation of the buckling load pcr,pipe of the steel shell between the ring stiffeners.

355
Eq. (1)). The critical pressure pcr,Ring leads to stresses equal to the yield strength fy in one
point of the critical cross section.
pffiffiffiffiffiffiffiffiffiffiffi
bm ¼ tv þ1:56  R  ts ð1Þ

The second design check (N2) verifies the critical pressure pcr,pipe for the pipe between the
rings stiffeners. The assumptions for the boundary conditions for the pipe between the ring
stiffeners shows Figure 2. The edges of the pipe (cylindrical shell) can be assumed as radial
and longitudinal fixed, because the ring stiffeners are encased in concrete. For this design
check the formulae for the critical pressure in Eurocode EN 1993-1-6 can be used. The rigid
radial support of the surrounding rock is omitted in this design check.
A more complex determination of the critical pressure pcr,Ring for the ring and the pipe shows
Figure 3. This solution is based on the assumptions of Feder (1971) and the pipe between the
rings is also considered in this solution. The pipe and the ring are buckling at the same time and
with the same radial deformations in longitudinal direction. For this reason, the external pres-
sure can be calculated for the entire cross section (pipe and ring; Ages and Iges in Figure 3) with
the theory of Amstutz (1970) in the form of Taras et al. (2007). After that, to determine the
critical pressure pcr,Ring based on the circumferential stresses in the ring and in the shell, the
normal forces and bending moments have to be split for the ring and the shell (pipe) separately,
as shown in Figure 3. For the ring, the critical stress always occurs at the outmost point of the
ring section. The critical pressure again leads to stresses equal to the yield strength fy in the ring
or in the shell. The equations for this procedure to get pcr,Ring are given in Eq. (2) to (13). The
design check N2 in Figure 2 is still necessary. This complex determination of the critical pres-
sure pcr,Ring is the reference solution for the comparison with the FE-Analysis in section 3.
The equations to determine the pressure pges = p0 ∙ eRing are given in Eq. (2) to (9) (with the
Theory of Amstutz (1970) in the form of Taras et al. 2007). The variables to describe the
shape of the dent are a, b, c and α. They are shown in Figure 1. Figure 3 shows the other
parameters, which are needed.

a  λ  sinðλαÞ
b¼ ð2Þ
sinðαÞ

Figure 3. Critical external pressure pcr,Ring of Amstutz/Feder for a steel pipe with ring stiffeners (verifi-
cation N1); assumptions for interaction of ring and steel pipe.

356
 
c ¼ a  λ2  1  cosðλαÞ ð3Þ

λ  tanðαÞ ¼ tanðλαÞ ð4Þ


sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
pges  Rm 3
λ¼ 1þ ð5Þ
E  Iges

with: E. . . modulus of elasticity for steel; 210 000 N/mm²


 
π  Rm  2  E  Iges
 pges  Rm  a  λ  1  cosðλαÞ  pges þ þ j  π ¼ a  A  a2  ðB þ C Þ
E  Ages Rm 3
ð6Þ

with: j. . .gap between steel liner and concrete


A, B, C are given in Eq. (7) to (9)
 
1
A¼ λ  ½λα cosðλαÞ  sinðλαÞ
ð7Þ
λ
"  #
1 2 sin ðλαÞ
2
2 sin ðλαÞ
2  2  2 3
B¼ αþλ α þλ þ 2α λ  1 cos ðλαÞ  4 sinðλαÞ cosðλαÞ  λ  λ þ λ
3
4Rm sin ðαÞ2 tanðαÞ 4
ð8Þ
" #
1 sin ðλαÞ2 2 sin ðλαÞ
2
C¼ λ2 α  λ sinðλαÞ cosðλαÞ þ λ2 α  λ ð9Þ
4Rm sin ðαÞ2 tanðαÞ

The maximum normal force and bending moment for the entire cross section is given by Eq.
(10) and Eq. (11).

E  Iges  
Mges;max ¼ 2
 a  λ2  1  ð1  cosðλαÞÞ ð10Þ
Rm
    
Nges ¼ pges  Rm  a  1 þ λ2  1 cosðλαÞ ð11Þ

With Figure 3 the maximum stresses in the ring and in the shell at the outmost point can be
calculated with Eq. (12) and (13).
 
Nges  bm ering Mges;max  Iv Iges
σ;v;max ¼ þ  fy ð12Þ
Av Wv

with Wv ¼ Iv =ðts þ hv  zs Þ
   
Nges 1  bm ering Mges;max  Is Iges
σ;s;max ¼ þ   fy ð13Þ
As ering  bm  ts 2 =6

With these equations, the solution pges/eRing = pcr,Ring can be determined iteratively, using cer-
tain values of the parameter a to determine the related values α, λ, pges, b and c. Therefore, the
value of a is increased steadily, until one of the two Equations (12) and (13) reaches the yield
strength fy. That defines the critical pressure pcr,Ring.

357
Figure 4. FE-model for numerical studies of the load carrying capacity of steel liners with ring stiffeners
under external pressure; assumed initial imperfections of the FE-model.

2.2 FEM-analysis of the ring stiffened steel liner


Three dimensional Finite-Element-calculations are necessary to model the complex buckling
behavior of the ring stiffened steel liner. As mentioned earlier, it is a combined buckling of the
ring itself and the pipe between the rings.
The FE-model shows Figure 4. Only a quarter of the ring stiffened steel liner needs to be
modelled, because of the double-symmetrical geometry, load and buckling behavior (see
Figure 1). The entire ring stiffened steel liner is surrounded by a rigid surface without any gap
( j = 0). The formulated contact interaction between the ring stiffened steel liner and the rigid
surface allows only deformations toward the center of the pipe. To get the minimum critical
external pressure, no friction between the steel liner and the rigid surface was considered.
The FE-model of the ring stiffened steel liner has a very fine mesh with linear continuum
elements with reduced integration. The steel liner (pipe) has 6 elements over the thickness and
250 elements in circumferential direction (half of the pipe). The choice of the element type and
element number is based on a parametric study. Quadratic elements provide nearly the same
critical external pressure, but with much longer calculation time and numerical difficulties
with the contact interaction (Brantweiner 2018).
An initial imperfection for the numerical calculation is given by an out of roundness with
a maximum deviation from the perfect shape of u = 0.3 ∙ ts (see Figure 4). The out of round-
ness extends across 50 degrees in circumferential direction. This is a common used out of
roundness for the design. Also other types of imperfections were investigated. These results
can be found in Unterweger et al. (2017).
An ideal elastic – ideal plastic material law was used for steel. The calculated maximum
strains were controlled for the critical external pressure and they were quite small (ɛmax about
1%). The numerical calculation is a so called GMNIA (Geometrically and materially non-
linear analysis with imperfections).
The next section shows the results of the analytical solution and the FE-calculations for six
different built ring stiffened steel liners.

3 RESULTS – COMPARISON OF ANALYTICAL AND FEM-ANALYSIS

Table 1 shows the investigated steel liners with ring stiffeners (case F1 to F6). These steel
liners are some examples of built ring stiffened steel liners with a large variation of the pipe
thickness ts, the ring distance eRing and the radius R. To get better insights into the buckling
behavior, additional ring distances eRing,mod out of the usual built range were investigated.
They are specified underneath Table 1.

358
Table 1. Dimensions of representative examples of ring stiffened steel liners.
case hv tv da R ts R/ts eRing eRing/R As/Av Av/Ages Iv/Iges

- mm mm mm mm mm - mm - - - -
F1 150 15 5030 2507,5 15 167 1750* 0,70 0,32 0,25 0,97
F2 150 22 2444 1211 22 55 750* 0,62 0,35 0,47 0,98
F3 200 25 2500 1219 62 20 1500* 1,23 0,15 0,34 0,83
F4 209 16 3338 1659,5 19 87 1167 0,70 0,38 0,35 0,99
F5 200 25 5540 2760 20 138 1000* 0,36 0,39 0,51 0,99
F6 350 25 9070 4517,5 35 129 2000* 0,44 0,28 0,40 0,99

*eRing,mod: F1: 875, 3500, 7000, 10000; F2: 1500, 2270, 4800; F3: 4500; F5: 2500; F6: 4500

Figure 5 shows the typical buckling behavior of ring stiffened steel liners. In the Finite-
Element-Analysis (FEA) also single lobe buckling occurs, as presumed in the analytical solu-
tion. In the red zones in Figure 5, the Von-Mises equivalent stresses reach the yield strength fy
(355 N/mm² for both cases). This indicates a combined buckling of the ring and the pipe
between the rings for case F6 with eRing = 4500 mm. F1 in Figure 5 represents a case with
a very large distance of the ring stiffeners (eRing = 10 000 mm). Here, buckling only occurs in
the pipe and the ring stiffeners can´t increase the critical external pressure of the pipe itself,
due to the large distance of the ring stiffeners. If a significant increase of the critical pressure
of the pipe itself is requested, then the distance of the ring stiffeners has to be quite small
(eRing/R around 0.3 to 1.0; see Unterweger et al. 2017). To compare the analytical solution
with the FEA, also no gap was assumed (j = 0) and the out of roundness u (see Figure 4) has
to be taken into account. This was done by increasing the radius R in the analytical solution
to R* (see Figure 4), based on Eq. (14).

ðR  sinð25 ÞÞ2 þ ½R  ð1  cosð25 ÞÞ


2 þ u2  2  R  u  ð1  cosð25 ÞÞ
R ¼ ð14Þ
2  ½R  ð1  cosð25 ÞÞ  u

The results of the FEA are illustrated exemplarily for the case F1 with four different dis-
tances between the rings. Figure 6 shows the load-deformation curve of the FEA and the
critical pressures of the analytical solutions. On the x-axis the maximum radial deformation
wmax of the pipe between the rings is plotted. The related deformation of the ring is signifi-
cantly smaller. The following critical pressures of the analytical solution are plotted in
Figure 6:

Figure 5. Buckling behaviour of the FE-model of steel liner with radial stiffeners under external pres-
sure (stresses and deformations at ultimate load).

359
Figure 6. Load carrying behavior under external pressure for case F1, with variation of the distance
eRing between stiffeners (material S355).

– critical pressure pcr,Ring of the ring and pipe (design check N1; see Figures 2 and 3; pcr,Ring
Amstutz/Feder)
– critical pressure pcr,pipe of the pipe between the rings (design check N2; see Figure 2; based
on Eurocode 1993-1-6)
– critical pressure of the unstiffened pipe pcr,0 (solution of Amstutz (1970) in the form of
Taras et al. 2007)
The Table 2 shows the critical pressures of the six investigated cases, with different ring distances
eRing. The different critical pressures are shown for case 1 in Figure 6. Also different yield strengths
fy were investigated. Table 2 shows the critical pressures for the lowest (fy = 355 N/mm²) and for
the highest (fy = 890 N/mm²) investigated yield strength.

4 DISCUSSION OF THE RESULTS

As expected, the critical pressure of the FEA and the analytical solution decreases with
increasing distances of the ring stiffeners (see Figure 6 and Table 2). If the distance of
the rings is very large (eg. eRing = 7 000 mm for F1), the critical pressure of the FEA is
equal to the critical pressure of the unstiffened pipe pcr,0 (see Figure 6). This confirms
the analytical solution of Amstutz (2007) for unstiffened pipes. Especially for that large
distances of the ring stiffeners, the solution for the pipe between the rings pcr,pipe (N2)
underestimates the critical pressure, because the rigid support in radial direction is omit-
ted. If the distance between the rings is short, the critical pressure pcr,pipe approaches the
critical pressure of the FEA, but still underestimates it (see Figure 6 and Table 2). The
critical pressure pcr,Ring overestimates the critical pressure of the FEA (see Figure 6). If
the distance between the rings increases, the differences between the solutions also
increase.

360
Table 2. Comparison of critical external pressure.
fy = 355 N/mm² fy = 890 N/mm²

case eRing pcr,Ring pcr,pipe pFEA pcr,0 pcr,Ring pcr,pipe pFEA pcr,0
- mm N/mm² N/mm² N/mm² N/mm² N/mm² N/mm² N/mm² N/mm²
F1 875 1,79 1,54 1,70 0,54 2,88 2,23 2,63 0.70
1750 1,28 0,93 1,06 1,95 0,93 1,07
3500 0,89 0,44 0,61 1,33 0,70 0,77
7000 0,62 0,22 0,54 0,92 0,70 0,59
F2 750 8,17 6,08 6,66 3,40 15,21 12,38 12,54 5.27
1500 5,95* 5,01 4,84 10,77 8,14 7,36
2270 5,58 4,24 4,13 8,70 5,33 6,19
4800 3,98 2,46 3,61 5,98 5,27 5,58
F3 1500 15,12* 18,06 15,52 13,62 28,04 41,81 31,99 24.81
4500 12,42 15,59 13,20 19,45 28,95 26,75
F4 1167 3,86* 3,10 3,79 1,63 8,41 4,88 5,07 2.37
F5 1000 2,47* 2,04 2,55 0,75 4,84 3,39 4,39 1.02
2500 1,89 1,13 1,31 2,91 1,13 1,36
F6 2000 2,30 2,07 2,11 0,84 3,77 3,18 3,55 1.16
4500 1,59 1,21 1,23 2,43 1,21 1,37

* Stress in the shell section is decisive for the critical pressure (see Eq. (13)).

The usage of high strength steels is only useful, if the slenderness of the pipe isn’t too high
(as with every stability problem). If the results of case F1 (very slender pipe R/ts = 167) with
a ring distance of eRing = 1750 mm in Table 2 are compared, nearly the same critical pressures
for S355 and S890 are obtained. For case F2 (R/ts = 55) the usage of high strength steel is
useful and much higher critical pressures can be obtained.
Figure 7 shows the values of Table 2 for the different analytical solutions related to the crit-
ical pressure of the FEA (p_FEA). Despite the high calculation effort of the analytical solu-
tion, especially for the critical pressure pcr,Ring, the difference between the analytical solution
and the FEA is quite large. In most cases the critical pressure pcr,Ring overestimates the critical
pressure of the FEA. The main reason for that, is the assumption, that the ring and the pipe
between are buckling at the same time, with the same radial deformations in longitudinal dir-
ection. For that reason the analytical solution pcr,Ring overestimate the load carrying capacity
of the pipe between the rings. That´s also the reason for the larger difference between pcr,Ring
and pFEA for larger ring distances.
In Figure 7 in most cases the critical pressure pcr,pipe underestimates the pressure of the
FEA. So the minimum of the two pressures pcr,Ring and pcr,pipe leads to a design on the safe
side, but with a very high calculation effort and sometimes uneconomic results.

5 CONCLUSION AND OVERVIEW OF A NEW DESIGN MODEL

As the comparison of the analytical and numerical determined critical external pressure in sec-
tion 4 shows, the minimum of both independent analytical solutions (N1 and N2) leads to
a design on the safe side, but especially the critical pressure of the ring pcr,Ring has a limited
accuracy, even though the calculation effort is very high. Because of that, a simplified design
model, calibrated by the presented results of the FEA, is developed. The design concept stays
the same. Still the critical pressure of the ring, but in a simplified way, and of the pipe between
the rings needs to be determined. After that, the critical pressure is determined with
a factor f. This factor f interpolates the critical pressure of the ring and of the pipe between
the rings, calibrated by the FEA. This improves the accuracy and makes the design procedure
easier to handle (Unterweger et al. 2017).

361
Figure 7. Critical external pressures of the analytical solutions related to the FEA (p_FEA), for all stud-
ied cases; a.) fy = 355 N/mm², b.) fy = 890 N/mm²; designation on x-Axis: case, eRing, fy.

REFERENCES

Amstutz E., 1970, Buckling of pressure-shaft and tunnel linings, Journal of Water Power, pp. 391–399.
Berti D. et al., 1998, Buckling of steel liner under external pressure, Technical Forum, Journal of Energy
Engineering, Vol. 124, pp. 55–89.
Brantweiner, P., 2018, Beulen von ringversteiften Druckschachtpanzerungen unter Außendruck, Master
Thesis, Institute of Steel Structures, TU Graz.
EN 1993-1-6, 2017, Eurocode 3 - Design of steel structures - Part 1-6: Strength and Stability of Shell
Structures.
Feder, G., 1971, Zur Stabilität ringversteifter Rohre unter Außendruckbelastung, Schweizerische Bauzei-
tung, 89. Jahrgang, Heft 42, pp. 1043–1051.
Jacobsen S., 1974, Buckling of circular rings and cylindrical tubes under external pressure, Journal of
Water Power, pp. 401–407.
Taras, A. Greiner, R., 2007, Zum Gültigkeitsbereich der Bemessungsformeln für Druckschachtpanzerun-
gen unter Außendruck, Ernst & Sohn, Stahlbau 76, Ausgabe 10, pp. 730–738.
Ullmann F., 1964, External water pressure designs for steel-lined pressure shafts, Journal of Water
Power, part one pp. 298–305 and part two pp. 338–342.
Unterweger H. Ecker A., 2017, Beulen von stählernen Druckschachtpanzerungen unter Außendruck,
Research report, 211 pages, not published, Institute of Steel Structures, TU Graz.

362
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Numerical investigation of steel built-up columns composed of


track and channel cold-formed sections

M.A. El Aghoury
Structural Engineering & Construction Management Department, Future University, Cairo, Egypt

E.A. Amoush & A.M. El Hady


Civil Engineering Department, Higher Technological Institute, Tenth of Ramadan City, Egypt

S.M. Ibrahim
Structural Engineering Department, Faculty of engineering, Ain Shams University, Cairo, Egypt

ABSTRACT: In this paper, a relatively new cold formed built-up cross section is proposed.
The cross-section is composed of double lipped channels assembled with flanges of double
back to back track sections. The nominal thicknesses used range from 1 mm to 2 mm. The
sections are assembled using enough bolt interconnectors either at the flanges or at the webs,
depending on the sectional configurations. In the present study, columns are subjected to axial
compressive loads although the cross section is suitable for bi-axial load application. Numer-
ical Finite Element model that accounts for both material and geometrical nonlinearities is
developed using ANSYS software. Previously tested built-up cold-formed columns are used in
the present model verification. Then the model is used to investigate the strength of the built-
up columns having different parameters such as the aspect ratios of cross-section components
and the overall slenderness ratios with various column lengths. The initial geometric imperfec-
tions of each specimen were considered. The FE models were further used to quantify the
effect of the fastener interconnector layouts on the combined action and buckling behavior of
cold-formed steel built-up columns. In general, modes of failure for different models were
local buckling and interactive overall distortional buckling.

1 INTRODUCTION

Built-up cold-formed section (CFS) members are used in lightweight steel framing to sustain
higher loads. CFS built-up members cross-section is commonly symmetric which having
higher strength and resistance versus to out-of-plane movements. CFS built-up elements are
considered as a good economical alternative regarding single profiles. Thin-walled CFS mem-
bers are preliminarily failed by local plate buckling and cross-section distortion if not inte-
grated in the design. Built-up members buckling is not necessarily comparable to that noticed
in the individual components. Previous published studies which related to this paper scope are
briefly mentioned below.
Georgieva I. et. al. (2012), evaluated experimentally innovative cross-section shapes for
built-up columns from cold-formed steel (CFS) profiles. They used four single profiles (sigma,
zed, channel and track) to assemble and proposed different four built-up shapes. The direct
strength method (DSM) was used in the design of the suggested built-up columns, where those
columns shapes were proposed to exclude various buckling effects.
Also, Georgieva I. et. al. (2012), used the direct strength method in order to predict the
axial compression capacity of innovative cross-section shapes for built-up columns from cold-
formed steel (CFS) profiles. They obtained on highly stable members with reduced sensitivity
to initial imperfections after excluding various buckling mechanisms. Their study shown that,

363
DSM could be extended for composed members which are sensitive to global-distortional
buckling interaction. Eventually, they concluded that a good agreement with experimentally
obtained normal capacities and suggested that a similar design methodology could be adopted
in structural standards for built-up CFS columns.
Li Y et. al. (2014), carried out a series of axially-compressed tests on built-up box shape
section columns which composed of double channel sections connected via self-drilling screws
at flanges. They investigated different failure modes such global, local and distortional buck-
ling behaviors of single and built-up sections components. They concluded that, the proposed
design method was matched well to the test result and the provisions of AISI for built-up
members were also reliable.
Liu X. and Zhou T. (2017), tested a total of 18 uniaxially loaded compression cold-formed
built-up T-section (CFBUT) columns with different two sectional dimensions and three length
types. The test result showed that the main failure modes of long columns and intermediate
long columns were flexural-torsional buckling mode while the failure mode of short columns
was distortional buckling mode.
Meza F. and Becque J., (2017), tested a total of 20 cold-formed built-up stub columns with
four different cross-sectional geometries having fixed end conditions. The cross-sections com-
posed of flat plate, lipped and un-lipped channels where its nominal thicknesses ranging from
1.2 mm to 2.4 mm. Two assembly options were used, the first using M6 bolts and the second
using M5.5 self-drilling screws. It was concluded that the connectors behavior did not have
a significant effect on the ultimate capacity of CFS built-up stub columns.
Kesawan S. et. al. (2017), tested a total of 45 stub built-up cold-formed steel hollow flange
sections columns. They tested the built-up columns with both hollow flange I-shape (HFI) and
channel sections with/or without stiffened web elements using either single steel sheet or three
steel elements. It was concluded that, the behavior of HFI sections with intermittent fastening
along web to flange connections approximately similar to the continuously welded sections.
M. El Aghoury et. al. (2016), studied the axial strength of columns consist of back-to-back
cold formed sigma sections experimentally and numerically. They found that the failure modes
of short columns were governed by distortional buckling of the flanges. However, in columns
with intermediate heights the failure mode is the interactive distortional overall buckling.
In this paper, a numerical analysis and verification models of previously tested built-up
cold-formed columns are carried out. A relatively new built-up columns cross-section com-
posed of double lipped channel assembled with flanges of double back-to-back track sections
using bolts and subjected to axial compressive load is suggested. Different parameters such as
cross-section component aspect ratios and overall slenderness ratios with various column
lengths are considered. The FE models are further used to quantify the effect of the fastener
interconnector layouts on combined action and buckling behavior response of cold-formed
steel built-up columns.

2 COMPARISON WITH EXPERIMENTS

The results of the F.E. model are compared with the experimental results reported in Meza
F. and Becque J. (2017). Section and material properties reported in Meza F. and Becque
J. (2017) and listed in Table 1, have been used in the F.E. analysis. SHELL181 element is used
to simulate the steel sections while element BEAM4, LINK10 and BEAM44 are used to
model the bolts. The end conditions of the column are taken as fixed-fixed. Figure 1 shows
good agreement between load-displacement relationship for test result and finite element
results along with local buckling failure mode of the column.
Table 1. Nominal dimensions of column components (mm) - Meza F. and Becque J. (2017).
Column Section Web/Plate Width Flange Lip Thickness E (GPa) σ0.2% (MPa)

SC1 T15414 154 54 – 1.4 207 609


SC1 P20024 200 – – 2.4 195 437

364
Figure 1. Column failure modes and load-displacement relationship.

3 PARAMETRIC STUDY

The main aim of this study is to investigate the behavior of axially loaded built-up columns
composed of double lipped channels and double tracks as shown in Figure 2. The webs of
the channels are connected with the track flanges using M6 quality 5.6 bolts. Lipped chan-
nels sections have 40 mm flange width with 15 mm lip. Two different thicknesses 1.25 mm &
2.00 mm are used for the channel. Channel web depth, B, has two value 112 mm and
188 mm. In addition, the double track cross-section dimensions are chosen as shown in
Figure 2. The track cross section is kept the same for all samples with depth 250 mm and
thickness 1.0 mm. Further, to study the built-up columns behavior, different parameters
such overall slenderness, λ, are ranged from 60 to 200 and width-to-depth ratio, B/D equals
0.45 & 0.75, channel flange width-to-thickness ratios equals 90 & 95 as listed in Table 2. In
addition, effect of variation of the end interconnectors spacing, ES, and intermediate inter-
connector spacing, IS, parameters effects are discussed. The studied models are labelled such
that 120C90T-0.45. The first three digits indicates that the overall slenderness ratio, λ,
is 120, the letter C indicates that the flange is channel section, while the following

Figure 2. Built-up column cross-section dimensions and interconnectors arrangement.

365
Table 2. Studied models.
Model λ H B t* B/D t B/t
(mm) (mm) (mm) (mm)

60C90T-0.45 60 1900 112 1.25 0.45 1.0 90


120C90T-0.45 120 3700 112 1.25 0.45 1.0 90
150C90T-0.45 150 4460 112 1.25 0.45 1.0 90
200C90T-0.45 200 6220 112 1.25 0.45 1.0 90
60C95T-0.75 60 3420 188 2.00 0.75 1.0 95
120C95T-0.75 120 6840 188 2.00 0.75 1.0 95
150C95T-0.75 150 8340 188 2.00 0.75 1.0 95
200C95T-0.75 200 11340 188 2.00 0.75 1.0 95

t* = Channel thickness t = Track thickness

number “90” indicates the web of the channel width -to-thickness ratio, (B/t) equals 90, the
letter “T” indicates the web shape is track section and the decimal number “0.45” indicates
that the cross section width-to-depth ratio, (B/D) equal 0.45.

4 MATERIAL PROPERTIES

The material properties of the steel sheets were modelled as elastic-plastic, using bi-linear elas-
tic material behavior as shown in Figure 3-a. Different steel material properties were assigned
to the flat portions and the corner regions. The elastic behavior was defined using a Poisson’s
ratio of 0.3 and Young’s modulus, E and the yield stress, Fy, of the steel material are con-
sidered as 210 GPa and 240 MPa, respectively. The transfer bolts (M6) of grade (5.6) has yield
stress, Fy, 300 MPa.

5 FINITE ELEMENT MODELING

FE models of built-up column were developed using the ANSYS (2012) software. The compo-
nents (channels, tracks and plates) of the built-up columns were modelled using thin-shell

Figure 3. Finite element model.

366
elements. Thin-shell element with four nodes and six degrees of freedom at each node,
SHELL181, is used to simulate the built-up column components and thick shell element is used
to model the end loaded plates. BEAM4 is used to simulate the bolt shank as spring element.
The spring element has two nodes and six degrees of freedom at each node. LINK10 and
BEAM44 elements with two nodes have three and six degrees of freedom at each node; respect-
ively. These elements with full bending stiffness are used to simulate the head and nut of the
bolts. The model of bolts was calibrated by Jakab (2009), to provide the stiffness of the bolts in
connections. The initial overall geometric imperfection is taken as L/1000 at mid-height point of
the column as geometrically perfect FE model generated in ANSYS. Both large deflection and
elasto-plastic material behavior have been incorporated in the non-linear finite element model.
The ends conditions of the column are treated as pined column conditions. The loaded end of
the column model is prevented from both translations in X, Z directions and rotation about
Y axis. However, its unloaded end is prevented from translations in the three directions
X, Y and Z and rotation about Y axis. The model is loaded with vertical load act through the
center of gravity of the column cross-section as shown in Figure 3-b. The load is gradually
increased through sequential load steps up to failure. Arc-length solution technique is used in
the analysis to capture any post buckling behavior specially in short and intermediate specimens.

6 DISCUSSION OF THE RESULTS

6.1 Column failure modes


The results of the column models having overall slenderness ratios, λ, ranged from 60 to 200
and cross section width-to-depth ratio, B/D = 0.45 & 0.75 are presented. The failure modes
are categorized by local buckling, (LB) for short columns with overall slenderness ratios,
λ = 60, and width-to-depth, B/D = 0.45 & 0.75, as shown in Figure 4-a. In addition, for col-
umns with overall slenderness ratios, λ = 120, and width-to-depth, B/D = 0.45 & 0.75, the
failure modes are local-overall distortional buckling, (L-OBD) as shown in Figure 4-b. For

Figure 4. Buckling failure mode shapes for short, medium and long built-up columns.

367
Table 3. Finite element results.
Model name IS ES Pu (kN) PFEM/Py Failure Mode
60C90T-0.45 120 mm — 220.5 0.781 LB
120C90T-0.45 120C90T-0.45 — 192.6 0.681 L-ODB
150C90T-0.45 2.4 D 2.0 D 135.1 0.477 ODB
200C90T-0.45 2.4 D 2.0 D 86.3 0.305 ODB
60C95T-0.75 180 mm — 305.2 0.709 LB
120C95T-0.75 180 mm — 275.4 0.625 L-ODB
150C95T-0.75 2.4 D 2.0 D 191.1 0.640 ODB
200C95T-0.75 2.4 D 2.0 D 111.4 0.259 ODB

*LB : Local buckling.


*L–ODB : Local - Overall Distortional buckling.
*ODB : Overall Distortional buckling.

long columns with λ = 150 & 200, and B/D = 0.45 & 0.75, the overall distortional buckling,
(ODB) is the mode of failure as shown in Figure 4-c & 4-d.

6.2 Column strength


The flange and web slenderness ratio of built-up column section are controlling the behavior
of short column. Width-to-depth ratio of column cross section is controlling the behavior of
long column. The results of the columns with specific values of, IS, and, ES, are presented in
Table 3 and plotted in Figure 5. The ultimate loads increase by increasing the ratio, B/D, for
columns having the same slenderness ratio, λ. The effect of changing the spacing between
interconnectors is considered in this study. For short column, the effect of changing IS, is con-
sidered with values 90, 120 and 240 mm, (i.e 0.75 B, B and 2 B), as shown in Figure 6-a. The
ultimate loads of specimens 60C90T-0.45 are 220.5 kN and 206.5 kN when, IS, taken 120 mm
and 240 mm; respectively as shown in Figure 6-a. Also, for long column different, IS, are
varied with constant ES = 2.5 D. The difference between ultimate capacities of specimen
200C90T-0.45 (long column) when, IS, considered from 0.5 D to 4.8 D, was within 7% as
shown in Figure 6-b.

6.3 Load-displacement relationship


The columns structure response of the studied specimens was obtained by examining their load
deformation plotting. Therefore, the applied load, P, is normalized with respect to the squash

Figure 5. Strength (Pu/Py) ratio and overall slenderness (λ) relationship for built-up columns with
B/D = 0.45& 0.75 ratios.

368
load, Py = gross area times Fy, and plotted versus the lateral displacement of web at mid height
(point 2). Moreover, the axial shortening measured at mid of loading plate at (point 1). Figure
6-a and 6-b represent load-lateral displacement relationship for columns with overall member
slenderness ratios λ = 60, and 200; respectively for different IS values. The ultimate strength is
approximately equal for different interconnector spacing. Meanwhile, for short column
60C90T-0.45, it is noted that columns with interconnector spacing equals 120 mm (B) and
240 mm (2B) are more ductile than column with IS 90 mm as shown in Figure 6-a. From the
results, it is concluded that in long column, IS can be taken as 2.4 D which results in a drop in
the ultimate load of merely 3% compared to column with IS 0.5 D .
Load-displacement relationships in Figure 7 for long columns 150C90T-0.45, 150C95T-0.75,
200C90T-0.45 and 200C95T-0.75, indicate that, the width-to-depth ratio plays an important
role in the ultimate capacity and deformation of this columns. Medium length columns such
120C90T-0.45 and 120C95T-0.75 exhibit more ductile behavior compared to short columns
such as 60C90T-0.45 and 60C95T-0.75 as shown in Figure 8. This is attributed to the change in
failure modes from local buckling in short columns to local-overall distortional buckling in
intermediate columns.

Figure 6. Load-lateral displacement relationships for columns 60C90T-0.45 & 200C90T-0.45.

Figure 7. Load-displacement relationships for columns with λ = 150 & 200.

369
Figure 8. Load-lateral displacement relationships for columns with λ = 60 & 120.

7 CONCLUSIONS

In this paper, the axial strength of built-up columns composed of double lipped channel
assembled with flanges of double back to back track sections has been investigated numeric-
ally. It is found that, the failure modes of short columns were governed by local buckling.
However, for columns with intermediate height the failure mode is the interactive local-overall
distortional buckling. Moreover, failure mode is overall distortional buckling for long
column. For column having the same slenderness ratio, the ultimate load increase by increas-
ing column cross section width to depth ratio, B/D. Moreover, for long column the end inter-
connector spacing and the intermediate interconnector spacing, is recommended to be taken
as two times, two times and half the cross-section depth; respectively. Further, numerical
results reveal that the ultimate loads are not significantly changed by varying the spacing
between interconnectors.

REFERENCES

Georgieva I., Schueremans L., Pyl L. and Vandewalle L., 2012. Innovative cross-section shapes for
built-up CFS columns. Experimental investigation. Twenty-First International Specialty Conference on
Cold-Formed Steel Structures, St. Louis, Missouri, USA, October 24–25.
Georgieva I., Schueremans L., Vandewalle L. and Pyl L., 2012. Design of built-up cold-formed steel col-
umns according to the direct strength method. Procedia Engineering 40, 119–124.
Li Y., Li Y., Wang S., Shen Z., 2014. Ultimate load-carrying capacity of cold-formed thin-walled col-
umns with built-up box and I section under axial compression. Thin-Walled Structures 79, 202–217.
Liu X. and Zhou T., 2017. Research on axial compression behavior of cold-formed triple-lambs built-up
open T-section columns. Journal of Constructional Steel Research 134, 102–113.
Meza F. and Becque J., 2017. Experimental and numerical investigation of cold-formed steel built-up
stub columns. EUROSTEEL 2017, September 13–15,Copenhagen, Denmark.
Kesawan S., Mahendran M., Dias Y. and Zhao W.B., 2017. Compression tests of built-up cold-formed
steel hollow flange sections. Thin-Walled Structures 116, 180–193.
M. El.Aghoury, M.T. Hanna, E.A.Amoush 2016. Axial stability of columns composed of combined
sigma CFS. Proceedings of the Annual Stability Conference Structural Stability Research Council
Orlando, Florida, April 2016.
David C. Fratamico D.C., Torabian S., Zhao X., Kim J.R. R., Schafer B.W., 2018. Experiments on the
global buckling and collapse of built-up cold-formed steel columns. Journal of Constructional Steel
Research 144, 65–80.
ANSYS User Manual Revision 12.0, 2012. ANSYS®, Inc., Canonburg, Pennsylvania, USA.
Gabor Jakab, 2009. Analysis and design of cold-formed C-section members and structures. PhD Disser-
tation, Budapest University of Technology and Economics.

370
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Fatigue failure of skew beam grid steel bridges—causes and


assessment
M.A. El Aghoury
Structural Engineering & Construction Management Department, Future University, Cairo, Egypt

I.M. El Aghoury
Department of Structural Engineering, Ain Shams University, Cairo, Egypt

Amr M. El Hady
Higher Technology Institute of 10th of Ramadan, Egypt

ABSTRACT: The aim of this research is to investigate the causes of failure of Mansoura-
Belkas roadway bridge junction. This junction is one of 4 similar junctions between the bridge
and the crossing two highway roads. The junction consists of skew beam grid steel bays. The
structural system of the grid consists of 3 main girders connected to each other by highly
skewed cross girders connected by skew web stiffeners. Due to the subsequent overloaded
trucks passing over the junction for a considerable period; the failure of two out of three main
girders occurred near the main girder splices (which are not at the mid-span).
By investigating the design capacity of the girders, it was evident that the failure wasn’t just
due to overloading. Since the failure of the bridge took place only 8 years after its construction,
a fatigue life assessment was conducted. The assessment showed that the fatigue detail category of
the highly skew web stiffener connection was highly reduced which needed a thorough investiga-
tion. Thus, a finite element model is developed to accurately model the stress transition zone near
the splices close to the skew web stiffeners at the connections between main girders and the cross
girders. A small parametric study was also conducted to investigate the effect of different param-
eters on the stress concentration factors. Several observations, conclusions and some practical
recommendations for designers are given.

1 INTRODUCTION

Engineers need to continuously update their knowledge and keep up with developments in their
fields of specialization. Bridge engineering continuously evolves and engineers gain valuable
insights from the study and analysis of bridge failures. Bridges are important projects constructed
mostly to relief the traffic congestion at busy intersections and to cross over obstacles presented
by roadway, railway or waterways. Failure of bridges isn’t tolerated by any means as it may
cause loss of life as well as many resources. Each bridge failure has its unique causes and features,
thus the causes of failure cannot be simply generalized (Choudhury & Hasnat 2015). However,
there are some common observations that may be noted. Many bridge failures, especially in steel
bridges, are attributed to bad details such as highly skewed connections and details with sudden
change of geometry or the state of stress. Moreover, as the bridges age and the traffic volumes
grow together with heavier loads of vehicles the useful life of the bridge is reduced. Additionally
old bridges which were poorly designed or didn’t have sufficient safety built into the original
design suffer from these effects. Khan, Ayub and Qadir, (2014) discussed the effect of overloaded
vehicles on the performance of a highway medium span bridge girder in Pakistan. Also, El
Aghoury (2018) studied the effect of overloading on Belkas bridge in Egypt. Hassan, Elsawaf and
Abbas, (2017) studied the current conditions and problems of existing metallic bridges in Egypt
as several steel bridges that form an important part of the transportation network already

371
exceeded their intended design life. They identified the most frequently encountered structural
problems and defects during an inspection campaign of a group of metallic bridges in Egypt.
Reported problems included, but were not limited to, corrosion, fatigue cracks, and permanent
deformations in structural elements.
One of the most important aspects of steel bridge design is accounting for the effect of
fatigue. However, some fatigue details can be miss-categorized and in some cases can be the
source of failure of some critical elements which can cause failure of the whole bridge struc-
ture. Song and Lee, (1999) analyzed the effect of single tensile overload on the crack behavior
by using fatigue crack opening behavior. The unusual growth pattern of short crack after the
single tensile overload was applied, was explained by the variation of crack opening stress.
Fisher et al., (2001) performed forensic and analytical examinations of the failed and
remaining spans of the Hoan Bridge (also known as the Milwaukee Harbor Bridge). They
reported that two of the three girders exhibited full depth fractures, leaving the span near col-
lapse. The report showed that all three girders had web cracks that were initiated from the
crack-like geometric condition that resulted from the intersection of the shelf plate and the
transverse connection plate. These plates had welded connections with intersecting and over-
lapping welds. Brittle fractures were found to develop at every web crack examined without
any detectable fatigue crack extension or ductile tearing at the origin of the crack.
Okelo, (2017) performed a three-dimensional (3D) finite element analysis (FEA) on two
units of the I-345 Bridge using the commercially available software LUSAS. The model was
used to determine the magnitude of the distortion-induced stresses, the fatigue life, and the
effectiveness of the retrofit strategies. The results, in the form of contour plots for the as-built
and repaired connections, showed high stress concentrations at crack initiation points, which
concurs with the bridge inspection reports. The used retrofit methods, when properly carried
out, provide a full restoration of the service life of the connections. The author proposed
a relationship to predict the fatigue life for distortion-induced cracking.
From the previous review of relevant literature, it can be concluded that there is never
a single cause of bridge failure. This work is focused on a special case study in which overload-
ing of a bridge and fatigue crack propagation combined with some unfavorable detailing are
the main factors contributing to the failure of the bridge.

2 PROBLEM STATEMENT AND RESEARCH MOTIVATION

2.1 Introduction
On the 26th of April, 2015, the composite steel span (between axes 13 and 14) of the Belkas
Bridge in Mansoura, Egypt, suddenly collapsed as shown in Figure 1(a). A study by El
Aghoury (2018) reported that the bridge was subjected to repetitive overloads from heavy
trucks in the direction coming from a nearby port. It is worthy to note that the failure didn’t

Figure 1. (a) The two failure locations along the span (b) MG-02 failure near splice.

372
Figure 2. Bridge Plan for the failed bay (Failure locations on MG-01 and MG-02 are encircled).

occur at the mid-span section which has the maximum straining actions. Rather, the failure
occurred near one of the field splices at one-thirds of the span from the maximum moment
side as indicated in Figure 1(a). After the sudden failure of main girder number 1 (MG-01)
the second girder broke at an identical detail as shown in Figure 1(b). The bridge remained
standing in this position but highly distorted until it was demolished.

2.2 Belkas bridge description


The bridge superstructure consisted of 17 reinforced concrete spans along with 4 steel compos-
ite spans. The composite spans cross over main roads or water channels. The bridge road
width is 17 ms equally divided between the two directions separated by a longitudinal expan-
sion joint along the bridge. The failed skewed composite steel bay (between axes 13 and14)
consists of three girders having a span of 41.5m. Figure 2 shows the plan of the bridge with
the locations of failure circled in red.
The steel girder depth is 1560 mm; while this depth is 600 mm near supports. Bridge girders are
rested on neoprene bearings. Steel main girders are connected transversely via highly skewed steel
cross beam (inclined 37° from the bridge axis) to achieve the grillage action of the bridge. The
bridge deck slab is rested on the girders with projections of 1.0 and 1.5 m outside the outer and
inner girders; respectively as shown in Figure 3. It is important to note that the configuration of
the slab and the available roadway lanes is not symmetric with respect to the three main girders
as it is shifted more towards MG-01. The thickness of the deck slab is 250 mm. The composite
action between the R.C. slab and the steel girders is achieved using steel shear connectors in the
form of channel sections. Each girder is divided into 3 parts and spliced using bolted field splices
for webs and flanges. The lengths of these parts are 15.0m, 11.9m and 15.0m; respectively.

2.3 Bridge failure description


One interesting feature of the failure that is worth noting is that the failure did not occur at the
location of maximum moments. Moreover, Figure 1 shows that both MG-01 and MG-02 failed
at the same exact detail (adjacent to the splice). This points to the possibility that there must be
something special about this detail and location, which is not exactly at the girder’s mid-span.
Several factors can contribute to this failure, but what is clear in this bridge is that the cross
girder stiffeners are very close to the splice plate and are highly skewed. In the following

Figure 3. Bridge Half Elevation (Top) and Cross Section (Bottom).

373
Figure 4. The CSI Bridge 3D finite element model.

sections, a full bridge finite element model will be described to investigate the straining actions
near this detail in order to pinpoint the causes of the failure.

3 FULL BRIDGE FINITE ELEMENT MODELING

A complete finite element model was developed for the Belkas Bridge using the commercial
software “CSI Bridge” to determine the straining actions at the above mentioned detail as
indicated in Figure 4. The model accounts for the change in stiffness due to member offsets.
Elastomeric bearings are modelled using link elements. The slab’s composite action is simu-
lated using body constraints between the slab nodes and the underlying shell elements repre-
senting the main girders. Stiffeners are modelled as shell elements at their exact locations.
Regarding the materials, the modulus of elasticity, E and the yield stress, Fy, of the steel
material are taken as 210000 MPa and 360 MPa, respectively and an elastic-perfectly plastic
yield plateau is assumed. A yield stress, Fy = 640 MPa is used to model the splice bolts which
are of grade 8.8 and have a diameter of 16mm. In addition, the modulus of elasticity of the
concrete slab is taken as 28000 MPa. Regarding the bridge loading, it has been reported as
mentioned in section 2.1 that the loads passing over this bridge during the few years preceding
the failure have been considerably heavier than design loads. The reported dimensions and
loads for the two trucks causing the failure are shown in Figure 5. Both trucks were modeled
in CSI Bridge and applied to their corresponding lanes in a multi-step static analysis to simu-
late their movement on the bridge. A uniform Load of 0.5 t/m2 is assumed to fill the rest of
the loading lane to simulate other traffic passing simultaneously with the trucks.
The finite element model was loaded using the “Fissured Rock Truck (B)” on the Left Lane
and the “Gravel Truck (A)” on the Right lane. The maximum straining actions on the three
main girders due to the aforementioned loading scheme are listed in Table 1:
According to the Egyptian code of practice, the left main girder, MG-01, is incapable of
supporting the loads present on the bridge at the incident of failure. A demand-to-Capacity
ratio (D/C) = 127% was calculated according to the Egyptian code. However, this cannot be
the reason for failure as the stresses (statically) induced in the cross calculated from the results
of the FE model did not even reach the yield stress. It can be observed that the bending
moment at the splice location is 91% of the moment at mid-span. So, by plotting the straining
actions on each girder due to dead loads and live loads, it was observed that the skewness of
the cross girders resulted in a big variation in axial forces induced on the main girder with the
largest variation at the section where the failure occurred as shown (plots from CSI Bridge) in
Figure 6. Since the bridge is accurately modelled using actual beam offsets and with its realis-
tic skewness and bearings, the axial forces generated by the live loads in the girder exhibited

Figure 5. Truck A: Gravel Rock Truck (Left), Truck B: Fissured Rock Truck (Right) (Dims. in meters).

374
Table 1. Straining Actions due to the failure loading (SID = Super Imposed Dead ).
M Dead M SID M L+I V Dead V SID V L+I

MG-1 5470 kN.m 1510 kN.m 12,300 kN.m 570 kN 150 kN 1420 kN
MG-2 5470 kN.m 1790 kN.m 8560 kN.m 600 kN 180 kN 1260 kN
MG-3 4800 kN.m 1870 kN.m 6760 kN.m 480 kN 200 kN 680 kN

Figure 6. Envelope of total normal forces along the span of MG-01 (Values are not shown for clarity).

large fluctuations from tension to compression, most likely due to the bridge skew, as shown
in Figure 6. It is to be noted that the variation in axial forces is maximum at the location of
the failed splice. The results of the analysis also indicated that the distribution of cross girder
reactions on the main girder is non-uniform and has its greatest fluctuation range at the loca-
tion of failure. This drove the authors to create a detailed finite element model for the failed
girder to trace the sources of stress concentrations as will be discussed in the following section.

4 OUTER MAIN GIRDER FINITE ELEMENT MODELING

A comprehensive model was created for the connection under the effect of the above acquired
straining actions using the finite element based software ANSYS 12.0 (2012). The developed
finite element model shown in Figure 7 is used to study the different parameters and detail
configurations that might have influenced the state of stresses causing fatigue at the failed sec-
tion. All steel plates, including section flanges, web and stiffeners are modelled using shell
elements (SHELL181) which has four nodes and six degrees of freedom at each node. Large
deformations were accounted for in the analysis and the model was calibrated with the
deformations obtained by the CSI Bridge model. Newton-Raphson iteration technique is used
in solving the nonlinear equations.
To model the splice connection of the bridge main girder; BEAM4 element is used to model
the bolt shank. LINK10 and BEAM44 are used to simulate the heads and nuts of bolts.
CONTA178 is used to model the sliding between the girder steel plates and the splices. Fitted
bolts are used to carry the shear and bearing loads. The effect of the bearing is modelled by
the compression only Spar elements; shear is transferred by the Shaft element. The model is
calibrated by (Jakab 2009), to provide the same stiffness to the connection. To check the valid-
ity of the model used for the splice a finite element model is prepared and verified by (El Hady
2019) using (ANSYS 12.0) using the experimental work reported by Chung (Chung, K.F.;
Lau 1999). The boundary conditions considered for modeling the bridge outer main girder
were adapted to simulate the same boundary conditions as in the total bridge model but with
lateral restraint added for slab above the girder to simulate its continuity. The straining
actions from the CSI bridge model are applied to the main girder at its connections to the
cross girders. The load is incrementally increased through successive load steps.

375
Figure 7. Full Shell elements ANSYS 3D Model with splices.

5 SENSTIVITY STUDY

A sensitivity study was performed to investigate the possible causes of failure and to assess the
details near the failed section under different loading conditions using different possible details.

5.1 Stress distribution with and without splice


It became necessary to investigate the effect of introducing the splice at this location and to
assess the effect of its presence in this location on the stress distribution. Figure 8 shows a plot
of longitudinal normal stresses on the outer main girder for both cases (with and without the
splice). In both cases, the stress plot shows a sudden variation of stresses at the mid-web zone
near the cross girder. However, this variation is sharper and more abrupt when the splice is
present. This is because the splice region increases the stiffness the web plat in this region.
The investigation was further extended to plot the stress distribution along the web height
for both the inner and outer faces of the web plate in order to monitor the change in stresses
across its thickness. Two extreme loading cases were obtained (from the CSI Bridge model) at
the location of the failure section from the straining actions of the cross girder, namely max-
imum negative and maximum positive end moments. Figure 9 shows the plot of longitudinal
normal stresses for both faces due to the two studied extreme cases of loading. It is interesting
to note that the stress range on the outer face at the lower flange of the cross girder (approxi-
mately mid-height of the web of the main girder) is higher than the stress range in the lower
web zone (where the highest normal stresses are expected).

5.2 Effect of splice plate thickness


The effect of splice plate thickness was studied using three thicknesses: 20, 30 and 40 mm, to be
able to quantify the possible increase in stresses beyond the splice plate zone as shown in
Figure 10. It has been noticed that with the smaller thickness (20mm), the loads were trans-
ferred more through the web, while with the 30mm thickness and more, the stresses are totally
transferred by the flange. The splice plate thickness did not influence the stresses along the fail-
ure section.

Figure 8. Full Shell elements ANSYS 3D Model without splices (Left) and with splice (Right).

376
Figure 9. Stress Distribution along the web height at location of failure for both outer and inner web faces.

Figure 10. Stress Distribution around the splice region for splice thicknesses 20, 30 and 40 mm.

5.3 Discussion of results


The CSI Bridge model indicated that the bending moments and normal stresses at the splice
location are approximately 91% of those at the mid- span. Consequently, the stress ranges
due to live loads at the location of failure were slightly lower than those near the mid-span.
However, the failure of the bridge happened near the splice for both girders (MG-01
and MG-02). The zone near the web splice showed a high variation in stress gradient and
a non-uniform distribution of normal stresses along the web section as shown in Figure 9.
This must have changed the fatigue category of such detail (creating a web splice close to
a loaded transverse stiffener). It can be easily proved that the outer girder can suffer from the
initiation of internal cracks at this location due to the high load reversals at this section.
Figure 11 illustrates the cross section of MG-01 showing signs of a kink in the web. This can
indicates that this location might have been the source of crack initiation, then the crack
started propagation with successive overloads during the bridge’s short life span.

Figure 11. Cross section of MG-01 at failure section with kink at possible location of crack initiation.

377
After failure of MG-01, the dead loads were transferred to the other two main girders.
Surprisingly, MG-02 failed at the same detail (which indicates that it might have had some
fatigue cracks too). This leads us to recommend experimentally study this connection to accur-
ately categorize such detail in fatigue and to accurately list it in codes of practice.

6 CONCLUSIONS

A study was performed to investigate the possible causes of failure of one of the steel spans of
Mansoura-Belkas Bridge. The failed span is composed of three main girders having the same
cross section, however the bridge configuration is not symmetric in section causing one of the
main girders to receive significantly higher straining actions due to live loads than the other
two girders. The bridge is highly skew and it has been continuously reported that very heavy
trucks have been passing over the bridge during the past few years before failure.
Bridge splices are usually designed to maintain the bridge section shear and moment capacity.
The effect of splice plate thickness was studied and showed no significant effect on the distribution
of stresses along the failure section. However, having the splice near the skew cross-girder stiffener
caused more stiffening to web and increased the stress variation at the mid-web zone near the
cross girder. This created a zone of high stress gradients which consequently caused failure at this
section (which is near the neutral axis of the main girder which is usually not checked for fatigue).
In summary, it is important to design bridges taking into consideration any possibilities of
overloading as the repetition of these kinds of loads can initiate fatigue cracks. Therefore, it is
important to wisely choose the location of the splice to be away from locations of high stress
concentrations from bracings or lateral cross girders, especially if they are skew. Moreover, it
is highly recommended to study this detail in future to properly categorize it in its appropriate
fatigue detail.

ACKNOWLEDGMENT

The authors are grateful for the generous financial and technical support of AGECS Research
Center (ARC). The revision of Prof. Mohamed Sobhy and the help and support of Eng. Hay-
tham Zaghloul and Eng. Mohannad Emad in FEA results extraction is highly appreciated.

REFERENCES

El Aghoury, I., 2018. Failure of Belkas Roadway Steel Bridge, Lessons Learned. In CSCE Small and
Medium Span Bridges. Quebec City: CSCE, p. 169 (1–11).
ANSYS, 2012. ANSYS.
Choudhury, J.R. & Hasnat, A., 2015. Bridge collapses around the world : Causes and mechanisms.
IABSE-JSCE Joint Conference on Advances in Bridge Engineering-III, (August), pp. 978–984.
Chung, K.F.; Lau, L., 1999. Experimental investigation on bolted moment connections among cold
formed steel members. Engineering Structures, 21, pp. 898–911.
Fisher, J.W. et al., 2001. Hoan Bridge Forensic Investigation Failure Analysis Final Report., (June 2001),
p. 103.
El Hady, A.M., 2019. Behaviour of Fixed Bases Connections for Cold-Formed Steel Frames. Ain Shams
University.
Hassan, M.M., Elsawaf, S.A. & Abbas, H.H., 2017. Existing metallic bridges in Egypt: current condi-
tions and problems. Journal of Civil Structural Health Monitoring, 7(5), pp. 669–687.
Jakab, G., 2009. Analysis and design of cold-formed C-section members and structures. Budapest Univer-
sity of Technology and Economics.
Khan, S.U., Ayub, T. & Qadir, A., 2014. Effect of overloaded vehicles on the performance of highway
bridge girder: A case study. Procedia Engineering, 77, pp. 95–105.
Okelo, R., 2017. Fatigue life prediction for distortion-induced cracking of steel bridges. International
Journal of Steel Structures, 17(2), pp. 801–820.
Song, S.-H. & Lee, K.-R., 1999. Analysis of short and long crack behavior and single overload effect by
crack opening stress. KSME International Journal, 13(12), pp. 865–878.

378
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Lateral torsional buckling of hybrid steel–glass beams

M. Eliasova & I. Pravdova


Faculty of Civil Engineering, CTU in Prague

ABSTRACT: Due to the intensive progress and research in the field of glass structures,
mechanical properties of glass have been distinctively improved and possibilities of using glass
for load-bearing elements are now advanced. The disadvantage of glass is a relatively small
tensile strength, which determines the load-bearing capacity of structural elements such as
beams, walls, railings. Therefore, an effort is made to improve the behaviour of glass beams
by adding another material (steel, wood, carbon fibre) and to create a hybrid structure while
preserving the transparency. If glass or hybrid steel glass beams are used as vertical supporting
fins for glass facades, the loss of lateral torsional buckling cannot be neglected. This paper
deals with lateral torsional buckling of hybrid steel-glass beams consisting of a glass web and
steel flanges. The presented research was focused on full scale experiments whose results were
subsequently used to the verification of the numerical and analytical model of hybrid beam.

1 INTRODUCTION

In modern architecture, the importance of glass is growing, as emphasis is placed on the transpar-
ency and lightness of the structures. Therefore, glass is not used only as a filling material for win-
dows but it is also increasingly used for load-bearing structural elements, which carry, in addition
to their own weight, loads of wind, snow or imposed load. Glass is a material with high compres-
sive strength but low tensile strength. It behaves flexibly until it breaks with brittle fracture, which
occurs suddenly without prior warning. This must be taken into account when installing, operat-
ing and designing structural glass elements, details especially. Thus, the safe design of these struc-
tural elements requires a different approach. The structures should have not only sufficient load-
bearing capacity, i.e. first crack in the glass pane, but they should also be able to carry the load
after the failure. This residual load-bearing capacity of the glass beam can be increased by adding
other material such as steel, stainless steel, wood or concrete, (Louter, 2011).
Research carried out at the Faculty of Civil Engineering of the CTU in Prague was focused
on hybrid beams consisting of glass webs and steel flanges with adhesive bonding between the
steel and glass. The combination of these two materials permits to utilize the strength and duc-
tility of the steel and to keep up the transparency and esthetic appearance. Hybrid beams can
be installed vertically and horizontally and therefore they can be used as facade fins or as
members of high transparent roof or floor structure. If hybrid beams are used as vertical
beams supporting glass facades, consideration should be given to the loss of lateral torsional
buckling that may occur in the load case of the wind suction.
Although glass has been used more and more often on load-bearing structures recently,
there is no European standard for the design of glass constructions, taking into account the
loss of lateral torsional buckling. The stability of glass beams is only dealt with in Australian
Standard AS 1288-2006, in Italian document CNR-DT 210/2013 and Guidance for European
Structural Design of Glass Components, (Feldmann, 2014). The loss of lateral torsional buck-
ling of hybrid steel-glass beams has not been thoroughly investigated yet, only pilot experi-
ments were carried out at RWTH Aachen. The research presented in this paper was aimed at
the experimental investigation of stability problems of hybrid steel-glass beam subjected to
bending.

379
2 PERFORMED EXPERIMENTS

2.1 Shear tests of adhesive bonding


The joint between the glass web and steel flanges plays a crucial role in task of the structural
behavior of hybrid beam. Chosen adhesive has to ensure an adequate stiffness but on the
other hand, it must be soft enough allowing a compensation of different temperature elong-
ation of steel and glass and reduction resp. redistribution of stress peaks or other constraints.
Shear modulus of adhesive is important characteristics and it has a high influence on the load-
bearing capacity, because the stiffness of the bonded connection depends on the stiffness of
the adhesive and geometry of the joint. Based on the results of material tests, the acrylic adhe-
sive SikaFast 5211 NT was selected, (Pravdová, 2016).
Small scale steel – glass connection tests were performed for selected adhesive. Test speci-
mens were assembled from two steel plate with thickness 25 mm and geometric dimensions
75 x 50 mm, which were joined by two float glass panes with thickness 19 mm and dimensions
110 x 50 mm, see Figure 1. Resulting bonded area was four times 50 x 50 mm with bond line
thickness 3 mm. The test specimen was loaded by tension with displacement rate 1 mm/min to
reach the shear stress in the adhesive layer. Shear stress – strain relation with representative
curves is shown in Figure 2. It is obvious that the adhesive is stiff with linear dependence until
the shear stress of 3.5 MPa.

2.2 Load-bearing capacity of hybrid steel-glass beam


In order to determine the load-bearing capacity of a hybrid steel-glass beam with adhesive bond-
ing between glass web and steel flanges, full-scale experiments were performed. During the test,
the dependence of the shear modulus of used adhesive on the glass normal stress was obtained.
The beam was simply supported on a span of 4 m and subjected to a four-point bending test. The
load introducing points were at a distance of 2.4 m from each other, where also lateral supports
were arranged to avoid lateral deflection. The test specimens no. 01, 02 were continually loaded at
the rate of 50 N/s until the beam collapse. Schematic test setup is shown in Figure 3.
The hybrid beam 4.25 m long was made of steel flanges S235 with dimensions 60 x 8 mm, which
were attached to a glass web made of single-layer thermally toughened glass of 19 mm thick and
290 mm height, Figure 4. The joint between the web and the flanges consisted of a 3 mm thick
adhesive layer SikaFast 5211 NT that was able to compensate for manufacturing tolerances.
During the tests, the applied force, vertical deformation in the middle of the beam span,
mutual flange and web displacement at the end of the beam, and transverse displacement of
the flanges and the web in the middle of the span were recorded. Strain gauges were mounted
on the beam for indirect stress measurement. Eight strain gauges were placed on the steel

Figure 1. Test set-up.

380
Figure 2. Shear stress – strain relation.

Figure 3. Four-point bending test set-up of hybrid beam.

flanges and eight strain gauges were situated on the glass web in the middle of the span and at
the supports. The results of the experiments are summarized in Table 1.

2.3 Lateral torsional buckling of hybrid steel-glass beam


Test specimens for experiments aimed at the loss of lateral torsional buckling of hybrid beam
varied from the first set of both cross-section geometry and length. The hybrid beam with the
length of 4.75 m consisted of 40 x 8 mm flanges of steel grade S235, to which a glass web was
attached by adhesive bonding, see Figure 5. The web was made of single-layer thermally
toughened glass 10 mm thick and 290 mm high. For a 3 mm glued joint between the web and
flanges, a two-component acrylic adhesive SikaFast 5211 NT was used again. Before the
experiments, the geometrical (global and local) beam imperfections were measured using the
Surphaser 25HSX panoramic scanner. Figure 6 illustrates the imperfections of horizontal sec-
tions of three beams. It is worth mentioning that the maximum imperfection of the glass web
was very small (<L/1500), although thermally toughened glass was used.
All 3 test specimens (no. 03, 04 and 05) of hybrid beam were simply supported with 4.5 m span
and subjected to four-point bending test with a load introduction distance 2.9 m. The lateral dis-
placement and rotation about the longitudinal axis of the beam was prevented in the places of
load introduction, see Figure 7. The test specimen was loaded by controlling the force value with
the loading rate 50 N/s until the collapse. During the tests, the applied force, the vertical deform-
ation in the middle of the beam span, the lateral displacement of the flanges and the web in the
middle of the span were recorded as well as the rotation of flanges towards the web in the middle
of the span and at the supports. In addition, 16 strain gauges for indirect stress measurement were
installed on the beam. Position of the strain gauges and linear potentiometric transducers is
shown in Figure 7. Lateral torsional buckling of the beams was lost in all tests. Figure 8 illustrates

381
Figure 4. Geometrical dimension – load-bearing capacity (test specimen no. 01 and 02).

Table 1. Summary of results – load-bearing capacity of the hybrid beam.


Max. force Max. moment Max. stress in Max vertical
Test specimen F [kN] M [kNm] glass [MPa] deflection [mm] Failure mode

01 81.27 32.5 41.61 11.0 adhesion


02 137.20 54.9 79.45 22.0 failure of the glass

the images from a test taken with a high-frequency camera. All results for individual hybrid
beams are summarized in the Table 2.
It is obvious from Table 2 that in the experiments focused on the loss of transversal and
torsional stability, the maximum tensile stress in the glass of all samples reached 46 MPa.
Compared to beam 02 held against loss of transverse and torsional stability, see Table 1, in
which a maximal glass stress was 79 MPa, in case of hybrid beam with the loss of lateral tor-
sional buckling the normal stress in glass web is 44% lower.

3 ANALYTICAL AND NUMERICAL MODEL OF HYBRID STEEL-GLASS BEAM

3.1 Möhler method


For the design of hybrid steel-glass beams, the Möhler method can be used provided that the
adhesive bonding between web and flanges behaves linearly. Using this method, it is possible to
determine the stress redistribution over a cross-section composed of different materials. However,
most polymer adhesives have a nonlinear shear stress strain relation. Therefore, the Möhler
method needs to be modified, (Netušil, 2015). In order to calculate the load-bearing capacity of
hybrid beams, it is necessary to create a dependence of the shear modulus of the applied adhesive
on the tensile stress in the glass web. The elastic shear modulus of the adhesive was sought for
different values of the applied bending moment, so that the glass stress calculated by the Möhler
method corresponded to the experimentally determined values at the same points. Based on this
comparison, it was verified that the shear modulus of elasticity of the acrylate adhesive SikaFast-
5215 NT is high at the reached level of normal stress in glass web (47 MPa). The adhesive bond is

382
Figure 5. Geometrical dimension – lateral torsional buckling.

Figure 6. Imperfections of horizontal section along the beam length for 3 test specimens.

stiff and there is no need to take into account flexibility of the adhesive to determine load-bearing
capacity and critical moment.

3.2 Determination of the critical moment


The critical moment of a hybrid steel-glass beam can be determined according to the following
equation
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

π2 ðEIz Þeff π2 ðEIw Þeff
Mcr ¼ L 2 ðGI t Þ eff þ L 2 ; ð1Þ

where (EIz)eff = Eglass Iz,glass + Esteel Iz,steel;


(GIt)eff = Gglass It,glass + Gsteel It,steel;
(EIw)eff = Eglass Iw,glass + Esteel Iw,steel.

383
Figure 7. Four-point bending test set-up of hybrid beam aimed at lateral torsional buckling.

Figure 8. Progress of the deflection during the test.

Table 2. Summary of results – Lateral torsional buckling of the hybrid beam.


Max.
Test Max. force moment Lateral Lateral Max. vertical Max. stress in
Spec. F [kN] M [kNm] displ.* [mm] displ.** [mm] deflection [mm] glass [MPa]

03 50.35 20.14 11.3 89.5 16.4 42.75


04 48.79 19.52 10.0 50.7 14.3 41.53
05 46.63 18.65 16.6 46.6 15.3 46.09

* Lateral displacement of the glass in the middle when the maximum force was reached.
** Lateral displacement of the glass in the middle when the collapse was reached.

Eglass/Esteel = Young’s modulus of elasticity of glass/steel; Gglass/Gsteel = is the shear


modulus of glass/steel and L = length of the beam between of restraint points.
Several models have been created in the software LTBeam. First, model A corresponded to
the hand calculation. It was therefore only the central part of the beam with the length 2.9 m,
simply supported and loaded with end moments, see Figure 9. The lateral displacement and
rotation of the beam around the horizontal axis of the beam was prevented at both edges.
Rotation around the vertical axis of the beam was not prevented. Subsequently, model B was
created, which differed from model A only by transverse supports. At the edges of the mod-
eled part of the beam, not only the transverse displacement and rotation around the horizon-
tal axis of the beam was prevented, but the rotation about the vertical axis of the beam was
prevented as well.
Furthermore, the whole beam was modeled in three variants: model C, model D and model
E, which differed in the type of the supports and boundary conditions. Since the transverse sup-
port in the experiments was neither completely free nor rigid in rotation around the vertical axis
of the beam, model E was created corresponding to the performed experiment 02 of hybrid
beam. To tune the model E, measured lateral displacement was used at the edge and in the
middle of the beam span. The torsional stiffness around the vertical axis of the beam was
selected in the LTBeam so that the model showed the same lateral displacements corresponding
to the experiment. Thus, one support was set with a spring of 150 kNm, the other with 75 kNm.

384
Figure 9. LTBeam model of middle part of the beam, on the left – schematic setup, on the right - results.

The comparison of all models with the experiment is shown in Figure 10. Based on
these results, it is clear that the critical moment calculated on the simplified model of
the mid-section (models A and B) came out lower than on the whole hybrid beam the
model (model C and D). Thus, the design of the critical moment on a simplified model
is on the safe side.

3.3 Numerical model of hybrid steel-glass beam


Numerical model of hybrid beam was created in two computational programs - in ANSYS,
version 11.0 and RFEM 5.05. In both cases, the beam was modeled with the initial imperfec-
tion. The deformation was sinusoidal with an amplitude of 2.5 mm (L/1800), which was the
imperfection of the beam recorded during laser scanning. The boundary conditions were
chosen according to the experiment
The ANSYS hybrid beam model was created using the 3-D eight-node SOLID185 elements.
The material model of the glass was chosen linearly isotropic with a modulus of elasticity E = 70
GPa and a Poisson coefficient ν = 0.23. In the case of steel flanges, it was assumed in the model
that the yield strength was not exceeded, i.e. that the behavior of the material will be linearly elas-
tic. Therefore, a linear isotropic material with a modulus of elasticity E = 210 GPa and a Poisson
coefficient ν = 0.3 was chosen. To describe the behavior of the SikaFast-5215 NT acrylate adhe-
sive, a multilinear isotropic material model with an initial elastic modulus of 260 MPa and
a Poisson coefficient of 0.4 was chosen.
In RFEM 5.05, a numerical model of a hybrid beam was created using 2-D elements.
The material model of glass and steel was isotropic linearly elastic. Also, the isotropic
linearly elastic material model was chosen for the adhesive, because the material charac-
teristics of the adhesive did not change during the experiments. The modulus of elasticity
was 260 MPa, the shear modulus of elasticity was 93 MPa, and the Poisson coefficient
was 0.4.
Figure 11 shows the bending moment dependence on the horizontal displacement of
the top edge of the glass web. The results of both numerical models are compared with
experiments.

3.4 Maximal bending moment of hybrid beam with loss of lateral torsional buckling
The maximal bending moment of a hybrid steel-glass beam with loss of the lateral torsional
buckling can be determined by the following equation

Mb ¼ χLT Mk ; ð2Þ

where χLT = reduction factor for lateral torsional buckling; Mk = characteristic value of the
bending moment.

385
Figure 10. Moment - horizontal displacement relation: comparison of the experiment with critical
moments.

The design procedures for glass beams are currently based on the buckling curves used for
steel structures in determining the reduction coefficient of lateral torsional buckling.
A buckling curve “c” is recommended for safe and reliable design of glass beams, (Feldman,
2014). The same approach has been used to design hybrid beams. In order to find the value of
the buckling coefficient χLT, it is necessary to determine the relative slenderness as
qffiffiffiffiffiffi
λLT ¼ Mcr ;
Mk
ð3Þ

where Mcr = critical moment of hybrid steel-glass beam.


Relative slenderness was determined based on the critical moment that most closely
matched the experiments (software LTBeam, model E), as shown in Figure 10. Figure 12
shows the moment dependence on the horizontal displacement of the upper edge of the glass
web. The graph shows all the calculated maximal bending moments of the hybrid beam for
each buckling curve. Comparing to the results of the experiments, it is evident that a “c” buck-
ling curve, the same as for pure glass beam, can be used for a conservative and safe design of
hybrid steel glass beam with adhesive bonding.

4 CONCLUSIONS

Experimental research of hybrid beams has demonstrated sufficient load-bearing capacity


of hybrid beams with the influence of lateral torsional buckling. The results of the
experiments were compared with both numerical models in ANSYS and RFEM. By
means of a numerical model it was possible to demonstrate a significant effect of the
initial imperfections on the beam behaviour. Furthermore, the critical moments of the
hybrid beam were calculated using the LTBeam software for different variants of bound-
ary conditions.
Based on the experimental research, it has been shown that it is possible for a safety
design of hybrid steel-glass beams to use the same approach as for a pure glass beam,
i.e. to use the buckling curve “c” from EN 1993-1-1 “Design of steel structures”. All
results are valid for the two-component SikaFast 5211 NT acrylic adhesive, which sig-
nificantly affects the behaviour of the beam composed of steel flanges and glass web
with adhesive bonding.

386
Figure 11. Moment - horizontal displacement relation: comparison of the experiment with numerical
models.

Figure 12. Moment - horizontal displacement relation: comparison of the experiment value with max-
imal bending moments

ACKNOWLEDGEMENTS

This research was supported by grant no. GA16-17461S of the Czech Science Foundation.

REFERENCES

Feldmann, M. & Kasper, R. & et al. 2014. Guidance for European structural design of glass components.
Luxembourg. Publications Office of the European Union.
Louter, Ch. 2011. Fragile yet ductile. Zutphen: Wöhrmann Print Service.
Netušil, M. & Eliášová, M. 2015. Design and evaluation of bonded composite glass beams. ICE Proceed-
ings of the Institution of Civil Engineers. Structures and Buildings. 168(7), p. 490-499.
Pravdová, I. & Machalická, K. & Eliášová, M. 2016. Steel-glass structural elements with a new gener-
ation of adhesives. In Challenging Glass 5; Proc. Intern. Conf. Ghent, Ghent University.

387
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

A yielding criterion for seismic gusset plates in tension

M.D. Elliott & L.H. Teh


School of Civil, Mining and Environmental Engineering, University of Wollongong, Australia

ABSTRACT: Gusset plate yielding has been recognised as a complementary ductile yielding
mechanism in the balanced design procedure proposed for special concentrically braced
frames (SCBF). The procedure aims to maximise the frame’s drift capacity. To achieve the
aim, accurate models for determining the yield resistance of relevant components is desirable.
Thicker than necessary gusset plates have been found to limit the drift capacity. Based on the
authors’ previous finding regarding the Whitmore tension resistance of bolted connections,
this paper proposes a more rational yielding criterion for bolted gusset plates in tension. The
proposed block shear yielding criterion not only facilitates the application of the balanced
design procedure to SCBF with bolted braces in tension, but also ensures a reliable ductile
yielding mechanism.

1 INTRODUCTION

In seismic areas, special concentrically braced frames (SCBF) are used to sustain cyclic inelas-
tic yielding of the brace during earthquake (AISC 1997, Astaneh-Asl et al. 2006, Lehman
et al. 2008, Uang & Bruneau 2018). One important aspect in the design of SCBFs is the con-
nection details. When a welded connection is used between the brace and the gusset plate, the
weld is subject to premature fracture thus reducing the frame’s drift capacity. Such an undesir-
able failure mode can be prevented using the balanced design procedure (Lehman et al. 2008).
The balanced design procedure (BDP) balances the primary yielding mechanism such as brace
yielding with secondary yielding mechanism such as gusset plate yielding, thus increasing the
system ductility.
The BDP method was illustrated by Roeder et al. (2011) through a design example involv-
ing a brace welded to the gusset plate. The ultimate and yielding capacities of the gusset plate
were determined using the Whitmore concept (Whitmore 1952). However, the Whitmore con-
cept has been demonstrated to be unviable for bolted gusset plates (Elliott & Teh 2019a).
In this paper, the yielding resistance of a bolted gusset plate is determined using the block
shear model described by Teh & Deierlein (2017). The local ductility characteristics of the
bolted gusset plates resulting from the Whitmore and block shear yielding criteria are com-
pared against each other.

2 ALTERNATIVE CRITERIA FOR FRACTURE AND YIELDING

Consider the bolted gusset plate depicted Figure 1. Teh & Deierlein (2017) have given the
block shear ultimate resistance Bu as
h n oi
Bu ¼ Fu Ant þ 0:6Fu Aev ¼ Fu ðnl  1Þ ðg  dh Þ þ 1:2 ðnr  1Þp þ e1  ð2nr  1Þ=4dh t ð1Þ

in which Fu is the material tensile strength, Ant is the net tension area, Aev is the effective shear
area, nl is the number of bolt lines in the direction of loading (equal to 2 in the figure), g is the

388
Figure 1. Geometry of a bolted gusset plate subject to block shear.

gauge, dh is the bolt hole diameter, nr is the number of bolt rows perpendicular to the loading
direction (equal to 4 in the figure), p is the pitch, e1 is the end distance and t is the plate thick-
ness. The first term in the first line of Equation (1) represents the tensile resistance, while
the second term is the shear resistance of the block. The tension and shear resistance planes in
the block shear mode are indicated in the figure. The effective shear area Aev is the mean
between the gross and the net shear areas defined in the specification (AISC 2016).
The following block shear yielding resistance By is proposed by replacing the material ten-
sile strength Fu in Equation (1) with the yield stress Fy

By ¼ Fy Ant þ 0:6Fy Aev


h n oi ð2Þ
¼ Fy ðnl  1Þ ðg  dh Þ þ 1:2 ðnr  1Þp þ e1  ð2nr  1Þ=4dh t

The block shear yielding resistance By is therefore always lower than the block shear ultimate
resistance Bu, as logically expected.
The Whitmore effective width Ww for determining the tension capacity of a bolted gusset plate
is illustrated in Figure 2. The ultimate Whitmore tension capacity Nu is

Nu ¼ Fu Ae ¼ Fu Ww t ¼ Fu ½ðnl  1Þ ðg  dh Þ þ f2 ðnr  1Þp tan 30o  dh g


t ð3Þ

The Whitmore yielding resistance Ny is computed using the gross Whitmore width Wg (Asta-
neh-Asl 1998, AISC 2017)

Figure 2. The whitmore effective width.

389
Ny ¼ Fy Wg t ¼ Fy ½ðnl  1Þg þ 2 ðnr  1Þp tan 30o
t ð4Þ

Because the Whitmore yielding resistance Ny is computed assuming the area lost to the bolt
holes is still available for yielding resistance, it is anomalously higher than the ultimate Whit-
more tension capacity Nu for many bolted gusset plates (Elliott & Teh 2019b). Such
a condition renders it impossible to apply the Balanced Design Procedure to the SCBF with
bolted braces.

3 LOCAL DUCTILITY RESULTING FROM DIFFERENT YIELDING CRITERIA

The ductility of the brace to the gusset plate connection affects the seismic performance of an
SCBF (Astaneh-Asl et al. 1985, Foutch et al. 1987). However, there have been no formal
measures proposed in the literature or specifications for determining the local ductility
required of a bolted gusset plate under tension in an SCBF. The present work employs simple
measures similar to those used for other (primary) seismic resisting members or systems.
Figure 3 shows the typical load-deflection graph of a bolted gusset plate undergoing the
block shear failure mode. The yielding displacement Δy is the displacement corresponding to
the yield resistance Ry, the ultimate displacement Δu corresponds to the ultimate resistance Ru,
and the maximum displacement Δmax is the displacement where the load drops back to the
yield resistance Ry. Two ductility measures are defined in this paper, called the zultimate duc-
tility factor μu and the maximum ductility factor μmax.
The ultimate ductility factor μu is defined as

Δu
μu ¼ ð5Þ
Δy

and the maximum ductility factor μmax as

Δmax
μmax ¼ ð6Þ
Δy

The ductility factors resulting from the block shear and the Whitmore criteria are given in
Table 1. The specimens were tested by Aalberg & Larsen (1999), Huns et al. (2002) and

Figure 3. Response of a bolted gusset plate failing in block shear.

390
Table 1. Ductility factors under different yielding criteria.

e1 p g dh t Fy Fu μu μmax
Spec nr nl
(mm) (mm) (mm) (mm) (mm) (MPa) (MPa) By Ny By Ny

T-1 50 60 65 21 8.4 2 3 373 537 6.1 6.1 11.5 11.5


T-7 38 47.5 47.5 19 2 5.5 7.5 14.6 20.2
T-9 3 2 5.0 5.6 8.9 10.1
T-11 4 4.1 3.8 9.5 8.7
T1B 38 76 51 21 6.6 3 2 336 450 11.6 11.6 27.5 27.5
T2B 25 51 4 9.3 3.6 12.2 4.2
4U 38 76 102 6.8 2 2 317 415 3.1 3.5 4.1 4.6
8U 4 3.6 3.4 4.6 4.3
12U 6 3.0 2.6 4.0 3.4
14U 7 3.6 3.1 6.9 5.7
16U 8 3.0 2.5 6.0 4.6

Mullin & Cheng (2004). It can be seen that, whether the ultimate ductility factor μu or the
maximum ductility factor μmax is used as the ductility measure, both criteria allow for ductile
performance of the bolted gusset plates. All ultimate ductility factors μu are greater than 3,
except for Specimens 12U and 16U computed under the Whitmore criterion.
The load-deflection graph of Specimen 16U is shown in Figure 4. It can be seen that, even
for this specimen, significant shear yielding deformation followed the tensile fracture with the
sustained loads being greater than the block shear yielding resistance.
The Whitmore yielding resistance Ny of Specimen 16U is 6% higher than its block shear
yielding resistance By. The reverse is true for Specimen T-7 tested by Aalberg & Larsen (1999),
whose block shear yielding resistance is 11% higher as shown in Figure 5. However, it can be
seen that the block shear criterion still allows for a ductile yielding mechanism of the bolted
gusset plate.
The highest ultimate and maximum ductility factors μu and μmax are found for Specimen
T1B tested by Huns et al. (2002), which coincidentally has the same Whitmore and block
shear yielding loads as evident from Table 1. The reason for the high ductility factors of

Figure 4. Specimen 16U (Mullin & Cheng 2004).

391
Figure 5. Specimen T-7 (Aalberg & Larsen 1999).

Specimen T1B is its high aspect ratio so it benefits from the ductile nature of shear yielding of
the block. The load-deflection graph of the specimen is shown in Figure 6.
Figure 4 through 6 show that the block shear yielding criterion results in more consistent
post-yield ductility responses of the bolted gusset plates. The Whitmore yielding resistance can
be either too close to the ultimate resistance (see Figure 4) or too low (see Figure 5).

Figure 6. Specimen T1B (Huns et al. 2002).

392
4 CONCLUSIONS

The block shear yielding criterion is more rational than the Whitmore yielding criterion for
use in the balanced design procedure of bolted SCBFs for a few reasons. The block shear
yielding resistance is always lower than the block shear ultimate resistance, as logically
expected and desired in seismic design. On the other hand, due to the inclusion of areas lost to
the bolt holes in determining the Whitmore yielding resistance, it can be computed to be
higher than the Whitmore tension resistance and complicate the design. Furthermore, the
block shear yielding criterion results in more consistent post-yield ductility responses of bolted
gusset plates of various configurations. It facilitates efficient designs of bolted connections
between tension braces and gusset plates that provide a secondary yielding mechanism to
a special concentrically braced frame.

ACKNOWLEDGMENTS

This research has been conducted with the support of the Australian Government
Research Training Program Scholarship for the first author, administered by the Univer-
sity of Wollongong. The authors would also like to thank the Sustainable Building
Research Centre at the Innovation Campus of the University of Wollongong for the use
of its facilities.

REFERENCES

Aalberg, A., & Larsen, P.K. 1999. Strength and Ductility of Bolted Connections in Normal and High
Strength Steels, Report N-7034, Dept. of Structural Engineering, Norwegian University of Science
and Technology, Trondheim, Norway.
AISC 1997. Seismic Provisions for Structural Steel Buildings, American Institute of Steel Construction,
Chicago IL.
AISC 2016. Specification for Structural Steel Buildings, ANSI/AISC 360-16, American Institute of Steel
Construction, Chicago IL.
AISC 2017. Design Examples – Companion to the AISC Steel Construction Manual, 15th ed., American
Institute of Steel Construction, Chicago IL.
Astaneh-Asl, A. 1998. Seismic behaviour and design of gusset plates, Steel Tips, Structural Steel Educa-
tional Council, Moraga CA.
Astaneh-Asl, A., Cochran, M.L., & Sabelli R. 2006. Seismic detailing of gusset plates for special concen-
trically braced frames. Steel Tips, Structural Steel Educational Council, Moraga CA.
Astaneh-Asl, A., Goel, S.C., & Hanson R.D. 1985. Cyclic out-of-plane buckling of double angle bracing.
J. Struct. Eng., 111 (5): 1135–1153.
Elliott, M.D., & Teh, L.H. 2019a. The Whitmore tension section and block shear. J. Struct. Eng., 145
(2): 04018250.
Elliott, M.D., & Teh, L.H. 2019b. Criterion for bolted gusset plate yielding in special concentrically
braced frames. submitted to J. Struct. Eng.
Foutch, D.A., Goel, S.C., & Roeder, C.W. 1987. Seismic testing of full-scale steel building – Part I.
J. Struct. Eng., 113 (11): 2111–2129.
Huns, B.B.S., Grondin, G.Y., & Driver, R.G. 2002. Block Shear Behaviour of Bolted Gusset Plates, Struc-
tural Engineering Report No. 248, University of Alberta, Edmonton, AB.
Lehman, D.E., Roeder, C.W., Herman, D., Johnson, S., & Kotulka, B. 2008. Improved seismic perform-
ance of gusset plate connections. J. Struct. Eng., 134 (6): 890–901.
Mullin, D.T., & Cheng, J.J.R. 2004. Ductile gusset plates – Test and analyses, Pacific Structural Steel
Conference, Long Beach, California, American Institute of Steel Construction.
Roeder, C.W., Lumpkin, E.J., & Lehman, D.E. 2011. A balanced design procedure for special concen-
trically braced frame connections. J. Construct. Steel Res., 67 (11): 1760–1772.
Teh, L.H., & Deierlein, G.G. 2017. Effective shear plane model for tearout and block shear failure of
bolted connections. Eng. J. AISC 54 (3): 181–194.

393
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Experimental calibration of centrally loaded built-up battened


compression members

G.M. El-Mahdy
The British University in Egypt (BUE), El-Sherouk City, Cairo, Egypt

ABSTRACT: There is a large diversity in the requirements for designing built-up compression
members between international design codes. Most codes, including the North American stand-
ards and specifications specify the use of an equivalent or modified slenderness ratio. In general,
all North American standards and specifications agree on the need of using a modified slender-
ness ratio, but differ in the factor used to multiply the local slenderness ratio. Theoretically it is
difficult to estimate the exact value of this factor, and using experimental methods does not cap-
ture the exact value of this factor either. Hence, a different approach is needed to determine the
value of this factor. In this paper an experimental method to determine the exact value of the
modified slenderness ratio of a built-up compression member by comparing it to the slenderness
ratio of a solid column is given. The results are further illustrated using finite element modeling.

1 INTRODUCTION

Built-up compression members composed of two steel sections interconnected along the open
web, as shown in Figure 1, exhibit an increase in flexibility due to the effect of shear when
compared to solid columns with the same cross-sectional inertia. This leads to a decrease in
compression resistance of the member. In general, international codes, standards, and specifi-
cations handle this problem by using an equivalent or modified slenderness ratio as outlined
by Beedle (1991). In most codes, this modified slenderness ratio is in the form of the square
root of the square of the integral slenderness ratio plus the square of the slenderness ratio of
one main member between interconnectors multiplied by a factor, as given in Equation 1.
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Λm ¼ Λ2o þ ðKi Λi Þ2 ð1Þ

where Λm = modified slenderness ratio about the axis that passes through the open web; Λo =
slenderness ratio of the integral column about the same axis; Λi = slenderness ratio of the
main members between interconnectors; Ki = factor that represents the end conditions of the
main members between interconnectors.
A second requirement given in international codes, standards, and specifications is to limit
the slenderness ratio of the main members between interconnectors, as given in Equation 2.
However, this limit differs from code to code (Beedle 1991).

Λi <bi Λo ð2Þ

where bi = factor which is less than 1.


The theoretical background to using a modified slenderness ratio is to account for the effect
of shear (Timoshenko and Gere 1961). The purpose of limiting the effective length factor of
the main members in between interconnectors is to prevent the occurrence of simultaneous
local and global buckling (Bažant and Cedolin 1994).

394
Figure 1. Built-up compression members with: (a) perforated cover plates; (b) lacing bars; (c) batten plates.

This paper presents a brief review of the development of the modified slenderness ratio in the
AISC specifications (AISC 2005, 2010 & 2016), as well as, the requirements of the Canadian
Standard S16-14 (CSA 2014), and the Egyptian Code of Practice for the design of steel struc-
tures (ECP 2001 & 2007). Previous experimental results are presented to demonstrate the limita-
tions of using them to determine the exact factor, Ki, in the modified slenderness ratio formula.
An experimental example of the proposed calibration method is given backed by the results of
a 3D finite element model.

2 NORTH AMERICAN AND INTERNATIONAL CODES, STANDARDS, AND


SPECIFICATIONS

The concept of using an equivalent or modified slenderness ratio was first introduced into the
first edition of the AISC LRFD Specification in 1986 (AISC 1986) and was developed by
Zahn and Haaijer (1987) based on experimental results conducted by Zandonini (1985) and
Astaneh et al. (1985). In this edition of the AISC – LRFD Specification for Structural Steel
Buildings, the modified slenderness ratio was taken as:
a) For intermediate connectors that are snug-tight bolted,

  ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
s   2ffi
KL KL 2 a
¼ þ ð3Þ
r m r o ri

b) For intermediate connectors that are welded or pretensioned bolted,


  s ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
  2
KL KL 2 a
with a=ri > 50 ¼ þ  50 ð4Þ
r m r o ri
   
KL KL
with a=ri <50 ¼ ð5Þ
r m r o

where (KL/r)m = modified column slenderness of the built-up member; (KL/r)o = column slender-
ness of built-up member acting as a unit; a/ri = largest column slenderness of individual compo-
nents where a = distance between connectors (see Fig. 1) and ri = minimum radius of gyration of

395
individual components. Equations 3-5 were mainly specified for built-up members with the main
members interconnected using perforated cover plates or lacing with tie plates at the ends as
shown in Figure 1 (a) and (b). Using batten plates, Figure 1 (c), to interconnect the main members
was not specifically covered.
Based on the analysis of battened columns by Bleich (1952), Aslani and Goel (1991a) pro-
posed a new formula for the modified slenderness ratio which was introduced into the second
and third editions of the AISC-LRFD specification, as well as, in the 2005 specification
(AISC 2005). This new formula replaced Equations 4 and 5 and took into account the separ-
ation ratio between the main members, α. Hence for intermediate connectors that are welded
or pretensioned bolted the modified slenderness ratio became:

  ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
s   2
KL KL 2 α2 a
¼ þ 0:82 ð6Þ
r m r o 1 þ α2 rib

where α = separation ratio equal to h/2rib; h = distance between centroids of individual com-
ponents perpendicular to the member axis of buckling (see Fig. 1); rib = radius of gyration of
an individual component relative to its centroidal axis parallel to the member axis of buck-
ling. Although this formula was derived based on Bleich’s (1952) analysis of battened col-
umns, the AISC specification (AISC 2005) did not specifically cover battened columns.
Duan and Chen (1988), based on the same experimental data used by Zahn and Haaijer
(1987) proposed a simpler empirical formula for the modified slenderness ratio, which was
first adopted by the Canadian Standard S16.1-M89 (CSA 1989) and has been specified by all
later editions of this standard. This simpler formula takes the modified slenderness ratio as:

  ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
s   
KL KL 2 Km a 2
¼ þ ð7Þ
r m r o ri

where Km = effective length factor for the main members between interconnectors and is
taken as 1.0 for snug-tight bolted connections and 0.65 for welded and pretensioned bolted
connectors. However, based on test results conducted by Temple and Elmahdy (1995, 1996)
and Elmahdy (2008a) it was found that the above mentioned values for Km were applicable to
built-up members that are in contact or separated by filler plates, but for built-up members
composed of two interconnected shapes separated by batten plates (or lacing bars) the Km
factor should be taken as 1.0.
Finally, as more test data became available (Zandonini 1985; Astaneh et al. 1985; Aslani
and Goel 1991b; Temple and Elmahdy 1993, 1995 & 1996; Sherman and Yura 1998; and Lue
et al. 2004) Sato and Uang (2007) developed Equation 7 using a statistical analysis to derive
more values for the Km or Ki factor which was adopted in the 2010 AISC specification (AISC
2010 & 2016). These specifications specified Equation 3 for snug-tight bolted connectors, and
the following for intermediate connectors that are welded or are connected by means of pre-
tensioned bolts:
   
KL KL
with a=ri <40 ¼ ð8Þ
r m r o
  s  2  2ffi
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
KL KL Ki a
with a=ri 440 ¼ þ ð9Þ
r m r o ri

where Ki = 0.50 for angles back-to-back


= 0.75 for channels back-to-back
= 0.86 for all other cases.

396
The Egyptian Codes of Practice (ECP 2001 & 2007) also use a form of Equation 9 but with a Ki
value of 1.25 for battened columns and 1.0 for laced columns, which is much more conservative
than North American standards and specifications.
The EuroCode 3 (EC3 2003) is the only code that has not yet adopted the concept of using
a modified slenderness ratio; instead it uses a continuous (smeared) shear stiffness of the column.

3 PREVIOUS EXPERIMENTAL RESULTS

Figure 2 plots the experimental results of various previous researches using the Ki factors given
in Equation 9 comparing the experimental results with the AISC column curve (AISC 2016)
and the Euler curve. It can be seen from the scatter of results, especially in Figures 2(b) and (d)
that the exact value of the Ki factor cannot be verified by conducting direct compression tests
on struts. Also, from Figures 2(a), (c), (e), and (f), it is hard to decide which curve should be
used to find the exact value of the Ki factor. This is in addition to any uncertainties in the meas-
ured experimental failure load. This leads to the need of further research into the appropriate
factor for Ki that should be used for the modified or equivalent slenderness ratio formula.

4 PROPOSED CALIBRATION METHOD

Determining the exact modified slenderness ratio experimentally for each type of interconnector
and its type of connection has proven to be difficult using direct compression tests due to the
many uncertainties involved. To eliminate the redundancies involved in determining the slender-
ness ratio from a measure of the load carrying capacity of the built-up compression member,
a comparison of the equivalent slenderness ratio to a fixed integral slenderness ratio is proposed.
This is to be done by loading the member and seeing which mode of buckling governs, and at
which integral slenderness ratio both modes of buckling govern, which confirms that the equiva-
lent slenderness ratio is equal to the integral slenderness ratio. As the purpose of this calibration
method is to determine the effect of shear on a certain type of interconnector and its connection
to the main member it is recommended that double channel sections be used to eliminate the
occurrence of flexural-torsional buckling, which occurs in angles, as shown in Figure 3(a).

Figure 2. Previous experimental results with Ki factors from the 2010 AISC specification.

397
Figure 3. Details of experimental specimens.

Similarly, double I-sections can also be used. The results of this method can also be used to
assess the equivalent slenderness ratio of built-up compression members composed of steel
shapes that have more than one axis passing through an open web, such as a four-corner angle
configuration. Its results may also be useful for built-up members with only one axis of symmetry
where one of the modes of buckling is flexural-torsional buckling, as in the case of double back-
to-back angles. Although, the effect of local buckling about an axis that is not parallel to the
overall buckling axis, as in the case of using angle sections, requires further consideration, how-
ever, the effect of shear on these members can be accounted for.
The different types of interconnectors used commonly in practice that need to be calibrated
are lacing bars, batten plates, button plates as in the case of intermittent fillers, and perforated
cover plates. In addition to these types of interconnectors, the different types of connection
for each type of interconnector, must also be calibrated such as fully welded connections, par-
tially welded connections, high-strength pretensioned bolted connections, snug-tight bolted or
riveted connections, as well as, the different number of connecting bolts for a connection.
The test specimens had a cross section as shown in Figure 3(a) which consisted of two
C75x6 channel sections placed toe-to-toe with a separation of 7.56 mm. This gave an equal
moment of inertia, I, about both the X- and Y-axes of 1.34x106 mm4 and a radius of gyration,
r, of 29.6 mm. The length of the specimens, L, was 3552 mm giving an integral slenderness
ratio L/r of 120 about both axes assuming pinned end conditions.
In order to range the modified slenderness ratio about the axis that passes through the open
web, the Y-axis, to a value greater than, equal to, and less than the slenderness ratio of the
integral member about the axis that passes through the closed web, the X-axis, for this config-
uration of double channels, the end condition about the former axis, the Y-axis, was taken as
fixed and the end condition about the latter axis, the X-axis, was taken as fully pinned, as
shown in Figure 3(b). Also, the number of interconnectors was ranged from one to four to
find a pivotal modified slenderness ratio, as represented schematically in Figures 4 and 5.
Figure 3(c) shows the test set-up for the pivotal specimen with 2 interconnectors. The

Figure 4. Schematic details of test specimens.

398
Figure 5. Schematic representation of governing slenderness ratio calculated using a Ki factor of 0.86.

Table 1. Mode of failure and experimental buckling load of test


specimens.
Pexpt.
No. of interconnectors Buckling mode
(kN)

1 Y-axis 109.3
2 Combined X- and Y-axis 208.1
3 X-axis 217.7
4 X-axis 212.3

interconnectors used were batten plates having dimensions of 63.5x63.5x3.18 mm, which were
fully welded along all sides to the main members giving a rigidly connected connection. The
minimum radius of gyration of the C75x6 channel sections was 10.1 mm. Naturally, finding
the exact length at which there is an exact pivot in the equivalent slenderness ratio may take
a few trials. However, as the results obtained using this method can be applied to built-up
compression members with similar types of interconnectors and connections and with any
number of interconnectors to obtain a more accurate value of the equivalent slenderness ratio,
this may still be worth the extra experimental testing required.
Tension tests conducted on coupons taken from the web of the channel sections showed that
the channel sections had an average yield stress, Fy, of 351 MPa and an average modulus of
elasticity, E, of 208 GPa. The interconnector material had a yield stress of Fy = 595 MPa and
a modulus of elasticity of E = 217 GPa.
Table 1 shows the buckling mode and experimental buckling load obtained during testing.
It can be seen that the specimen with two interconnectors was the pivotal specimen and hence
a further study of this specimen is needed using a finite element model.

5 FINITE ELEMENT VERIFICATION

To further study the behavior of the pivotal specimen with two interconnectors a nonlinear
finite element analysis was conducted varying the initial imperfections xo and yo in the X- and
Y- directions, respectively, to see if there was any imperfection sensitivity with regards to the
governing mode of buckling. The finite element analysis was conducted using ABAQUS
(1995), a finite element package. The channel elements and end plates were modeled using
8-node, doubly curved shell elements with reduced integration suitable for thick elements, and
the batten plates were modeled using 4-node shell elements suitable for thin elements. The dis-
placed configurations of the specimen with two interconnectors are shown in Figure 6.
For the finite element model shown in Figure 6 the initial imperfections xo and yo were varied
from 0.1 mm to 3.55 mm (L/1000) in both the X- and Y-directions, respectively. The matrix of
results was documented in Elmahdy (2008b). Figure 7 shows the plots of the imperfection sur-
face for this specimen. The failure load varied from 220 kN to 184 kN and the buckling mode
pivoted from buckling about the X-axis to buckling about the Y-axis as determined by the max-
imum displacement in the X-direction and the maximum displacement in the Y-direction. Hence
this specimen is a good specimen to set the criteria for determining the value of Ki in the modi-
fied slenderness ratio formula. Equating the modified slenderness ratio of this specimen about
the Y-axis to the integral slenderness ratio about the X-axis (120) gave a value for Ki of 0.95.

399
Figure 6. Displaced configuration of finite element model of specimen with two interconnectors.

Figure 7. Finite element analysis imperfection surface.

This shows that the Sato and Uang’s (2007) presumption of using a Ki value of 0.86 for double
channel members placed toe-to-toe given in the AISC specification is un-conservative and not
accurate. Further research is needed to determine the correct Ki values for different types of
interconnectors and their connections.

6 CONCLUSIONS AND RECOMMENDATIONS

Determining the exact modified slenderness ratio for a certain type of interconnector and its con-
nection to the main member from simply using the failure load of a test specimen has proven to
be impossible, as this failure load in itself has many other factors it is dependent on. The new
proposed method to determine the modified slenderness ratio for a certain type of interconnector
and its connection to the main member by comparing it to a fixed integral slenderness ratio is
a more accurate method of determining the Ki factor in the modified slenderness ratio formula.
It means more experimental tests; however, the results obtained can be used to tabulate the Ki
factor for different types of interconnectors and their connection to the main members. It is
believed that by using this method all types of interconnectors and their connection to the main
members may be evaluated experimentally and tabulated for further use in developing design
standards or verifying the load carrying capacity of existing built-up compression members.

ACKNOWLEDGEMENT

The author would like to thank NSERC for funding this research, and the University of
Windsor for hosting this research, as well as Prof. M.C. Temple for all his advice and support.

400
REFERENCES

ABAQUS 1995. ABAQUS/Standard User’s Manual Vol. 1-2, ABAQUS/Standard Example Problems
Manual Vol. 1-2, ABAQUS Theory Manual, and ABAQUS/Post User’s Manual, Version 5.5. USA:
Hibbit, Karlsson, and Sorensen, Inc.
AISC, 1986. ANSI/AISC 360-86 Specification for structural steel buildings. Chicago: American Institute
of Steel Construction.
AISC, 2005. ANSI/AISC 360-05 Specification for structural steel buildings. Chicago: American Institute
of Steel Construction.
AISC, 2010. ANSI/AISC 360-10 Specification for structural steel buildings. Chicago: American Institute
of Steel Construction.
AISC, 2016. ANSI/AISC 360-16 Specification for structural steel buildings. Chicago: American Institute
of Steel Construction.
Aslani, F. & Goel, S.C. 1991a. An analytical criterion for buckling strength of built-up compression
members. Engineering Journal, AISC 28(4): 159–168.
Aslani, F. & Goel, S.C. 1991b. Stitch spacing and local buckling in seismic resistant double-angle braces.
Journal of Structural Engineering, ASCE 117(8): 2442–2463.
Astaneh, A.A., Goel, S.C. & Hanson, R.D. 1985. Cyclic out-of-plane buckling of double-angle bracing.
Journal of Structural Engineering, ASCE 11(5): 1135–1153.
Bažant, Z.P. & Cedolin, L. 1991, Stability of structures: elastic, inelastic, fracture, and damage theories.
New York: Oxford University Press, Inc.
Beedle, L.S. (ed.) 1991. Stability of metal structures: a world view (2nd ed.). Bethlehem: Structural Stabil-
ity Research Council.
Bleich, F. 1952. Buckling strength of metal structures. New York: McGraw Hill Book Company, Inc.
CSA, 1989. CAN/CSA-S16.1-M89 Design of steel structures. Rexdale: Canadian Standards Association.
CSA, 2014. CAN/CSA-S16-14 Design of steel structures. Rexdale: Canadian Standards Association.
Duan, L. & Chen, W.F. 1988. Design rules of built-up members in load and resistance factor design.
Journal of Structural Engineering, ASCE 114(11): 2544–2554.
EC3 2003. EuroCode 3: Design of steel structures. Brussels: European Committee for Standardization.
ECP, 2001. Egyptian code of practice for steel construction and bridges (allowable stress design). Cairo:
Housing and Building National Research Center.
ECP, 2007. Egyptian code of practice for steel construction (load and resistance factor design). Cairo:
Housing and Building National Research Center.
Elmahdy, G.M. 2008a. Contemporary theories of built-up compression members. Proceedings of 2008
Annual Stability Conference, SSRC: 19–38.
Elmahdy, G.M. 2008b. Innovative method for calibrating built-up compression members. Proceedings of
the Annual Conference, CSCE: 1028–1038.
Lue, D.M., Yen, T., Liu, J.L. & Hsu, Y.T. 2004. Experimental investigation for buckling strength of
double-channel columns. Proceedings of 2004 Annual Stability Conference, SSRC: 397–416.
Sato, A. & Uang, C.M. 2007. Modified slenderness ratio for built-up members. Engineering Journal,
AISC 44(3): 269–280.
Sherman, D.R. & Yura, J.A. 1998. Bolted double angle compression members. Journal of Constructional
Steel Research 46(1-3): 470–471.
Temple, M.C. & Elmahdy, G. 1993. Buckling of built-up compression members in the plane of the
connectors. Canadian Journal of Civil Engineering 20(6): 895–909.
Temple, M.C. & Elmahdy, G. 1995. Local effective length factor in the equivalent slenderness ratio. Can-
adian Journal of Civil Engineering 22(6): 1164–1170.
Temple, M.C. & Elmahdy, G. 1996. Slenderness ratio of main members between interconnectors of
built-up compression members. Canadian Journal of Civil Engineering 23(6): 1295–1302.
Timoshenko, S.P. & Gere, J.M. 1961. Theory of elastic stability, New York: McGraw Hill Book
Company, Inc.
Zahn, C.J. & Haaijer, G. 1987. Effect of connector spacing on double angle compressive strength.
Proceedings Materials and Member Behavior, ASCE: 199–212.
Zandonini, R. 1985. Stability of compact built-up struts: experimental investigation and numerical
simulation. Constuzione Metalliche (No. 4): 202–224.
Ziemian, R.D. (ed.) 2010. Guide to stability design criteria for metal structures (6th ed.). Hoboken: John-
Wiley & Sons Inc.

401
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Stainless steel fillet weld tests

N. Feber, M. Jandera, L. Forejtova & L. Kolarik


Czech Technical University in Prague, Prague, Czech Republic

ABSTRACT: The correlation factor for determining the design resistance of fillet welds
should be taken as 1,0 for all stainless steel families according to Eurocode 3 (EN 1993-1-4,
2006). The aim of this study was to investigate the value of the correlation factor for austenitic
grade EN 1.4301, based on experiments. This article describes 5 uniaxial tests of stainless steel
weld connections. The specimen contains two plates of 15 mm thickness connected by four
longitudinal fillet welds. The tests were performed with the load parallel to the weld causing
only shear stress in the fillet weld. Every specimen was tested to investigate the strength, strain
and failure mode. During the tests, for some specimens, deformations were measured using
digital image correlation technique (DIC) to obtain the strain field around the welds and dis-
placement between the plates.

1 INTRODUCTION

The development of steel structures aim is to produce more effective construction. Nowadays,
labour costs make a significant value in total price. Stainless steel offers a considerable reduc-
tion in maintenance costs, because of almost no need to repair the surface treatment. More-
over, it is a structural load-bearing material with great mechanical properties, superior
ductility, and strain hardening characteristic.
Stainless steel structures are increasingly used in the building industry due to their corrosion
resistance, aesthetic appearance, and favorable properties. In fact, the corrosion resistance is
the ability of the steel to form a thin, invisible, impermeable, and renewable passivating layer
on the surface, which gives these steel grades resistance to electrochemical corrosion in oxida-
tion environment. This ability is achieved by chromium addition (it is necessary for the steel
to contain at least 11.7% of the chromium alloy).
Furthermore, both open and hollow cross-sections are made of stainless steel, nevertheless,
those sections are mostly welded. Hence, it is important to investigate the weldability and
strength welding connection. A various study shows that the most of stainless steel grades
have very good weldability by almost all commonly used welding methods (Mathers, 2019),
provided that specific processes are followed.
This paper is focused on the load-bearing capacity of the stainless steel welded joints. The
aim of this work is to determine the value of the correlation coefficient based on experiments
and existing analytical eurtions.

2 LITERATURE REVIEW

The weldability of high-alloy stainless steels depends on their structure, which is related to the
chemical composition. Most stainless steels are materials with very good weldability, but it is
necessary to choose the welding method and procedure correctly regarding their structure.
The choice of additional materials and protection of the welding also plays a major role. Gen-
erally speaking, the additional material should have the same or very similar chemical com-
position as the basic material.

402
It is important to mention that the main disadvantage of the austenitic stainless steel is
susceptibility to intergranular corrosion, which might lead to hot-cracking (in the case of
increasing contents of elements S, P, Si, Ti, Nb in the structure). It occurs owning the deg-
radation of grain boundaries by Cr. (Ogawa & Tsunetomi, 1982) (Whillock & Dunnett,
2004). However, several methods have been used to prevent or minimalize the intergranular
corrosion of austenitic stainless steel. For instance, welding without preheating (maximal
thermal input up to 15 kJ/cm and interpass temperature up to 150°C, i.e. reduce the heat
input as much as possible) or adding strong stabilizing elements such as niobium or titanium
in the stainless steel.
According to valid Eurocode 3 Design of steel joints (EN 1993-1-8, 2005), the design resist-
ance of the fillet weld is sufficient if equations 1 and 2 are both satisfied.
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
fu
σw;Ed ¼ σ2? þ3  ðτ2? þτ2II Þ  ð1Þ
βw γM2

0:9f u
σ?  ð2Þ
γM2

where: fu = ultimate tensile strength of the weaker part of the two base metals; βw = correlation
factor; σ? = normal stress perpendicular to the weld throat plane; τ? = shear stress perpendicu-
lar to the weld throat axis; τII = shear stress parallel to the weld throat axis; γM2 = partial safety
factor for the connection resistance.
As can be seen from equation 1, the fillet weld capacity is reduced by correlation factor
βw. In practice, the calculation procedure for stainless steel weld capacity is almost the same
as for carbon steel, namely the von Mises stress in the weld (see equation 1) comparison
with the ultimate design resistance. Correlation factor βw for carbon steel rises from 0.8 to
1.0 depends on the steel grade. However, the design procedure for stainless steel is regulated
by Eurocode 3 Supplementary rules for stainless steels (EN 1993-1-4, 2006), which states
that the value of the correlation factor for all grades of stainless steels should be assumed to
be 1.0 (see Table 1).
According to Eurocode: 1) (EN 1993-1-8, 2005); 2) (EN 1993-1-4, 2006)
In the work of Fortan, et al., (2018) was proved that the average overestimation of welded
stainless steel joints is 27% for transverse welds and 12% for longitudinal welds. It follows that
existing design rules need to be updated and improved.
Another important research of Rao & Deivanathan (2014) is aimed to examines the tensile
and bending resistance of a weld (TIG method).
Also, a series of experimental studies on the strength of welded joints in stainless steel
constructions (Yang, et al., 2017) was carried out on round bars samples. Six tests with
transverse weld and five samples with longitudinal weld were performed to determine the
capacity of the welds, mode and the angle of failure. Tensile and bending resistance of the
weld provided by the TIG method was examined (Rao & Deivanathan, 2014). The influence
of the TIG method on austenitic steel joints was described in (Niagaj, 2006).

Table 1. Comparison of strengths and correlation factors of materials.


Steel grade S2351) S3551) 1.43012)

fy [MPa] 235 355 210


fu [MPa] 360 490 520
βw [-] 0.8 0.9 1
fu/(βw·γM2) [MPa] 360 436 416

403
3 EXPERIMENTAL INVESTIGATION

3.1 Welding process and procedure


For the experiments carried out for this work, chromium-nickel austenitic steel (AISI 304 or EN
1.4301) has been chosen, which contains 17-19% Cr, 8-10.5% Ni and less than 0.08% C. Generally,
that is the most common stainless steel grade widely used in chemical, medical, gastronomy,
machinery industry and newly in civil engineering due to its sufficient resistant to corrosion in
such surroundings as water, weak alkalis, weak acids, and the city’s atmosphere), high ductility,
and excellent forming properties. Furthermore, austenitic stainless steels retain their structure at
a normal temperature and below the freezing point (Ptáček, 2002). For purpose of this study, sam-
ples were welded manually using the tungsten inert gas (TIG) method (i.e. Method 141 according
to Standard EN (ISO 4063, 2001) with the additional material OK AUTROD 316LSi on diameter
of 1 mm, which has the most suitable chemical composition of the resulting metal for the next
experiments, as it can be seen in Tables 2 and 3.

3.2 Test programme


In the course of the work, 5 tensile welding tests of stainless-steel plates (EN 1.4301 grade)
were performed. Each sample consists of two plates (of size 250 x 140 mm, 15 mm thick) con-
nected by four longitudinal filler welds of 3 mm throat thickness and 50 mm length. As was
mentioned, welding is done manually using the TIG method with the OK AUTOD 316LSi
additional material from ESAB(http://www.esab.cz, 2019). The specimens were bolted to jigs
designed for this purpose (see Figure 1). The tests were performed in a tensile device with

Table 2. Chemical composition of welding material (http://www.esab.cz, 2019).


Mn [%] Mo [%] Cu [%] C [%] Si [%] Ni [%] Cr [%]

1.8 2.6 0.12 0.01 0.9 12.2 18.4

Table 3. Mechanical properties in tensile (http://www.esab.cz, 2019).


Tensile strength Rm [MPa] Yield strength Re [MPa] Extension A [%]

440 - 560 340 - 440 26 - 37

Figure 1. The test setup.

404
Figure 2. Stress-strain curve for the tested material EN 1.4301.

a maximum capacity of 300 kN (predicted strength according to Eurocode 3 (EN 1993-1-4,


2006) of the welded connection Fpred = 144.107 kN as can be seen in equation 5, which results
from equations 3 and 4). The specimens were loaded by axial force causing the shear stress
parallel to the weld throat axis. The load was applied quasi-statically with speed 0.462 mm/
min to yield strength and then 16.5 mm/min to full specimen failure.

pffiffiffi fu
τII  3 ¼ n  ð3Þ
βγM2
F
τII ¼ ð4Þ
aw lw
f u aw l w 520  3  50
Fpred ¼ n pffiffiffi ¼ 4 pffiffiffi ¼ 144107N ð5Þ
βw γM2  3 1:0  1:25 3

where: fu = ultimate tensile strength of the weaker part of the two base metals; βw = correlation
factor; σ? = normal stress perpendicular to the weld throat plane; τ? = shear stress perpendicu-
lar to the weld throat axis; τII = shear stress parallel to the weld throat axis; γM2 = partial safety
factor for the connection resistance.
In this study, the direction of the weld loading was chosen based on experiment result from
the previous research (Fortan, et al., 2018), where was shown that the average overestimation
of the longitudinal welds is lower than the transversal. The difference bin weld strength between
the two loading directions was also noticed in the literature (Kuhlmann, et al., 2008). Hence, it
was decided to investigate more precisely the weld strength in the longitudinal direction,
In order to get information about the deformations, several specimens were measured optic-
ally using stereo digital image correlation (DIC) systems, but only for two of four welds (i.e.
two cameras for one weld, in total four cameras per measurement). This system is able to
measure and visualize full-field, 3-Dimensional measurements of shape, displacement, and
strain (The VIC-3D 8 System, 2019).
One of the essential parameters for design welded joints is the ultimate strength of the base
material. Consequently, the mechanical properties of the base material (EN 1.4301) were deter-
mined from three coupons which were cut in the same direction as welded specimen were tested
in tension. As can be seen in Figure 2, the tested material exhibits a considerable ductility and
strain hardening(fy = 288.93 MPa; fu = 598.9 MPa). Fillet welds strength was calculated in
accordance with standard (EN 1993-1-4, 2006)1, herein mechanical properties of the filler material
(which were shown in Table 3) were not considered for calculation.

3.3 Welds geometry


Owing to the fact, that welds were provided manually, precise throat thickness along the
entire weld cannot be guaranteed. In the preliminary calculation, the weld size of 3 mm was
considered, this value is also may called “nominal” size of the weld. To evaluate the

405
Figure 3. DIC weld geometry in 3D (left) and longitudinal section of the weld (right).

experiment, it was necessary to determine the exact weld size of each specimen. Before the
experiment, weld geometry measurements were done using a measuring gauge. However, due
to the imprecise angle between welded plates, this method is not enough accurate for the pur-
pose of this study. After completion of the experiment, repeated measurements of the weld
were performed at 10 points along each weld using a digital caliper.
Additionally, digital image correlation (DIC) was used to verify the weld geometry. High-
resolution cameras were captured weld geometry at each pixel in three-dimensional space. The
resulting images were evaluated using VIC-3D software (The VIC-3D 8 System, 2019). Unfor-
tunately, due to the time-consuming nature of this method, it was applied only for one sample
(CS05). The average weld size of specimen CS05 according to DIC is 3.12 mm, compared to
caliper measurement (3.29 mm see in Table 4 below) difference is approximately 5%. The dis-
advantage of this method lies in determining the coordinate system of such a complicated
surface.
It is important to mention that besides the correlation factor, the strength of the weld
is also affected by penetration effect. On the basis of visual inspection of the specimens
before and after the experiment, can be assumed that the penetration is insignificant. It
has been also shown in the literature (Niagaj, 2006) that traditional TIG welding has
a much smaller penetration depth than A-TIG. Thus, the weld penetration effect in this
study was neglected, only the nominal size of the weld was taken into account in prelim-
inary weld capacity calculation.

3.4 Test results


In all cases, sample failure occurred at the weld, specifically perpendicular to the weld throat
plane. Failure has mostly ductile feature. Typical failure mode can be seen in Figure 4. The
experimental-to-predictable strength ratios are on the safe side with significant reserve for
every specimen.
The results of the experiments are shown in Table 4. By measuring the weld throat over the
whole length of the weld, the average volume is considered in the evaluation for each speci-
men. For the purpose of evaluation, the partial safety factor γM2 was taken as 1.0. The correl-
ation factor was determined according to equation 6, which results from equations 3 and 4).

f u  aw  l w
βw ¼ n pffiffiffi ð6Þ
γM2 F exp  3

Based on the calculated correlation factors for the tested specimens, the suggested correlation
factor was determined statistically (see equation 7) based on 5 tests.

406
Figure 4. Typical failure mode.

Table 4. Overview of the test results.


aw lw Aw Fexp βw Fexp/Fpred
Name [mm] [mm] [mm2] [N] [-] [-]

CS01 3.0648 48.21 591.1 247 972.90 0.824 1.721


CS02 2.5842 47.72 493.3 228 714.30 0.745 1.587
CS03 3.1222 47.86 597.7 247 972.90 0.833 1.721
CS05 3.2933 49.875 657.0 265 910.30 0.854 1.845
CS06 3.1304 49.13625 615.3 256 310.80 0.830 1.779

Rk ¼ Rm þ k  s ¼ 0:817 þ 2:33  0:037 ¼ 0:903 ð7Þ

where: Rm = mean value of the adjusted test results; k = appropriate coefficient for 5 tests;
s = standard deviation.
During the test, deformations were recorded for the calculation of the stresses in the vicinity
of the joint, the force at fault and the mode of failure. Several samples will also be subjected
to digital image correlation (DIC) deformation measurements during the experiments.

4 CONCLUSION

The correlation factor has an important influence on the designed capacity of the stainless steel
welded joints. For the specified test data, the evaluation ended up as βw ¼ 0:903 for inless steel
(EN 1.4301) longitudinal fillet weld performed by traditional TIG method. Based on investiga-
tion herein, it can be said that the relevant approach in Eurocode (EN 1993-1-4, 2006) leads to
slightly conservative design of welded stainless steel joints and less conservative correlation
factor may be suggested. More specimens are prepared for tests and will be presented on the
conference.

ACKNOWLEDGMENT

This study was performed with the support of TA ČR (TJ01000045), aimed to support the
scientific development of young scientists.

407
REFERENCES

EN 1993-1-4, 2006. Eurocode 3: Design of Steel Structures - Part 1-4: General Rules - Supplementary
Rules for Stainless Steel. Brussels: European Commission for Standardization (CEN).
EN 1993-1-8, 2005. Eurocode 3: Design of steel structures - Part 1-8: Design of joints. Brussels: European
Committee for Standardization (CEN).
EN ISO 4063, 2001. Welding and allied processes - Nomenclature of processes and reference numbers.
Brusel: European Committee for Standardization (CEN).
Fortan, M., Dejans, A., Debruyne, D. & Rossi, B., 2018. Experimental Study on the Strength of Stainless
Steel Fillet Welds. International Conference on Experimental Mechanics, pp. 1–6.
http://www.esab.cz, 2019. OK AUTROD 316LSi. [Online][Accessed 2 02 2019]
Kuhlmann, U., Günther, H.-P. & Rasche, C., 2008. High-strength steel fillet welded connections. Steel
Construction 1, Issue 1, pp. 77–84.
Mathers, G., 2019. Welding of austenitic stainless steel. [Online] Available at: https://www.twi-global.com
/[Accessed 31 01 2019].
Niagaj, J., 2006. Use of A-TIG method for welding of titanium, nickel, their alloys and austenitic steels.
Welding International, pp. 516–520.
Ogawa, T. & Tsunetomi, E., 1982. Hot Cracking Susceptibility of Austenitic Stainless Steels. Welding
research supplement, pp. 82–93.
Ptáček, L., 2002. Nauka o meteriálu II. ISBN 80-7204-130-4 ed. Brno: Cerm, c1999 — 350 s.
Rao, A. V. & Deivanathan, R., 2014. Experimental Investigation for Welding Aspects of Stainless Steel
310 for the Process of TIG Welding. Procedia Engineering, Issue 97, pp. 902–908.
The VIC-3D 8 System, 2019. s.l.:https://www.correlatedsolutions.com.
Whillock, G. O. H. & Dunnett, B. F., 2004. Intergranular corrosion testing of austenitic stainless steels in
nitric acid solutions.
Yang, L. et al., 2017. Strength of duplex stainless steel fillet welded connections. Journal of Constructional
Steel Research, Issue 152, pp. 246–260.

408
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Experimental investigation of compressed stainless steel


angle columns

Aljoša Filipović, Jelena Dobrić, Zlatko Marković, Milan Spremić & Nenad Fric
Faculty of Civil Engineering, University of Belgrade, Republic of Serbia

Nancy Baddoo
Steel Construction Institute, UK

ABSTRACT: This paper presents the experimental programme of an extensive research project
that deals with analysing the behaviour of compressed stainless steel columns with equal angle
sections. The research focuses on different material and geometrical parameters of pin-ended
angle columns produced from austenitic and lean-duplex stainless steel alloys. To assess the influ-
ences of initial imperfections on column compressive resistance caused by different manufacturing
procedures, three types of angle products are included: cold-formed, hot-rolled and laser-welded
sections. In order to obtain both elastic and inelastic ultimate responses, including flexural and
torsional-flexural buckling modes, the lengths and cross-section geometries of specimens are
selected to cover non-slender sections and slender sections covering the entire wide slenderness
range. The experimental programme covers all relevant tests: 15 tensile coupon tests, 15 stub-
column tests, 48 overall buckling tests on slender columns and 2 residual stress tests, including
measurements of initial geometric imperfections. The purpose of these experiments is to provide
a reliable database to determine design rules for angle members and extend the scope of proced-
ures currently codified in EN 1993-1-4 (2015), not only to conventional angle profiles, but also to
contemporary products such as laser-welded angles.

1 INTRODUCTION

Despite high initial material cost, the use of stainless steel structural elements in modern steel
construction has been increasing due to their high corrosion resistance, good strength, tough-
ness and fatigue properties, excellent ductility in combination with strain hardening character-
istics, and full recycling possibility, which favorably affects life cycle costs.
Stainless steel angle sections are manufactured by cold forming, hot rolling and arc or laser
welding. While cold-rolled and press-braked angles are all widely available in austenitic, duplex
and ferritic stainless steel grades, hot-rolled and welded angle elements are commonly produced
from austenitic stainless steels. Stainless steel hot-rolled angles are available in sizes from
15 x 15 x 2 mm to 150 x 150 x 20 mm (Montanstahl, 2019). Laser-welded angles, classified as sharp
edged profiles, are made by employing powerful laser techniques without the use of filler material.
The section dimensions are from 50 x 50 x 5 mm up to 300 x 300 x 20 mm (Montanstahl, 2019).
The specific geometry of an angle section featured in its asymmetry and non-coincidence of
the shear centre with the section’s centroid causes complex structural responses of axially com-
pressed angle columns. The position of the shear centre at the intersection of angle legs has
impact on a negligible warping torsional stiffness. This consequently leads to lower resistance
to torsion, wherefore the overall buckling mode of equal-angle columns is a flexural-torsional
mode in the low and intermediate slenderness range, and minor axis flexural mode in the inter-
mediate and high overall slenderness. It should be noted that the flexural-torsional buckling
resistance of short angle columns tends to the torsional buckling resistance which is actually
the local buckling mode in the case of equal-angle members.

409
Significant experimental and analytical research on the behaviour of compressed carbon steel
angle members has been conducted in the past decades to develop comprehensive design guide-
lines with a wide range of fabricated products and material properties including both mild and
high-strength steel materials (Al-Sayed and Bjorhovde, 1989; Popovic et al., 1999; Adluri and
Madugula, 2002; Young, 2004; Shi et al., 2012; Bhilawe and Gupta, 2015; Cao et al., 2015;
Landesmann et al., 2017). On the other hand, the number of investigations into their stainless
steel counterparts is still limited. De Menezes et al. (2019) presented experimental data for aus-
tenitic stainless steel angles under compression and developed new buckling curves including
the imperfection factor and limiting slenderness. Zhang et al. (2019) experimentally and numer-
ically studied the flexural-torsional buckling behaviour of fixed-ended cold-formed stainless
steel equal-angle columns. The derived database was utilized to evaluate the accuracy of the
codified design rules in Europe, America and Australia/New Zealand. Liang et al. (2019) per-
formed a similar research approach to hot-rolled austenitic stainless steel equal-angle columns.
The lack of experimental data has resulted in the design of axially compressed stainless
steel angle members not being explicitly covered by the current design method stated in EN
1993-1-4 (2015).
This paper briefly shows a comprehensive experimental investigation of the buckling behav-
iour of stainless steel equal-leg angle section columns including material tests, initial geometric
imperfection measurements, stub-column tests, overall buckling tests and residual stress meas-
urements. The obtained experimental data on compressive column capacities are compared
with the predicated design data in accordance with existing provisions for centrically com-
pressed stainless steel members in EN 1993-1-4 (2015). The objective of the research is to
evaluate and quantify differences between structural responses of hot-rolled, laser-welded and
cold-formed stainless steel angle columns influenced by different residual stress distributions
and material properties. The effects of imperfections related to out-of-straightness and end
conditions on the column strength are also assessed.

2 EXPERIMENT

2.1 Material tests


Tensile coupon tests were carried out to obtain the material properties of the two studied
stainless steel alloys. The basic material of hot-rolled and laser-welded specimens was austen-
itic stainless steel grade EN 1.4301 (X5CrNi18-10). The cold-formed specimens were press-
braked from flat strips of lean-duplex stainless steel EN 1.4162 (X2CrMnNiN21-5-1).
Three flat coupons were cut from both legs of hot-rolled and laser-welded specimens in
the longitudinal direction. The material tests on cold-formed angle specimens included two
flat coupons longitudinally cut from the legs and two corner coupons used within the
boundary of the internal radius of the cross-section’s corner region. All coupons were cut by
a water jet cutter to decrease heating of the material during preparation. The coupons were
tested in accordance with EN ISO 6892-1 (2016). The obtained engineering stress–strain
curves for all three types of angle products are provided in Figure 1.
The results showed that the measured yield strengths of austenitic stainless steel both for
hot-rolled and laser-welded angles exceed the nominal value for stainless steel sections given
in EN 1993-1-4 (2015) by a margin of 52% and 39%, respectively. The margin for duplex stain-
less steel used for cold-formed angles is lower, up to 7% for flat coupons. The enhanced yield
strength of the corner regions is approximately 36% higher in comparison with the strength of
flat materials. Relative to austenitic stainless steel, ductility of duplex stainless steel is approxi-
mately 33% lower, while the yield strength is about 80% higher.

2.2 Stub column tests


The stub column tests were performed in accordance with clause A.3.2.1, EN 1993-1-3 (2006).
Three specimens of each size of angle products (hot-rolled, laser-welded and cold-formed) were
tested in pure axial compression to assess their cross-section resistance. Table 1 gives the

410
Figure 1. Engineering stress-strain curves.

specimen designations, nominal dimensions and full cross-section slenderness λp . The slender-
ness λp is taken as the square root of the ratio of the yield stress fy to the full cross-section elastic
buckling stress σcr that was determined using a linear buckling analysis in the Abaqus FE soft-
ware package (2012).
The end plates of the testing machine were fixed flat and parallel. End shortening of the speci-
mens was monitored by longitudinal displacement transducers (LVDTs) mounted on four
points and placed on the upper plate of the testing machine. Four linear electrical strain gauges
(SGs) were affixed to each specimen at its mid-height. The compressive load was applied using
a strain-controlled Amsler hydraulic testing machine with the strain accretion rate of 0.001 s-1.
Figure 2 depicts the experimental setup and failure mode of lase-welded specimen.
The failure mode of hot-rolled and laser-welded angle specimens was inelastic local buckling
of the legs, while cold-formed angles had an elastic local buckling response at the ultimate
limit state. For all specimens, buckling was localised in the middle part of the specimen height
and characterised by the torsion deformations of angle legs (see Figure 2).
Figure 3 compares the EN 1993-1-4 (2015) Class 3 limits for hot-rolled, laser-welded and
cold-formed angle sections in compression with the experimental data. In Figure 3, the ultim-
ate load Nc,u obtained in the experiment was normalized by the corresponding cross-section
yield load and plotted against the slenderness parameters h/(t∙ε). The cross-section yield load
of hot-rolled and laser-welded angles was taken as the product of the cross-sectional area
A and the yield strength fy. To allow for the higher strength material in the corner regions,
enhanced average yield strengths were used for the section yield loads of cold-formed angles.
It can be seen from Figure 3 that hot-rolled and laser-welded angle specimens meet criteria for
Class 3, while the cold-formed angles are classified as slender Class 4 cross-sections.

2.3 Overall buckling tests


To investigate the compressive response of the slender pin-ended angle columns, a total of 48
specimens were tested. Considering the type of steel product, cross-sections’ dimensions and

Table 1. Nominal geometry and section slenderness of specimens in the stub column tests.
Specimen geometry
Designation Stainless Section Number of
of specimen steel Angle Leg width Thickness Length slenderness repeated
series grade product c (mm) t (mm) L (mm) λp tests

AHR 100 1.4301 Hot-rolled 100 10 300 0.44 3


AHR 60 1.4301 Hot-rolled 60 6 180 0.43 3
ALW 100 1.4301 Laser-welded 100 10 300 0.43 3
ALW 60 1.4301 Laser-welded 60 6 180 0.43 3
ACF 80 1.4162 Cold-formed 80 4 240 1.15 3

411
Figure 2. Test setup for stub column specimens.

Figure 3. Assessment of Class 3 slenderness limits for angle-section in compression, EN 1993-1-4.

specimens’ lengths, the specimens are divided into 12 tested series, each including four
repeated tests. The nominal geometrical dimensions of the specimens including the overall
slenderness ratio about the minor principal axis are reported in Table 2.
The specimens were tested under concentric compression loading using a hydraulic testing
machine with a capacity of 5000 kN. The pin-ended bearings were designed to allow rotations
about the minor axis, while restraining major axis rotations as well as twist rotations. All tests
were performed monotonically using displacement control with a proper loading rate not
exceeding 0.01 mm/s up to the maximum load capacity, and then continued up to approxi-
mately 70% of the maximum load before stopping the test.
The specimens were positioned so that their minor principal axis coincided with the axis of
the support rotation, affecting the bending of the specimen about the minor axis. To record
both flexural and torsional movements of the specimen cross-sections, four LVDTs were
attached perpendicular to each leg at mid-height and near to the ends of all tested columns. The
displacement in the expected buckling plane perpendicular to the minor principal axis was also
measured with additional LVDTs at the mid-height of specimens. Besides, four LVDTs were
mounted on the bottom bearing plate to record its rotations and displacements. Axial strains
were monitored by SGs attached around the perimeter surface of the cross-section at the

412
Table 2. Nominal geometry and slenderness ranges of specimens in the overall buckling tests.
Specimen geometry
Stainless
Designation of steel Angle Leg width Thickness Length Overall Repeated
specimen series grade product c (mm) t (mm) L (mm) slenderness tests

AHR 100-500 1.4301 Hot-rolled 100 10 500 26 4


AHR 100-1500 1500 77 4
AHR 100-2500 2500 128 4
AHR 60-800 1.4301 Hot-rolled 60 6 800 70 4
AHR 60-2000 2000 170 4
ALW 100-500 1.4301 Laser-welded 100 10 500 25 4
ALW 100-1500 1500 76 4
ALW 100-2500 2500 127 4
ALW 60-800 1.4301 Laser-welded 60 6 800 68 4
ALW 60-2000 2000 172 4
ACF 80-1000 1.4162 Cold-formed 80 4 1000 65 4
ACF 80-2000 2000 130 4

specimens’ mid-height. A calibrated load cell C6A Force Transducer was used to measure the
applied load. A data acquisition system was used to record the applied load, LVDTs and SGs
readings during the tests.
Depending on the overall slenderness ratios, width-to-thickness ratios of the legs and initial
eccentricity conditions, the specimens failed in flexural, torsional-flexural, and coupled flex-
ural/torsional-flexural modes. Typical deformed shapes of selected specimens after buckling
and corresponded experimental results presented in diagrammatic form as load–axial strain
curves are shown in Figures 4–6.
The local buckling mode characterized by twisting of the cross-section was dominant for
failure of hot-rolled and laser-welded specimens in the low slenderness range, λv = 25–26, as
displayed in Figure 4. However, in the intermediate slenderness range, λv = 68–77, the inelastic
flexural buckling about the minor principal axis was notified for these types of angle products.
Contrary, the overall compressive capacity of the cold-formed angles, λv = 65 was limited by
the capacity of the individual section elements. It should be emphasized that local buckling,
which for this case was coupled with overall flexural buckling about the minor axis, is the
same as the critical bucking mode of short columns (see Figure 5). The largest number of the
most slender hot-rolled and laser welded angle specimens, λv = 127–172, had pure flexural

Figure 4. Failure mode and strain distribution of specimen AHR100-500-3.

413
Figure 5. Failure mode and strain distribution of specimen ACF80-1000-3.

Figure 6. Failure mode and strain distribution of specimen ALW100-2500-1.

buckling about minor axis, as shown in Figure 5. However, three specimens failed due to the
interaction of flexural and torsional-flexural buckling modes. It was found that the value and
direction of the eccentricity of the applied axial load or bearing reaction related to the sec-
tion’s centroid strongly affect the ultimate buckling response of equal-angle slender columns.
The eccentricity direction applied towards the leg tips increases the stress level at the free leg
edges, causing earlier local buckling and reducing the compressive capacity of the angle
column. Conversely, the eccentricity direction applied towards the angle corner leads to
higher buckling resistance of angle columns. Additionally, the interaction between the initial
twist and bow geometric imperfections together with initial eccentricity conditions can
increase the twisting tendency of the column and lead to a much higher failure load than the
tested columns which were filed by flexural buckling about the minor axis. In the case of cold-

414
Figure 7. Comparison between experimental results and design predictions for hot-rolled and laser-welded
specimens.

formed stainless-steel angles, λv = 130, it was observed that the influence of section slenderness
on the ultimate flexural buckling response about the minor axis is not explicitly evident.
The generated experimental data on hot-rolled and laser-welded specimens, failed by pure
flexural buckling about the minor axis, are compared with predicted data obtained in accord-
ance with the method for the flexural buckling resistance of hot-rolled carbon steel angle col-
umns given EN 1993-1-1 (2006). The comparisons are presented in Figure 7. Material
properties obtained from the tensile coupon tests, limiting slenderness λ0 = 0.2, and partial
safety factor γM1 = 1.0 were employed in calculations of column buckling resistance NFb,u,EC.
Using buckling curve b for hot-rolled specimens, the mean experiment-to-predicted buckling
load ratio NFb,u,test/NFb,u,EC is 1.2 and the Coefficient of Variation (CoV) is 13.4%, while in
the case of the laser-welded tested angles, the mean experiment-to-predicted buckling load
ratio NFb,u,test/NFb,u,EC is 1.0 and the CoV is 6.0%. This is an indication of the good agreement
between the the experiment and design, but with unsafe strength predictions for several data.
This indicates that further improvements of design rules are required.

3 CONCLUSIONS

An experimental testing programme was conducted to investigate the buckling behavior of


stainless steel equal-angle columns. The investigation emphasizes the differences between the
structural responses of hot-rolled, laser-welded and cold-formed angles, taking into account
influences of material properties and initial imperfections. Besides, this investigation fills the
gap created by the lack of experimental results and explicit design guidelines for stainless steel
equal-angle columns.
In the stub column tests, the failure mode of the compressed hot-rolled and laser-welded
angle section was governed by inelastic local buckling, which occurred at a stress value that is
up to 51% higher than the measured yield strength. Cold-formed specimens failed by local
buckling prior to reaching the cross-section yield load.
Depending of the overall slenderness ratios and width-to-thickness ratios of of the angle
legs, the tested columns failed in flexural and torsional-flexural buckling modes. The influence
of initial eccentricity conditions on column strengths was indicated in the case of most slender
hot-rolled and laser-welded specimens failed due to the coupling of flexural and torsional-
flexural buckling. Local buckling of cold-formed specimens was observed during the tests.
The obtained experimental results were utilised to evaluate the accuracy of the existing
design rules for carbon steel angles, set out in EN 1993-1-1 (2006). The evaluation results gen-
erally indicated good agreement between the experiment and design. Experimental buckling
strengths of most hot-rolled are higher than the corresponding predicted values for design

415
column curve b. On the other hand, several data for laser-welded specimens are lower than
corresponding design values. However, the experimental database generated in this research is
limited in defining explicit design provisions. This indicates a need to carry out quantitative
numerical parametric studies supported by experiments and theoretical analyses to develop
more suitable design approaches for stainless steel equal-leg angle section columns.

ACKNOWLEDGEMENTS

This investigation is supported by the Serbian Ministry of Education, Science and Techno-
logical Development through the TR-36048 project. The authors are grateful to companies
Montanstahl ag Switzerland, Vetroelektrane Balkana Belgrade, Armont SP Belgrade, Institute
for Testing of Materials Belgrade, Institute for Materials and Structures Faculty of Civil
Engineering University of Belgrade, ConPro Novi Sad, Energoprojekt Industrija PLC Bel-
grade, Vekom Geo Belgrade, CO-Designing, Peri Oplate Belgrade, North Engineering Subo-
tica, Amiga Kraljevo, Mašinoprojekt kopring PLC Belgrade, Sika Belgrade, DvaD Solutions
Belgrade and Soko Inžinjering Belgrade for their support.

REFERENCES

ABAQUS User Manual, Version 6.12, Providence RI, USA: DS SIMULIA Corp 2012.
Adluri, S.M.R. and Madugula, M.K.S. 2002. Flexural Buckling of Steel Angles: Experimental
Investigation, Journal of Structural Engineering 122(3): 309–317.
Al-Sayed, S.H. and Bjorhovde, R. 1989. Experimental study of single angle columns, Journal of Construc-
tional Steel Research 12(2): 83–102.
Bhilawe, J.V. and Gupta, L.M. 2015. Experimental Investigation of Steel Equal Angle Subjected To
Compression, Engineering Structures and Technologies 7(2): 55–66.
Cao, K., Guo, Y.J. and Zeng, D.W. 2015. Buckling behavior of large-section and 420 MPa high-strength
angle steel columns, Journal of Constructional Steel Research 111: 11–20.
EN 1993-1-1 Eurocode 3: Design of steel structures - Part 1-1: General rules and rules for buildings 2006.
Brussel: CEN European Committee for Standardization.
EN 1993-1-3 Eurocode 3: Design of steel structures – Part 1-3: General rules – Supplementary rules for
cold-formed members and sheeting 2006. Brussel: CEN European Committee for Standardization.
EN 1993-1-4 Eurocode 3: Design of steel structures - Part 1-4: General rules - Supplementary rules for
stainless steels 2015. Brussel: CEN European Committee for Standardization.
EN ISO 6892-1, Metallic materials—Tensile testing Part 1 : Method of test at room temperature 2016.
Brussel: CEN European Committee for Standardization.
Landesmann, A., Camotim, D, Dinis, P., Cruz, R. 2017 Short-to-intermediate slender pin-ended
cold-formed steel equal-leg angle columns: Experimental investigation, numerical simulations and
DSM design, Engineering Structures 132: 471–493.
Liang, Y., Jeyapragasam, V., Zhang, L., Zhao, O. 2019. Flexural-torsional buckling behaviour of
fixed-ended hot-rolled austenitic stainless steel equal-leg angle section columns, Journal of Construc-
tional Steel Research 154: 43–54.
de Menezes, A. A., Pedro, P., de Lima, L., da Silva, A. 2019. Experimental and numerical investigation
of austenitic stainless steel hot-rolled angles under compression, Journal of Constructional Steel
Research 152: 42–56.
Montanstahl AG, https://www.montanstahl.com (date accessed 25.02.2019.).
Popovic, D., Hancock, G. J. and Rasmussen, K. J. R. 1999. Axial compression tests of cold-formed
angles, Journal of Structural Engineering 125: 515–523.
Shi, G., Liu, Z., Ban, H., Zhang, Y., Shi, Y., Wang, Y. 2012. Tests and finite element analysis on the
local buckling of 420 MPa steel equal angle columns under axial compression, Steel and Composite
Structures 12(1): 31–51.
Young, B. 2004. Tests and Design of Fixed-Ended Cold-Formed Steel Plain Angle Columns, Journal of
Structural Engineering 130(12): 1931–1940.
Zhang, L., Tan, K. H. and Zhao, O. 2019. Experimental and numerical studies of fixed-ended
cold-formed stainless steel equal-leg angle section columns, Engineering Structures 184: 134–144.

416
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Accurate and efficient account of geometrical imperfections in


Koiter analysis of elastic solid-like shells

G. Garcea, F.S. Liguori, L. Leonetti, D. Magisano & A. Madeo


Dipartimento di Ingegneria Informatica, Modellistica, Elettronica e Sistemistica, Università della Calabria,
Rende (Cosenza), Italy

ABSTRACT: The Koiter method recovers the equilibrium path of an elastic structure using
a reduced model, obtained by means of a quadratic asymptotic expansion of the finite element
model. Its main feature is the possibility of efficiently performing sensitivity analysis by
including a-posteriori the effects of the imperfections in the reduced non-linear equations. The
state-of-art treatment of geometrical imperfections is accurate only for small imperfection
amplitudes and linear pre-critical behavior. This work enlarges the validity of the method to
a wider range of practical problems through a new approach, which accurately takes into
account the imperfection without losing the benefits of the a-posteriori treatment. A mixed
solid-shell finite element is used to build the discrete model. Some numerical tests validates the
proposal.

1 INTRODUCTION

Load-carrying capabilities of thin-walled beams and shells are often determined by buckling,
which often occurs for loads much lower than the failure loads of materials. As a consequence
of modal buckling interaction Garcea et al. (2016), shell-like structures may exhibit a very
unstable post-buckling behavior and may be highly sensitive to geometrical imperfections. In
light of this an imperfection sensitivity analysis Flores and Godoy (1992) becomes mandatory.
It consists in seeking the so called worst (detrimental) imperfection cases, which are the
shapes of the geometrical imperfections associated with the minimum limit load. The Monte
Carlo simulation, generally adopted to this end, may require thousands of equilibrium path
evaluations. The use of composite structures, which require an optimization of materials
Genoese et al. (2018) and layup Henrichsen et al. (2016), further complicates the design pro-
cess. Standard path-following approaches, aimed at recovering the equilibrium path for
a single loading case and assigned imperfections, are not suitable for this purpose because of
the high computational burden of the single run. For these reasons, the FE implementation of
Koiter’s theory of elastic stability Rahman and Jansen (2010); Rizzi et al. (2013) has recently
become more and more attractive. The Koiter method consists of the construction of
a reduced model, in which the FE model is replaced by its second order asymptotic expansion
using the initial path tangent, m buckling modes and the corresponding second order modes.
Once the reduced model is built, the equilibrium path can be obtained by solving the non-
linear reduced system of m equations in m þ 1 unknowns, which represent the modal ampli-
tudes and the load factor. The coefficients of the reduced system are evaluated using strain
energy variations up to the 4th order. Shell structures can require a very large number of FE
DOFs to avoid significant discretization errors, while m is usually at most on the order of 10
to 20.

417
The solution of the reduced system has a very low cost, negligible compared to the construc-
tion of the reduced model, because of the small value of m. However the most attractive fea-
ture of the Koiter approach consist in the possibility of including a-posteriori the effects of the
imperfections in the reduced model of the structure without imperfections, built once and for
all, by simply adding some energy terms in the reduced system. In this way, a very fast imper-
fection sensitivity analysis, which can consider a very large number of imperfections in
a reasonable computational time, is obtained, making possible a Monte Carlo simulation
Liguori et al. (2018, 2019).
In Magisano et al. (2017b); Leonetti et al. (2018b); Magisano et al. (2018), the Koiter
method has been implemented exploiting the non-linear Cauchy continuum based on a Green
strain measure. In this way, adopting the mixed Hellinger-Reissner variational formulation,
the strain energy has a 3rd order only polynomial dependence on the FE DOFs with the zero-
ing of all the fourth order strain energy variations. The resulting asymptotic formulation
appears accurate, efficient and simple.
Nevertheless, the state-of-the-art a-posteriori account of geometrical imperfections is based
on the hypothesis of linear pre-critical behaviors and small imperfection amplitudes, that
leads to additional terms in the reduced system which are just linear in the load factor. As
a consequence, inaccuracies occur even for small pre-critical non-linearities and significant
imperfection amplitudes, considerably limiting the application of the method.
The goal of this paper is to overcome these inaccuracies and to propose a new accurate
treatment of the geometrical imperfections, which coherently takes into account the effects of
the geometrical imperfection up to the 2nd order, without losing the advantages of the a-pos-
teriori account.

2 THE KOITER ASYMPTOTIC STRATEGY FOR MIXED SOLID-SHELL FE


MODELS

The asymptotic approach is recalled briefly. It is particularized for a strain energy which has
a cubic polynomial dependence from the configuration variables as occurs for the nonlinear
Cauchy continuum based on a Green strain measure, when a hybrid solid-shell FE model is
employed. Isogeometric formulations can also be adopted Leonetti et al. (2018a).

2.1 The current implementation of the asymptotic algorithm


Consider a slender hyperelastic structure subject to conservative nominal loads ^
p proportion-
ally increasing with the amplifier factor λ. The equilibrium is expressed by the virtual work
equation

0 ½u
δu  λ^
p δu ¼ 0; u 2 U ; δu 2 T ð1Þ

where u 2 U is the field of configuration variables, ½u


denotes the strain energy, T is the tan-
gent space of U at u and a prime is used to express the Frechèt derivative with respect to u.
Due to the assumed 3rd order polynomial dependence of ½u
on u, it can be exactly
replaced with its 3rd order Taylor expansion from a given configuration u ¼ u0 , that is
 
1
0 ½u
δu :¼ 0 0 þ 00 0 ðu  u0 Þ þ 000 ðu  u0 Þ2 δu; 8δu 2 T ; ð2Þ
2

where a subscript denotes the point in which the quantities are evaluated, i.e. 00 ≡ 0 ½u0
and
so on, while 000 is constant with u.
The method starts with the evaluation of the fundamental path uf ½λ
assumed as analytical in
λ and approximated with its tangent in the equilibrium configuration ðu0 ; λ0 ¼ 0Þ as
uf ¼ u0 þ λ^u. It is evaluated through a first order Taylor expansion in λ of Eq. (1), that is

418
000 ^
uδu  ^
pδu ¼ 0; 8δu 2 T : ð3Þ

With the adopted linear extrapolation in λ of uf , the bifurcation condition becomes

00 ½uf ½λ

_vi δu ≡ ð000 þ λ000 ^


uÞ_vi δu ¼ 0 8δu 2 T ð4Þ

where v_ i and λi are the bifurcation modes and loads. Note that the expression in Eq. (4) is
exact Magisano et al. (2017b) and provides the m bifurcation loads and modes, orthogona-
lized according to

000 ^
uv_ i v_ k ¼ δik ð5Þ

with δik the Kronecker symbol.

2.1.1 The reduced model of the perfect structure


The space U is decomposed as a direct sum of the critical subspace V and its orthogonal com-
plement W, defined as
( )
X
m
U ¼ V W; V¼ v: v¼ ξ i v_ i ; W ¼ fw : 000 ^
uv_ i w ¼ 0g ð6Þ
i¼1

where ξ i , with i ¼ 1    m are the buckling mode amplitudes. The space of admissible configur-
ations, following a Galerkin approach, is limited to

ud ¼ uf ½λ
þ v½ξ i
þ w½λ; ξ i
ð7Þ

where the corrective term w 2 W is assumed to be at least quadratic in λ and ξ i and the com-
pact notation f ½ξ i
is used to denote the dependence of function f on all the ξ i .
Using a Ritz-Galerkin approach the equilibrium equation is imposed assuming v_ i and δw as
test functions, and the configuration defined by ud , that is

rw ½λ; ξ i
≡ f0 ½ud
 λ^
pgδw ¼ 0; rk ½λ; ξ i
≡ f0 ½ud
 λ^
pg_vk ¼ 0: ð8Þ

From the condition rw ½λ; ξ i


¼ 0 and using a Taylor expansion up to the 2th order in
λ; ξ 1 ;    ξ m we obtain the quadratic correctives (see Magisano et al. (2016))

1 ^ 1X ^^ δw ¼ 000 ^
00b w u2 δw

w ¼ λ2 w ξ ξ wij 00 000 _ _ 8δw 2 W ð9Þ
2 2 ij i j  b wij δw ¼ vi vj δw

where the subscript b denotes quantities evaluated in λb ^


u and λb is a suitable reference value of
the bifurcation load.
From the condition rk ½λ; ξ i
¼ 0 we obtain the reduced nonlinear system which defines the
equilibrium path

λ2 X
m X m
ξiξj X m
ξiξj ξh
rk ½λ; ξ i
≡ μk ½λ
þ ðλk  λÞξ k  ξ i Cik þ Aijk þ Bijhk ¼ 0 ð10Þ
2 i¼1 i;j¼1
2 i;j;h¼1
6

where

419
Aijk ¼ 000 v_ i v_ j v_ k Cik ¼ ^
00b w
^ wik
ð11Þ
Bijhk ¼ 00b ðwij whk þ wih wjk þ wik wjh Þ 1 2 000 ^2 _
μk ½λ
¼ 2 λ  u vk :

Eqs. (10) are an algebraic nonlinear system of m equations in the m þ 1 variables λ; ξ 1    ξ m


that, due to the small size of the system, can be efficiently solved using the arc–length scheme.

2.1.2 Standard a-posteriori account of geometrical imperfections


Small imperfections, expressed by an initial displacement ~ u, can easily be considered in the
asymptotic analysis. In the current proposal Magisano et al. (2016) the following coefficients
μk :¼ λ000 ^u~
~ uv_ k ð12Þ

are added to Eq. (10), that is


rk þ ~
μk ¼ 0 ð13Þ

and the reduced model is corrected adding ~


u to the expression (7)

ud ¼ ~
u þ uf ½λ
þ v½ξ i
þ w½λ; ξ i
: ð14Þ

Once the steps in Eqs. (3), (4), (9), (11) of the analysis have been performed, once and
for all, small imperfections in the geometry can be taken into account by adding a few
additional terms in the expression of rk . The computational extra-cost is negligible since
just the reduced nonlinear equations Eq. (10) have to be solved again for each new
imperfection. In this way the method allows a low cost imperfection sensitivity analysis.
However, comparisons with standard path-following analyses show that the accuracy of
this approach is limited to small imperfection amplitudes and structures with an almost
linear pre-critical behavior.

2.2 FEM implementation of the asymptotic approach


Denoting with a bold symbol the discrete FEM counterpart of the continuum quantities, and
referring to the solid-shell finite element model presented in Magisano et al. (2016), the con-
struction of the reduced model of the perfect structure consists of the following steps.
1. The fundamental path defined by Equation (3) becomes in FE format

u;
uf ½λ
¼ u0 þ λ^ u ¼ ^
K0 ^ p; K0 ≡ K ½u0
ð15aÞ

and requires the solution of a linear system to evaluate the initial path tangent ^
u.
2. The buckling modes and loads are obtained by the following eigenvalue problem

u
Þ v_ ¼ 0
K½λ
_v ≡ ðK0 þ λK1 ½^ ð15bÞ

where K0 and K1 are obtained from the following energy equivalence


δ uT K0 δ u : ¼ 000 δu2 , δuT K1 δ u ¼ 000
0^ uδu2 .
^
^ 2 W are obtained by the solution
3. The m  ðm þ 1Þ=2 quadratic correctives FE vectors wij , w
of the linear systems (i ¼ 1 . . . m; j ¼ i . . . m)

Kb wij þ pij ¼ 0 ; ^
^ þ p00 ¼ 0 ;
Kb w 8w 2 W ð15cÞ

in which Kb ≡ K ½λb
, pij ; p00 are defined as a function of modes v_ i and ^
u as

420
δ wT pij ¼ 000
bv _ j v_ j δw δ wT p00 ¼ 000
b^ u2 δw:

4. Evaluation of the coefficients in Eq. (11) of reduced equilibrium system in Eq. (10) as a sum
of finite element contributions.
The evaluation of the equilibrium path, to be repeated for each imperfection, requires
the following steps: 1) evaluation of ~μk ¼ λ000 ^
u~
uv_ k ; 2) solution of the reduced system in
Eq. (13) and drawing of the equilibrium path according to Eq. (14).

3 AN ACCURATE A-POSTERIORI ACCOUNT OF GEOMETRICAL


IMPERFECTIONS

The Koiter algorithm is now reformulated in order to coherently consider the presence of geo-
metrical imperfections, removing the hypothesis of linear pre-critical behavior.

3.1 The strain energy and the equilibrium path of the structure with geometrical imperfection
Using a Hellinger-Reissner variational principle the mixed strain energy i ½u
is expressed, as
usual in a FE context, as a sum of element contributions

XZ  1 T 1

I ½u
≡ t ðρ½d
 ρ½d
Þ  t Cρ t dOe :
T
ð16Þ
e Oe 2

ρ½d
and t are the vectors collecting the generalized strains and stresses components
for the given structural model, Oe is the finite element domain and d is the displacement field
and C 1
ρ the compliance matrix of the structural model. The
strain energy of the structure
without imperfections ½u
is obtained by zeroing the terms ρ ~ d in Equation (16).
Denoting with a symbol δ the variation of d and t, the first variation of I ½u
becomes

PR n
~ o
I ½u
0 δu ¼ Oe δt T ð ρ ½ d
 ρ d  C1 T 0
ρ tÞ  t ρ ½d
δd dOe
e ð17Þ
¼ ð½u
0  0 ½~
u
Þδu

that is the difference between the perfect and imperfect first order strain energy variation,
being

XZ

0 ½~
u
δu :¼ δtT ρ ~
d dOe ð18Þ
e Oe

u (which has ~
the first variation of the perfect structure evaluated in ~ t ¼ 0).
The equilibrium path is obtained from the following condition

ð½u
0  0 ½~
u
 λpÞδu ¼ 0 8δu ) s½u
 ~
p  λp^ ¼ 0: ð19Þ

p and the imperfection vector ~


where the internal force vector s½u
, the load vector ^ p are defined
by the energy equivalences

T
sT δ u≡0 ½u
δu ; ^ p δu ; ~
p δu≡0 ½~
T
p δu≡^ u
δu ; 8 δu: ð20Þ

Eq. (19) can be solved using standard path-following techniques Magisano et al. (2017a,b) for
an assigned imperfection ~
u.

421
3.2 The new reduced model with geometrical imperfection
The space of admissible configurations is obtained by adding an additional term which repre-
sents the initial imperfection, assumed to be a linear combination of the displacement part of
the buckling modes  ui , to the configuration field of the perfect structure that is

X
n
ud ½λ; ξ i ; ~ξi
¼ ~ u þ v½ξ i
þ w½ξ i ; ~ξi ; λ

u þ λ^ with ~
u¼ ~ξi 
ui : ð21Þ
i¼1

Note that, unlike the reduced model in Equation (14), now the quadratic correctives w½ξ i ; ~ξi ; λ

depend on the geometrical imperfection amplitudes ~ξi .


The residual Equation (19) after 3rd order Taylor series, and then without any truncation
error, becomes

1 1
ð00k ðv þ wÞ þ ðλ  λk Þ000 ^
uðv þ wÞ þ λ2 000 ^
u2 þ 000 ðv þ wÞ2 þ 000 ~
uðλ^
u þ v þ wÞÞδu ¼ 0:
2 2
ð22Þ

Note that the equilibrium condition for the structure with no imperfection is regained
for ~u ¼ 0.

3.2.1 Projection of the equilibrium equation in the space W


The corrective field w 2 W is obtained by projecting Equation (22) in direction δw, i.e. assum-
ing δu ¼ δw, and expanding it in Taylor series up to the second order in the asymptotic param-
eters ðλ; ξ i ; ~ξi Þ. The complete procedure is reported in Garcea et al. (2017) to which we refer
for a deeper discussion. The residual equation simplifies as
 
~ 00 1 2 000 2 1 000 2 000
~rw ½λ; ξ i ; ξi
≡ b w þ λ  ^ u þ  v þ ~ uðλ^
u þ vÞ δw ¼ 0: ð23Þ
2 2

Remembering the expression of v½ξ i


, the quadratic correctives of the imperfect structure are
sums of the correctives for zero imperfections of Eq. (9) and of the additional contribution
due to the geometrical imperfection

1 ^ 1X
w½ξ i ; ~ξ i ; λ
¼ λ2 w
^þ ξξw ~
€ ij þ w ð24Þ
2 2 i;j i j

where
X X X
w ~
~ ¼ λw
^þ ~_ i
ξiw with ~
^ :¼
w ^
~ξj w
~j; ~_ i :¼
w ~_ ij :
~ξj w ð25Þ
i j j

The terms in Eq. (25) can be evaluated, once and for all in the perfect structure step of the
Koiter analysis, being known the imperfection basis, as

^
00b w
~ i δw ¼ 000 ^
u
ui δw ~_ ij δw ¼ 000 v_ i 
00b w uj δw 8δw 2 W: ð26Þ

3.2.2 The new reduced equations with geometrical imperfection


Substituting the expression of w and v previously obtained, using the mode normalization con-
dition in Equation (5) and maintaining terms in λ; ξ i ; ~ξi until the 3rd polynomial order, the kth
equilibrium equation becomes

422
~rk ½λ; ξ i
≡ rk ½λ; ξ i
þ ~
μk ½λ; ξ i
¼ 0; k ¼ 1m ð27Þ

with rk ½λ; ξ i
¼ 0 the kth reduced equilibrium equation in Equation (10) and the new imperfec-
tion factor ~μk defined as

P 1X
^
 
~μk ≡ ξ i λ000 vi w
~ v_ k þ ~_ j v_ k þ 000 vj w
ξ i ξ j 000 vi w ~_ i v_ k þ 000 ~
uw € ij v_ k
i 2 ij
X ð28Þ
^ 1 ^
þλ000 ~
uð^ ~ Þ_vk þ
uþw ξ i 000 ~
uð_vi þ w ~_ i Þ_vk þ λ2 000 ~ ^ v_ k :
uw
i
2

It is possible to observe that the only change, with respect to the standard reduced system in
subsection 2.1.2, regards the imperfection coefficient ~ μk which is now more complex than the
one used in Equation (11), which only maintains the linear contribution in λ, that is
~μk ¼ λ000 ^u~uv_ k , while the quadratic terms in λ and the terms in ξ are neglected, leading to
inaccuracy as the pre-critical non-linearity increases.
Furthermore, note that the proposed reduced model assumes the following final expression

X 1 X 1 ^
ud ½λ; ξ i ; ~ξi
¼ ~
u þ λð^ ~
^Þ þ
uþw ~_ i Þ þ
ξ i ð_vi þ w € ij þ λ2 w
ξiξj w ^: ð29Þ
i
2 ij
2

The new correctives can be seen as a correction to the fundamental path tangent and the buck-
ling modes of the perfect structures in order to take into account the geometrical
imperfection.

3.3 FEM implementation of the proposed algorithm


The construction of the reduced model of the perfect structure presented in subsection 2.2 is
completed by adding the evaluation of the new corrective after Eq. (15c)

~_ ij þ ~
Kb w pij ¼ 0 ^
~þ~
Kb w p0i ¼ 0 ; 8w 2 W ð30Þ

where δw T ~pij ¼ 000


b  p0i ¼ 000
vj u_ j δw, δ wT ~ b^ u
ui δw. The imperfection coefficients ~
μk are evaluated
using the expression (28) instead of (12). Once the reduced non-linear system (27) is solved,
the equilibrium path is traced according to (29).

4 NUMERICAL RESULTS

In this section some benchmarks are considered in order to test the accuracy of the proposed
a-posteriori account of geometrical imperfection. A comparison with the different approaches
is made. In particular, the numerical results report:
– the solution of the full FE model non-linear equations, obtained using a standard path-
following technique, denoted as Riks and considered the reference solution;
– the solution obtained through the Koiter method including the imperfection a-priori in the
model, which means that the reduced model is re-constructed for each imperfection while
~μk ¼ 0, denoted as K0 ;
– the solution obtained through the Koiter method using the reduced model of the perfect
structure, built once and for all, and taking into account the imperfection a-posteriori in
the standard way, denoted as Klin ;
– the solution obtained through the Koiter method using the reduced model of the perfect
structure, built once and for all, and taking into account the imperfection a-posteriori
according to the new proposal, denoted as Kquad .

423
The geometrical imperfection is given as a linear combination of the displacement shapes of
the buckling modes and its maximum displacement components, denoted as ~ umax .

4.1 Simply supported plate


The first example regards a simply supported and uniformly compressed plate whose geom-
etry, load and boundary conditions are reported in Figure 1. The imperfection shape is pro-
portional to the first buckling mode reported in the same figure.
Figure 1 shows the equilibrium paths obtained with the different methods. In this case the
proposed Kquad approach provides results very similar to reference Riks ones, even for a large
imperfection magnitude, while the standard Klin approach gives a result which is completely
wrong. In this case the energy terms associated with w~_ and w
^
~ are large also for small values of
the imperfection amplitude due to the membrane hyperstaticity of the plate.

4.2 Compressed simply supported channel section


A simply supported channel section, whose geometry and material properties are reported in
Figure 2 is now analysed: the imperfection is the displacement shape of the first buckling
mode (flexural). The structure exhibits buckling mode interaction phenomena.
In Figure 3 it is reported how the limit loads change with the amplitude of the first
imperfection, while Figure 4 shows the equilibrium paths for ~ umax ¼ 5t. The good behav-
iour in the evaluation of the limit point of Kquad is evident while Klin presents significant
errors in the equilibrium path evaluation notwithstanding the fairly accurate value of the
limit load.

Figure 1. Simply supported plate: geometry and equilibrium paths for ~


umax ¼ t.

Figure 2. Compressed simply supported channel section with material properties.

424
Figure 3. C-section: limit load versus imperfection magnitude for flexural imperfection.

Figure 4. C-section: equilibrium paths for first buckling shape imperfection and umax ¼ 5t.

5 CONCLUSIONS

A new strategy to include, a-posteriori, geometrical imperfections in Koiter analysis is pro-


posed in this paper. The main idea is to correct the linear modes of the perfect structure
reduced model using additional corrective modes, which take into account the imperfections.
The reduced system of the imperfect structure is obtained starting from the system of the per-
fect structure by adding some terms, which coherently consider the geometrical imperfection
up to the second order. In this way, the Koiter method with a-posteriori account of the imper-
fections becomes accurate even for pre-critical non-linearities and significant imperfection
amplitudes, making the approach suitable for a wide range of practical problems.

REFERENCES

F.G. Flores and L.A. Godoy. Elastic post-buckling analysis via finite element and perturbation tech-
niques. Part 1. International Journal for Numerical Methods in Engineering, 33(9):1775–1794, 1992.

425
G. Garcea, R. Gonçalves, R. Bilotta, D. Manta, R. Bebiano, R. Leonetti, M. Magisano, and
D. Camotim. Deformation modes of thin-walled members: A comparison between the method of gen-
eralized eigenvectors and generalized beam theory. Thin-Walled Structures, 100:192–212, 2016. ISSN
0263-8231. doi: 10.1016/j.tws.2015.11.013.
G. Garcea, F. S. Liguori, L. Leonetti, D. Magisano, and A. Madeo. Accurate and efficient a posteriori
account of geometrical imperfections in Koiter finite element analysis. International Journal for Numer-
ical Methods in Engineering, 112(9):1154–1174, 2017. doi: 10.1002/nme.5550.
Alessandra Genoese, Andrea Genoese, Nicola Luigi Rizzi, and Ginevra Salerno. Force constants of BN,
SiC, AlN and GaN sheets through discrete homogenization. Meccanica, 53(3):593–611, Feb 2018.
ISSN 1572-9648. doi: 10.1007/s11012-017-0686-1.
S.R. Henrichsen, P.M. Weaver, E. Lindgaard, and E. Lund. Post-buckling optimization of composite
structures using Koiter’s method. International Journal for Numerical Methods in Engineering, 2016.
doi: 10.1002/nme.5239.
Leonardo Leonetti, Francesco Liguori, Domenico Magisano, and Giovanni Garcea. An efficient isogeo-
metric solid-shell formulation for geometrically nonlinear analysis of elastic shells. Computer Methods
in Applied Mechanics and Engineering, 331:159–183, 2018a. ISSN 0045-7825. doi: 10.1016/j.
cma.2017.11.025.
Leonardo Leonetti, Domenico Magisano, Francesco Liguori, and Giovanni Garcea. An isogeometric
formulation of the Koiter’s theory for buckling and initial post-buckling analysis of composite shells.
Computer Methods in Applied Mechanics and Engineering, 337:387–410, 2018b. ISSN 0045-7825. doi:
10.1016/j.cma.2018.03.037.
F.S. Liguori, A. Madeo, D. Magisano, L. Leonetti, and G. Garcea. Post-buckling optimisation strategy
of imperfection sensitive composite shells using Koiter method and monte carlo simulation. Composite
Structures, 192:654–670, 2018. ISSN 0263-8223. doi: 10.1016/j.compstruct.2018.03.023.
F.S. Liguori, G. Zucco, A. Madeo, D. Magisano, L. Leonetti, G. Garcea, and P.M. Weaver. Postbuck-
ling optimisation of a variable angle tow composite wingbox using a multi-modal Koiter approach.
Thin-Walled Structures, 138:183–198, 2019. ISSN 0263-8231. doi: 10.1016/j.tws.2019.01.035.
D. Magisano, L. Leonetti, and G. Garcea. Koiter asymptotic analysis of multilayered composite struc-
tures using mixed solid-shell finite elements. Composite Structures, 154:296–308, 2016. doi: 10.1016/j.
compstruct.2016.07.046.
D. Magisano, L. Leonetti, and G. Garcea. How to improve efficiency and robustness of the Newton
method in geometrically non-linear structural problem discretized via displacementbased finite
elements. Computer Methods in Applied Mechanics and Engineering, 313: 986–1005, 2017a. ISSN
0045-7825. doi: 10.1016/j.cma.2016.10.023.
D. Magisano, L. Leonetti, and G. Garcea. Advantages of the mixed format in geometrically nonlinear
analysis of beams and shells using solid finite elements. International Journal for Numerical Methods in
Engineering, 109(9):1237–1262, 2017b. doi: 10.1002/nme.5322.
D. Magisano, K. Liang, G. Garcea, L. Leonetti, and M. Ruess. An efficient mixed variational
reduced-order model formulation for nonlinear analyses of elastic shells. International Journal for
Numerical Methods in Engineering, 113(4):634–655, 2018. doi: 10.1002/nme.5629.
T. Rahman and E.L. Jansen. Finite element based coupled mode initial post-buckling analysis of
a composite cylindrical shell. Thin-Walled Structures, 48(1):25–32, 2010. doi: 10.1016/j.tws.2009.08.003.
N.L. Rizzi, V. Varano, and S. Gabriele. Initial postbuckling behavior of thin-walled frames under mode
interaction. Thin-Walled Structures, 68:124–134, 2013. ISSN 0263-8231. doi: 10.1016/j.tws.2013.03.004.

426
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Ductility of different types of shear connectors—experimental


and numerical analysis

N. Gluhović, M. Spremić, B. Milosavljević, Z. Marković & J. Dobrić


Department of materials and structures, Faculty of Civil Engineering, University of Belgrade, Republic of
Serbia

ABSTRACT: Different types of shear connectors, which are used for achievement of steel-
concrete composite action, have been developed during decades. Their suitability for various
range of structures and loadings is determined by shear resistance, ductility of shear connectors,
obtained failure mechanisms and fatigue resistance. Headed studs are widely used for composite
construction of buildings and bridges. Perforated shear connectors, mechanically fastened shear
connectors or bolted shear connectors with mechanical couplers are alternative solution for
achievement of composite action between steel and concrete. Results of experimental and numer-
ical analysis of these types of shear connectors are used for comparative analysis and presented in
this paper. Presented analysis included comparison of stiffness and different failure mechanisms
for these types of shear connections, with focus on the ductility at different load levels. Besides,
results are compared with recommendation given in EN 1994-1-1 for ductile shear connectors.

1 INTRODUCTION

Headed studs are the most commonly used type of shear connectors. They are extensively used
due to highly automatized process of installation and comprehensible design recommendations
given in various standards (EN 1994-1-1, 2004), (ANSI/AISC 360-05, 2005), (JSCE, 2009). The
main disadvantages of headed studs are their low fatigue resistance and specific conditions
regarding required electricity, weather and temperature conditions, when installed in-situ. Devel-
opment of different types of shear connectors represent alternative solution to the traditionally
used headed studs, considering their shear resistance, stiffness and ductility. Besides, usage of
different types of shear connectors reduces the construction time and overall construction cost.
Headed studs, perforated shear connectors, mechanically fastened shear connectors and bolted
shear connectors with mechanical couplers are compared and presented in this paper. Experi-
mentally and numerically gained data are used for comparative analysis. Three main characteris-
tics of these connectors: shear resistance, stiffness and ductility are compared and presented
herein. Analysis of shear connectors’ deformation at different load levels and comparison with
current recommendations given in (EN 1994-1-1, 2004) is in the focus of this paper.

2 EXPERIMENTAL RESEARCH

Experimental investigation of headed studs, perforated shear connectors, mechanically fas-


tened shear connectors and bolted connectors with mechanical couplers was performed in
Laboratory of Materials, at the University of Belgrade, Faculty of Civil Engineering. All
shear connectors were analysed through standard push-out tests, according to (EN 1994-1-1,
2004). Headed studs, perforated shear connectors and mechanically fastened connectors were
analysed through group arrangement of connectors positioned in envisaged openings of pre-
fabricated concrete slabs, denoted with ST, PF and HSF test series, respectively. Bolted shear
connectors with mechanical couplers were examined in full-depth concrete slabs casted in

427
Figure 1. Push-out specimens’ layout.

place and designated as BC test series. Push-out specimens’ layout of all examined shear con-
nectors is given in Figure 1.
Four headed studs with 16.0 mm diameter and 100 mm height were positioned on each steel
beam flange of ST test series, eight headed studs in total. Transversal distance between headed
studs, relative to the shear force direction, was 80.0 mm, while longitudinal distance was
100.0 mm (Spremic et al., 2013), which is within the range of minimal recommended distances
according to (EN 1994-1-1, 2004). One perforated shear connector with 6.0 mm thickness and
two holes with 35.0 mm diameter was positioned on each beam flange of PF test specimen,
two perforated connectors in total. Perforated shear connector, analysed through own experi-
mental investigation, was adopted with 76.0 mm high and 180.0 mm length. HSF test series
was conducted with four X-HVB 110 shear connectors, as main representative of mechanically
fastened shear connectors, which were installed on each beam flange, eight connectors in
total. Each X-HVB 110 shear connector was fastened to the steel base material with two cart-
ridge fired pins and positioned at minimal recommended distances (ETA-15/0876 Assessment,
2016), 50.0 mm and 100.0 mm of transversal and longitudinal spacing, respectively (Gluhović
et al., 2017). BC test series was conducted with two 16.0 mm diameter bolted shear connectors
with mechanical couplers, which were positioned on each beam flange at 100.0 mm of trans-
verse distance (Milosavljević, 2014), four bolted shear connectors in total. Number of exam-
ined test specimens within one test series, geometry and main material properties of
experimentally analysed push-out specimens are presented in Table 1.
All concrete slabs were reinforced with standard reinforcement layout with 10 mm diameter
ribbed bars and R500 grade, as shown in Figure 1. Two horizontal ribbed bars, in upper and
bottom reinforcement layer, were positioned between shear connectors of ST and HSF test
series. These horizontal ribbed bars were cut at the contact layer between prefabricated con-
crete slab and infill concrete of PF test specimen due to installation restrictions of analysed
perforated shear connector, as shown in Figure 1. Connecting surface of steel flanges was

Table 1. Geometry and material properties of push-out specimens.


Specimens’ geometry (mm) Material properties(MPa)

slab opening spacing shear connector concrete connector

Series nsp1 w/h/t2 weo /heo3 ls /ts4 nsc /dsc (tsc) /hsc (lsc)5 fc,cube fub

ST 4 600/650/120 240/240 100/80 8/16 (-)/100 (-) 42.0 520


PF 1 600/650/120 140/240 - 2/(-) 6/76 (180) 46.0 355
HSF 4 600/650/140 240/240 100/50 8/(-) 2/110 (-) 38.8 432
BC 3 400/550/300 - -/100 4/16 (-)/100 (-) 32.6 907

1 nsp - number of specimens within test series, 2 w/h/t - concrete slab width/height/thickness; 3 weo/heo - envis-
aged opening width/height; 4 ls/ts - longitudinal/transversal distance between connectors; 5 nsc - total
number of shear connectors per specimen, dsc - connector diameter, tsc - thickness, hsc - height, lsc – length

428
greased in order to avoid effects of bond to the concrete slab. Prior to concreting, inner sur-
faces of prefabricated slabs’ openings were cleaned and treated with the layer of concrete glue
as the continuing layer between new and old concrete. After the installation of shear connect-
ors, envisaged openings were filled in horizontal position with three-fraction concrete with
shrinkage reduction admixtures. Concrete slabs of BC test series with bolted shear connectors
were casted in place. For all analysed test series, after three days, half assembled specimens
were turned and second assembling phase was performed in the same way as the first one. Hot
rolled steel profile HEB 260 was used for all test specimens. The loading regime is adopted
according to (EN 1994-1-1, 2004). Force controlled cycling loading was applied in 25 cycles
ranging from Fmin to Fmax, corresponding to approximately 5% and 40% of expected shear
resistance. After the cyclic loading, failure loading was applied in one step, with constant dis-
placement rate, such that failure does not appear in less than 15 minutes. Longitudinal slip
and uplift between steel profile and concrete slab and separation between concrete slabs were
measured with inductive displacement transducers.

3 EXPERIMENTAL AND NUMERICAL ANALYSIS

Abaqus software package (ABAQUS, 2009) was used for numerical analysis of all analysed
push-out tests series. Quarter of the real specimens were modelled using double vertical sym-
metry conditions. Numerical FE models are presented in Figure 2. Displacement controlled
failure loading was applied in reference point at the top of the steel profile web. None uniform
mass scaling was used for explicit dynamic analysis.
The measured material properties, which are presented in Table 1, were used in FE analysis.
For FE models of shear connectors, structural steel and reinforcement the isotropic plasticity
model was used and material properties were calibrated with the experimental results of material
coupons. Material ductile damage and shear damage models were used for shear connectors,
according to (Pavlović et al., 2013). Besides, ductile damage material model was used for steel
profile corresponding to the material properties of steel grade S275. Concrete damage plasticity
was implemented in FE models of all push-out series. Stress-strain curves were defined according
to recommendations given in (EN 1992-1-1, 2004) for part of stress-strain curve up to strain εcu1
= 3.5 ‰. For strain εc > 3.5 ‰ the value of stress was defined according to (Pavlović et al., 2013).
Comparison of numerically obtained force-slip curves with experimental results of push-out
tests is presented in Figure 3. Force-slip curves of ST, HSF and BC specimens represent aver-
aged results of four and three push-out specimens, respectively, as given in Table 1. As shown
in Figure 3a and Table 2, satisfying agreement between experimental and numerical results is
obtained for all analysed test specimens. Also, good prediction of initial stiffness and descend-
ing branch of the force-slip curves is achieved.
Based to the experimentally and numerically gained data, the highest ultimate shear
resistance is achieved for group arrangement of four headed studs with 16 mm diameter

Figure 2. Numerical FE models.

429
Figure 3. Comparison of experimental and FE analysis.

Table 2. Experimental vs. FE analysis.


Shear resistance (kN) Characteristic slip capacity (mm)

Experimental FEA Ratio Experimental FEA Ratio

Series Pult,exp Pult,FEA Pult,FEA /Pult,exp δuk,exp δuk,FEA δuk,FEA /δukexp

ST 726.00 732.10 1.01 6.50 6.65 1.02


PF 441.87 435.3 0.99 6.75 8.67 1.28
HSF 332.67 347.10 1.04 9.66 8.30 0.86
BC 356.62 381.70 1.07 5.23 4.37 0.84

(ST test series) and is approximately 60% higher in comparison to the analysed PF test
specimen. Besides, PF test specimen achieved approximately 20% higher ultimate shear
resistance in comparison to the tests series with mechanically fastened (HSF test series)
and bolted shear connectors (BC test series), while these two test specimens achieved simi-
lar ultimate shear resistances. Characteristic value of slip capacity, δuk, higher than
6.0 mm is achieved for ST, PF and HSF test series, as given in Table 2, which is limiting
value according to (EN 1992-1-1, 2004) to consider shear connector as ductile.
Besides, comparison of stiffness and slip capacity of all analysed test series at the loading
level, which represent 70% of achieved ultimate shear resistance, is presented in Figure 3b.
This loading level refers to the structure behaviour at service loads or serviceability limit state,
SLS. Headed studs and perforated shear connector achieved the smallest slip capacity corres-
ponding to SLS, 0.49 and 0.43 mm, respectively, conforming to the largest stiffness. The lar-
gest slip capacity at same loading level is achieved for bolted shear connectors, approximately
1.00 mm, corresponding to the lower achieved stiffness, as shown in Figure 3b.
The same failure mechanisms are obtained during experimental and numerical analysis of
all analysed shear connectors. Failure of concrete obtained through numerical FE models of
all push-out test specimens is presented in Figure 4, corresponding to 90% of achieved ultim-
ate shear resistance. Dominant failure mechanism of headed studs and bolted shear connect-
ors with mechanical couplers is shearing failure of connectors. Mechanically fastened shear
connectors obtained a different failure mechanism, pull-out and shear failure of cartridge fired
pins with shear connector deformation. The dominant failure mechanism of perforated shear
connector is failure of concrete, considering global failure of the concrete slab and shearing
failure of the concrete dowels. Besides, the highest damage of concrete is obtained for push-
out specimen with perforated shear connector. Damage of the concrete slab through appear-
ance of longitudinal crack starts to propagate at early stages of specimen’s failure loading.
Moreover, local failure of concrete slabs in surrounding zone of shear connectors is obtained
for ST, HSF and BC test series.

430
Figure 4. Failure of concrete corresponding to 90 % of ultimate shear force – FEA.

4 DUCTILITY OF SHEAR CONNECTORS

“Ductile connectors are those with sufficient deformation capacity to justify the assumption
of ideal plastic behaviour of the shear connection in the structure. A connector may be taken
as ductile if the characteristic slip capacity δuk is at least 6mm” (EN 1994-1-1, 2004). Analysis
of shear connectors’ ductility is performed through numerical FE analysis. Comparison of
shear connectors’ deformation is performed for three different levels of failure loading: load
corresponding to 70% of ultimate shear resistance, i.e. SLS, load corresponding to 6.0 mm of
total slip, and load corresponding to 96% of ultimate shear resistance, prior to failure, as pre-
sented in Table 3.
Difference between shear connector displacement at the connector root (displacement at con-
tact with steel profile) and top of the shear connector embedded in concrete, in the direction of
the shear force, is considered as connector deformation capacity. Deformation of headed studs,
perforated shear connectors and X-HVB shear connectors at 6.0 mm of total slip is shown in
Figure 5. Shear failure of bolted shear connectors with mechanical couplers is achieved for load-
ing level corresponding to the total slip lower than 6.0 mm, as shown in Figure 3a.
Characteristic value of slip capacity of headed studs is higher than 6.0 mm, as presented in
Figure 3a. Besides, for loading level corresponding to 6.0 mm of total slip, difference between
headed studs’ deformation at two observed points (U2root -U2top) is approximately 5.78 mm,
as presented in Table 2. Considerable smaller deformation of shear connector head (U2top) is
related to low degree of concrete failure. According to (EN 1994-1-1, 2004), headed studs with
diameter 16<d<25 mm could be classified as ductile connector and results of experimental and
numerical analysis presented herein confirm this statement.
Experimentally and numerically achieved characteristic value of slip capacity of analysed
perforated shear connector is also higher than recommended limiting value given in (EN
1994-1-1, 2004), as shown in Figure 3a and Table 2. However, obtained deformation of per-
forated shear connector at 6.0 mm of total slip is limited to approximately 2.5 mm, as given
in Table 3. Significant failure of concrete obtained for PF push-out specimen resulted in
lower deformation capacity of perforated shear connector. Moreover, exception of

Table 3. FE analysis of shear connector deformation.


slip at SLS shear load 6.0 mm slip slip at 96%ULS shear load
U2roo U2to U2root-U2top U2root U2top U2root-U2top U2root U2top U2root-U2top
Specimen [mm] [mm] [mm] [mm] [mm] [mm] [mm] [mm] [mm]

ST 0.49 0.00 0.49 6.03 0.25 5.78 3.86 0.13 3.73


PF 0.43 0.19 0.24 6.02 3.55 2.47 1.79 0.66 1.13
HSF 0.70 0.00 0.70 6.06 0.50 5.56 3.46 0.50 2.96
BC 1.00 0.00 1.00 - - - 3.60 0.01 3.59

431
transversal reinforcement bar above shear connector and in the concrete dowels of shear
connector holes influenced higher damage of concrete and lower deformation capacity of
this type of shear connector. Similar behaviour of perforated shear connector is obtained for
three analysed levels of failure loading, as given in Table 3. Moreover, influence of reinforce-
ment bars’ layout in concrete slabs and concrete dowels on overall behaviour of perforated
shear connectors is analysed by (Spremić et al., 2017).
Mechanically fastened shear connectors obtained similar deformation capacity in compari-
son to the headed studs, at total slip corresponding to 6.0 mm. As shown in Figure 5 and
Table 3, displacement of shear connector is uniform over it’s height, which is related to low
degree of concrete failure. Deformation capacity of X-HVB shear connectors for various load-
ing levels is directly related by ductility of cartridge fired pins, as shown in Figures 5 and 6.
According to the (ETA-15/0876 Assessment, 2016) X-HVB shear connectors should be con-
sidered as ductile. Results obtained from experimental and numerical analysis of X-HVB 110
shear connectors presented herein, confirm this statement.
Deformation capacity of four analysed shear connectors at loading level corresponding to 96 %
of ultimate shear force is presented in Figure 6. Perforated shear connector achieved the lowest
deformation at this loading level, in comparison to the other types of shear connectors. Besides,
numerically obtained failure mechanisms clearly indicate that shear failure of headed studs, mech-
anically fastened and bolted shear connectors are in direct relation to the low degree of concrete
failure and higher deformation capacity of shear connectors. Moreover, headed studs and mech-
anically fastened shear connectors can be considered as ductile according to recommendations
given in (EN 1994-1-1, 2004), which should be considered as main advantage of these types of
shear connectors in comparison to the analysed perforated and bolted shear connectors. Perfor-
ated shear connector which is analysed in this paper can not be classified as ductile, considering
that it’s deformation capacity is significantly lower than minimal recommended 6.0 mm.

Figure 5. Deformation of shear connectors at the 6.0 mm of total slip.

Figure 6. Deformation of shear connectors at the 96% of ULS.

432
5 CONCLUSIONS

Standard push-out tests of four different types of shear connectors were performed. Results of
experimental and numerical analysis were compared and presented in this paper. Following
conclusions can be drawn:
Headed studs, perforated shear connector and mechanically fastened shear connector posi-
tioned at envisaged openings of prefabricated concrete slabs obtained characteristic value of
slip capacity higher than 6.0 mm, based on the results of push-out tests and FE analysis.
Bolted shear connectors with mechanical couplers achieved lower characteristic value of slip
capacity than 6.0 mm, which is limiting value defined in (EN 1994-1-1, 2004).
Deformation of headed studs and X-HVB shear connectors, obtained as difference between
connector’s displacements at two characteristic points, is approximately in a range between
5.5 mm and 6.0 mm, for loading level corresponding to total slip of 6.0 mm. Perforated shear
connector resulted in approximately 50 % lower deformation capacity.
Behaviour of perforated shear connector is mainly reflected through global failure of con-
crete slabs, failure of concrete dowels and development of initial cracks even for low loading
levels. Exception of transversal reinforcement bars above the perforated shear connector sig-
nificantly influenced obtained failure mechanism and deformation capacity.
Considering obtained slip to failure curve for analysed push-out specimen with perforated
shear connector and it’s deformation, approximately 2.5 mm slip is the result of the connector
deformation and another part up to total slip of 6.0 mm is the result of the concrete failure.
Therefore, it can be concluded that this type of shear connector cannot be considered as duc-
tile. Moreover, deformation capacity of perforated shear connector is directly related to the
geometry of connector and specimen’s layout.
This investigation is the part of TR-36048 project supported by the Serbian Ministry of
Education, Science and Technological Development.

REFERENCES

ABAQUS User Manual 2009. Providence, RI, USA: DS SIMULIA Corp.


ANSI/AISC 360-05 An American National Standard - Specification for structural steel buildings 2005.
Chicago, Illinois: American Institute of Steel Construction.
EN 1992-1-1: Eurocode 2: Design of Concrete Structures - Part 1-1: General rules and rules for buildings
2004. Brussels: CEN European Committee for Standardization.
EN 1994-1-1: Eurocode 4: Design of composite steel and concrete structures - Part 1-1: General rules and
rules for buildings 2004. Brussel: CEN European Committee for Standardization.
European Technical Assessment ETA-15/0876 2016. Berlin: Deutsches Institut fur Bautechnik.
Gluhović, N., Marković Z., Spremić M., Pavlović M. 2017. Experimental investigation and specific
behaviour of X-HVB shear connectors in prefabricated composite decks, In Dirk Jesse (ed.) The 8th
European Conference on Steel and Composite Structures; Proceedings of Eurosteel 2017, 2080–2089,
Copenhagen, 13–15 September 2017. Ernst & Sohn.
Milosavljević, B. 2014. Theoretical and experimental research of the behaviour of renforced concrete and
steel element connection by rebar couplers (in Serbian). University of Belgrade, Faculty of Civil Engin-
eering, PhD thesis.
Pavlović, M., Marković Z., Veljković, M., Buđevac, D. 2013. Bolted shear connectors vs. headed studs
behaviour in push-out tests, Journal of Constructional Steel Research 88: 134–149.
Spremic, M., Markovic, Z., Veljkovic, M., Budjevac, D. 2013. Push-out experiments of headed shear
studs in group arrangements, Advanced Steel Construction 9 (2): 139–160.
Spremić, M., Gluhović, N., Marković, Z., Dobrić. J., Filipović. A. 2017. Comparison of headed studs
with perfobond shear connectors - experimental and numerical analysis, In Dirk Jesse (ed.) The 8th
European Conference on Steel and Composite Structures; Proceedings of Eurosteel 2017, 2237–2246,
Copenhagen, 13–15 September 2017. Ernst & Sohn.
Standard specifications for steel and composite structures 2009. Tokyio: JSCE Japan Society of Civil
Engineers.

433
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

GBT-based semi-analytical solutions for the elastic/plastic stability


analysis of stainless steel thin-walled columns exposed to fire

R. Gonçalves & R. Marçalo Neves


CERIS and Departamento de Engenharia Civil, Universidade Nova de Lisboa, Portugal

D. Camotim
CERIS, DECivil, Instituto Superior Técnico, Universidade de Lisboa, Portugal

ABSTRACT: This paper presents and illustrates the application of an efficient Generalized
Beam Theory (GBT) semi-analytical solution procedure to determine elastic and plastic bifur-
cation loads (linear stability analysis) of stainless steel thin-walled members subjected to uni-
form compression and exposed to fire. Besides global (flexural or flexural-torsional) buckling,
the proposed GBT formulation allows for cross-section deformation and therefore makes it
possible to capture local and distortional buckling. The temperature effect is taken into
account using the material law for stainless steel specified in Eurocode 3 part 1-4 (CEN 2006)
and Annex C of part 1-2 (CEN 2009). For plastic buckling, the material tangent moduli are
obtained using both J2 small-strain incremental and deformation plasticity theories. For illus-
trative purposes, the procedure is applied to columns with a lipped channel cross-section.

1 INTRODUCTION

Although initially more expensive than conventional carbon steel, stainless steel can be com-
petitive because of its increased fire resistance, lower maintenance needs, higher corrosion
resistance, better aesthetic appearance and lower life-cycle cost (SCI 2017). The fire behaviour
of stainless steel members has been the subject of research of recent studies (Gardner &
Baddoo 2006, Ng & Gardner 2007, Uppfeldt et al. 2008, Lopes et al. 2012). However, Accord-
ing to Eurocode 3, these members should be checked using the buckling interaction formulas
for carbon steel members, which have been shown to be imprecise and even unsafe in some
cases (Lopes et al. 2010). Moreover, studies concerning the behaviour of members with slender
sections (Class 3 or 4 according to Eurocode 3), susceptible to local and/or distortional buck-
ling, are still lacking.
To increase the knowledge on the behaviour of thin-walled stainless steel structures in case
of fire, project “Fire design of stainless steel structural elements—StaSteFi” was launched in
2018 (see the acknowledgements for further details). This paper reports the first activities car-
ried out in the context of this project, which aimed at developing a fast and accurate tool to
calculate elastic and plastic buckling (bifurcation) loads/modes of thin-walled stainless steel
columns (uniformly compressed members) exposed to fire and undergoing global/local/distor-
tional buckling. The tool is based on a semi-analytical approach that relies on Generalized
Beam Theory, a thin-walled bar theory that efficiently accounts for cross-section arbitrary in-
plane and out-of-plane (warping) deformation through the consideration of so-called “cross-
section deformation modes” (see, e.g., Schardt 1989, Camotim et al. 2010). The intrinsic non-
linear stress-strain law of stainless steel, including temperature effects, is taken into account
using appropriate tangent elastic moduli, based on both J2 (von Mises) small-strain incremen-
tal and deformation plasticity theories. For illustrative purposes, the tool is applied to assess
the elastic and plastic buckling behaviour of thin-walled lipped channel columns made with
steel grade 1:403 and subjected to fire.

434
The notation in this paper follows that introduced previously, which relies on a simple
vector/matrix form of the equations (Gonçalves et al. 2010b, Gonçalves & Camotim 2011,
2012). A derivative is indicated by subscript commas (e.g., f;x ¼ ∂f =∂x), δ designates a virtual
variation, Δ is an incremental variation to the buckled state, d is a generic variation and super-
scripts ðÞM and ðÞB are used for plate-like membrane and bending terms, respectively.

2 GBT SEMI-ANALYTICAL BIFURCATION EQUATION

2.1 Kinematics and strain


For an arbitrary thin-walled member, as shown in Figure 1, local axes for each wall are set,
(x, y, z), defining the member axis, wall mid-line and thickness directions, respectively.
According to the classic GBT kinematic description, Kirchhoff’s thin plate assumption is
assumed and the displacement vector Uðx; y; zÞ for each wall is expressed as
2 3 2 3
Ux ð
u  zwÞT ;x
Uðx; y; zÞ ¼ 4 Uy 5 ¼ 4 ðv  z
w;y ÞT  5; ð1Þ
Uz zwT 

where uðyÞ, vðyÞ, w  ðyÞ are column vectors containing the deformation mode displacement
components along x, y and z, respectively, and ðxÞ is a column vector that contains the amp-
litude functions of each deformation mode along the beam length, which constitute the prob-
lem unknowns. The deformation mode displacement components are obtained from the so-
called “GBT cross-section analysis”, which is explained in, e.g., Gonçalves et al. 2010b,
Bebiano et al. 2018, and is implemented in the GBTUL program, freely available at www
.civil.ist.utl.pt/gbt.
The non-null small strain components for each wall are grouped in vector ε, which is
straightforwardly obtained from the displacements and reads

2 3 2 3
εxx ð
u  zwÞT ;xx
6 7
ε ¼ 4 εyy 5 ¼ 4 ðv;y  z w;yy ÞT  5; ð2Þ
γxy u;y þ v  2z
ð T
w;y Þ ;x

where the terms with/without z correspond to bending/membrane deformation, respectively.

Figure 1. Arbitrary thin-walled member geometry and local coordinate systems.

435
If null membrane transverse strains (εM yy ¼ 0) are assumed, which is acceptable in most
beam-type problems, then v;y ¼ 0. Furthermore, if Vlasov’s assumption is adopted (null
membrane shear strains, γM xy ¼ 0), which is acceptable for open sections, then 
u;y þ v ¼ 0.
These two assumptions are essential to reduce the number of deformation modes (hence
the problem DOFs) while ensuring that accurate results are obtained in a wide range of
problems.
For the calculation of bifurcation loads, the Green-Lagrange (non-linear) longitudinal
strains are required. Only the membrane component needs to be retained, which reads

1 T T 
Exx ≈ Exx
M
¼ εxx þ w
;x vv þ w  T ;x þ T;xx 
uuT ;xx : ð3Þ
2

u
In the latter equation, the term with u T may be discarded without significant loss of accuracy
(Gonçalves et al. 2010a). The relevant Green-Lagrange strains are therefore grouped in
h iT
vector E T ¼ Exx εyy γxy .

2.2 Stresses and constitutive laws


For the stresses, a plane stress state is assumed and the incremental constitutive relations
between the Green-Lagrange strains and second Piola-Kirchhoff stresses are written as

dS ¼ C t dE; ð4Þ

where S T ¼ ½Sxx Syy Sxy


and C t is the tangent elastoplastic constitutive matrix for the case
under consideration.
If null transverse membrane strains are assumed, the membrane and bending stresses are
separated to avoid over-stiff solutions, leading to

dS M ¼ C M
t dE ; dS ¼ C t dE ;
M B B B
ð5Þ

where C Bt ¼ C t , whereas C M
t is calculated for dSyy ¼ 0. If Vlasov’s assumption is further
M

enforced, C t is also calculated for dSxy ¼ 0, leading to a simple uniaxial law dSxx
M M M
¼ Et dExx
M
,
where Et is the uniaxial tangent modulus.
The small-strain J2 (von Mises) theory is adopted, with associated flow rule and isotropic
strain hardening. In the following expressions, E and G are the elastic (initial) Young and
shear moduli,  is Poisson’s ratio, Es is the secant modulus and H 0 ¼ Et =ð1  Et =EÞ. In this
case the tangent constitutive matrix for the bending terms (plane stress) is given by (Gonçalves
et al. 2010a)
2 3
Ctxy
Ctxx 0
Ct ¼ 4 Ctyy 0 5 ð6Þ
Sym: 0 Gt ;

where the coefficients read, for the incremental (flow) theory

E 2 þ 4EH 0 4E 2 þ 4EH 0
Ctxx ¼ ; Ctyy ¼ ; ð7Þ
ð5  4ÞE  ð 2  1Þ4H 0 ð5  4ÞE  ð 2  1Þ4H 0

2E 2 þ 4EH 0
Ctxy ¼ ; Gt ¼ G ð8Þ
ð5  4ÞE  ð 2  1Þ4H 0

and, for deformation theory,

436
E 2 þ ð1 þ 3es ÞEH 0
Ctxx ¼ ; ð9Þ
ð3es þ 2  4ÞE  ð4 2  3es  1ÞH 0

4E 2 þ 4EH 0
Ctyy ¼ ; ð10Þ
ð3es þ 2  4ÞE  ð4 2  3es  1ÞH 0

2E 2 þ 4EH 0 E
Ctxy ¼ 0
; Gt ¼ Gs ¼ ; ð11Þ
ð3es þ 2  4ÞE  ð4  3es  1ÞH
2 2  1 þ 3es

where es ¼ E=Es . For the membrane terms the same law applies unless dεM yy ¼ 0 is assumed, in
which case dSyyM
¼ 0, Ctxx ¼ Et for both theories and Gt is given by G (incremental theory) or
Gs (deformation theory).
For stainless steel members subjected to fire, the material law specified in Eurocode 3 part
1-4 (CEN 2006) and Annex C of part 1-2 (CEN 2009) is adopted. In the example presented in
Section 3, the cross-section thickness is smaller than 6 mm. For such small values it may be
assumed that the temperature is uniform over the member volume. An austenitic steel grade
1:4301 is considered, in which case one has, for a cold-rolled strip product with t  6 mm,
E ¼ Ea ¼ 200 GPa,  ¼ 0:3, fy ¼ 230 MPa, fu ¼ 540 MPa and the uniaxial law under fire con-
ditions reads

Ea;θ ε
σ¼ ; ðε  εc;θ Þ; ð12Þ
1 þ aεb
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
σ ¼ f0:2p;θ  e þ ðd=cÞ c2  ðεu;θ  εÞ2 ; ðεc;θ 5ε  εu;θ Þ; ð13Þ

where θ is the temperature, Ea;θ ¼ Ea kE;θ , f0:2p;θ ¼ fy k0:2p;θ , and the coefficients ki and a-e, as
well as the tangent modulus Et , are given in Annex C of EC3 part 1-2 (CEN 2009), as
a function of the temperature θ (their values and expressions are not reproduced here due to
lack of space). For illustrative purposes, Figure 2 shows the uniaxial law for the stainless steel
grade considered, for temperatures ranging between 20 and 1000 ⁰C.

Figure 2. Uniaxial constitutive laws for the stainless steel adopted.

437
2.3 Bifurcation equation
In a linearised bifurcation analysis, the fundamental path is determined assuming geometric
linearity and the bifurcation equation corresponds to an “initial stress” problem, obtained
from the linearization of the virtual work equation in the direction of an incremental configur-
ation change,

ΔðδWð ¼ 0; λÞÞ ¼ 0; ð14Þ

where λ is the loading parameter. Since the displacement vector U is a linear function of  and
;x , only the internal part of the virtual work is non-null and the bifurcation equation reads
Z
 
 M ΔδE M dV0 ¼ 0;
δεT Ct Δε þ λS ð15Þ
xx xx
V

where V is the beam initial volume, S  M are the membrane longitudinal normal stresses for
xx
λ ¼ 1 and the first term must be separated into membrane and bending terms if null transverse
membrane strains are assumed, as previously discussed.
In the following derivations, null transverse membrane strains/stresses are assumed. The inte-
gration of Equation 15 along the cross-section leads to a standard eigenvalue problem of the form
2 3T 2 32 3
Z δ B 0 D2 Δ Z
4 δ;x 5 4 0 D1 0 54 Δ;x 5dx þ λ δT;x X Δ;x dx ¼ 0; ð16Þ
L δ;xx DT2 0 C Δ;xx L

where L is the member initial length and the GBT modal matrices are given by
Z Z  
t3 t3
B¼ Ctyy  ;yy w
w  T;yy dy; C ¼ u
Et t uT þ Ctxx w w T dy; ð17Þ
S 12 S 12
Z  
t3
D1 ¼ Gt tð u;y þ vÞT þ w
uy þ vÞð  ;y w
 T;y dy; ð18Þ
S 3
Z Z
t3  M tðvvT þ w
D2 ¼ Ctxy w  ;yy w T dy; X¼ S xx w T Þdy; ð19Þ
S 12 S

where S is the cross-section mid-line and t is the wall thickness. In these expressions, the mem-
brane terms are affected by t, whereas the bending terms are multiplied by t3 .

2.4 Semi-analytical bifurcation equation


For simply supported members under uniform stress states, sinusoidal amplitude functions of
the form Δk ¼ k sinðnπx=LÞ constitute the exact solutions, where n is the number of half-
waves of the buckling mode and k is the deformation mode amplitude. Substituting the exact
solution into the bifurcation Equation 16 leads to
 
nπ2 L2
C þ D1  D2  D2 þ 2 B þ λX  ¼ 0:
T
ð20Þ
L2 nπ

This equation is quite efficient from a computational point of view, since the number of
DOFs equals the number of deformation modes included. Each bifurcation load λ (eigenvalue)
is associated with a buckling mode  (eigenvector) whose elements correspond to the partici-
pation of each GBT deformation mode.

438
Since the GBT matrices depend on λ through the tangent moduli, an iterative strategy is
necessary to calculate the bifurcation loads. For the calculation of the critical load (the lowest
bifurcation load), the procedure increases λ and calculates the tangent moduli at each step
until jλcr  λj5TOL, where λcr is the critical load parameter obtained from Equation 20 at
each step. To increase the speed of the procedure, the GBT matrices are initially stored without
the tangent moduli and are updated at each step by multiplying each one with the correspond-
ing moduli.

3 ILLUSTRATIVE EXAMPLE

For illustrative purposes, columns with the lipped channel section shown in Figure 3 are ana-
lysed. The material properties correspond to those given in the previous section.
For the calculation of the GBT cross-section deformation modes, three/four equally
spaced intermediate nodes are included in the flanges/web, respectively, leading to 48
deformation modes. However, only the so-called conventional modes (Bebiano et al.
2018) are relevant for the buckling problem under consideration and therefore are the
only ones included in the analyses: 4 rigid-body modes, 2 distortional modes and 12
local-plate modes. This means that the bifurcation problem has only 18 DOFs and it is
worth mentioning that all these deformation modes comply with the εM yy ¼ 0 and γxy ¼ 0
M

assumptions. The in-plane and warping displacements of the first 12 modes are displayed
in Figure 3.
First, a buckling analysis is carried out assuming a linear elastic material law with
Et ¼ Ea;θ ¼ Ea kE;θ . The graphs in Figure 4 show the the critical stress as a function of the
buckling mode half-wavelength (i.e. n ¼ 1, the so-called “signature curve”) and the tempera-
ture, as well as the GBT modal participations and the mid-span cross-section deformed con-
figuration associated with the elastic buckling mode for selected half-wavelengths (top of the
figure). These results show a rather small variation of the critical stresses for 0  θ  700,
since kE;θ ¼ 0:71. However, for 1000 °C one has kE;θ ¼ 0:20 and the drop in the critical stres-
ses is significant. The modal participation graphs show that the mode transitions occur for
smaller half wavelengths as the temperature increases. Nevertheless, the trend of the buckling
modes is similar (from small to large half wavelengths): (i) local (7 + 9), (ii) symmetric

Figure 3. Lipped channel geometry (the dimensions correspond to the wall mid-line) and shapes of the
first 12 GBT deformation modes.

439
Figure 4. Lipped channel columns: signature curve, buckling modes and modal participation diagram
for elastic behaviour.

distortional (5), (iii) anti-symmetric distortional-flexural-torsional (6 + 2 + 4), (iv) flexural-


torsional (2 + 4) and finally (v) minor axis flexural.
Consider now the plastic buckling case. Figure 5 shows the results obtained for both the incre-
mental (left) and deformation (right) theories. Although both theories yield virtually identical
results for θ ¼ 20 °C, the differences increase with the temperature, with deformation theory pro-
viding the lowest critical loads, as expected (see, e.g., Gonçalves & Camotim 2004, 2007). Con-
cerning the mode participation diagrams, in all cases the nature of the buckling mode changes
with the half wavelength in a manner similar to that obtained for the elastic cases. However, it
should be noted that the incremental theory predicts mode transitions for smaller half wave-
lengths, most remarkably for θ ¼ 1000 °C (bottom left graph), in which case the local mode (7 +
9) does not appear in the graph. Finally, attention is called to the fact that the “kink” appearing
in the critical stress curve for the incremental theory, θ ¼ 1000 °C and L ¼ 1000 mm, is not due
to a change in the buckling mode shape but rather to a discontinuity of the tangent moduli for
εxx ¼ εc;θ , due to the particular form of the uniaxial constitutive law adopted, given by Equations
12-13.

440
Figure 5. Lipped channel columns: signature curve and modal participation diagram for plastic
behaviour.

4 CONCLUSION

This paper reported the first activities carried out in the context of project “StaSteFi”. A fast
and accurate numerical tool to assess the elastic and plastic local/distortional/global bifurcation
behaviour of thin-walled stainless steel columns exposed to fire was presented. The tool is based
on a GBT-based semi-analytical approach and therefore is capable of handling, efficiently and
accurately, cross-section arbitrary in-plane and out-of-plane (warping) deformation with a very
small computational cost. Moreover, the intrinsic non-linear stress-strain law of stainless steel,
including temperature effects, can be straightforwardly taken into account using the appropriate
tangent elastic moduli pertaining to J2 small-strain incremental and deformation plasticity the-
ories. For illustrative purposes, the elastic and plastic bifurcation behaviour of thin-walled
lipped channel columns made with steel grade 1:403 and subjected to fire was assessed.
Future developments, which are already under way and will be reported in the near future,
include extending the proposed GBT formulation to a finite element setting, in order to
handle arbitrary loading and support conditions.

ACKNOWLEDGEMENTS

The authors gratefully acknowledge the financial support of FCT (Fundação para a Ciência
e a Tecnologia, Portugal), through Project “StaSteFi—Fire design of stainless steel structural
elements”.

441
REFERENCES

Bebiano, R., Camotim, D. & Gonçalves, R. 2018. GBTUL 2.0 – a second-generation code for the GBT-
based buckling and vibration analysis of thin-walled members. Thin-Walled Structures 124: 235–253.
Camotim, D., Basaglia, C., Bebiano, R., Gonçalves, R. & Silvestre N. 2010. Latest developments in the
GBT analysis of thin-walled steel structures. In E. Batista, P. Vellasco, L. Lima (eds.), Proceedings of
International Colloquium on Stability and Ductility of Steel Structures, Rio de Janeiro, 8-10 Septem-
ber 2010 33–58 (Vol. 1).
CEN 2006. Eurocode 3 - Design of steel structures - Part 1-4: General rules - Supplementary rules for stain-
less steels. Brussels, Belgium: European Committee for Standardization.
CEN 2009. Eurocode 3 - Design of steel structures - Part 1-2: General rules - Structural fire design. Brus-
sels, Belgium: European Committee for Standardization.
Gardner, L. & Baddoo, N. 2006. Fire testing and design of stainless steel structures. Journal of Construc-
tional Steel Research 62: 532–543.
Gonçalves, R. & Camotim, D. 2004. GBT local and global buckling analysis of aluminium and stainless
steel columns. Computers & Structures 82(17-19): 1473–1484.
Gonçalves, R. & Camotim, D. 2007. Thin-walled member plastic bifurcation analysis using Generalised
Beam Theory. Advances in Engineering Software 38(8-9): 637–646.
Gonçalves, R., Le Grognec, P. & Camotim, D. 2010a. GBT-based semi-analytical solutions for the plas-
tic bifurcation of thin-walled members. International Journal of Solids and Structures 47(1): 34–50.
Gonçalves, R., Ritto-Corrêa, M. & Camotim, D. 2010b. A new approach to the calculation of
cross-section deformation modes in the framework of Generalized Beam Theory. Computational
Mechanics 46(5): 759–781.
Gonçalves, R. & Camotim, D. 2011. Generalised Beam Theory-based finite elements for elastoplastic
thin-walled metal members. Thin-Walled Structures 49(10): 1237–1245.
Gonçalves, R. & Camotim, D. 2012. Geometrically non-linear Generalised Beam Theory for elastoplastic
thin-walled metal members. Thin-Walled Structures 51: 121–129.
Lopes, N., Vila Real, P., Simões da Silva, L. & Franssen, J-M. 2010. Axially loaded stainless steel col-
umns in case of fire. Journal of Structural Fire Engineering 1(1): 43–59.
Lopes, N., Vila Real, P., Simões da Silva, L. & Franssen, J-M. 2012. Numerical analysis of stainless steel
beam-columns in case of fire. Fire Safety Journal 50: 35–50.
Ng, K. & Gardner, L. 2007. Buckling of stainless steel columns and beams in fire. Engineering Structures
29(5): 717–730.
Schardt, R. 1989. Verallgemeinerte Technische Biegetheorie. Berlin: Springer-Verlag.
SCI 2017. Design Manual for Structural Stainless Steel, 4th Edition, Ascot, UK: The Steel Construction
Institute.
Uppfeldt, B., Outinen, T. & Veljkovic, M. 2008. A design model for stainless steel box columns in fire.
Journal of Constructional Steel Research 64(11): 1294–1301.

442
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

A comparative analysis on the stability and ultimate strength


of steel plated girders with planar and corrugated webs

A. González, L. Vallelado & M.A. Serna


Department of Structural and Mechanical Engineering, University of Cantabria, Spain

ABSTRACT: The paper presents a finite element study on buckling loads and ultimate
strength of plated girders with trapezoidal corrugated webs and the results are compared with
those obtained for girders with planar webs, with or without stiffeners. The comparative ana-
lysis is focused on three main points: local buckling of the web due to shear forces; ultimate
strength of the girder under shear forces and bending moments; and global stiffness and duc-
tility post-buckling behavior. Results show that the use of corrugated web significantly
increases buckling loads and overall ultimate strength when compared with unstiffened planar
web. However, these benefits are at the cost of reducing ductility. Moreover, similar structural
properties to those obtained with corrugation may be attained using stiffened planar webs
with a reduced number of stiffeners.

1 INTRODUCTION

The use of sinusoidal and trapezoidal corrugated steel plates as webs in welded girders has
seen a growing research interest in the last two decades, as has been pointed out by Hasan
et al (2017). From a structural engineering point of view, it seems clear that corrugation has
the benefit of increasing local buckling loads. A numerical study on shear capacity of plate
girders with trapezoidally corrugated webs was developed by Luo & Edlund (1996). In the
same period, a comprehensive experimental study of built-up girders with trapezoidal corru-
gated web was presented by Elgaaly et al. (1997) and Elgaaly & Seshadri (1998). More
recently, Jáger et al. (2015, 2017a, 2017b and 2017c) have developed a complete experimental
and theoretical research on the interaction of shear, bending and path loading. From
a theoretical point of view, the stress state of I-section girders has been studied by Abbas et al.
(2006, 2007a, 2007b), who have proposed a direct analytical method for their analysis.
Leblouba et al. (2017a and 2017b) have presented a normalized shear strength formulation for
trapezoidal corrugated webs. Centered on sinusoidal corrugation, Pasternak & Kubieniec
(2010) have carried out an experimental research and have proposed new formulations for
patch loading and lateral torsional buckling resistance.
In this context, the paper presents a study on the buckling loads and ultimate strength of
plated girders with trapezoidal corrugated webs and results are compared with those obtained for
girders with planar webs, with or without stiffeners. Results are obtained using finite element ana-
lysis. Trapezoidal shape and girder geometry are taken from some of the elements used by Hassa-
nein & Kharoob (2013a and 2013b) in their analysis of bridge girders with corrugated webs.
The comparative analysis presented in the paper is focused on the following points: local
buckling of the web due to shear forces; ultimate strength of the girder under shear forces and
bending moments; and global stiffness and ductility post-buckling behavior. Results show
that corrugated webs provides a significant increase in shear buckling resistance when com-
pare with planar plates of the same thickness. However, corrugated webs are prone to fragile
post buckling behavior. On the other hand, planar webs develop membrane stresses and pro-
vide a more ductile collapse. Moreover, ultimate strength of stiffened enough planar web gir-
ders may be equivalent to that of corrugated webs.

443
In what follows, first we outline the problem statement. Then buckling results are presented,
showing the number of stiffeners needed for a planar web in order to reach local buckling
load equivalent to that of corrugated web. Next, ultimate strength analysis of girders with cor-
rugated web are presented and compared with those obtained for planar web girders with 0, 1,
2 and 3 intermediate stiffeners.

2 PROBLEM STATEMENT

The paper has two main objectives. The first one is to determine elastic buckling load and
ultimate strength of girders with trapezoidal corrugated web. The second is to obtain a clear
picture of the beneficial effect of using corrugated web by comparing buckling load and ultim-
ate strength results with those obtained for girders with planar webs, with or without stiff-
eners. All analysis are performed using finite element models with an I-section simple
supported girder subjected to a load applied at the span center. The trapezoidal corrugated
web is defined in Figure 1, where general geometric parameters are defined.

2.1 Buckling analysis


In order to have a simple picture of the beneficial effect of using trapezoidal corrugated webs,
our buckling analysis is centered on a simple question: how many stiffeners would be needed
so that shear buckling load of a planar web would be equal to that of a corrugated web?.
For this particular analysis we have worked with a trapezoidal wave with fixed
a1 = a2 = 325 mm and a corrugation angle a that varies from 25º to 45º. On the other hand,
planar web is supposed to have equally separated stiffeners, being aS the intermediate distance
(Figure 2).
The elastic buckling analysis is performed for web thickness, tw, of 6, 8, 10, 12, 14, 16 and
18 mm, and a flange thickness tf = 8tw, so that shear web buckling develops. I-section dimen-
sions h and b are taken as 1000 and 400 mm respectively. The girder half span S is the distance
corresponding to two trapezoidal waves, that is S = 4(a1+ a4) with a4 = a2cosa (Figure 2).
For each web thickness and corrugation angle, buckling load is first determined for the cor-
rugated web. Then, following an iterative process in which as is used as variable, the equiva-
lent distance between stiffeners is determined so that corrugated web and stiffened planar web
have the same buckling load.

2.2 Ultimate strength analysis


For the ultimate strength analysis a fixed trapezoidal wave geometry is used with the following
dimensions: a1 = a2 = 325 mm, a4 = 275 and a3 = 175 mm. Half span is taken as 4800 mm,

Figure 1. Geometric parameters.

444
Figure 2. Corrugated web and stiffened planar web dimensions.

Figure 3. Corrugated web and stiffened planar web girders scheme.

which corresponds to 4 trapezoidal waves (Figure 3). Three values for web thickness tw (6, 9
and 12 mm), and two for I-section high h (1600 and 2400 mm) are considered. Flange thick-
ness tf and wide b are taken as 50 and 500 mm respectively. Regarding planar web girders,
four cases are considered, corresponding to the number of intermediate stiffeners: 0, 1, 2
and 3.
For all analysis, structural steel S355, with the properties defined by the Eurocode 3, is
used. Following Driver et al. (2006) study, imperfections are taken for each girder using the
first buckling mode scaled by a factor equal to the corresponding web thickness. The non-
linear finite element analysis is performed using ANSYS Academic Mechanical and the force-
displacement curve is followed until no further equilibrium point is achieved in the numerical
integration.

3 BUCKLING RESULTS

Buckling analysis results are summarized in Figure 4. It can be seen that, in order to have
similar buckling load, tw = 6 mm girders with planar web need to have intermediate stiffeners
at distances as between 1,20a1 and 1,23a1 corresponding the smallest value of distance as to
the higher corrugation angle. Distance as increases with web thickness, so that for tw equal to

445
Figure 4. Stiffeners separation for equivalent buckling loads.

18 mm as varies from 1,40a1 to 1,73a1. Consequently, for a corrugation angle of 45º the stiff-
ener intermediate distance is 390 mm for 6 mm web thickness and 455 mm for 18 mm.
These results show that folds on corrugated webs have the effect of non-rigid stiffeners.
Buckling modes usually affect to more than one panel of the trapezoidal wave. On the other
hand, since stiffeners used for planar webs have a thickness equal to 8tw, buckling modes for
planar webs only affects one panel.
From the results it is clear that corrugation has a significant and positive effect on elastic
shear buckling resistance.

4 ULTIMATE STRENGTH RESULTS

Ultimate strength results are presented in a graphical way by the deformed shape at the point
of maximum load and the force-displacement curve. All deformed shapes are shown multiply-
ing the actual deformation by a factor of 5.

4.1 Girders with corrugated web


Ultimate strength of girders with trapezoidal corrugated web directly depends on web thick-
ness and I-section high. Figure 5 shows that girders with 6 and 9 mm web thickness have
a failure mode related to shear web strength. On the other hand, the failure of girders with
a web thickness of 12 mm is produced by bending moment affecting flanges at mid span.
Force-displacement curves are shown in Figure 6, where first number in captions refers to
section high, the second to web thickness, and the letter C to corrugated web. It can be seen
that shear failure of corrugated webs, which happens to be the case for 1600-6-C, 1600-9-C,
2400-6-C and 2400-9-C, leads to linear and fragile behavior, with a sudden folding of the web
at ultimate strength. This fast change in web geometry stops the numerical integration. In con-
trast, when bending moment failure takes places (1600-12-C and 2400-12-C) a ductile process
follows the linear elastic force-displacement curve.

4.2 Girder with planar web


As indicated above, four cases are considered for each planar web thickness. Figure 7 shows
deformed shapes at maximum load, with a scale factor of 5, and force-displacement curves for

446
Figure 5. Deformed shapes for corrugated web girders.

Figure 6. Force-Displacement curves for corrugated web girders.

1600 mm high girders. The number of stiffeners is indicated by the digit “n” in the caption
“XXXX-X-PnS”. It can be seen that shear failure and ductile behavior are present for all web
thicknesses. Similar results are obtained for 2400 mm high girders (Figure 8).
As expected, force-deformation curves indicate that initial girder stiffness (initial slope)
does not depend on the number of stiffeners but on the web thickness. The number of stiff-
eners does influence girder ultimate strength for each tw.

4.3 Comparison
FEM results show that the use of corrugated webs significantly increases ultimate strength
when compared with unstiffened planar webs. This comparative benefit diminishes and may

447
Figure 7. Deformed shapes and strength curves for planar web girders (h = 1600 mm).

be canceled with the number of stiffeners. For the particular case of 1600 mm high section,
planar web girders with three stiffeners have similar values of ultimate load, and better initial
rigidity, than corrugated web girders. For 2400 mm high section, ultimate strength of corru-
gated web girders surmounts that of planar web girders with three stiffeners, but initial rigidity
is lower. Figure 9 compares force-displacement curves of corrugated web girders and planar
web girder with three stiffeners. Regarding failure mode, it is worth noting that corrugated
web introduces flange failure possibility at lower loads, due to the increase of flange buckling
length with web corrugation.
Another interesting result concerns to post-buckling resistance of girders with planar web.
Results show that ultimate strength for planar web girders are usually higher than elastic
buckling loads. This post-buckling resistance comes from the development of tension stresses

448
Figure 8. Deformed shapes and strength curves for planar web girders (h = 2400 mm).

as has been shown by Höglund (1997). Consequently, plane web girders have a ductile failure
behavior while corrugated web girders have a fragile collapse. On the other hand, failure load
for corrugated web girders are usually lower than the corresponding elastic buckling load, and
have a fragile mode.

449
Figure 9. Comparison of force-displacement curves.

5 CONCLUSIONS

A finite element study on buckling loads and ultimate strength of plated girders with trapez-
oidal corrugated webs have been presented. Results have been compared with those obtained
for girders with both stiffened and un-stiffened planar webs.
Buckling analysis have shown that corrugation significantly increases buckling resistance.
Moreover, trapezoidal wave with higher corrugation angle have higher buckling load. Results
indicate that corrugation beneficial effect diminishes with plate thickness. The paper has also
presented a study on the number of intermediate rigid stiffeners needed for planar webs in
order to have the same buckling loads as the corrugated ones.
Ultimate strength analysis have provided force-displacement curves for a set of corrugated
web girders, varying I-section high and web thickness. These results have been compared with
those corresponding to planar web girders with up to three intermediate rigid stiffeners.
Results show that corrugated web girders have higher ultimate strength than planar web gir-
ders. For a 2400 mm high section and 6 mm web thickness, ultimate load for a corrugated
web girder doubles that of a planar web girder. Corrugated web failure has shown to be fra-
gile, in contrast with planar web failure which is ductile.
Results have demonstrated that the use of intermediate stiffeners significantly improves
ultimate strength of planar web girders. In fact, it has been shown that for 1600 mm high sec-
tion, planar web girders with three stiffeners have similar resistance than corrugated web gir-
ders. For the 2400 mm high girders, three stiffeners are not enough to reach equivalent results,
particularly for 6 and 9 mm web plate thicknesses.
Finally, results for girders with 12 mm thick corrugated web have shown that corrugation
diminishes girder resistance to flange buckling, something related to the increase of flange
buckling length with corrugation.

REFERENCES

Abbas, H.H., Sause, R. & Driver, R.G. 2006. Behavior of corrugated web I-girders under in-plane loads.
Journal of Engineering Mechanics 132 (8): 806–814.
Abbas, H.H., Sause, R. & Driver, R.G. 2007a. Analysis of flange transverse bending of corrugated Web
I-girders under in-plane loads. Journal of Structural Engineering 133 (3): 347–355.
Abbas, H.H., Sause, R. & Driver, R.G. 2007b. Simplified analysis of flange transverse bending of corru-
gated web I-girders under in-plane moment and shear. Engineering Structures 29 (11): 2816–2824.
Driver, R.G., Abbas, H.H. & Sause, R. 2006. Shear behavior of corrugated web bridge girder. Journal of
Structural Engineering 132 (2): 195–203.
Elgaaly, M., Seshadri, A. & Hamilton, R.W. 1997. Bending strength of steel beams with corrugated
webs. Journal of Structural Engineering 123 (6): 772–782.
Elgaaly, M. & Seshadri, A. 1998. Steel built-up girders with trapezoidally corrugated webs. Engineering
Journal 35 (1): 1–10.

450
Hasan, Q.A., Wan Badaruzzaman, W.H., Al-Zand, A.W. & Mutalib, A.A. 2017. The state of the art of
steel and steel concrete composite straight plate girder bridges. Thin-Walled Structures 119: 988–1020.
Hassanein, M.F. & Kharoob, O.F. 2013a. Behavior of bridge girders with corrugated webs: (I) real
boundary conditions at the juncture of the web and flanges. Engineering Structures 57: 554–564.
Hassanein, M.F. & Kharoob, O.F. 2013b. Behavior of bridge girders with corrugated webs: (II) shear
strength and design. Engineering Structures 57: 544–553.
Hoglund, T. 1997. Shear buckling resistance of steel and aluminium plate girders. Thin-Walled Structures
29 (1-4): 13–30.
Jáger, B., Dunai, L. & Kövesdi, B. 2015. Girders with trapezoidally corrugated webs subjected by com-
bination of bending, shear and path loading. Thin-Walled Structures 96 (art. no. 4603): 227–239.
Jáger, B., Dunai, L. & Kövesdi, B. 2017a. Flange buckling behavior of girders with corrugated web Part
I: Experimental study. Thin-Walled Structures 118: 181–195.
Jáger, B., Dunai, L. & Kövesdi, B. 2017b. Flange buckling behavior of girders with corrugated web Part
II: Numerical study and design method development. Thin-Walled Structures 118: 238–252.
Jáger, B., Dunai, L. & Kövesdi, B. 2017c. Experimental investigation of the M-V-F interaction behavior
of girders with trapezoidally corrugated web. Engineering Structures 133: 49–58.
Leblouba, M., Barakat, S., Altoubat, S., Junaid, T.M. & Maalej, M. 2017a. Normalized shear strength
of trapezoidal corrugated steel webs. Journal of Constructional Steel Research 136: 75–90.
Leblouba, M., Junaid, M.T., Barakat, S., Altoubat, S. & Maalej, M. 2017b. Shear buckling and stress
distribution in trapezoidal web corrugated steel beams. Thin-Walled Structures 113: 13–26.
Luo, R. & Edlund, B. 1996. Shear capacity of plate girders with trapezoidally corrugated webs. Thin-
Walled Structures 26: 19–44.
Pasternak, H. & Kubieniec, G. 2010. Plate girders with corrugated webs. Journal of Civil Engineering and
Management 16 (2): 166–171.

451
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Experimental testing of plastic buckling of moderately thick


circular rings under uniform lateral loading

F. Guarracino
University of Naples “Federico II”, Naples, Italy

ABSTRACT: Evaluation of the critical external uniform loading for a circular ring occurs
frequently in a wide range of structural problems and as a consequence it has been extensively
treated in a large number of works. Since the available results in literature result often scat-
tered and not always consistent, a number of hoc ad experimental tests have been carried out
by means of a recently patented apparatus in order to highlight the main features of the plastic
collapse mechanism for rings of moderate thickness. The experiments are reported and inter-
peted with respect to a variety of factors and it is found that the shape of the initial imperfec-
tion and the seal friction can both affect the collapse and produce a variation in the value of
the collapsing pressure.

1 INTRODUCTION

In deep and ultra-deep waters the diameter of trunk lines coupled with the hydrostatic pressure
tends to lead to failure of the pipeline by external collapse. This failure mode is an instability
phenomenon that is governed by the geometry of the pipeline and its material properties. Fail-
ure of a pipeline section is dependent on many factors and as such the collapse pressure is very
difficult to determine. Finite element techniques can allow the estimation of collapse where the
actual geometry and material properties are known but in reality shape and properties vary
along the pipe length, around the circumference as well as through the wall thickness.
To enable effective design work to be performed, and proven during service, it is necessary
to ensure that the mechanical and dimensional properties of the pipe are controlled within
very strict boundaries. More recently, and in collaboration with SAGE (Middle East to India
Pipeline), atest method has been developed that allows the determination of pipe collapse
pressures through the application of external pressure to rings of pipe materials (Roberts
et al., 2010).
This test method has been demonstrated to correlate well with full scale collapse data and it
the manufacturing procedure qualification tests (MPQT) of offshore pipelines projects has
recently started to rely on it. This approach is capable of providing valuable insight into the
collapse behaviour of offshore pipelines and has practical, financial and logistic benefits..
In fact full-scale tests performed using complete pipe joints are very expensive and only
a limited number of test facilities can perform them. Moreover, the experimental setting must
be accurately designed and controlled, since the boundary conditions can adversely affect the
results under some circumstances (Guarracino et al., 2008b and 2009; Guarracino, 2011).
The test apparatus and the results from these tests are analysed in this work with respect to
a variety of factors such as the shape of the initial imperfection and the seal friction during the
loading procedure.

2 TESTING APPARATUS AND PROCEDURE

Figure 1 shows a section through the ring test equipment (Roberts et al., 2010).

452
Figure 1. Typical four-point bending test arrangement.

The test ring (12) is held between two effectively rigid plates, (16 and 20). The two halves of
the pressure test chamber 16 and 20 are provided with locating spigots (19) that correspond to
locating holes with associated seals (21) to allow location of the two halves. O-ring seals (20)
are located in the top and bottom sections. The central void (26) inside the test ring (12) is
vented to atmosphere through a bleed hole (28) which is of sufficiently large diameter to also
provide access for any strain gauge instrumentation (not shown) on the inner cylindrical sur-
face of the test ring (12). The plates (16 and 20) are bolted together and the pressure is intro-
duced into the space marked (22). In fact, the plates (16 and 20) are engaged by the test ring
(12) to form an annulus accessible by a supply of pressurised hydraulic test fluid through an
appropriate inlet port (24). The screws (30) extend through holes (32) in the top section (16)
and pass through the void (26) to engage in threaded bores (34) in the bottom section (20).
The pressure is retained in the space by the O-rings between the surface of the plate and the
edge of the ring. The O-rings allow the ring to move relative to the plates while still maintain-
ing a pressure seal.
The inner surface of the ring is open to the atmosphere. The ring is fitted with four strain
gauges equally-spaced around the inner surface of the ring. Four displacement transducers are
also attached to the inner surface of the ring at positions between the strain gauges (see Figure 2).
The test procedure is as follows: rings are positioned in the bottom plate of the equipment
and then the top plate is placed over it and bolted in position. The instruments are connected
and set to zero. The pressurising fluid is introduced into the space between the outer surface
of the rings and the test plate and the pressure is gradually increased in a controlled manner.
As the applied load reaches the collapse limit, the pressure suddenly reduces and with con-
tinued addition of fluid the ring collapses.

Figure 2. Assembled ring test equipment (left), lower plate with the test ring positioned in the test
cavity (middle) and strain gauges (A-B) and displacement transducers (TN130-139) (right).

453
3 SOME EXPERIMENTAL RESULTS

The results from four different circular beams machined at Tata Steel some years ago (Tata
Steel, 2011) from the same pipe of diameter D = 457.2mm and thickness t = 31.75mm, are
shown in Figure 3, where the displacements are referred to four transducers placed at 0, π/2, π
and 3π/2 positions, see Figure 3.
The first three cases, namely (a), (b) and (c), show a limit pressure consistently ranging from
67.6 to 69.5 MPa, while the last one, (d), shows a limit pressure of 55.4 MPa.
This kind of behaviour has been repeatedly found and reported, to various degrees, also in
additional instances, such as the Oman India test program (Nash et al., 2013) and the South
Stream Offshore Pipeline FEED (Selker et al., 2014) and the need to understand the possible
causes of such findings has led to the present investigation into the mechanics of the problem.
The attention is focused on two different parameters, i.e. initial out-of-roundness of the ring
and the friction between O-rings and the plates of the test rig.

4 ANALYSIS OF INITIAL IMPERFECTIONS

Out-of-roundness (OOR) in many design codes is defined in terms of a parameter involving the
maximum, minimum and nominal diameters of a structure. For example, in DNV-OS-F101
Submarine Pipeline Systems (Det Norske Veritas, 2000) the ovality parameter is given by

Dmax Dmin
ð1Þ
Dnom

where Dmax, Dmin and Dnom are, respectively the maximum, minimum and nominal diameters
of the pipe.

Figure 3. Measured radial displacements against the applied pressure for four different rings machined
from the same pipe (D = 457.2mm, t = 31.75mm).

454
Figure 4. Sample OOR imperfection shapes: two-waves (left) and three-waves (right).

The analyses conducted in some previous works (Fraldi et al., 2011a, 2011b, 2013, 2014;
Guarracino, 2019)) suggest that it is important to be able to quantify not only the degree of
out-of-roundness of pipes and rings if an accurate estimate of their strength is to be attained.
In particular with reference to Figure 4, it was found that in the case of UOE formed pipes
a three-wave shape component is very likely to play an important role, along with the trad-
itional elliptical shape, in the collapse of the structural element under external pressure.
Out-of-roundness was measured using a lase FARO® scanning arm (FARO Technologies,
2009) and the influence of mixed-mode imperfections was analysed by means of a commercial
Finite Element (FE) package (ANSYS, 2007). In general, numerical investigation of the non-
linear behaviour of structures must follow the equilibrium path, identifying and computing
the singular points like limit or bifurcation points, whose secondary branches in the equilib-
rium path must be examined and followed and this procedure can be adversely affected by
any kind of approximation, as shown even in the simplest examples (Guarracino, 2007; Guar-
racino et al., 2008a). To overcome difficulties with limit points, displacement control tech-
niques were introduced and for this reason the modified arc-length method was used to follow
the load-deformation path (Forde et al., 1987).
The influence of initial imperfection modes varies with their relative amplitude and offset,
but the increment in the buckling load deriving from the presence of the second mode (three-
waves) does not result always proportional to its weight and provides the value of 67.764
MPa, in line with the results from Figure 3 (a)-(c), when the contribution from the lowest
buckling mode (two-waves) is very low or nearly negligible.
Figures 5 and 6 show both the load-displacement plot and the deformation at collapse from
FE analyses for the limit cases of pure two and pure three-waves initial imperfection.
From the numerical simulations it can be inferred that the circular beams tested in cases
(a)-(c) of Figure 3 are likely to have been shaped in the manufacturing process with
a predominant three-waves imperfection.

Figure 5. Results from FE analysis (predominant two-waves imperfection).

455
Figure 6. Results from FE analysis (predominant three-waves imperfection).

5 EFFECTS OF FRICTION BETWEEN O-RINGS AND THE PLATES


OF THE TEST RIG

In order to quantify the radial restraint induced by the O-rings, reference is made to push-out
tests performed for the South Stream Ring Collapse Test Program (Selker et al., 2014). The
South Stream Offshore Pipeline project was aimed to the installation of a pipeline from the
Russian shore of the Black Sea, through Turkish waters, to the Bulgarian shore, in water
depths ranging up to 2200 m. A large number of ring specimens, manufactured from base plate
material produced by several plate mills, were supplied from Europe and Asia. All the speci-
mens were characterised by an external diameter (OD) equal to 812.8 mm (32 in) and a nominal
wall thickness equal to 39 mm. Execution of the ring tests was performed by C-FER at their
test facilities in Edmonton, Canada, under careful procedure and quality control.
The push-out testing means that global O-ring friction properties were obtained through
measurements of hydraulic ram force versus push distance. In other words, an hydraulic con-
trolled force was applied to a side of the ring inside the testing device and record of force-
displacement plots was taken.
The tests were performed for several values of hydrostatic pressure but a considerable scat-
ter was found in the results, so that it is here believed that the most convenient approach to
represent the possible restraint due to the friction between the seals and the ring sample may
consist in making reference to a modified Coulomb’s law of friction. Consequently, contact
friction directed in the direction opposite to the ring displacement inside the test rig can be
described, for each of the O-rings, by a relationship of the type

τ  μ s σn when no slip occurs


ð1Þ
τ ¼ μ k σn when slip occurs

μs and μk being the static and kinetic coefficient of friction, respectively, and σn the normal
pressure exerted on the top and bottom surfaces of the tested ring. An exponential decay rela-
tionship can be assumed to link the static and kinetic coefficient of friction as function of the
slip velocity, γ_ eq ,

μ ¼ μs þ ðμs  μk Þeψ_γeq ð2Þ

where ψ is a scalar.
The test setup has been reproduced in its governing aspects by means of a commercial
Finite Element package (Simulia, 2014) and parametric analyses have been performed in
order to investigate the influence of friction on the collapse.

456
The testing device has been modelled using two stiff plates, as shown in Figure 7. The upper
plate is allowed to move in the z-direction, so that when a sealing vertical load is applied,
a vertical pressure is transmitted to the tested ring. The plates measure 1200 × 1200 mm.
Top and bottom plates are linearly elastic. The material model shown in Figure 8 has been
implemented for the ring under test.
The adopted contact model (so-called “hard-contact”) conforms to Equation (1) and also
implies that:

1. the surfaces transmit no contact pressure unless the nodes of the two surfaces are in contact;
2. no penetration is allowed at each constraint location;
3. there is no limit to the magnitude of contact pressure that can be transmitted when the sur-
faces are in contact.

In order to setup a sensitivity assessment, two pairs of friction coefficients (static/kinetic) have
been adopted, i.e. 0.2/0.16 and 0.4/0.32.
The loading is applied in subsequent steps: first, a vertical pressure of 50 kPa is applied on
the top surface of the upper plate; second, a hydrostatic pressure is progressively applied to
the lateral surface of the test ring. The maximum value of the hydrostatic pressure, pmax , is
60 MPa.
In such a manner the lowest kinetic coefficient of 0.16 gives origin to a value of the ram
force equal to 24 kN, which corresponds to the highest value found during the pull-out tests
(Selker et al., 2014). The results of the analyses are shown in Figure 9.
It is immediate to notice that in the frictionless case the collapse pressure results equal to
33.6 MPa, which is the result provided by the analytical formulation proposed in Fraldi et al.,
2011a, for a two-wave imperfection shape and a very low value of initial imperfection. In the
case of friction coefficients equal to 0.2/0.16, the collapse pressure results equal to 34.8 MPa,
which approximately corresponds to the lowest value found in the South Stream Ring Col-
lapse Test Program. In the case of friction coefficients equal to 0.4/0.32, the collapse pressure
results equal to 40.8 MPa, which approximately corresponds to the highest value found in the
South Stream Ring Collapse Test Program. Incidentally, this is also about the result provided
by the analytical formulation proposed in Fraldi et al., 2011a,for a three-wave imperfection
shape.
At this stage it can be concluded that both the shape of the initial imperfection and the
effects of friction between the sealing O-rings and the plates of the testing machine are capable
to justify a variation in the measured values of the buckling pressure, as the one found in
many of the reported experimental tests.

Figure 7. FE model of the testing device.

457
Figure 8. Ramberg-Osgood fit of the coupon test results (E = 221 GPa,  = 0.3).

Figure 9. Results from FE analyses: influence of friction.

6 CONCLUSIONS

The ring test procedure recently used in manufacturing procedure qualification tests (MPQT)
of offshore pipelines projects has been presented, together with some experimental results.
The attention has been focused on two different parameters, i.e.the initial out-of-roundness of
the ring and the friction between O-rings and the plates of the test rig.
In particular, with respect to the friction it has been shown that the variation in lateral
restraint forces from the O-rings is capable of causing some changes in the calculated collapse
pressures which are not proportional to their value. Therefore, it is possible that large
increases in the friction resistance force which might occur due to a fault in the test equipment
could cause a significant change in the collapse pressure and this fact could lead to estimating
pressure collapse values that would be erroneously large and non-conservative.
This is confirmed by the heterogeneity in some early experimental results, which is likely to
be due to the combination of the factors analysed here, the initial out-of-roundness of the ring
and the friction between O-rings and the plates of the test rig, and it is concluded that further

458
experimental investigations, which are becoming progressively available, must be analysed in
depth for a conclusive assessment of the procedure.
In fact, it can be pointed out that from a general standpoint the ring tests, if carefully set up
and interpretated, are capable to account in an integral manner for the effects on the collapse
pressure from all the various factors of material and geometric variability around and through
the pipe wall, factors which are still partially or completely ignored in many current design
guidelines.

ACKNOWLEDGEMENTS

The present study has been motivated and supported over the years by Verderg Ltd. The
author gratefully acknowledges the support of Dr Aldo Giordano for the development and
analysis of the FE model of Section 5.

REFERENCES

ANSYS, Inc. 2007. ANSYS 11.0 User’s Documentation. Canonburg, PA 15317, USA.
Det Norske Veritas 2000. Design Standard OS F101 Offshore Pipelines.
FARO Technologies. 2009. FaroArm® Manual. Lake Mary, FL 32746, USA.
Forde W.R.B., Stiemer S.F. 1987. Improved Arc Length Orthogonality Methods for Nonlinear Finite
Element Analysis. Computers & Structures 27: 625–630.
Fraldi M., Guarracino F. 2011. An improved formulation for the assessment of the capacity load of cir-
cular rings and cylindrical shells under external pressure. Part 1. Analytical derivation. Thin-Walled
Structures 49: 1054–1061.
Fraldi M., Freeman R, Slater S, Walke AC, Guarracino F. 2011. An improved formulation for the
assessment of the capacity load of circular rings and cylindrical shells under external pressure. Part 2.
A comparative study with design codes prescriptions, experimental results and numerical simulations.
Thin-Walled Structures 49: 1062–1070.
Fraldi M., Guarracino F. 2013. Towards an accurate assessment of UOE pipes under external pres-
sure: Effects of geometric imperfection and material inhomogeneity. Thin-Walled Structures 63:
147–162.
Fraldi M., Guarracino F. 2014. Stability analysis of circular beams with mixed-mode imperfections
under uniform lateral pressure. Advances in Mechanical Engineering Article number 294507. doi:
10.1155/2014/294507.
Guarracino F. 2007. Considerations on the Numerical Analysis of Initial Post-Buckling Behaviour in
Plates and Beams. Thin-Walled Structures 45: 845–848.
Guarracino F., Walker A.C. 2008. Some Comments on the Numerical Analysis of Plates and
Thin-Walled Structures. Thin-Walled Structures 46: 975–980.
Guarracino, F., Fraldi, M., Giordano, A. 2008. Analysis of testing methods of pipelines for limit state
design. Applied Ocean Research 30: 297–304.
Guarracino, F., Walker, A.C., Giordano, A. 2009. Effects of Boundary Conditions on Testing of Pipes
and Finite Element Modelling. Int. J. Press. Ves. Piping 86: 196–206.
Guarracino, F. 2011. A Simple Formula for Complementing FE Analyses in the Estimation of the Effects
of Local Conditions in Circular Cylindrical Shells. CMES – Comp. Mod. Eng. Sci. 72: 167–184.
Guarracino, F. 2019. Remarks on the stability analysis of some thin-walled structures in the
elastic-plastic range. Thin-Walled Structures 138: 208-214.
Nash I., Carr P. 2013. The production and testing of meidp line-pipe for 3500m application. In Proc. 23rd
Int. Offshore Pipeline Technology Conference, Anchorage, Alaska, 14–23.
Roberts P.M., Walker A.C. 2010. Method and apparatus for pipe testing. United States Patent
no.20100212405.
Selker R., Ramos P.M.C., Liu P. 2014. Interpretation of the South Stream Ring Collapse Test Program
Results. In Proc. 24th Int. Ocean and Polar Engineering Conference, Busan, South Korea 2: 88–95.
Simulia. 2014. ABAQUS 6.1.3 User’s Manual. Dassault Systèmes, Providence, Rhode Island, USA.
Tata Steel 2011. Ring Collapse Testing Data. Internal Report. Tata Steel Tubes: Rotherham, UK.

459
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Ultimate bending resistance of pipes: Testing arrangements and


design approaches. A multi-year perspective

F. Guarracino
University of Naples “Federico II”, Naples, Italy

ABSTRACT: The design of submarine pipelines operating in very deep water or in complex
environments which can face geological hazards relies on accurate test results for the measure-
ment of maximum strain under bending loading. Over the years several test results have
shown apparently anomalous values of axial tensile and compressive strains in comparison to
the values that would be expected on the basis of simple theory. This fact has important con-
sequences for the efficacy of the design factors derived using experimental results. Examples
of anomalous test results are given and the cause of the differences between the strain values
obtained in the tests are explained on the basis of nonlinear effects. Conclusions are drawn
with respect to limit state for the design of onshore and offshore pipelines.

1 INTRODUCTION

Pipelines are vital components in the energy systems of all economically developed countries
and are designed to sustain the effects of a wide range of loading conditions resulting from
internal/external pressure and bending during installation and operations (Guarracino et al.,
1999). Much work has been devoted to evaluate the capacity of sustaining considerable exter-
nal pressures, as in the case of deepwater pipelines, a phenomenon which is affected by
a number of stability problems (Guarracino, 2019) and to this purpose a ring test procedure
has been recently developed for use in manufacturing procedure qualification tests (MPQT).
Pipeline design has traditionally been based on a limiting stress approach but since 1996
a limit state code has been developed by DNV (Det Norske Veritas, 2000). The use of the
limit state approach provides a more comprehensive basis for the calculation of the ultimate
conditions for pipes subjected simultaneously to pressure and bending loads. One aspect that
is incorporated in the code is the concept of ultimate states of pipe with ‘load-controlled’ and
‘displacement-controlled’ loading. In the former, the bending deformations are directly
affected by the variation of the applied moments. The ultimate state thus relates to
a maximum moment condition. In the ‘displacement controlled’ condition the deformations
are controlled, say by bending the pipe round a prescribed surface, and once the pipe is in
contact with the surface additional loading, of whatever kind, could not reduce the local
curvature of the pipe. This form of loading is associated with a minimum curvature, i.e.
a maximum local axial strain.
In both these cases, the ultimate state of the pipeline deformation or loading is calculated
using a model that describes the characteristic ultimate moment or strain related to the geom-
etry and material properties of the pipe. The characteristic values are then modified using
design factors that produce the design values of allowable ultimate moment or strain. The
models used to determine the characteristic levels of moment or strain are generally chosen as
those describing as closely as possible mean values of these variables measured during tests.
The design factors are calculated using statistical descriptions of the scatter of test results com-
pared to the mean values together with the statistical descriptions of the variables composing
the particular model, e.g. material strength, modulus etc. The calculation to determine the
design factors also include a ‘calibration’ stage in which the values of the design factors are

460
adjusted to ensure that the results from applying the code formulation and factors prescribe
levels of the ultimate conditions that conform to industry-wide acceptable levels for the prob-
ability of failure of the pipe.
In the process described above, it is generally assumed that the scatter of tests results from
minor and usually random variations in the variables included in the design model. In the case
of a pipe, these variations would generally relate to the differences in the geometries of the test
pipes from their corresponding nominal values, say pipe wall thickness or out-of-roundness,
and material properties. The design model is intended to provide a description of the test con-
ditions that account for systematic differences between one test and another.
In the present work the apparently anomalous results from tests on pipes, which have been
observed over the years are analysed and explained with the aid of finite element simulations.
The potential influence that these anomalies might have on the process of providing design
calculation guidance using the limit state method is highlighted.

2 TEST ARRANGEMENT AND EXPERIMENTAL RESULTS

By and large, testing a section of a circular cylindrical shell in purely bending loading is car-
ried out on the basis that the test specimen deforms according to simple bending beam theory,
accounting, where necessary, for the von Kármán/Brazier ovalistation effect (von Kármán,
1911; Brazier, 1927). Primarily this implies that while the material remains elastic the applica-
tion of purely bending moment will induce maximum tensile and compressive strains that are
identical in magnitude. A typical test rig for a medium diameter pipe, of about 700 mm diam-
eter, is shown in Figure 1. The test rig applies a four-point bending condition with the central
section of the test pipe assumed to be subjected to bending action only, with no, or at most
very little, shear or axial forces.
From the limit state point of view the two relevant conditions are the maximum moment
and corresponding strain, for load-controlled conditions of design, and the strain at which the
reduction of load-bearing capacity first occurs, which relates to displacement-controlled
design conditions. Following the attainment of that strain, as the loading is further applied,
the pipe develops a very local form of buckling which, under the action of external pressure,
can propagate catastrophically (Palmer et al., 1975).
Since the pipe is assumed to be an extremely simple structural element, and the simple beam
theory holds true, it has been common practice to assume that the axial strains have identical
values in tension and compression and that the strains can be calculated directly from the
curvature or the vertical displacements of the central section of the pipe. The ultimate strain
values from tests in which the pipe has been loaded to the point of local buckling have usually
been inferred from measurements of the deformations or through strain gauges attached to
the test pipe to measure axial strains directly.
Several years ago tests (Ellinas et al., 1985) were carried out on 152 mm diameter pipe to
determine the minimum curvature to which the pipe could be deformed prior to local buckling
occurring. An arrangement similar to that in Figure 1 was used, and strain gauges to measure
axial and circumferential strains were attached at intervals of 100mm apart along the central
test section. Very thick steel collars were used to protect the pipe at the supports and at the

Figure 1. Typical four-point bending test arrangement.

461
loading points. The collars were machined to fit very closely around the pipe to ensure no
localised loading was applied to the pipe wall. As a result, the pipe was fully prevented from
ovalising at all these points. In the design of the test rig it was assumed that a central test sec-
tion of about 5D would suffice to ensure end effects due to the loading conditions would
diminish to a negligible level along the major part of that section.
The results showed that the axial strains were fairly uniform along the length of the test
section but there were significant differences in the averaged values of the compressive strains
compared to the tensile strains.
At that time the evident anomaly between the measured strains with the expected values
vis-à-vis the simple bending theory was not followed up, and even after checking that the
strain gauges were correctly positioned and the instrumentation was functioning properly the
cause of the anomaly was not further investigated.
Few years later, proving tests were carried out on a section of 609 mm diameter pipes containing
a thin liner made from a corrosion resistant material (Walker et al., 2003). The purpose of the tests
was to determine accurately the level of strain to which the pipe could be bent before the liner
buckled locally. The test arrangement of Figure 1 had a loading arm 2 m long to create the
moment in the central section of the test pipe. The test section was arranged to be 3.5 D. The load
was applied to the test pipe using straight bars and loose yokes around part of the pipe circumfer-
ence. Saddles were used at the loading points and at the supports, see Figure 2.
In this case not only the pipe section was not prevented from ovalising but, on the contrary,
the localised actions at the loading points and at the supports were such to cause a local
degree of ovalisation. A number of axial strain gauges were attached along the top and
bottom centre lines of the pipe at intervals from the support points. The values of strain were
monitored as the load values were progressively increased. Figure 3 shows plots of the values
for the top and bottom gauges averaged along the test sections and plotted against the corres-
ponding value of applied load.
The investigation determined that the cause lay in the effect of the imposed ovalisation
applied by the saddles at the load points. This result pointed to a proposal for the modifica-
tion of the loading application in which the loads were applied, not through local stiffening of
the pipe wall or saddles, but by means of a shaft through the neutral axis of the pipe, as
shown in Figure 4.
The test pipe was fitted with strain gauges, as before, and also gauges to measure the ovality
of the pipe. The values of the axial strains measured by the gauges along the test section of the

Figure 2. Saddles at the supports and load application points.

462
Figure 3. Averaged strain values plotted against corresponding values of applied loading (Walker
et al., 2003): Maximum Ratio of Tensile to Compressive averaged strains = 1.28 (D = 609.6 mm, t =
18.9 mm, (D/t = 32), X65 material).

Figure 4. Test arrangement with modified support and load application points.

pipe were very uniform. As expected, with the modified loading and support arrangement, the
averaged measured values of compressive strains agreed very closely with the corresponding
values of the tensile strains, see Figure 5.
This said it can be expected that if the ovalisation is prevented, as it is in the case of
applied stiffeners of buckling arrestors, the compressive stress will exceed the value of ten-
sile ones under bending, whereas if some sort of localised ovalisation is imposed, the ten-
sile stress will exceed the value of compressive ones.

463
Figure 5. Results from modified bend test.

3 FINITE ELEMENTS ANALYSES

Following the observation of the apparent anomaly in the variation of the tensile and com-
pressive strains compared with the values expected on the basis of the simple theory of bend-
ing, analytical (Guarracino, 2003 and 2011) and numerical (Guarracino et al., 2008 and 2009)
modelling has been carried out to investigate the root cause of the anomaly.
Essentially, it was found that during the test of a short section of pipe, the practical loading
and support arrangements can result in boundary conditions that may impose some degree of
ovalisation at the point of load application or, alternatively, decrease the development of the
natural ovalisation. It has generally been assumed that such boundary effects would have
a minor consequence on the deformations of the test pipe and would persist for only a short
distance along the test length. However, the investigation has shown a hitherto unsuspected
mechanism in which the imposition or the prevention of the ovalisation at the loading point
of the test pipe will set up an axial strain system that is additional to the usual axial strain
caused by simple bending. The conjunction of the two strain systems thus causes a difference
between the values of the axial compressive and tensile values. The interaction between the
ovalisation and the axial strain effects can be investigated using finite element modelling.
Figure 6 shows the finite element simulation of a test arranged as in Figures 1 and 2 on
a 609.6 mm diameter pipe (t = 18.9 mm).
To this purpose the commercially available non-linear finite element code, ANSYS® v.11.0
was employed (ANSYS, 2007). The FE strain distribution along the mid-section of the pipe
for an applied load of 1MN is shown in Figure 7 for elastic conditions and is in good agree-
ment with the experimental results of Figure 5. It is evident that the tensile strains at the top
of the pipe are about 1.25 times the compressive strains at the bottom of the pipe.
As a matter of fact, the rigid loading yokes at the supports and at the loading points induce
some degree of ovalisation on the pipe and this results in an increased value of the tensile
strains at the top of the pipe, as confirmed by the finite element analysis, see Figure 7.
Figure 8 shows the values along the pipe axis of the longitudinal strains at the top, at
the bottom and at the side of the pipe, as yielded by the finite element analysis. It may be

464
Figure 6. Outline of the finite element modelling of a test arranged as in Figures 1 and 2.

Figure 7. Finite element simulation of a test arranged as in Figures 1 and 2: strain values (moduli) at
the mid-section of the pipe – Applied Load 1MN.

seen that the effects of the loading and of the constraints propagate in a quite complex
manner along the full length. It is also evident in the figure that at the mid-span of the
model, where the bending moment can be considered constant, the ratio of the tensile
axial strain to the corresponding compressive strain is about 1.25 times. Since generally
this is the section of pipe that is assumed to be free from the boundary support effects
and to have strain levels pertaining to simple bending theory, it can be seen that the ana-
lysis confirms the anomaly observed in the tests, see Figures 4 and 5, and also confirms
that the presence of the test loading conditions will affect the axial strain levels at which
local axial buckling will be initiated in a test pipe.

4 ANALYTICAL FORMULAE AND DESIGN CODES

In order to provide a simple tool to evaluate the effect of imposed or prevented ovalisation,
reference can be made to the classical Ritz’s approach (Ritz, 1909) and to a modified set of
Donnell’s strain and curvature changes (Donnell, 1933).

465
Figure 8. Finite element simulation of a test arranged as in Figures 1 and 3: strain values along the pipe
axis – Applied Load 1MN.

The advantage of such a procedure lies in the extreme simplicity of its final expression, which
can give a meaningful physical insight into the parameters which govern the problem at hand
and offer a way to account for these phenomena in the currently available design formulae.
The end result can be summarised in the following formula, which provides the top and
bottom mid-surface strain on account of the deformation induced by two opposite forces, F,
acting along the vertical diameter at the mid-span
   
64 1  ν2 F 2βx β2 x2 β4 x4
ε¼ e 1 þ ð1Þ
E rt 2 24

where
p ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
π2 4
ð1  ν2 Þ
β¼ pffiffiffiffiffiffiffiffi ð2Þ
4 r3 =t

t is the thickness of the pipe, r is its mean radius and E and ν stand for the Young’s modulus
and the Poisson’s ratio, respectively. x is the coordinate in the axial direction of the cylinder,
starting at point of application of the forces F.
For a pipe characterised by D = 609.6 mm, t = 18.9 mm and X65 material, as the one of
Figure 3, for an applied load of 1.118 MN the absolute value of the top and bottom strains
calculated according to the simple bending theory would be 2:3504  103 . Equation (2) yields
the additional strain at the mid-span on account of the local deformation induced by the con-
centrated loads, that is 3:5062  104 . By adding this latter quantity to the tensile strain and
subtracting it from the compressive one, it follows that the ratio of tensile to compressive

466
strains is equal to 1.35, with a difference from the experimentally measured ratio of about
5.5% (see Figure 3). This can be considered a quite satisfactory result, bearing however in
mind that in the actual testing arrangement the supports and the applied loads are not oppos-
ite, as it is the hypothesis leading to Equation (1), but distant 2 m apart.
Equation (1) can be used to account for the distrurbance induced by the experimental setup
in conjuction with the indications from most industry codes such as, for example, DNV-OS-
F101, that gives the characteristic compressive bending strain capacity, εM,c as

t  
σh 1:5
εM;c ¼ 0:78  0:01 1 þ 5 α αgw ð3Þ
D σy h

where D = 2r, σh is the characteristic hoop stress, equal to σh ¼ ΔpðD  tÞ=2t, Δp is the jump
in pressure, σy is the yield strength of material, αh is the maximum yield to tensile ratio and
αgw is girth weld factor, equal to the unity in case of no weld.

5 CONCLUSIONS

The present work has presented a review of previous experimental and numerical findings
together a simple analytical expression which can provide a straightforward evaluation of the
effects of tesing arrangements or of stiffeners and buckle arrestors on the bending of pipelines
at the installation stage or in complex environments facing geological hazards.
It has been shown that the seemingly anomalous values of measured axial strain in many
tests can be explained very straightforwardly and a simple formulation which offers a physical
insight into the mechanics of the problem and can turn useful both at the design and at the
assessment stage of offshore pipelines has been introduced and briefly discussed.

REFERENCES

Ansys, Inc. 2007. ANSYS 11.0 User’s Documentation. Canonburg, PA 15317, USA.
Brazier, L.G. 1927. On the flexure of thin cylindrical shells and other thin sections. Proc. Roy. Soc. A
116: 104–114.
Det Norske Veritas 2000. Design Standard OS F101 Offshore Pipelines.
Donnell, L.H. 1933. Stability of Thin-Walled Tubes under Torsion. NACA Rep. No. 479.
Ellinas, C.P., Walker, A.C., Langfield, G.N., Vines, M.J. 1985. A Development in the Reeling Method
for Laying Subsea Pipeline. In Proc. 1st Petroleum Tech. Australian Conf., Perth, Australia.
Guarracino, F., Mallardo, V. 1999. A refined analytical analysis of submerged pipelines in seabed laying.
Applied Ocean Research 21: 281–293.
Guarracino, F. 2003. On the analysis of cylindrical tubes under flexure: theoretical formulations, experi-
mental data and finite elements analysis. Thin-Walled Structures 41: 127–147.
Guarracino, F., Fraldi, M., Giordano, A. 2008. Analysis of testing methods of pipelines for limit state
design. Applied Ocean Research 30: 297–304.
Guarracino, F., Walker, A.C., Giordano, A. 2009. Effects of Boundary Conditions on Testing of Pipes
and Finite Element Modelling. Int. J. Press. Ves. Piping 86: 196–206.
Guarracino, F. 2011. A Simple Formula for Complementing FE Analyses in the Estimation of the Effects
of Local Conditions in Circular Cylindrical Shells. CMES – Comp. Mod. Eng. Sci. 72: 167–184.
Guarracino, F. 2019. Remarks on the stability analysis of some thin-walled structures in the
elastic-plastic range. Thin-Walled Structures 138: 208–214.
Palmer, A.C., Martin, J.H. 1975. Buckle propagation in submarine pipelines. Nature 254: 46–48.
Ritz, W. 1909. Über eine neue Methode zur Lösung gewisser Variationsprobleme der mathematischen
Physik. Journal für die Reine und Angewandte Mathematik 135: 1–61.
von Kármán, Th. 1911. Ueber die Formänderung dünnwandinger Rohre, insbesondere federnder Ausgle-
ichrohre. Zeitschrift des Vereines deutscher Ingenieur 45: 1889–1895.
Walker, A.C., Holt, A., Guarracino, F., Wilmot, D. 2003. Test Procedure for Pipe and Pipeline Material.
In Proc. Offshore Pipeline Technology Conference, Amsterdam, Holland.

467
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

On the GNI analysis of simple thin-walled beams with using linear


buckling mode as geometric imperfection

M.Z. Haffar, M.H. Taher & S. Ádány


Budapest University of Technology and Economics, Budapest, Hungary

ABSTRACT: In this paper lateral-torsional buckling of thin-walled beams is discussed. Analyt-


ical and numerical calculations are compared, including calculations for the critical moment, for
the buckling shape, and for the geometrically nonlinear analysis with initial geometric imperfec-
tions. Though the analysed beams are simple, with simple supports and uniform moment along
the beam length, various cross-sections are considered and both principal-axis and non-principal-
axis bending cases are analysed. The results demonstrate how the cross-section asymmetry and the
deviation of the bending from the principal axis influence the behavior.

1 INTRODUCTION

When a beam is subjected to non-principal-axis bending, the traditional design approach is to


decompose the moment into principal-axis components and then to superpose the behavior from
the results of the principal-axis cases. This approach is reflected in current design standards, such
as the Eurocode 3, see CEN (2005). However, there is a trend nowadays to build more and more
complete (finite element) models, to apply realistic loads, to (even if approximately but) directly
consider imperfections, and to perform nonlinear analysis in order to get the load bearing capacity
of the structure. In nonlinear analyses of complex models the decomposition into principal-axis
bending components would be rather artificial, and cannot reasonably be completed in many
cases. Thus, it is not enough to consider doubly-symmetrical cross-sections subjected to principal-
axis bending only, but it is important to understand LTB in more general situations, including
beams with mono-symmetric or asymmetric cross-sections.
In this paper elastic behavior is studied. First, the linear buckling problem is discussed, via
the solution of the corresponding generalized eigen-value problem, which leads to critical
moment values and buckled shapes. Then, geometrically nonlinear analyses are completed
with an initial geometric imperfection, which type of analysis is widely referred as GNI ana-
lysis. For the initial imperfection the first lateral-torsional buckling mode is employed.
In this paper numerical studies are shown: the results of analytical formulae and shell finite
element calculations are compared. The analytical formulae for the linear buckling problem are
taken from Glauz (2017), while for the GNI analyses the formulae are derived by the authors,
based on the work of Szalai & Papp (2010) and Agüero et al. (2015). The finite element calcula-
tions are performed by the so-called constrained finite element method (cFEM), which is essen-
tially a shell finite element calculation, but the method is developed so that modal decomposition
would be possible. This means that separation of the behavior modes is possible, by applying con-
straints. When a member is constrained into a deformation mode (e.g., to global mode), it is
enforced to deform in accordance with some mechanical criteria, characterizing for the intended
deformation (e.g., global) mode. For more details, see (Ádány 2017 and Ádány et al. (2017). The
great advantage of using cFEM is that it is easy to get lateral-torsional buckling separately from
other types of buckling (e.g. local-plate buckling or distortional buckling), even if the beam is
short and/or the member is thin-walled, the analysis of which cases is rather challenging (if possible
at all) by using a general purpose shell finite element method. The results prove the applicability of
cFEM, as well as demonstrate how the cross-section asymmetry and the deviation of the bending

468
from the principal axis influence the nonlinear behavior of beams subjected to lateral-torsional
buckling.

2 CRITICAL MOMENT

Examples for lateral-torsional buckling are shown and discussed, including critical moments,
buckled shapes, and the results of GNI analyses. Since we want to compare analytical and
numerical results, it is essential to have basic beam arrangement, namely: simply supports
(also known as fork supports) at both beam ends, and uniform moment along the length.
However, various cross-sections are considered, as shown in Figure 1, including doubly and
mono-symmetrical ones. Also, principal-axis and non-principal-axis bending cases are both
considered. In all the examples steel-like material properties are used with E = 210000 MPa
and ν = 0.3.
In this Section, first, critical moment results are presented. The analytical formulae are
taken from Glauz (2017). The cFEM calculations are completed in various options. An
important option is whether all the regular FE deformations are allowed or only certain (e.g.,
global) deformations. Another option is the consideration of the second-order effects: either
all the second-order strain terms from the Green-Lagrange strain tensor are considered, or
only a few terms. In this latter case, more specifically, only those terms are considered which
are work-compatible with the longitudinal normal stresses, and out of the 3 such terms only
(∂v=∂x)2 and (∂w=∂x)2 terms are considered and the (∂u=∂x)2 term is neglected, where u, v and
w are the translational displacements along the local x, y, and z-axis, respectively. (For the
interpretation of local axes, see e.g., Ádány, 2018).
Based on the above options, the following cFEM results will be presented here:
• cFEM #1: regular FE analysis, with all the terms of the Green-Lagrange strain tensor;
• cFEM #2: the deformations are constrained into the global modes (i.e., global modes are
allowed only), with all the terms of the Green-Lagrange strain tensor;
• cFEM #3: the deformations are constrained into the global and transverse extension modes
(i.e., global modes and transverse extension modes are allowed only), with all the terms of
the Green-Lagrange strain tensor;
• cFEM #4: the deformations are constrained into the global and transverse extension modes
(i.e., global modes and transverse extension modes are allowed only), with only two terms
of the Green-Lagrange strain tensor (as defined above).
Selected results of the comparison are summarized in Table 1. The results demonstrate that
without the constraining it might be demanding to identify the buckling mode that can reason-
ably be identified as LTB, (e.g., in some of the considered cases such mode has not been found
among the first 200 buckling modes), and even if such mode is identified, the corresponding
critical moment value might considerably deviate from the analytical solution. The results also
demonstrate that consideration of global modes lead to slightly (but systematically) increased

Figure 1. Cross-sections, coordinates, basic notations.

469
Table 1. Critical moments for selected cases in various options.
length αM Analytic cFEM #1 cFEM #2 cFEM #3 cFEM #4

cross-section mm degree kNm kNm kNm kNm kNm

1 (b = 50) 800 0 84.62 82.96 92.24 85.51 85.71


1 (b = 50) 2000 30 26.61 26.62 28.44 26.88 26.88
1 (b = 50) 5000 60 16.34 16.37 17.31 16.48 16.47
1 (b = 100) 800 90 2169 1578 2814 1821 1665
1 (b = 100) 2000 120 208.5 201.5 225.6 207.7 207.1
1 (b = 100) 5000 150 37.83 37.71 40.02 37.88 37.89
2 800 180 125.6 103.9 135.7 128.0 128.3
2 2000 210 64.26 58.45 66.36 64.67 64.69
2 5000 240 61.85 60.99 63.67 61.91 61.93
3 800 270 807.9 - 841.1 755.0 795.3
3 2000 300 62.61 - 68.43 62.34 62.63
3 5000 330 7.11 6.54 7.78 7.12 7.12

critical values due to the Poisson effect, as discussed e.g., in Ádány (2012). The best cFEM
option is #4, the results of which are practically identical to those of the analytical solutions,
with the only exception when the critical moment value is extremely high. Therefore, through-
out this paper cFEM #4 will be used.
Figure 2 demonstrates how the critical moment value is dependent on the inclination of the
moment vector. The plot illustrates that the critical moment can readily be calculated to any αM
angle and to any cross-section, and the critical value is (typically) highly dependent on the cross-
section shape.

3 BUCKLED SHAPE

In this Section the buckled shapes are briefly discussed. In Figure 3 sample buckled shapes are
shown. Though the 3D plots are qualitatively meaningful, it is better to describe them quanti-
tatively. Therefore, we introduce here the direction of the secondary displacements, d2, which
will be calculated at mid-length of the beam, and interpreted as shown in Figure 4. Similarly,
direction of the primary displacements, d1, is introduced, interpreted also in the figure. The
angle of the primary and secondary displacements will be referred by α1 and α2, respectively.

Figure 2. Critical moment in the function of αM.

470
Figure 3. LTB buckled shape samples.

Figure 4. Interpretation of the direction of primary and secondary displacements.

The direction of the primary displacement (which, in fact, is perpendicular to the neutral
axis) is well-known from university courses on strength of materials. The direction of the sec-
ondary displacements can analytically be derived. For this reason, we have made derivation,
considering arbitrary (i.e., symmetric or asymmetric) cross-section and arbitrary moment dir-
ection. It can be proved that the angles of the primary and secondary displacements satisfy the
following equations:
 π Iu
tan /1  /u  ¼ tanð/M  /u Þ ð1Þ
2 Iv
Iv
tanð/2  /u Þ ¼ tanð/M  /u Þ ð2Þ
Iu

where αu is the angle of the first principal axis to the Y-axis, and Iu and Iv are the principal
moment of inertias to the major and minor axis, respectively. It is to note that in all the here
presented examples the principal axes coincide with the Y and Z axes.
Figure 5 illustrates how the primary and secondary displacements are dependent on the
cross-section and on the αM angle. The presented results are obtained by analytical formulae,
however, they can also be determined from the finite element results. Results show
a practically perfect coincidence between the cFEM and analytical results. Figure 5 demon-
strates that the primary and secondary displacements are typically not perpendicular to each
other. It can be proved analytically that the primary and secondary displacements are perpen-
dicular only if the bending is principal-axis bending.

471
Figure 5. Relation of primary and secondary displacements in the function of αM.

4 GNI ANALYSIS

In this Section The results of GNI analyses are presented. GNI analysis means that first the
linear buckling problem is solved and buckling shapes are determined, then a selected buckling
shape (with some scaling) is employed as initial geometric imperfection and geometrically non-
linear (but elastic) analysis is conducted. Since the aim here is to study LTB, the first lateral-
torsional buckling mode is applied.
Analytical solution for the problem is available. For example, in Szalai & Papp (2010) ana-
lytical solution for LTB GNI is presented in the case of doubly-symmetrical cross-sections. In
Agüero et al. (2015) a more general GNI solution is presented. We have also completed ana-
lytical derivation for LTB, by considering arbitrary cross-section geometry and principal-axis
or non-principal-axis bending. All of these derivations are based on the same classic set of
mechanical assumptions, which are well-known from the background of the Ayrton-Perry for-
mula. All of these derivations lead to the same conclusion: in the course of GNI analysis the
displacements remain similar to the initial imperfect geometry, but the displacements are mag-
nified. Mathematically, the d displacement vector in a nonlinear analysis can be obtained by:

d ¼ Adini ð3Þ

where dini is the displacement vector of the initial imperfect geometry, and A is the amplifica-
tion factor, which, in the case of LTB, can be calculated as:
. M

A¼1 1 ð4Þ
Mcr

where M is the actual moment value, and Mcr is the corresponding critical moment value.
It is to underline that this analytical result predicts a symmetric bifurcation, since the ampli-
fication is not dependent on the direction of the initial geometry. Moreover, since the initial
shape is similar to the (first) buckling shape, this analytical solution suggests that the primary
and secondary displacement are perpendicular if the bending is principal-axis bending, while
they are not perpendicular if the bending is non-principal-axis bending.
GNI analysis can be completed by cFEM, too. In the current cFEM implementation the
initial geometric imperfection is realized by an equivalent initial loading. The applied non-
linear solution technique is a fairly simple (but rather robust) incremental procedure. Since
the solution algorithm is controlled by the loads, only the ascending branch of the load-
displacement curve can be obtained.
As the results will show, sometimes the numerical solutions are very similar to the analytical
prediction, but in other cases the numerical results are very different. The differences are the conse-
quences of the different mechanical assumptions, namely different assumptions on (i) the displace-
ments, and (ii) on how the stiffness modifying effect of the deformations/stresses are considered.

472
In the analytical solutions the primary displacements/deformations are intentionally disre-
garded, and the second-order displacements are assumed in a pre-selected form (e.g., in the basic
case of LTB, sinusoidal displacement distributions are assumed in the longitudinal direction of the
beam) The assumed displacements, hence, similar to the initial imperfect shape, therefore, the
result that the second-order displacements are similar to the initial geometry is more like an initial
assumption than a result. Moreover, according to the classic FE terminology, the stiffness-
modifying effects are included in the geometric stiffness matrix. In classic analytical derivations it
is assumed that the geometric stiffness matrix is linearly dependent on the load intensity. On the
other hand, when the finite element method is applied to perform the geometrically nonlinear ana-
lysis, there is no pre-assumed displaced shape, the primary and secondary displacements/deform-
ations are not separated, and the geometric stiffness matrix is re-calculated multiple times as the
load intensity is increasing, and the re-calculated geometric stiffness matrix is dependent on the
actual deformations/stresses, thus, usually it is not linearly dependent on the load intensity.
As the following examples demonstrate, the differences between the background assump-
tions of the analytical and numerical solutions are more or less visible in the results, depending
on the cross-section topology and dimensions, as well as on the loading.
In all the here presented examples the member is 2000 mm long, the initial imperfection is
L/1000, hence 2 mm at the middle of the beam. (Note, the 2 mm is the total transverse transla-
tion of the beam.) Both positive and negative imperfection are considered. It is to note that
the results are dependent on the beam length, sometimes not only quantitatively but also
qualitatively. This aspect, however, is not discussed in this paper.
First, the simplest problem is considered: the cross-section is doubly-symmetric (CS #1), the
bending is principal-axis bending and around the strong axis of the cross-section. In Figure 6 load-
displacement diagrams are shown for 3 flange width values, 50 mm, 100 mm, and 200 mm. The
load is non-dimensional in the figure, the actual bending moment is normalized by the correspond-
ing critical moment value. Numerical (cFEM) and analytical results are shown: the numerical
results are different for the 3 cross-sections, while the analytical one is independent.
As Figure 6 shows, the numerical and analytical results are practically identical if the flange
is narrow (i.e., 50 mm), but as the flange width is getting larger, the numerical results more
and more deviate from the analytical ones. In this case this deviation means that the finite
element GNI analysis predicts somewhat smaller displacements.
In Figure 7 results for CS #2 are shown. The cross-section is mono-symmetric, and 3 cases of
principal-axis bending are considered. It is obvious from the figure that the numerical and analyt-
ical results are considerably different, but the numerical results themselves are rather different,
too. It can be observed that if the bending takes place in the plane of symmetry, i.e., α M = 0
or αM = 180 (that is, the upper or lower flange is in compression, respectively), the load-
displacement curves are practically symmetrical, i.e., the bifurcation seems to be symmetrical
one. If the wider (upper) flange is in compression, the maximum load is nearly identical to

Figure 6. Load-displacement curves from GNI analysis, CS#1.

473
Figure 7. Load-displacement curves from GNI analysis, CS#2.

the critical load, however, if the narrower (lower) flange is in compression, a considerable
post-buckling reserve exists.
However, if the bending plane is perpendicular to the plane of the symmetry (αM = 90), the
behavior is clearly asymmetric, i.e., the bifurcation seems to be asymmetric, even if the bend-
ing is principal-axis bending. In the actual problem: if the initial imperfection is positive, the
maximum load is considerably smaller than the critical load, while if the initial imperfection is
negative, the maximum load is slightly but clearly larger than the critical load. In this case the
secondary displacements are hardly changing during the loading.
In Figure 8 the results of another mono-symmetrical cross-section (CS #3) are presented. The
obvious difference between CS #2 and CS #3 is the position of the shear center: in CS #2 it is
located in the web (relatively close to the centroid), in CS #3 it is located outside the web (rela-
tively far from the centroid). The non-linear load-displacement curves are partly similar, partly
different from those of CS #2. The results suggest that even if the bending is in the plane of the
symmetry, the behavior might be symmetric or asymmetric. There are some common features
of CS #2 and CS #3: (i) in some cases there is considerable post-critical reserve, (ii) the behav-
ior is asymmetric if the bending plane is perpendicular to the plane of the symmetry.
Finally, Figure 9 shows some results for non-principal-axis bending. The cross-section is
a common doubly-symmetrical cross-section (CS #1, with b = 100 mm), and the bending just
slightly deviates from the major-axis bending. The figure proves that the behavior is always asym-
metric. It is also shown that the behavior is gradually (though rapidly) changing as αM is
increasing.

Figure 8. Load-displacement curves from GNI analysis, CS#3.

474
Figure 9. Load-displacement curves from GNI analysis, CS#1, non-principal-axis bending.

5 SUMMARY

In this paper lateral-torsional buckling of thin-walled beams is discussed. Numerical examples


are shown for the calculation of the critical moment and buckling shape, as well for the solu-
tion geometrically nonlinear analysis with using the buckled shape as initial geometric imper-
fections. The examples illustrate some important features of the behavior. The most
important findings are as follows.
The classic analytical solution for the geometrically imperfect second-order analysis gives
reasonably precise results only in some very special cases, but cannot be applied in general.
The results suggest that some post-buckling reserve might exist in the case of lateral-torsional
buckling (at least if the material is assumed to be perfectly elastic). The nonlinear behavior
might be symmetric (i.e., independent of the direction of the initial geometric imperfection),
but only in relatively special cases. In general, the nonlinear behavior is asymmetric. Finally,
the results well demonstrate that the constraint finite element method is rather beneficial to
apply in the analysis of lateral-torsional buckling of beams, since it can handle virtually any
cross-section, any support and loading conditions, still, it directly provides the lateral-
torsional buckling solution separately from local or distortional buckling.

REFERENCES

Ádány, S. 2012. Global Buckling of Thin-Walled Columns: Analytical Solutions based on Shell Model,
Thin-Walled Structures, Vol 55, pp 64–75.
Ádány, S. 2018. Constrained shell Finite Element Method for thin-walled members, Part1: constraints
for a single band of finite elements, Thin-Walled Structures, Vol 128, July 2018, pp. 43–55.
Ádány, S., Visy, D. & Nagy, R. 2018. Constrained shell Finite Element Method, Part 2: application to
linear buckling analysis of thin-walled members, Thin-Walled Structures, Vol 128, July 2018, pp.
56–70.
Agüero, A, Pallares, F.J. & Pallares, L. 2015. Equivalent geometric imperfection definition in steel struc-
tures sensitive to lateral torsional buckling due to bending moment, Engineering Structures, Vol 96,
pp. 41–55.
CEN. 2005. EN 1993-1-1:2006 - Eurocode 3: Design of steel structures - Part 1-1: General rules and rules
for buildings, Brussels, Belgium.
Glauz, R.S. 2017. Elastic lateral-torsional buckling of general cold-formed steel beams under uniform
moment, Thin-Walled Structures, Vol 119, pp. 586–592.
Szalai, J. & Papp. F. 2010. On the theoretical background of the generalization of Ayrton-Perry type
resistance formulas, Journal of Constructional Steel Research, Vol 66, pp. 670–679.

475
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

U-shaped steel plate dissipative connection for concentrically


braced frames

J. Henriques
Faculty of Engineering Technology, CERG, University Hasselt, Belgium

L. Calado
Instituto Superior Técnico, Universidade de Lisboa, Lisboa, Portugal

C.A. Castiglioni
Department of Architecture, Built environment and Construction Engineering, Politecnico di Milano,
Milano, Italy

H. Degée
Faculty of Engineering Technology, CERG, University Hasselt, Belgium

ABSTRACT: In concentrically braced frames, the use of dissipative connections allows to effi-
ciently dissipate the seismic energy in earthquake scenarios and subsequently, to reduce the costs
in the rehabilitation of the structure. To this end, U-shaped steel plates are simple and efficient
connection components where significant dissipation of seismic energy can take place through the
inelastic flexural deformation of the plate. This paper presents experimental results on the isolated
U-shaped steel plate connections and on single-story concentrically braced frame (real scale)
including the U-shaped steel plate to connect braces to adjacent members. The executed tests con-
sidered both monotonic and cyclic loading. The results highlight the efficiency of the U-shape steel
plate to dissipate the energy input through inelastic deformations. On the other hand, the cyclic
tests show potential fatigue behavior, as the deformation capacity is significantly reduced with
repeated loading and increasing stress amplitude, requiring thus specific attention in practical
design situations.

1 INTRODUCTION

Concentric braced frames (CBF) are commonly used in Europe as seismic resistant steel struc-
tures. Strong seismic activity has shown (Bertero V et al., 1994) significant damage in the
building structure, namely in the braces usually designed as dissipative members. Subse-
quently, the repair of the structure involves the strengthening or replacement of damaged and
buckled braces which requires considerable skill, material and labor cost (Morelli F et al.,
2017). The most suitable solution to this problem consists in the protection of the braces
through the use of anti-seismic devices, as dissipative connections.
In CBF with dissipative connections, the frame presents sufficient stiffness against lateral
displacements and incorporates flexible connections between braces and structure. These con-
nections provide ductility to the structure, increasing the dissipation of seismic energy, and
protect the braces from buckling. Moreover, as damage is concentrated in the connections,
the repair after strong seismic event becomes easier and less costly.
In the last decade, several seismic protection systems with innovative steel-based devices have
been developed (Vayas I et al, 2017). Within the research project INERD (Plumier, A. et al, 2004)
a dissipative connection consisting of a steel plate bent to a U shape was developed for applica-
tion in CBF, connecting the braces to the adjacent members (Figure 1). The main geometric

476
Figure 1. U-shape steel plate connection and its application.

variables of the U-shaped steel plate are: R – internal bend radius; e – thickness of the plate; α –
inner angle of the U-shaped (defined by the angle between brace and adjacent member); width of
the plate (out-of-plane dimension of the scheme in Figure 1-a); length of the straight parts;
number of bolts to perform the connection between plate and member. In service, the loading of
the U-shape steel plate connection varies according to the connection configuration. Accordingly,
two loading positions are possible (Figure 1-b): perpendicular and parallel. In fact the loading
position depends on the configuration of the connection between device and member which may
consist of single or double overlap connection or end-plate connection (see Figure 1-b).
In the present paper the results of the two experimental programmes on the U-shaped plate
connection, realized within the referred research project (Plumier, A. et al, 2004), are discussed.
One experimental campaign had focus on the mechanical characterization of the U-shaped plate
connection under monotonic and cyclic loading. These tests only considered tests on the isolated
connection. The second experimental programme considered the testing of full-scale CBFs incorp-
orating the referred connection. In the latter, only cyclic tests were performed.

2 EXPERIMENTAL TESTS ON ISOLATED U-SHAPED STEEL PLATE


CONNECTIONS

2.1 Test specimens, testing layout and testing programme


In order to experimentally assess the mechanical performance of the U-shaped steel plate connec-
tion, tests on the isolated connection device (Figure 2), at IST in Lisbon, Portugal, were per-
formed. The U-shaped steel plate was design to be the governing component and consequently,
the connected parts of the testing system were overdesigned. The tests considered the following
types of loading: i) static monotonic pull; ii) static monotonic push; iii) cyclic loading following the
ECCS cyclic loading protocol (ECCS, 1986); iv) cyclic loading with constant amplitude, where
four different displacement amplitudes were tested. In total, for each test specimen, 7 tests were
executed. In Table 1 is summarized the described testing programme. The testing variables con-
sisted in the following: i) internal bend radius (R); ii) plate thickness (e); iii) angle between con-
nected members (α) defining the angle between the U-shaped steel plate “legs”; iv) the loading
position. The plate width was equal for all test specimens (b = 160 mm).

2.2 Results
From the experimental tests on the isolated U-shaped steel plate, force deformation curves were
obtained as illustrated in Figure 3-a). The chart includes the results of both static loading tests and
of the cyclic loading test, following the ECCS cyclic loading protocol (ECCS, 1986), for test speci-
men Test 6. In Figure 3-b) are compared, for the same test specimen, the absorbed energy and the
total number of cycles achieved in each cyclic loading test. These results highlight the following:
i) the connection strength and stiffness is higher when the U-shaped steel plate is “compressed”
between the connected members (Mon Push vs Mon Pull); ii) under cyclic loading, the

477
Figure 2. Experimental tests on isolated U-shaped connection (Plumier, A. et al, 2004).

Table 1. Summary of the experimental programme on isolated U-Shaped steel plate connection.
Test
Specimen Load
ID R [mm] e [mm] α [º] Position Nº of Tests

Test 1 100 25 45 For all test specimens 2 tests with monotonic load-
Test 2 100 25 50 ing (1 Pull + 1 Push) and 5 tests with cyclic load-
Test 3 100 30 50 ing (1 ECCS,
Test 4 125 30 50 Δ = 20 mm,
Test 5 125 25 30 Δ = 40 mm,
Test 6 125 25 45 Δ = 60 mm,
Δ = 80 mm)
Test 7 125 25 50
Test 8 125 25 30
Test 9 125 25 39
Test 10 125 25 45
Test 11 125 30 39

deformation capacity of the connection is reduced; iii) under cyclic loading, the connection is sens-
ible to fatigue, with the increase of the cycles amplitude, the absorbed energy and the number of
completed load cycles decreases. These observations can be generalized for the other test speci-
mens. The detailed results and conclusions on the full experimental programme can be accessed in
(Plumier, A. et al, 2004) and (Henriques, J. et al, 2018).

3 EXPERIMENTAL TESTS ON CBF INCLUDING THE U-SHAPED STEEL PLATE


CONNECTION

3.1 Test specimens, testing layout and testing programme


At Laboratorio di Prove e Materiali of Politecnico di Milano, Italy, tests on a single story CBF
(Figure 4), incorporating the U-shaped steel plate to perform the connection between the diagonal
braces and the adjacent members, in this case the columns, were executed. The U-shaped steel
plate was design to be the governing component and consequently, connected devices and frame
members were overdesign. In particular, the braces were designed as continuous and considering
half-length of the diagonal in the computation of the buckling length (Constanzo S et al., 2019).
The tests on the single story CBF consider only the cyclic loading protocol. A summary of the
testing programme is given in Table 2. As for the tests on the isolated U-shaped steel plate, the
testing variables are the bend radius (R), the plate thickness (e) and the angle between the con-
nected members (α).

478
Figure 3. Results of the experimental tests on isolated U-Shaped steel connection.

Figure 4. Experimental tests on CBFs incorporating U-Shaped steel plate connection.

Table 2. Summary of the experimental programme on CBFs incorporating U-Shaped plate connection.
Test
α Load
Specimen R [mm] e [mm] Nº of Tests
[º] Position
ID

CBF 1 125 25 50 1 test using a cyclic quasi-static loading according to the


CBF 2 100 25 50 ECCS cyclic loading protocol (ECCS, 1986)
CBF 3 125 30 50
CBF 4 100 30 50
CBF 5 100 30 40
CBF 6 100 25 40
CBF 7 125 30 40
CBF 8 125 25 40

3.2 Results
The tests on CBFs allowed to obtain the force-interstory drift curves (cyclic loops) reflecting the
behavior of the frame using the U-Shaped steel plate to connect bracings to adjacent members, in
this case the columns. In Figure 5-a) are illustrated the results of two tests, namely CBF 2 and
CBF 6. These test specimens are similar being the main difference the loading position of the
U-shaped steel plate, and, because of limitation on the geometric dimensions of the frame due test
layout, the angle between connected members (50º and 40º, respectively). The curve shows the
global response is symmetric contrary to the behavior of the U-shaped steel plate connection (see
Figure 3-a) under push or pull loading. This was expected has the CBF is symmetric, and there-
fore, under service, two connections work under push loading and the other two work under pull-
loading. The curves show a very similar behavior, with a slightly stiffer response on the CBF
frame with perpendicular loading on the U-shaped steel plate connection (CBF6). In Figure 5-b),

479
Figure 5. Results of the experimental tests on CBFs incorporating U-Shaped steel plate connection
(CBF2 vs CBF6).

for the same test specimens, are shown the cumulative energy absorbed by the CBF and only by
the 4 U-shaped steel plate connections. Though not all tests on the CBFs could be taken up to
failure, due to the limitations on the experimental testing set-up, the referred test specimens
achieved the same number of cycles (25). Hence, comparing the cumulative absorbed energy one
can observe that the specimens with perpendicular loading absorb more energy. This is because of
the higher strength and stiffness of the connection under these loading conditions. These results
are in line with the results on isolated connections (Henriques, J. et al, 2018). In the same chart is
also included the percentage of energy absorbed by these connections with respect to the total
energy absorbed by the CBF. Here a difference is noticed. In CBF2, the referred connections
absorbed approximately 82% of the total energy, while in CBF6 approximately 90% of the energy
was absorbed by the connections. In any case, it is clear that the connections accomplish the fore-
seen function, and work as dissipative component of the CBFs.

4 ANALYSIS OF RESULTS

4.1 Static vs cyclic tests


In order to compare the performance of the U-shaped steel plate connection under monotonic
and cyclic loading, the ratio between the deformation capacity obtained in both types of tests
were computed. For the cyclic loading only the tests following the ECCS cyclic loading protocol
was considered. In Figure 6 are plotted the computed ratios. For each test specimens, two ratios
were determined, as the deformation capacity under push loading (compression) and pull loading
(Tension) were, for some cases, remarkably different. It is clear that under cyclic loading the
deformation capacity is significantly reduce. The average reduction is 50%. This indicates that
under such load conditions, the connection is susceptible to experience low cyclic fatigue effect.

4.2 Isolated vs CBF tests


In order to compare the behavior of the U-shaped steel plate on the isolated tests and when
incorporated in the CBF, an extrapolation of the frame behavior was performed using the results
of the experimental tests on the isolated connection. To accomplish this task, an mechanical
model was proposed to reproduce the CBF behavior. The model is described in (Henriques,
J. et al, 2018). In Figure 7-a) are compared the force-lateral deformation curves of the CBF frame
for both cases. It is clear that there is a stiffer response on the tests of the CBF in comparison to
the extrapolated behavior. The difference is justified by the fact that in the referred mechanical
model, the connections between columns and beam are assumed perfectly hinged and therefore
neglecting the “frame effect”. The latter has been identified and quantified in (Kanyilmaz A et al.,
2018) comparing the cyclic performance of two CBFs with two different construction detailing: i)
one frame constructed using actual (common) connections between the structural members (such
as fin plates and gusset plates); ii) other frame constructed using perfectly hinged connection
between members. The frame effect depends on the connection detailing (see Figure 4-b) therefore
within the experimental programme on CBFs (Plumier, A. et al, 2004), an experimental test was

480
Figure 6. Comparison between the maximum deformation of the U-Shape steel plate connection under
cyclic (ECCS cyclic loading protocol) and monotonic loading.

Figure 7. U-Shape steel plate response vs CBF response.

performed on a frame without any bracing allowing to determine the frame stiffness. The frame
stiffness was then introduced in the above referred mechanical model, and the response of CBF
extrapolated. In Figure 7-b) are given the results including the “frame effect”. The agreement
between tests and extrapolation is now excellent. Furthermore, comparing both charts, it is pos-
sible to observed that the U-shaped connection is the main source of deformation within the
CBF, and therefore the dissipative component.

5 FATIGUE BEHAVIOR

The comparison between deformation capacity of the connection within the monotonic and cyclic
loading tests highlighted the susceptibility of the U-shaped steel plate connection to low fatigue.
Hence, using the test results fatigue design curves were derived and proposed following the
S-N line approach prescribed in the EN 1993-1-9 (CEN, 2005). The detail of the used approach is
given in (Henriques, J. et al, 2018). The proposed equations, best fit and design curve, are given in
(1) and (2), respectively. In Figure 8 are compared the tests results with the proposed S-N lines.
The results show a good agreement and highlight the susceptibility of the U-Shape steel plate con-
nection to low cycle fatigue.
Log N ¼ 12:00  3 Log S ð1Þ

Log N ¼ 11:53  3 Log S ð2Þ

Figure 8. Best fit and low cycle fatigue design curves.

481
6 CONCLUSIONS

In this paper a dissipative connection, to be used between bracing and adjacent members in CBFs,
has been presented. The connection consists of a steel plate bend in a U-shape, bolted to the con-
necting members, and is design to dissipate energy through inelastic deformation. The proposed
connection is very practical given the simplicity to execute. In the paper, the mechanical behavior
of the referred connection was investigated by means of experimental tests realized within the
scope of the research project INERD. The experimental programme considered tests on the iso-
lated connection under monotonic and cyclic loading, and test on full-scale CBF incorporating the
connection. Besides the type of loading, several geometric parameters were varied within both
experimental programmes, namely: i) the load direction with respect to the U-plate (parallel or
perpendicular, in tension or compression) ii) the load type, i.e. monotonic or cyclic, iii) the thick-
ness of the U-shaped steel plate, iv) the U-shaped steel plate bent radius and v) the angle formed
by the U-shaped steel plate “legs”(which is directly related to the global geometry of the frame).
The main outcome of tests results can be summarized as follows:
■ The connection can achieve high deformation capacity and therefore successfully dissipate
significant seismic energy.
■ Under cyclic loading, the connection demonstrates to be susceptible to low cyclic fatigue
“effect”.
Subsequently, S-N line curves for the fatigue design of the connection have been proposed in
line the Eurocode approach. The comparison of the proposed equations with experimental
results demonstrated their accuracy.

ACKNOWLEDGEMENTS

The research leading to these results has received funding from the European Union’s Research
Fund for Coal and Steel (RFCS) research programme under CEC agreement N° 7210-PR-316
(The INERD project) and the grant agreement N° 709343 (INNOSEIS).

REFERENCES

Bertero V et al. (1994). Performance of steel building structures during the Northridge earthquake. Univer-
sity of California at Berkeley.
CEN. (2005). EN 1993-1-9: Eurocode 3 – Design of steel structures - Part 1-9: Fatigue. Brussels, Belgium:
European Committee for Standardization.
Constanzo S et al. (2019). Proposal of design rules for ductile X-CBFS in the framework of Eurocode 8.
Earthquake Engineering and Structural Dynamics, 48 (1): 124–151.
ECCS. (1986). Recommended testing procedure for assessing the behaviour of structural steel elements
under cyclic loads. Brussels: ECCS – Technical Committee 1 – Structural Safety and Loadings – Tech-
nical Working Group 1.3 – Seismic Design.
Henriques, J. et al. (2018). Dissipative connections with U-shaped steel plate for breaces of concentrically
braces frames. Bulleting of Earthquake Engineering, Submitted.
Kanyilmaz A et al. (2018). An adjusted design approach for concentrically braced frames in low-to-
moderate seismicity areas. Bulletin of Earthquake Engineering, 16:4159–4189.
Morelli F et al. (2017). Seismic behaviour of an industrial steel structure retrofitted with self-centering
hysteretic dampers. Journal of Constructional Steel Research, 157–175.
Plumier, A. et al. (2004). TThe INERD project, Two innovations for Earthquake resistant design. Brussels:
European Commission.
Vayas I et al. (2017). Innovative Anti-seismic Plates and systems. Ioannis Vayas (Ed.), Research Fund for
Coal and Steel, INNOSEIS Project RFCS-02-2015, ECCS press.

482
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Critical buckling load on transversally and longitudinally stiffened


steel plate girders subjected to patch loading

J. Herrera & R. Chacón


Department of Civil and Environmental Engineering, Universitat Politècnica de Catalunya, Barcelona, Spain

ABSTRACT: It has been investigated that the patch loading resistance model available in
European design rules may lead to an underestimation for plate girders with closely spaced trans-
verse stiffeners. Predicting models have been proposed for this case, as well as the case of girders
with longitudinal stiffeners. However, when the problem is dealt with coupled, the critical buck-
ling load estimation is affected by the geometrical limitations b1/hw ≤ 0,3 and 0,05 ≤ b1/a ≤ 0,3.
Outside these boundaries, the literature does not provide any information on how to proceed.
While it is recognized that the vast majority of realistic girders have panels with proportions
among these limits, in some cases, these conditions are not fulfilled. The focus of this investigation
is on a practical estimation of the critical buckling load on girders with proportions outside of
these boundaries based on a numerical parametric analysis of girders including both transverse
and longitudinal stiffening.

1 INTRODUCTION

Steel plate girders have been studied profusely in last decades and its usage in civil engineering
is vast. Numerous structures worldwide, both buildings and bridges, are nowadays totally or
partially assembled with steel girders which may be designed with symmetric-, non-symmetric
I-sections or box girders.
In bridges, I-shaped steel girders are very common. These girders are normally assembled
with slender plates for the web and stocky plates for the flanges, which leads to a high flexural
capacity for a relatively low weight. However, issues concerning the stability of the plates may
appear. Occasionally, these elements must be stiffened both transversally and longitudinally by
welding additional plates that enhance their overall and local behavior. Constructional guides
covering the vast majority of topics dealing with stability and plate girders have been presented
by Dowling et al. (1992), Galambos (1998), Dubas & Gehri (1986) and Beg et al. (2012).
In bridges assembled with the incremental launching method, patch loading is one major verifi-
cation to be considered at design stages. Launching implies that all cross-sections of the plate
girder pass over temporary supports or piers. Thus, concentrated forces act on both transversally
stiffened and unstiffened sections. In these cases, an alternative to increasing the web thickness is
to stiffen the web plate with equally spaced vertical stiffeners or a combination of vertical and
longitudinal elements. Although longitudinal elements are initially conceived for bending and
shear capacity, the overall resistance to other plate buckling phenomena is generally improved.
Extensive experimental, numerical and theoretical investigations related to patch loading in
both unstiffened and longitudinally stiffened panels have been published. For the case of unstif-
fened panels, summary works have been presented by Lagerqvist & Johansson (1995, 1996) and
Chacón et al. (2012). The particular case of transversally stiffened girders has also been studied by
Chacón et al. (2013a, 2013b, 2017). For longitudinally stiffened elements, extensive work has been
presented by Graciano et al. (2003a, 2003b, 2003c, 2015) in which both critical buckling and ultim-
ate resistance of such elements subjected to patch loading are analyzed. The concepts developed in
Lagerqvist & Johansson (1995) and Graciano and Johansson (2003) were implemented in the

483
harmonization of EN1993-1-5 guidelines (2006) for plate girders subjected to patch loading, using
the classical χ-λ form, common to other instability problems.
Particularly concerning critical buckling loads, Graciano & Lagerqvist (2003) presented for-
mulae for the critical buckling coefficients of longitudinally stiffened girder webs. These for-
mulae were later scrutinized by Clarin (2007). Finally, simplified versions of such proposals
were included in the EC3 Part 1-5 rules. The current expressions for the calculation of the
buckling critical load are presented in Equations (1)–(3):

t:3
Fcr ¼ 0:9kf E w
ð1Þ
hw
 2
hw b1 pffiffiffiffi
Kf ¼ 6 þ 2 þ 5:44  0:21 γs ð2Þ
a a
 3  
Ist a b1
γs ¼ 10:9  13 þ 210 0:3  ð3Þ
hw t3w hw a

Where kf is the buckling coefficient and γs is the relative flexural rigidity of the longitudinal
stiffener.
It is worth noticing that the flexural rigidity of the longitudinal stiffener depends on the
aspect ratio of the directly loaded subpanel (b1/a) and that Equation (3) is only valid within
the ranges 0,05 ≤ b1/a ≤ 0,3 b1/hw ≤ 0,3.
In this paper, the case is generalized to the case of girders with longitudinal stiffeners for
which the relation b1/hw ≤ 0,3 is achieved but the relation 0,05 ≤ b1/a ≤ 0,3 is not accomplished.
This situation is particularly likely for the case of girders with closely spaced transverse stiff-
eners. A numerical study aimed at pointing out conclusions related to this matter is developed.
A set of variations of other geometrical proportions such as web slenderness (hw/tw), the longi-
tudinal stiffener position (b1) and rigidity (tsl) are used in this numerical parametric study of
steel plate girders subjected to patch loading. The study is performed following an eigenvalue
extraction using a subspace procedure and was performed using a commercial FE-package.

2 NUMERICAL MODEL

The multi-purpose code Abaqus-Simulia (2016) has been systematically used as a simulation tool
in this research. The code has been widely contrasted and bench-marked in several plate-buckling
related phenomena. Also, the results have been compared using the software EBPlate 2.01 (2007)
which based upon Fourier series calculations, provides accurate values of elastic critical stresses
for a wide field of practical cases of rectangular plates subject to in-plane loading. The results
obtained with both tools coincided. The characteristics of the used FE-based numerical model
read:
– Geometrically, the girders are idealized with four-noded shell elements with reduced
integration.
– Materially, the steel is idealized as elastic (for buckling calculations).
– The stability analysis is performed following an eigenvalue extraction using a subspace pro-
cedure. 6 eigenvalues are extracted for each girder.

3 PARAMETRIC STUDY

The numerical model has been used systematically as a simulation tool in order to study steel
plate girders subjected to patch loading. The parametric study is used on a broader ongoing inves-
tigation where the main feature is the presence of longitudinal stiffeners in girders with closely
spaced transverse elements. To choose the geometric characteristics, conclusions provided by

484
a prior study by Chacón et al. (2019) are followed. Specific characteristics of the numerical study
are pointed out.
– The numerical study is developed in girders with hw = 2000 mm. The use of these longitu-
dinal elements is justified in girders with considerable web height.
– The position of the longitudinal stiffener ranges from b1/hw = 0,2 to b1/hw = 0,3 since these
values represent the optimal limits for an effective design to bending.
– As stated before, for the case of girders transversally and longitudinally stiffened simultan-
eously the critical buckling load estimation methods present in the literature propose
a geometrical limitation for the relations b1/hw ≤ 0,3 and 0,05 ≤ b1/a ≤ 0,3. To find informa-
tion outside these boundaries, the numerical study is divided in two groups.
– Group V1 has a geometrical relation of a/hw = 0.75. For this group the condition a/ly < 1,0
is achieved in all cases, however the condition 0,05 ≤ b1/a ≤ 0,3 is not achieved for most of
the cases. The present work focuses on the results of this group of beams.
– Group V2 has a geometrical relation of a/hw = 1.00. For this group the condition 0,05 ≤
b1/a ≤ 0,3 is always achieved however the condition for closely spaced transverse stiffeners
is not achieved in most of the cases.
– Other proportions are chosen from according realistic dimensions of steel girders. The web
slenderness is varied systematically for covering a broad range
The study includes variations of web thickness tw, longitudinal stiffener thickness tsl and the pos-
ition of the longitudinal stiffener b1, for both groups of beams. Tables 1 and 2 display the set of
variations as well as other magnitudes of interest. An extra set of girders following the same
characteristics but longitudinally unstiffened is included for comparison purposes. A total
amount of 6 Eigenvalue extractions (buckle) have been performed for each case.

4 PARAMETRIC STUDY

The numerical model has been Tables 3–6 display the results for critical buckling load obtained
both when applying the EN1993-1-5 formulation (Equations 1–3) and numerically from the
first eigenvalue of the analysis. For the sake of brevity, only the results for tsl = 20 mm are
shown. For this stiffener thickness, the stiffener can be considered “strong” for all cases, and as
stated in Kövesdi (2018), once this is achieved, the variations in resistance from increasing the
thickness are minimum. Table 3 shows the results obtained for unstiffened girders whereas
Tables 4, 5 and 6 show the results obtained when varying b1 = 400 mm; b1 = 500 mm and
Table 1. Variations of the numerical study.
Parameter Magnitudes

tw (mm) 6 8 10 12 15
tsl (mm) 10 20 30 40
b1 (mm) 400 500 600
a/hw 0.75 1

Table 2. Magnitudes
held constant.
hw (mm) 2000
bf (mm) 600
tf (mm) 60
Ss (mm) 750
bsl(mm) 110
tst (mm) 40
fyf (N/mm2) 355
fyw (N/mm2) 355
fys (N/mm2) 355

485
Table 3. EN1993-1-5 results for the set of unstiffened girders.
tw tsl b1 Fcr,EN Fcr, NUM
Number (mm) (mm) (mm) γs,corrected kF, EN (kN) (kN) kF, NUM

V1_000_C1 6 unstiffened – – 9,56 195,05 373,03 18,28


V1_000_C2 8 unstiffened – – 9,56 462,34 855,72 17,69
V1_000_C3 10 unstiffened – – 9,56 903,00 1628,82 17,24
V1_000_C4 12 unstiffened – – 9,56 1560,38 2751,84 16,85
V1_000_C5 15 unstiffened – – 9,56 3047,63 5207,31 16,33

Table 4. EN1993-1-5 results for the set of girders with b1=400mm.


tw tsl b1 Fcr, EN Fcr, NUM
Number (mm) (mm) (mm) γs,corrected kF, EN (kN) (kN) kF, NUM

V1_400_C6 6 20 400 12,48 13,94 284,53 592,50 29.03


V1_400_C7 8 20 400 12,48 13,94 674,44 1285,47 26.57
V1_400_C8 10 20 400 12,48 13,94 1317,26 2325,51 24.61
V1_400_C9 12 20 400 12,48 13,94 2276,22 3792,60 23.23
V1_400_C10 15 20 400 12,48 13,94 4445,74 6959,07 21.82

Table 5. EN1993-1-5 results for the set of girders with b1=500mm.


tw tsl b1 Fcr, EN Fcr, NUM
Number (mm) (mm) (mm) γs,corrected kF, EN (kN) (kN) kF, NUM

V1_500_C6 6 20 500 – – N.A. 669,47 32.80


V1_500_C7 8 20 500 – – N.A 1377,63 28.47
V1_500_C8 10 20 500 – – N.A 2406,51 25.47
V1_500_C9 12 20 500 – – N.A 3851,19 23.58
V1_500_C10 15 20 500 – – N.A 6978,42 21.88

Table 6. EN1993-1-5 results for the set of girders with b1=600mm


tw tsl b1 Fcr, EN Fcr, NUM
Number (mm) (mm) (mm) γs,corrected kF, EN (kN) (kN) kF, NUM

V1_600_C6 6 20 600 – – N.A 585,27 28.67


V1_600_C7 8 20 600 – – N.A 1244,07 25.71
V1_600_C8 10 20 600 – – N.A 2209,95 23.39
V1_600_C9 12 20 600 – – N.A 3561,03 21.81
V1_600_C10 15 20 600 – – N.A 6478,02 20.31

b1 = 600 mm respectively. Detailed scrutiny on the results obtained for Fcr anf KF (mutually
dependent) allow drawing the following observations.
– According to EN1993-1-5, for a given girder, longitudinally stiffened elements in
group b1=400 provide a greater buckling load Fcr than their unstiffened counterpart.
– For groups b1 = 500 and b1 = 600, Fcr could not be obtained using EN1993-1-5 since these
beams do not comply with the condition 0,05 ≤ b1/a ≤ 0,3.
– Using the numerical models, for a given girder, all longitudinally stiffened elements provide
a greater buckling load Fcr than their unstiffened counterpart.
– Using the numerical model, the difference in Fcr between girders with varying position of
the longitudinal element is not very high. The maximum value is consistently obtained for
girders with b1 = 500mm.
Table 7 displays frontal views of the 1st Eigenmode in 20 simulations. Numerical values are
also included. The following observations can be pointed out.

486
– In the cases when b1/a > 0.3 (b1 = 500 and 600 mm), the shape involves both subpanels (above
and below the longitudinal stiffener). An intertwined shape covering both subpanels is
observed.
– The elastic critical buckling loads increase with the web thickness as expected and it is
greater for longitudinally stiffened elements than for their unstiffened counterparts.
Results are also displayed in the form of behavioral trends. Figure 1 shows the results of the
numerical buckling coefficient kF,num, as a function of the position of the longitudinal stiffener
b1/a. The figure allows drawing the following observations.
– There is a clear change of tendency, although it seems to happen not exactly at
b1/a = 0.3, but at a slightly higher value (0.33). From this point onwards, the value of Fcr
starts to decrease although it still remains higher than its unstiffened counterpart in all cases.
– The decrease in slopes for each of the lines corresponding to each web thickness leads to
understand that as the web becomes stockier the contribution of the longitudinal stiffener
becomes less important.
Having confirmed that there is a characteristic point near the b1/a=0.3 boundary, an estimation of
a way in which the EN1993-1-5 formulae can be applied in a safe way for girders outside this
boundary is sought. As it is stated by Graciano (2002), the variation in the transition rigidity with
b1/a tends to be negligible after 0.30, so, one simple way to approach the estimation would be to
Table 7. Elastic buckling modes and shapes for the 1st Eigenvalue.
tw b1 = 400 b1 = 500 b1 = 600
(mm) Unstiffened (mm) (mm) (mm)

Fcr=373,03 kN Fcr=592,50 kN Fcr=669,47 kN Fcr=585,27 kN


8

Fcr=855,72 kN Fcr=1285,47 kN Fcr=1377,63 kN Fcr=1244,07 kN


10

Fcr=1628,82kN Fcr=2325,51 kN Fcr=2406,51 kN Fcr=2209,95 kN


12

Fcr=2751,84 kN Fcr=3792,60 kN Fcr=3851,19 kN Fcr=3561,03 kN


15

Fcr=5207,31kN Fcr=6959,07 kN Fcr=6978,42 kN Fcr=6478,02 kN

487
Figure 1. KF,num vs b1/a.

Table 8. Elastic buckling coefficient obtained with three different methods.


KF, prop1 KF, prop2
tw (using Equations (2) and (using Equation (2) considering
(mm) KF, NUM (3) considering b1/a=0.3) b1/a=0.33)

10

12

488
Table 9. Elastic buckling coefficient obtained with three different
methods.
tw (mm) b1/a Kf, num Kf,prop1 Kf,prop2

0 18.28 9.56 9.56


0.26 29.03 13.94 13.94
0.33 32.80 12.89 13.31
6 0.4 28.67 12.89 13.31
0 17.69 9.56 9.56
0.26 26.57 13.94 13.94
0.33 28.47 12.89 13.31
8 0.4 25.71 12.89 13.31
0 17.24 9.56 9.56
0.26 24.61 13.94 13.94
0.33 25.47 12.89 13.31
10 0.4 23.39 12.89 13.31
0 16.85 9.56 9.56
0.26 23.23 13.94 13.94
0.33 23.58 12.89 13.31
12 0.4 21.81 12.89 13.31
0 16.33 9.56 9.56
0.26 21.82 13.94 13.94
0.33 21.88 12.89 13.31
15 0.4 20.31 12.89 13.31

use Equations (2) and (3) considering the value b1/a = 0.3 as an upper bound. On the other hand,
since Figure 1 implies that the value for which tendency begins to shift is located at b1/a=0.33
a new calculation is performed, neglecting the term [210(0.3-b1/a)] from Equation (3) but using the
actual value of b1/a in Equation (2) limited to 0.33.
The results from both options as well as with the numerical results are shown both graphic-
ally (Table 8) and numerically (Table 9). In Table 8, the dotted line represents the results for
each studied panel whereas the red line displays a regression linear trend.
Closer inspection of Tables 8 and 9 allows drawing the following observations:
– For all cases, the linear tendency intersects the graphic for a value of b1/a equal or very
close to 0.33.
– As expected, in all cases the value of kF for both propositions is lower than the numerical
value obtained from the eigenvalue analysis.
– As expected, in all cases the value of kF for both propositions is lower than the numerical
value obtained from the eigenvalue analysis.
– The values for the calculated kF are closer to the numerical values as the web thickness
increases.
– Values of kF,num2 fit better than values of kF,num1 to the numerical values obtained
from FEM.

5 CONCLUSIONS

In this paper, the formulation for obtaining the critical buckling load of steel plate girders with
both transversal and longitudinal stiffening subjected to patch loading is revisited for potential
application in the particular case where the current geometrical limitation 0,05 ≤ b1/a ≤ 0,3 is
exceeded. Manifold conclusions can be drawn from this study.
– In its present form, EN-1993-1-5 provisions for steel girders both unstiffened and longitu-
dinally stiffened provide safe-sided results for the critical buckling load when compared to
the results obtained numerically. Assumptions such as conservative boundary conditions

489
are attributed to these differences. The results obtained herein reinforce those derived from
numerous previous publications.
– Numerical results show that longitudinally stiffened girders provide a greater critical buck-
ling load than unstiffened webs. The trend is observed for all the considered values of b1/a.
– The critical buckling load increases with b1/a up to a maximum values b1/a = 0.33, for
greater values of this parameter, the critical buckling load decreases.
– A simple approach has been assessed in order to estimate the critical buckling load for values
that exceed the condition 0,05 ≤ b1/a ≤ 0,3. The same formulae current available in the EN-
1993-1-5 provisions is used. The results are contrasted with those within the limits of the condi-
tion. Similar safe-sided ratios kf,num/kf,EN (and consequently Fcr,num, Fcr,EN) are observed.
– Results obtained suggest that the upper bound limit of the condition 0,05 ≤ b1/a ≤ 0,3 could
be extended up to a value b1/a = 0,33. A significant increase in critical buckling load may
be obtained for such girders.
– Further studies aimed at assessing ultimate load capacities as well as at assessing resistance
functions are necessary. An overall assessment of the EN1993-1-5 for transversally and lon-
gitudinally stiffened steel plate girders subjected to patch loading is presently being
addressed by the authors.

REFERENCES

Beg, D. Kuhlmann, U. Davaine, L. & Braun, B. 2012. Design of Plated Structures: Eurocode 3: Design of
Steel Structures, Part 1-5. Berlin: Ernst & Sohn.
Chacón, R. Braun, B. Kuhlmann, U. & Mirambell, E. 2012. Statistical evaluation of the new resistance
model of steel plate girders subjected to patch loading. Steel Construction. 5(1): 10–15.
Chacón, R. Mirambell, E. & Real, E. 2013a. Transversally stiffened plate girders subjected to patch load-
ing. Part 1. Preliminary study. Journal of Constructional Steel Research. 80(1): 483–491.
Chacón, R. Mirambell, E. & Real, E. 2013b. Transversally stiffened plate girders subjected to patch load-
ing. Part 2. Additional numerical study and design proposal. Journal of Constructional Steel Research,
80(1): 492–504.
Chacón, R. Herrera, J. & Fargier-Gabaldón, L. 2017. Improved design of transversally stiffened steel
plate girders subjected to patch loading. Engineering Structures. 150(11): 774–785.
Chacón, R., Mirambell, E. & Real, E. 2019. Transversally and longitudinally stiffened steel plate girders
subjected to patch loading. Thin-Walled Structures. 138(5): 361–372.
Clarin, M. 2007. Plate Buckling Resistance – Patch Loading of Longitudinally Stiffened Webs and Local
Buckling. Doctoral thesis. Luleå, Sweden: Luleå University of Technology, Division of Steel Structures.
Dowling, P. Harding, J. & Bjorhovde, R. 1992. Constructional Steel Design. An international guide.
London & New York: Elsevier applied science.
Dubas, P. & Gehri, E. 1986. Behaviour and design of steel plated structures. Zurich: ECCS-CECM-EKS.
Galambos, T. 1998. Guide to Stability design criteria for metal structures. New York: John Wiley and sons.
EBPlate 2.01. 2007. Elastic Buckling of Plates. CTICM, France
EN1993-1-5. 2006. Eurocode 3. Design of steel structures – Part 1-5: Plated structural elements CEN.
Graciano, C. 2002. Patch Loading Resistance of Longitudinally Stiffened Steel Girder Webs. Doctoral
thesis, Luleå, Sweden: Luleå University of Technology, Division of Steel Structures.
Graciano, C. 2003. Ultimate resistance of longitudinally stiffened webs subjected to patch loading. Thin-
Walled Structures. 41(6): 529–541.
Graciano, C. & Johansson, B. 2003a. Resistance of longitudinally stiffened I-girders subjected to concen-
trated loads. Journal of constructional Steel Research. 59(5): 561–586.
Graciano, C. & Lagerqvist, O. 2003b. Critical buckling of longitudinally stiffened webs subjected to com-
pressive edge loads. Journal of Constructional Steel Research 59: 1119–1146.
Graciano, C. 2015. Patch loading resistance of longitudinally stiffened webs- A systematic review. Thin-
walled Structures. 95(10): 1–6.
Kövesdi, B. 2018. Patch loading resistance of slender plate girders with longitudinal stiffeners. Journal of
Constructional Steel Research. 140: 237–246.
Lagerqvist, O. & Johansson, B. 1995. Resistance of plate edges to concentrated forces. Journal of Con-
structional Steel Research, 32: 69–105.
Lagerqvist, O. & Johansson, B. 1996. Resistance of I-girders to concentrated loads. Journal of Construc-
tional Steel Research 39(2): 87–119.

490
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Buckling behavior and strength of corroded steel shapes under


axial compression

K. Hisazumi
Nippon Steel Corporation, Futtsu, Chiba, Japan

R. Kanno
Kanazawa University, Kanazwa, Ishikawa, Japan
Nippon Steel Corporation, Futtsu, Chiba, Japan

ABSTRACT: Corrosion is one of the most serious threats to the sustainability of steel struc-
tures, especially for those exposed to the atmosphere and subjected to improper maintenance
management. Typical steel structures are bridges, transmission line towers, and conveyer-
supporting frames. As corrosion-related problems and accidents have been frequently
reported in recent years, the need for identifying the health condition of structures has
increased. In this context, the axial compressive strength of corroded steel shapes is investi-
gated with experiments and non-linear finite element analyses. The results indicate that the
behavior of the corroded members depends strongly on the degree of corrosion and shows
global flexural buckling, local buckling, section yielding, and combined failure modes.
A strength formula considering the interaction between global and local buckling is proposed,
showing that it can reasonably be used for estimating the strengths of corroded steel shapes.

1 INTRODUCTION

Deteriorating infrastructure has been a growing concern, particularly in developed countries


such as Japan, the US, and European countries. One of the major reasons for the deterior-
ation is corrosion, which becomes serious especially for steel structures directly exposed to the
atmosphere, such as bridges, transmission line towers, and conveyer-supporting frames at
plant facilities. In most cases, improper maintenance management caused by budgetary con-
cerns exacerbates the health condition of the structures. Therefore, identifying the quantitative
structural health condition at various stages of deterioration has become important for avoid-
ing serious accidents well in advance. However, effective methodologies have not been estab-
lished because of the inherent behavioral complexity of the corroded steel members.
Among the various structural steel systems, plate-girder-type structures seen in steel bridges
are relatively redundant, such that the systems do not have a sudden failure, even though cor-
rosion proceeds at some of the parts. However, truss-type structures, such as transmission line
towers and conveyer-supporting frames, need meticulous attention, because once one of the
members loses its strength, complete and sudden collapse of the entire structure results. This is
because truss frames are generally internally static. Therefore, the residual strength evaluation
of corroded truss frames becomes an urgent matter from a practical point of view.
As a result, this study focuses on steel truss frames consisting of relatively small hot-rolled
steel shapes, such as steel channels and angles. In the frames, the members are subjected pri-
marily to a tension or a compression force. In general, the compression strength becomes
smaller than the tension strength because of the buckling that occurred under compressive
stress; thus, compression behavior of the corroded steel shapes is of prime importance. Avail-
able research on corroded steel channels and angles under compression force are quite limited.
Only the compressive strengths of corroded steel angles were studied by those including

491
Beaulieu et al. (2010), Oszvald & Dunai (2012), and Oszvald et al. (2016). Since naturally cor-
roded steel members were difficult to obtain, artificially fabricated (by mechanical milling pro-
cess) or processed (with galvanic corrosion process) steel angles were used, which might not
represent the actual behavior of corroded steel shapes. Column strength formulae were pro-
posed for the corroded steel angles, but inconsistent findings were reported, particularly with
regard to the cross-sectional area related directly to the compression strengths: Beaulieu et al.
(2010) proposed using the mean area, while Oszvald & Dunai (2012) showed that the min-
imum area was a better parameter to use than the mean value.
With this background, the authors have conducted research on the compression behavior
and strength of corroded steel angles and channels by experiments and finite element (FE)
analyses, using naturally corroded specimens sampled directly from a truss frame with more
than 40 years of operation (Hisazumi et al. 2014, Hisazumi et al. 2018). This paper is based
primary on the authors’ previous research but is extended further, with additional findings
later obtained. In this paper, a compressive strength formula for corroded shapes is proposed,
providing its accuracy information for experimental and analysis results.

2 CORRODED STEEL SHAPES STUDIED

Corroded steel shapes were sampled from a typical steel truss frame consisting of upper and lower
chords, and secondary members, in which each member was either a channel shape or an angle
shape. The truss frame was in service at a plant facility for about 40 years, which was exposed
directly to a severe corrosive environment near the ocean. A total of 27 corroded steel specimens
were cut from the frame: 10 channels from the upper and lower chords, and 17 angles from the
diagonal members and struts. After sandblasting was made to remove the rust, surface profiles of
the specimens were precisely measured with a 3D laser displacement transducer with 1-mm grid.
Original (non-corroded) cross sections of the specimens were 125 × 65 × 6 × 8 mm (depth × width
× web thickness × flange thickness) for the channels and 50 × 50 × 4 (depth × width × thickness),
50 × 50 × 6, 65 × 65 × 6, and 75 × 75 × 9 for the angles, in which the thicknesses were defined at
the centers of the depth and width.
Figure 1a shows a longitudinal distribution of the sectional area, calculated based on the
measured profile for one of the corroded channels. Together with the sectional area, original
(non-corroded) and average areas are indicated. The location of the minimum area is specified
with an open circle. A perspective drawing of the specimen is also attached next to the area
distribution with thickness contours. As seen, the cross-sectional area varies significantly
along the corroded member: less corrosion at both ends and more in the middle in this case.
To quantify the degree of corrosion, a corrosion ratio R was introduced as the ratio of cor-
roded area to original area and calculated longitudinally along each specimen at 1-mm intervals.

Figure 1. Longitudinal distribution of sectional area and frequency distributions of corrosion ratios.

492
According to the definition, R approaches 100% as corrosion proceeds. The maximum and
average values of R for a corroded member are defined as Rmax and Rave, respectively. The
degree of corrosion of the specimens extend over a wide range: 8% to 74% for Rmax and 6% to
51% for Rave. Figure 1b indicates the frequency distributions of Rmax and Rave. Both distribution
characteristics are different from each other, indicating that corrosion did not occur uniformly
but rather localized and scattered. For the specimens with Rmax larger than approximately 40%,
corrosion pits and/or holes were observed. A total of 14 out of 27 specimens had corrosion pits
and/or holes, and in an extreme case, a 418 × 95 mm (longitudinal length × width) corrosion
hole was observed for a channel case with Rmax = 74%. More detailed information on the speci-
mens including geometries and the degree of corrosion can be found in Hisazumi et al. (2014).

3 AXIAL COMPRESSION TESTS AND STRENGTH EVALUATION

Uniaxial compression tests were conducted for all the specimens (Hisazumi et al. 2014). To
avoid uncertainty related to the end conditions, a complete fixed condition was chosen so that
the effective buckling length was one-half the total specimen length L (= L/2). Each specimen
had steel base plates welded to the ends. To minimize the eccentricity, grout was thinly placed
between the loading head and base plate at each end. Loading was applied under displacement
control and terminated after the maximum load and deformation were clearly identified.
Total length L of the specimens was 1700 mm for the channels, varying from 250 to
1000 mm for the angles. Measured yield strengths through coupon tests were between 301 and
338 MPa. An important parameter, normalized slenderness ratio λn was considered as follows:
rffiffiffiffiffiffiffi
λ σy
λn ¼ ð1Þ
π E

where σy = material yield strength; E = Young’s modulus; and λ = slenderness ratio. Since the
cross section varies along the specimen, λ is defined in this study using cross section properties
min as λ = ffi Lk/im, where im = radius of
at the location of the minimum cross-sectional areapAffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
gyration at the minimum cross-sectional area (¼ I min = Amin ); Imin = moment of inertia
around the weak axis of the original member; and Lk = effective buckling length (= L/2). The
normalized slenderness ratios λn of all the specimens ranged from 0.16 to 0.84, which might be
regarded as short and intermediate column.
Figure 2 shows typical load-displacement relationships and specimen photos taken after the
tests. Both axes in the graphs were normalized by the yield load (Py = σy Amin) and yield displace-
ment (δy = σy L/E). The figure includes two channels and two angles. The channels had larger
corrosion ratios Rmax of around 40%, whereas the angles had smaller Rmax of around 10%. As
seen, the specimens behaved similarly to each other: nearly linear behavior until the maximum
strength and then gradual decrease in strength. The degree of strength degradation was slightly
different between the specimens: the deterioration became smaller as Rmax decreased.
Through the observations and analyses, the following failure modes were identified:
1) global flexural buckling after full section yielding (Figure 2a), 2) flexural buckling before
full section yielding (Figure 2b), 3) combination of local and flexural buckling (Figure 2c),
and 4) local buckling (Figure 2d). The combined local and flexural buckling mode was the
most observed among the tested specimens. With the increase in the degree of corrosion
(Rmax), local buckling became more interacted and predominant.
Figure 3 indicates the relationships between ultimate strength Pmax and average cross-sectional
area Aave, and between Pmax and minimum cross-sectional area Amin. There is less correlation
between Pmax and Aave with a correlation coefficient of 0.67. On the other hand, a distinct relation-
ship exists between Pmax and Amin, with a correlation coefficient of 0.91. The implication here is
that the strength of a corroded steel member tends to be influenced more strongly by the weakest
part of the specimen (= the smallest cross-sectional area) than by its average property.

493
Figure 2. Typical load-displacement relationships and specimen photos after tests.

For the strength evaluation of the corroded specimens, a discussion is presented here on the
applicability of Johnson’s parabolic formula, which is one of the well-known formulas for
global flexural buckling strength. This attempt was made because the flexural buckling mode
was the one most frequently observed in the compression tests. Since the cross section of any
corroded member was not uniform along its length, the formula was applied using the cross-
section properties at the minimum area. This is because, as shown in Figure 3, a high correl-
ation with maximum buckling strength was identified. The formula applied is as follows:

Pcr ¼ Amin σcr ð2Þ


 
1  0:24λ n 2 σy λn 51:3
σcr ¼ ð3Þ
σy = λn 2 λn  1:3

where Pcr = estimated column strength; σcr = global flexural buckling strength given by Japan-
ese standards; and λn was defined in Equation (1), based on the cross-section properties at the
minimum sectional area.

Figure 3. Relationships between maximum strength and cross-sectional areas.

494
Figure 4a shows the relationship between the ratio of measured Pmax to calculated Pcr
(= Pmax/Pcr) and the maximum corrosion ratio Rmax for all the specimens. The mean
value of Pmax/Pcr is 0.93 and the coefficient of variation (COV) is 27.3%. There is
a relatively large scatter, but the formula provides reasonable strength estimates on the
safe side up to a maximum corrosion ratio Rmax of around 40%. However, when Rmax
becomes larger than 40%, the evaluation reveals unsafe conditions—the worst case cor-
responds to a Pmax/Pcr of approximately 0.6. This is possible because with an increase in
the degree of corrosion, local buckling might affect the strength to a larger degree and
decrease the effective cross-sectional area. For additional information, the relationship
between Pmax/Pcr and slenderness ratio parameter λn is shown in Figure 4b. As seen, the
relationship with λn is not clear, compared with that with Rmax.
Note in the above formula that the radius of gyration at the minimum cross-sectional area
(im) is needed, but its calculation becomes problematic because of complicated section geom-
etry. Based on a separate analysis of the geometry of the corroded specimens, it was found
that im might be replaced with the radius of gyration of an original member (= i). As a result,
σcr in Equation (2) could be determined as the buckling strength of the original member.

4 LOCAL BUCKLING STRENGTH EVALUATION

It is well-known that the ultimate strength of local buckling is strongly affected by the width-to-
thickness ratio, but owing to complex geometrical variations of corroded steel specimens, it is still
uncertain whether existing knowledge of local buckling strength might be applied. Thus, the
behavior and strength of corroded stub columns were studied with FE analyses using geometric
information obtained directly from the corroded specimens (Hisazumi et al. 2018).
Non-linear FE analyses were conducted on corroded stub column models, which were made
numerically by extracting a part of each corroded specimen, such that the minimum cross-
sectional area was located at the middle of the stub column. The length L of the stub column
was chosen to be three times the maximum section width. When the minimum area was located
near either end of the specimen, the stub column was taken from the end. Elasto-plastic large
deformation FE analyses were performed for 27 cases in total using commercial software called
MARC (Ver. 2014). Axial compressive loading was applied under displacement control with
a boundary condition that the upper and lower load applying heads did not rotate. Square-shell
elements of 2 × 2 mm mesh were used, and the thickness of the element was assumed to be
constant. True stress-true strain relationships obtained from tensile coupon tests were applied.
Since the definition of the width-to-thickness ratio is not straightforward because of the
variation in thickness, the following non-dimensional parameter (λp) was introduced:
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 b 12ð1   2 Þ σy
λp ¼ ð4Þ
π tave kE

Figure 4. Influence of maximum corrosion ratio Rmax and slenderness ratio parameter λn on Pmax/Pcr.

495
Figure 5. Definitions of plate elements.

where b = width of each plate element defined as a region excluding the corners (Figure 5);
tave = mean thickness of the corresponding plate element; and k = local buckling coefficient
(either 4.0 or 0.425, depending on the supporting condition). The maximum value of the width-
to-thickness parameter calculated at each 1-mm interval in a corroded specimen was defined as
λpc, which ranged between 0.52 and 4.06 for all the stub columns analyzed in this study. The
width-to-thickness parameters of the original (non-corroded) members varied from 0.51 to 0.8.
Figure 6 shows the non-dimensional load-displacement relationships of the three specimens
with red lines, including one lightly corroded column and two moderately corroded columns.
The load-displacement relationships of the plate elements in the section are also included in
the figure, in which the loads (i.e., resisting forces) of the plate elements (named Flange-L,
Flange-R, and Web) were calculated at the minimum cross-sectional area of the column.
Deformations and longitudinal stress contours are added to Figure 6. Solid triangles and open
circles in the figure indicate the locations of the maximum strengths for the entire stub col-
umns and plate elements, respectively.
Regarding the lightly corroded specimen shown in Figure 6a, the behavior was stable
enough such that the entire section yielded, and the entire column as well as the plate elements
in the section exhibited their maximum strengths at the same displacement simultaneously.
The moderately corroded specimens in Figures 6b and 6c exhibited similar behavior to the
lightly corroded specimen, except that they showed relatively larger strength degradation after
the maximum strengths were reached. An interesting observation in the moderately corroded
specimens was that although one of the plate elements in the section reached its maximum

Figure 6. Relationships between non-dimensional load and non-dimensional displacement (Reproduced


from Hisazumi et al. (2018) with permission of the ICASS2018 committee in China).

496
strength, the stub columns did not lose their strengths; the rest of the elements continued to
resist and reach their maximum strengths (as in the order of open circle number 1, 2, and 3 in
Figure 6b and 6c). Then, the entire stub column exhibited maximum strength.
It is reasonable to say that this behavior was caused by stress redistribution within the sec-
tion, suggesting that the well-known effective width theory for predicting the ultimate strength
of thin-walled members might be applicable, even to corroded steel members. Therefore, the
applicability of effective width theory to local buckling strengths of corroded stub columns was
examined. Effective width theory applied here was that proposed by the AISI Standard (2016):

Pe ¼ σy Ae ð5Þ
X
n
Ae ¼ ρi Ai ð6Þ
i¼1

Ai ¼ bi tiave ð7Þ

 1 λip 51:3
ρi ¼ 1  0:22=λip =λip ð8Þ
λip  1:3

where Pe = axial ultimate strength of the column section; Ae = the corresponding effective
area; ρi = effective cross-section ratio of the i-th plate element in the section; bi = width of the
i-th plate element defined in Figure 5; tavei = average thickness of the i-th plate element; and
λpi = non-dimensional width-to-thickness parameter of the i-th plate element.
Based on Equations (5) through (8), Pe is calculated in the following steps: at a cross section
given at an interval of 2 mm (= mesh spacing) in a corroded stub column, find the width-to-
thickness ratio parameter λpi for each plate element (i = 1 to n). Calculate the corresponding
effective cross-section ratios ρi by Equation (8) and the effective cross-sectional area Ae by
Equation (6). Then Pe is obtained with Equation (5). Repeating these steps, Ae for all the sec-
tions along the entire length at an interval of 2 mm, the minimum value of Pe becomes the
local buckling strength of the corroded stub column. Note that the minimum value of Pe was
found, in most cases, to occur at the location of the minimum sectional area through
a separate study. Therefore, Pe could be calculated only once, at the minimum sectional area.
Figure 7a shows the relationship between the strength ratio PFEM/Pe and the maximum cor-
rosion ratio Rmax, where PFEM is the maximum strength obtained by FE analyses. Since the
mean value of PFEM/Pe is 1.06 and the COV is 12.2%, it can be said that, overall, the max-
imum strength may be reasonably evaluated by the proposed formula shown in Equations (5)
through (8). Note that in Figure 7a, for the plots with Rmax larger than around 50%, some
scatter is found because of large corrosion holes. Failure modes for three cases are attached in
Figure 7a. In these cases, it is clearly understood that the sectional integrity or unity as one
single shape was completely lost, which falls of the scope of the proposed formula.

Figure 7. Accuracies of proposed formulae to analyses and test results.

497
5 BUCKLING STRENGTH FORMULA FOR CORRODED STEEL SHAPES

Combining the flexural buckling formula with the local buckling formula discussed so far, the
following coupled buckling strength formula for uniaxial compression strength is proposed:

0
Pcr ¼ Ae σcr0 ð9Þ

where σcr0 = global buckling strength of the original (non-corroded) member based on Equations
(1) and (3), where im can be replaced with i, and A′e = minimum effective area along the corroded
member based on Equations (5) through (8), where λpi should be calculated using σcr0 instead of
σy. As mentioned already, A′e can be calculated where the minimum sectional area Amin is found.
Solid square plots in Figure 7b show the accuracy of the proposed formula with respect to the
test results (Pmax), in which clearly out-of-scope cases (three cases mentioned related to Figure 7a)
are excluded. Compared with the corresponding data in Figure 4a (open plots in Figure 7b), the
overall accuracy increases with a mean value of 1.05 and a COV of 13.0%, suggesting that the
proposed formula may be used for strength evaluation of the corroded steel shapes. Considering
the complexity of corroded member geometry, the applicability of the proposed formula should
be limited within the ranges (e.g., slenderness ratio, width-to-thickness ratio, corrosion ratio, etc.)
of this research. Further research is clearly needed in the future.

6 CONCLUSIONS

A study was conducted on the axial compressive strength of corroded steel channels and
angles, including ones severely corroded, with experiments and non-linear FE analyses. After
measuring the precise 3D-geometry of the specimens, axial compression tests were conducted
on the specimens. The tests showed relatively simple to complex behavior, depending on the
degree of corrosion, such as local buckling, global flexural buckling, section yielding, and
their combined modes. Although the behavior was complex, it was suggested that the
strengths of the lightly and moderately corroded steel shapes were reasonably evaluated by
a typical global flexural buckling strength formula using sectional properties at the minimum
sectional area. In addition, as the degree of corrosion became larger, local buckling was found
to interact to a larger degree. Further non-linear FE analyses of corroded stub-columns
showed that effective width theory could be applied by using the mean thicknesses of the plate
elements. In conclusion, a strength formula considering global buckling, local buckling, and
section yielding interactions was proposed. It was identified that the formula would reason-
ably be used for strength evaluation of corroded steel shapes, but clearly, its applicability
should remain within the range of this study.

REFERENCES

AISI (American Iron and Steel Institute) 2016. North American Specification for the Design of Cold-
formed Steel Structural Members. Washington, D.C.: AISI.
Beaulieu, L.V., Legeron, F. & Langlois, S. 2010. Compression strength of corroded steel angle members.
Journal of Constructional Steel Research 66: 1366–1373.
Hisazumi, K., Kanno, R., Tominaga, T. & Imafuku, K. 2014. Axial compressive strength of severely cor-
roded channel and angle members used in truss structures. The 7th European Conference on Steel and
Composite Structures: EUROSTEEL’2014, Naples, Italy, vol. A: 393–398.
Hisazumi, K., Kanno, R. & Tominaga, T. 2018. Local buckling strength of corroded angle and channel
steel shapes and its evaluation using effective width theory. The Ninth International Conference on
Advances in Steel Structures: ICASS2018, Hong Kong, China.
Oszvald, K. & Dunai, L. 2013. Effect of corrosion on the buckling of steel angle members – experimental
study. Periodica Polytechnic, Civil Engineering 56/2: 63–75.
Oszvald, K., Tomka, P. & Dunai, L. 2016. The remaining load-bearing capacity of corroded steel angle
compression members. Journal of Constructional Steel Research 120: 188–198.

498
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Modal analysis of thin-walled members with transverse plate


elements using the constrained finite element method

T. Hoang & S. Ádány


Budapest University of Technology and Economics, Budapest, Hungary

ABSTRACT: In this paper a new version of the constrained finite element method is
reported. The method is essentially a shell finite element method, but mechanical con-
straints can be applied to enforce the analysed member to deform in accordance with
global, distortional or local deformations. The method already proved itself to be the
most general among similar methods with modal decomposition ability, but lately it has
further been generalized by extending its applicability to members with extra transverse
plates such as end-plates of transverse stiffeners. In this paper proof-of-concept examples
are presented.

1 INTRODUCTION

In the recent decades the numerical methods with modal decomposition ability gained
certain popularity in the analysis of thin-walled structural members (beams or columns).
Modal decomposition methods describe the displacements-deformations of the member
by practically meaningful modes, such as global mode (G), local mode (L), distortional
mode (D), and other modes (O). The deformation modes can also be considered as
a modal basis system, by the application of which the problem (e.g., buckling problem)
can be solved in a reduced deformation space, or general deformations can be identified.
This approach helps in understanding the behavior, and has been found especially useful
in understanding the buckling behavior.
In the case of the two most known methods, i.e., the generalized beam theory (GBT), see.
e.g., Bebiano et al. (2018), and the constrained finite strip method (cFSM), see e.g., Ádány
&Schafer (2014a, 2014b), the modal decomposition is based on thin-walled beam theory,
hence these methods can be applied to prismatic thin-walled beams-like members, such as
cold-formed steel members. Welded or hot-rolled steel members can also be considered as
thin-walled, however, traditional steel members are frequently supplied with transverse plate
elements, either as transverse stiffeners or as end-plates. From mechanical aspect the trans-
verse elements are out of the realm of beam theories, therefore, structural members with
added transverse plate elements are not yet handled by the existing modal decomposition
methods. However, there is a newer method, the constrained finite element method (cFEM),
as in Ádány (2018b) and Ádány et al. (2018), which is a shell finite element method, but can
perform modal analysis similar to GBT and cFSM. cFEM has a wide range of applicability,
including members arbitrary loading and support conditions, members with holes, or even
members with certain cross-section variations. Most recently, the cFEM has been extended to
cover members with transverse plate elements.
In this paper first the cFEM is briefly summarized, focusing on the aspect how it is extended
to handle transverse plates. Then proof-of-concept examples are shown. Various problems are
solved, namely first-order static analysis is performed then the displaced shapes are identified
(i.e., it is determined how the G, D, L and O modes are contributing to the displacements).
Then, various linear buckling problems are solved with and without transverse stiffeners, and
the results are discussed.

499
2 CFEM FOR MEMBERS WITH TRANSVERSE PLATES

The constrained finite element method (cFEM) is essentially a shell finite element calculation,
but the method is developed so that modal decomposition would be possible. In order to
maintain constraining ability, the longitudinal shape functions of the employed finite elements
are specially selected, but the shell elements can be used as any regular flat shell element. Sep-
aration of the behavior modes is realized by applying mechanical constraints. The subsequent
application of the properly selected mechanical criteria finally leads to an alternative basis
system of the displacement field defined by the finite element nodal degrees of freedom: the
practically useful feature of this alternative basis system is that the deformation modes are
separated.
If the constraints are applied to a member without any additional elements, the d displace-
ment vector can be expressed by the dM modal displacement vector and the RM constraint
matrix specific to a given ‘M’ deformation space, as:

d ¼ RM dM ð1Þ

If additional plates (e.g., end-plates or transverse stiffeners) are present, the degrees of free-
dom, hence the displacement vector can be partitioned: one partition contains the DOFs that
are located in the member (dm ), the other partition contains the DOFs that are located in the
additional stiffeners (ds ). It is the first partition that can be expressed by the modal basis
vector, while the second partition is not directly affected by the constraints, therefore the
whole displacement vector can now be expressed as:

dm RM 0 dM
¼ ð2Þ
ds 0 I ds

or in short:

d ¼ RM dM ð3Þ

The buckling problem, mathematically, requires the solution of a generalized eigen-value


problem, as follows:

Ke Φ  LKg Φ ¼ 0 ð4Þ

where Ke and Kg are the elastic and geometric stiffness matrices, L contains the eigen-values
(i.e., critical load multipliers), and Φ contains the eigen-vectors (i.e., buckled shapes). When
modal decomposition is applied to a linear buckling problem, a possible approach is to substi-
tute Equation 3 into Equation 4, which leads to another generalized eigen-value problem,
given in the reduced M deformation space, as follows:

T T
RM Ke RM ΦM  LRM Kg RM ΦM ¼ 0 ð5Þ

The modal decomposition might beneficially be used in two tasks. One is the calculation in
a reduced (i.e., constrained) deformation space, as described above. For example, the linear
buckling analysis can be done for the G deformations only, by applying Equation 5, meaning
that all the buckled shapes will be global (e.g., flexural, or lateral-torsional buckling). The cal-
culation in a constrained deformation space is realized by selecting only those modal basis
vectors that belong to the desired deformation mode space.
The other task of cFEM is modal identification. The objective is to determine the participa-
tion of the characteristic deformations modes (G, D, L, etc.) in a deformed shape of the
member. The deformed shape might come from a linear static analysis, or linear buckling

500
analysis, etc. A special basis system is necessary in which the various deformation modes are
separated. Such basis system is naturally provided by the cFEM basis functions. Mathematic-
ally this means that any deformation, defined by the displacement vector, can be expressed as
the linear combination of the basis vectors. Once the combination factors are calculated, the
magnitudes of the combination factors give the relative importance of the basis vectors in the
displacement vector to be identified. Since the basis vectors are separated into G, D, L, etc.
mode spaces, the relative importance of the G, D, L, etc. deformations can be calculated. For
more information, see Ádány (2018a). When additional plate elements (such as end-plates) are
present, modal identification can still be done, but only the dm partition of the displacement
vectors are to be identified.

3 LINEAR ANALYSIS WITH MODAL IDENTIFICATION

In this Section a box girder beam member is analyzed, with and without transverse stiff-
eners. The total width of the cross-section is 3600mm, while the dimensions of the
hollow part is 2000×1500 mm. The thickness is 30mm. If stiffeners are employed, the
stiffener is 500×1500mm with 30 mm thickness. The stiffeners are applied in pairs, inside
the box, attached to the webs. The total length of the beam is 20 m, the number (pair
of) of stiffeners is varying from zero to 5. The stiffeners are equally spaced along the
member length. The material is steel, with a Young’s modulus of 210000 MPa, and with
a Poisson’s s ratio of 0.3.
Concentrated forces are applied as loading. In the examples shown here the location of
the force(s) in the mid-length of the beam. The load position can be eccentric or concen-
tric in the transverse direction. When eccentric, one single force is acting at the junction
of the upper flange and one web, with a value of 950 kN. When concentric, two forces
are acting at the junctions of the upper flange and the webs, with a value of 475 kN for
one force.
Two kinds of supports are considered. In the case of Support-type #1, the end cross-
sections are supported along the cross-section line against translation in the direction perpen-
dicular to the (flange or web) plate. In the case of Support-type #2 only the junction points of
the webs and the lower flange are supported in the vertical and horizontal (transverse)
directions.
Linear static analyses have been performed, then the displaced shapes are identified: the
participations from the global, distortional, local, and other modes are calculated. Some
results are shown in Figure 1, both deformed shapes and participation plots. Since the mode
identification is performed cross-section by cross-section, in the participation plots the partici-
pations are shown as the function of the cross-section position along the length.
As can be expected, the behavior is dominated by global deformations. However, non-
negligible participations from distortional and local deformations can also be observed. It is
visible that the transverse stiffeners trigger local deformations, but the amount of local
deformations is strongly dependent on the parameters of the problem: for example local
deformations are much larger in the case of Support-type #1 compared to those of Support-
type #2. The amount and distribution of distortional deformations are also problem-
dependent: in the case of Support-type #1 the distortional participation is nearly constant
along the member length, up to 20–30%; while in the case of Support-type #2 the distortional
deformations are more pronounced in the vicinity of supports, and in general their amount is
smaller.
Finally, it is to observe that the eccentric or concentric load position has significant
effect: the last row of Figure 1 shows that the behavior is almost exclusively global if the
load is concentric, i.e., when there is no direct twisting effect from the load. In this case
even the presence of transverse stiffeners has little effect on the modal identification of
the deformations.

501
Figure 1. Sample displacement shapes and identification curves for the box-girder beam example.

502
4 PURE DISTORTIONAL BUCKLING OF HOLLOW-FLANGE MEMBERS

In this Section, hollow flange members are analyzed. Hollow flange beams have been exten-
sively studied during the recent years, due to their potential favorable behavior, and light
weight. From modal behavior point of view, hollow flange members have not been handled
by older modal decomposition methods, but by now they can also be analyzed, e.g., by the
cFEM, too. The here considered member is similar to that investigated in Anapayan et al.
(2011). The hollow flange C section dimensions are as follows: the total height is 250 mm, the
width is 75 mm, the depth of the hollow flange is 25 mm and the thickness is 2.5 mm. The
length of the member is 2 m. At each end of the member a 200×75 mm end-plate with 2.5 mm
thickness is applied. One single inner stiffener with the same dimension is also applied in vari-
ous position along the length of the member (the stiffener is 100mm to 1000mm distance from
one end of the section). The member is simple supported at both ends, according to Support-
type #1 in Section 3, i.e., the end cross-sections are supported along the cross-section line
against translation in the direction perpendicular to the (flange or web) plates. The material is
steel, as in the previous example.
Both bending problem (beam) and compression problem (column) have been analyzed. For
the beam, a concentrated force is applied at the middle of the beam (i.e., at mid-length of the
beam and at mid-height of the web). For the column, problem two compression forces have
been applied at the two ends, opposite in direction, uniformly distributed over the cross-
sections.
The linear buckling problems have been solved by applying various constraints, determining
the critical load values and the corresponding buckled shapes. Here only the first pure distor-
tional buckling modes are shown, see Figure 2. The results are shown for some selected posi-
tions of the single web stiffener.
The figures and numerical results show that if the stiffener is located close to a member end,
its effect is moderate. However, if it is located somewhere in the middle, its stabilizing effect is
(or can be) much more important, since the buckling length is forced to be relatively small.
The presented example also shows that the behavior is significantly influenced by the load-
ing. While in the case of pure compression the critical load is increasing by a maximum 60%,
the maximum increase in the actual beam problem is found to be more than 300%. In Figure 3
the critical values (relative to the case without transverse stiffeners) are shown in the function
of the position of the transverse stiffener along the beam or column length.

5 PURE LOCAL BUCKLING OF I-SECTION MEMBERS

In this Section the buckling behavior of an I-section beam is studied, with and without trans-
verse stiffeners. The I-section is symmetrical, similar to an IPE 300 profile, the flange size
being 150×14 mm, and web size being 300×10 mm. At each beam end there is an end-plate
with 150×300×10 mm dimensions. If stiffeners are employed, the stiffener are double-sided,
with 75×300 mm dimensions and with 10 mm thickness The length of the beam is 1 m, the
number (pair of) of stiffeners is varying from zero to 10. The stiffeners are equally spaced
along the member length. The material is steel, as in the previous examples.
The loading is one concentrated force, acting in the middle of the beam, at the junction of
the upper flange and the web. In the presented examples the ends are supported according to
Support-type #1 in Section 3, i.e., the end cross-sections are supported along the cross-section
line against translation in the direction perpendicular to the (flange or web) plate.
Linear buckling analyses have been performed, by applying various constraints. Here only
pure local buckling results are shown, see Figure 4, where selected buckled shapes are shown,
together with the corresponding critical force values. The effect of the second-order strain terms
(i.e., the elements of the Green-Lagrange strain matrix) is demonstrated. Pure local-plate buck-
ling solutions are calculated by systematically switching on and off the second-order terms of
the longitudinal strain, of the transverse strain, and of the shear strain. This might also be inter-
preted as if only the longitudinal normal stress (Sigx), or only the transverse normal stress

503
Figure 2. Pure distortional buckling results of hollow-flange members with web stiffeners.

Figure 3. The effect of stiffener position on pure distortional buckling of hollow-flange members.

504
Figure 4. Pure local buckling results of I-section beam with web stiffeners under mid-point force.

Figure 5. The effect of the number of stiffeners on pure local buckling critical load.

505
Table 1. Pure local buckling critical loads in various options.
‘Sigx’ ‘Sigy’ ‘Tauxy’ ‘All-stress’

Number of
Fcr [kN] relative Fcr [kN] relative Fcr [kN] relative Fcr [kN] relative
stiffeners

0 15502 100.0% 2419 100.0% 9419 100.0% 2466 100.0%


1 18122 116.9% 23159 957.4% 11166 118.5% 9874 400.5%
2 16600 107.1% 4313 178.3% 14422 153.1% 4170 169.1%
3 18942 122.2% 23694 979.5% 18112 192.3% 12674 514.0%
4 20067 129.4% 7340 303.5% 25042 265.9% 6893 279.5%
5 22392 144.4% 24982 1032.8% 31360 333.0% 16253 659.2%
6 24737 159.6% 11020 455.6% 39949 424.1% 10265 416.3%
7 27214 175.5% 26294 1087.1% 47988 509.5% 20264 821.8%
8 30197 194.8% 15303 632.7% 58789 624.2% 14229 577.1%
9 33032 213.1% 27531 1138.2% 63943 678.9% 23911 969.7%
10 36518 235.6% 20183 834.4% 81944 870.0% 18684 757.7%

(Sigy), or the in-plane shear stress (Tauxy) had been considered in the determination of the geo-
metric stiffness matrix of the generalized eigen-value problem, see Equation 4. It is to underline
that the results presented in Figure 4 are all first buckling modes of the constrained buckling
analysis, which demonstrates the efficiency and customizability of the cFEM method.
By looking at the All-stress options, there seems to be two different scenarios when the web
stiffeners are applied, depending on whether there is a stiffener right under the load or not. If
there is a stiffener under the load, the buckling looks like a shear buckling of the web panels
(due to the fact that the analyzed beam is fairly short). If there is no stiffener under the load,
buckling is concentrated under the load, which phenomenon is usually described as web crip-
pling. The two basic scenarios can also be observed in Figure 5 and in Table 1, where the crit-
ical forces are given in the function of the number of stiffeners. The critical load has an
increasing tendency with the number of stiffeners, but the graph shows a fluctuation depend-
ing on whether the number is odd (i.e., there is stiffener under the load), or even (i.e., there is
no stiffener under the load). Evidently, the higher critical load values belong to the odd
number of (equally spaced) stiffeners.
The figures also show the cases when only certain stress (or strain) components are con-
sidered. With this special feature of cFEM the local buckling behavior can be understood in
a more detailed way. As Figure 5 proves, the exact location of the transverse stiffeners is cru-
cially important from the viewpoint of web crippling (i.e., Sigy results), but does not seem to
be important from the viewpoint of buckling due to longitudinal normal stresses (Sigy results),
and from the viewpoint of shear buckling. It is also clear that the transverse stiffeners have
relatively small effect on the Sigy buckling results, i.e., not more than 20–30% (unless the stiff-
eners are extremely closely spaced). On the other hand, transverse stiffeners are very effective
against shear buckling.

6 SUMMARY

In this paper a recent extension of the constrained finite element method is reported. The
method is essentially a shell finite element method, but mechanical constraints can be applied
to enforce the analysed member to deform in accordance with global, distortional or local
deformations. The method already proved itself to be the most general among similar methods
with modal decomposition ability, but lately it has further been generalized by extending its
applicability to members with extra transverse plates such as end-plates of transverse stiff-
eners. In this paper the basic concept of considering the extra plates is briefly presented, then
proof-of-concept examples are provided. The examples include linear buckling analysis of

506
beam and column members with end-plates and transverse stiffeners, with I-shapes cross-
sections as well as with hollow-flange cross-sections. Moreover, linear static analyses of box-
girder beams with transverse stiffeners have been performed, then the displaced shapes are
identified. The results prove the applicability of cFEM for members with stiffeners, as well as
demonstrate the advantages of modal decomposition in understanding the complex behaviour
of thin-walled members.

REFERENCES

Ádány, S., Schafer, B.W. 2014a. Generalized constrained finite strip method for thin-walled members
with arbitrary cross-section: Primary modes, Thin-Walled Structures, Vol 84, pp. 150–169.
Ádány, S., Schafer, B.W. 2014b. Generalized constrained finite strip method for thin-walled members
with arbitrary cross-section: Secondary modes, orthogonality, examples, Thin-Walled Structures, Vol
84, pp. 123–133.
Ádány, S. 2018a. Modal identification of thin-walled members with and without holes by using CFEM,
Proceedings of the ICTWS 2018 conference, July 24–27, 2018. Lisbon, Portugal.
Ádány, S. 2018b. Constrained shell Finite Element Method for thin-walled members, Part1: constraints
for a single band of finite elements, Thin-Walled Structures, Vol 128, July 2018, pp. 43–55.
Ádány, S., Visy, D. & Nagy, R. 2018. Constrained shell Finite Element Method, Part 2: application to
linear buckling analysis of thin-walled members, Thin-Walled Structures, Vol 128, July 2018, pp.
56–70.
Anapayan, T., Mahendran, M., Mahaarachchi, D. 2011. Lateral-torsional buckling test of a new hollow
flange cjannel beam, Thin-Walled Structures, Vol 49, pp. 13–25.
Bebiano, R., Basaglia, C., Camotim, D., Gonçalves, R., 2018. GBT buckling analysis of generally loaded
thin-walled members with arbitrary flat-walled cross-sections, Thin-Walled Structures, Vol 123, pp.
11–24.

507
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Experimental verification of shear connection of thin-walled steel


built-up members

M. Horáček, J. Melcher & O. Ceh


Faculty of Civil Engineering, Institute of Metal and Timber Structures, Brno University of Technology,
Brno, Czech Republic

ABSTRACT: The current trend in the design of steel structures leads, due to the saving of
the material, to the frequent use of thin-walled cold formed steel sections. The majority of
thin-walled steel profiles has a mono symmetrical cross-section, with the bending stiffness of
the cross-section to the symmetry axis often higher than the bending stiffness to an axis per-
pendicular to the axis of symmetry. By a suitable choice of the cross-sectional shapes and their
shear connection (very often by means of bolted joints), the built-up thin-walled cross-sections
can be created. These built-up sections are usually double symmetrical and their bending stiff-
nesses to the axes of symmetry are of same order. The subject of this article is an experimental
investigation and evaluation of the shear connection of a pair of profiles with different
number of fasteners and their arrangement.

1 INTRODUCTION

This paper focuses on the problem of the bended built-up thin-walled steel profiles and their use
in the load bearing steel structures. The majority of thin-walled steel profiles has a mono sym-
metrical cross-section, with the bending stiffness of the cross-section to the symmetry axis often
higher than the bending stiffness to an axis perpendicular to the axis of symmetry. By a suitable
choice of the cross-sectional shapes and their shear connection (very often by means of bolted
joints), the built-up thin-walled cross-sections can be created. These built-up sections are usually
double symmetrical and their bending stiffnesses to the axes of symmetry are of same order. In
these cases, the question of ensuring full or partial shear connection of the single profiles arises.
For the investigation of the effect of ensuring the full or partial shear connection in the
built-up profiles consisting of two chords the thin-walled cold formed Sigma sections are used
(see Figure 1).

Figure 1. Cross section of single Sigma profile and built-up profile consisting of 2 Sigma profiles.

508
Table 1. Cross-sectional characteristics.
Characteristic Unit 2 × Single profiles (not connected) Built-up profile
2
Area [mm ] 2 × 736 = 1 472 1 472
Second moment of area to y axis [mm4] 2 × 2 293 850 = 4 587 700 4 587 700
Second moment of area to z axis [mm4] 2 × 193 550 = 387 100 1 108 400
Elastic section modulus to y axis [mm3] 2 × 30 585 = 61 170 61 170
Elastic section modulus to z axis [mm3] 2 × 6 048 = 12 096 20 405

The profile height is 150 mm, flange width is 50 mm and the material thickness is 2.5 mm.
The used Sigma profiles have continuous perforation in the web (2 lines of holes Ø11 mm,
spacing 50 mm). As a shear connection the slip-resistance bolted connection with M10 bolts
of grade 10.9 is used. The gap between the profiles is secured by the steel plates of thickness
6 mm. The bolts are tightened with the torque of magnitude 72 Nm, which causes in each bolt
the preloading force of magnitude 35.6 kN.

2 THEORETICAL ANALYSIS

2.1 Cross-sectional properties


In Table 1 the calculated cross-sectional properties of 2 single Sigma profiles (not connected)
and built-up profile considering the full shear connection between the profiles are listed. In
case of 2 single Sigma profile the ratio between the second moments of area to y and z axis is
approximately 1:12, by creating the built-up section this ratio changes to 1:4.

2.2 Material properties


Sigma beams are produced by cold forming from the steel sheets of declared steel grade
S355MC. The actual yield stress of the material determined from the tensile tests (coupon
tests) is 420 MPa.

3 EXPERIMENTAL VERIFICATION

3.1 Description of test set-up


The built-up profiles were tested as simple supported beams with span Lsupp = 2 850 mm
loaded by 4 point bending over the minor axis z. The load was realized by electro-hydraulic

Figure 2. Test set-up scheme.

509
Figure 3. Photo of test set-up.

cylinder. The force from cylinder is spread by means of stiff beam to the sections located
approximately in the thirds of the beam span (see test set-up scheme in Figure 2 and photo of
test set-up in Figure 3). At the beam ends (at the supports) and in the position of the load
application, the built-up profiles are stiffened by the bolted U profiles.

3.2 Measuring of deformations and stresses


During the loading, the vertical beam deflection and stresses in the middle of the beam span
are being measured. The vertical deflection is measured by means of the potentiometric draw-
wire sensor (Figure 4). Four strain gauges were installed over the built-up profile height
(Figure 5) for the measuring of the stress distribution in the beam mid span.

3.3 Process of loading


This paper summarized the results of the pilot tests. In total 4 loading cycles with 4 different
range of applied load were realized – first three in range of elastic stresses (Fmax = 5−10 kN),
the last one in the range of plastic stresses (Fmax = 16 kN). During the loading process the
relationship between the magnitude of applied force F and the vertical displacement in the
middle of the specimen span w is being recorded (Figure 6).

3.4 Bolts arrangement


In total 3 configurations, each with different bolts arrangement, were tested (see Figure 7). In
all configurations the bolts were installed also in the middle third of the beam span, where is
no shear (bolts are not loaded with the longitudinal shear, bolts are used only to secure the
geometry of the cross-section of built-up profile in the middle third of the beam span where is
the highest value of the bending moment).

510
Figure 4. Position of the potentiometric draw-wire sensor for vertical deflection measuring.

4 TEST RESULTS

4.1 Evaluation of the stress distribution


The Figure 8 shows the stress distribution in the profile mid span section measured in four
points over the built-up profile height (for description of the strain gauges T1–T4 position see
Figure 5).
From the graph follows, that within the load range 0–10 kN (loading cycles 1–3), the stres-
ses in the outer fibers are in the elastic zone and correspond to the stresses calculated accord-
ing to the theory of elastic behavior of the member subjected to the bending (with considering
section moduli for the most tensioned and compressed fibers of built-up section with full
shear connection). Within load range 0–10 kN the full shear connection between the coupled
profiles is sufficiently secured in all three configurations with different number of bolts and
their arrangement.

Figure 5. Position of strain gauges installed in the beam mid span.

511
Figure 6. Graph with F-w relationship.

In the case of applied load exceeding 10 kN, the stress in the outer fibers starts
increasing more rapidly. Since the stresses are still within the elastic zone, it indicates
the insufficient shear connection (the full shear connection between the profiles is not
secured for applied load exceeding 10 kN). The sudden drops in the stresses (first drop
approximately at load level 10.7 kN) point out the slip between the top and bottom pro-
file of the built-up section, which means slip of the slip-resistance bolted connection with
M10 bolts of steel grade 10.9.

4.2 Evaluation of the shear resistant of the slip-resistance bolted connection


From the test data follows, that the failure of the slip-resistance bolted connection occurred at
the load level 10.7 kN, which cause the constant shear force of magnitude 5.35 kN at the
outer thirds of the specimen span. The value of the longitudinal shear force between bottom
and top profile per millimeter length of the profile can be calculated as:

Vy  Sy 5:35  103  ð736  21Þ


Vl ¼ ¼ ¼ 74:6 N=mm ð1Þ
Iz 1; 108; 400

Figure 7. Description of bolts arrangement in three different configurations.

512
Figure 8. Test results (graph with σ-F relationship).

The total longitudinal shear force calculated for the outer third of specimen span is:

Fl ¼ Vl  l1=3 ¼ 74:6  925 ¼ 69; 005 N ¼ 69:0 kN ð2Þ

In the configuration 3 (this configuration of bolts arrangement was tested in the 4th loading
cycle) in total 6 bolts in each outer third of the beam span are installed). The shear force
acting on one bolt is:

Fs ¼ Fl =nb  ¼ 69:0=6 ¼ 11:5 kN ð3Þ

The characteristic slip resistance of the slip-resistance connection with M10 10.9 bolts can be
determined according to EN 1993-1-8 considering one friction surface and friction coefficient
μ = 0.2 (friction surface D = surface without treatment) as:

Fs;Rk ¼ ks  n  μ  Fp;C ¼ 1:0  1  0:2  35; 655 ¼ 7:13 kN; ð4Þ

The preloading force Fp,C = 35,6 kN is corresponding to the applied tightening torque of
magnitude 72 Nm. From the comparison of the characteristic slip resistance Fs,Rk and the
shear force Fs, acting on one bolt at the moment when the slip occurred during the testing, is
obvious, that the real slip resistance of one bolt is higher than the resistance calculated in
equation (4). The real slip resistance of one bolt is corresponding to the resistance calculated
according to equation (4) considering friction coefficient μ = 0.32. This value of friction coeffi-
cient conforms to the friction surface C (surfaced with removed rust). Since the tested Sigma
profiles were without any surface coating and they were stored in the dry conditions, the sur-
face of the tested specimens was without any rust and therefore the surface conditions meet
the definition of the friction surface C.

513
5 CONCLUSIONS

This article focuses on the possibility of creating a built-up sections composing of two single
profiles, screwed together by means of bolts designed as slip-resistance connections. An
advantage of these built-up sections is that these sections are mostly double symmetrical and
their bending stiffnesses to the axes of symmetry are of same order.
From the carried out pilot tests follows, that within the range of elastic stresses the full
shear connection between the single profiles can be achieved by reasonable number of slip-
resistance bolted connections without the need for special surface treatment (f.e. blasting).
This may in some specific cases effectively increase the buckling resistance of the compressed
members or the lateral torsional buckling resistance of the members subjected to the bending.
In the next phase of the testing program the torsional characteristics of these built-up sections
can be experimentally examined, especially due to the reason that the built-up section created
from two single Sigma profiles are partly closed sections, which have significantly higher tor-
sion stiffness than two single open sections.

ACKNOWLEDGMENT

This paper has been elaborated within the support of the project No FAST-S-18-5550 and
project reg. nr. LO1408 “AdMaS UP – Advanced Materials, Structures and Technologies,
supported by the National Sustainability Program I of the Ministry of Education, Youth and
Sports of the Czech Republic

REFERENCES

Wang, L. & Young, B. Beam tests of cold-formed steel built-up sections with web perforations. In Jour-
nal of Constructional Steel Research: Vol. 115, 18–33, December 2015.
Wang, L. & Young, B. Design of cold-formed steel built-up sections with web perforations subjected to
bending. In Thin-Walled Structures: Vol. 120, 458–469, November 2017.
Young, B. & Chen, J. Design of cold-formed steel built-up closed sections with intermediate stiffeners. In
Journal of Structural Engineering: Vol. 134, Issue 5, 727–737, 2008.
Abbasi, M., Khezri, M., Rasmussen, K. J. R. & Schafer, B.W. Elastic buckling analysis of cold-formed
steel built-up sections with discrete fasteners using the com-pound strip method. In Thin-Walled Struc-
tures: Vol. 124, 58–71, March 2018.
Fratamico, D.C., Torabian, S., Zhao, X., Rasmussen, K.J.R. & Schafer, B.W. Experiments on the global
buckling and collapse of built-up cold-formed steel columns. In Journal of Constructional Steel
Research: Vol. 144, 65–80, May 2018.
Vlasov, V.Z. 1962. Tenkostěnné pružné pruty (Thin-Walled Elastic Bars). Prague: National Publishing of
Technical Literature (SNTL).
Březina, V. 1962. Vzpzěrná únosnost kovových prutů a nosníků (Buckling Resistance of Metal Members).,
Prague: Czechoslovak Academy of Science Publishing
ČSN EN 1993-1-1Design of steel structures - Part 1-1: General rules and rules for buildings, Prague:
Czech standards institute, 2008.
ČSN EN 1993-1-8Design of steel structures - Part 1-8: Design of joints, Prague: Czech standards institute,
2006.

514
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Buckling analysis of circular arches with trapezoidal


corrugated web

J.R. Ibañez, R. Díez, C. López & M.A. Serna


Department of Structural and Mechanical Engineering, University of Cantabria, Spain

ABSTRACT: Circular arches with I-section may be produced by either curving steel profiles
or by cutting a curved web. An economic alternative solution to bending or cutting is to use
corrugated webs. Since a long rectangular corrugated plate may easily be curved, no cutting
process is needed and no material lost. The paper presents a FEM study on the stability of
a circular arch with trapezoidal corrugated web. Both non-laterally constraint arch and lat-
erally constraint arch cases are considered under two load types: gravitational load and pres-
sure load. The study is performed for two trapezoidal shapes, two flange thicknesses and
a wide range of web thicknesses. First buckling mode and the corresponding buckling load are
obtained in each analysis and results are compared with those obtained for a planar web arch.
The paper show that corrugated webs significantly improve web buckling, lateral torsional
buckling and out-of-plane flexural buckling resistances.

1 INTRODUCTION

Steel arches have structural applications in architectural buildings, industrial constructions


and bridge design. Arches with open I cross section may be produced by curving steel profiles.
When due to cross section dimensions the curving process is no feasible or costly efficient, the
arch morphology may be obtained by cutting a curved web that is then welded to curved
flanges. An alternative solution to bending or cutting is to use corrugated webs. Since, due to
its geometry, corrugated plates have almost no rigidity to axial forces, a long rectangular cor-
rugated plate may be easily curved and used as web, with no material lost as is the case when
using cutting to obtain curved planar webs. Consequently, corrugated webs may have a clear
economic benefit in the arch fabrication process.
Form a structural point of view, it is well know that corrugated webs significantly improve
local buckling resistance in girders subjected to bending. In fact, its behavior may be assimilated
to planar plate webs with intermediate vertical stiffeners. Hasan et al (2017) have presented
a state of the art paper that shows the increasing research interest on beams with corrugated webs
and the available different formulations. Elgaaly et al. (1997) and Elgaaly & Seshadri (1998) have
provided experimental results for built-up girders with trapezoidal corrugated webs. From
a theoretical perspective, Abbas et al. (2006, 2007a, 2007b) have studied the stress state of I-gir-
ders and proposed a direct analytical method for their analysis. Jáger et al. (2015, 2017a, 2017b
and 2017c) have developed a complete experimental and theoretical research on the interaction of
shear, bending and path loading. The behavior of bridge girders with corrugated webs have been
analyzed by Hassanein & Kharoob (2013a and 2013b). Finally, Leblouba et al. (2017a and
2017b) have presented a normalized shear strength formulation for trapezoidal corrugated webs.
Recently, the use of corrugated webs has been extended to arches. Based on previous
research developed by Bradford & Pi (2004), a team integrated by Chinese and Australian
researches have investigated the in-plane behavior of steel I-section circular arches with sinus-
oidal corrugated webs. In this context, Guo et al. (2016a and 2016b) have considered a pin-
ended circular arch under a uniform radial load and a uniform vertical load. Complementary,
Chen et al. (2017) have extended the previous theoretical and experimental results to arches

515
subjected to combine in-plane loads and have proposed a set of design equations that provide
lower bound predictions for arch strength.
In what follows, the paper present a buckling analysis of steel I-section circular arches
with trapezoidal corrugated web under uniform gravitational load and uniform pressure
load. Two trapezoidal shapes, with 40 mm and 80 mm trapezoidal wave depth, are con-
sidered. The buckling analysis is performed using FEM and considers both non-laterally
constraint and laterally constraint arches. First, the arch geometry and load cases are
described. Then, the different basic buckling modes are depicted. Finally, buckling results
for arches with corrugate webs are presented and compared with those obtained for arches
with planar web. Buckling load and buckling mode are obtained for a range of web thick-
nesses and three flange dimensions.

2 ARCH GEOMETRY AND LOAD CASES

Steel I-section circular arches with trapezoidal corrugated web are defined by means of the
following elements: a) circular medium line (Figure 1), b) I-section (Figure 2.a) and c) trapez-
oidal wave shape (Figure 2.b).
Our study deals with a circular arch geometry defined by a total length L = 10000 mm and
an angle Q = 120º. Consequently, R = 4775 mm, H = 8270 mm, and f = 2387 mm. Cross section
is defined by flange plate dimensions (b = 140 mm and tf = 12, 24 and 36 mm) and h = 412 mm,
where h is the flange center lines distance. For the corrugated web two trapezoidal wave
shapes are studied, with a geometry defined by a1 = 120 mm, a2 = 80 mm, and d = 40 mm and
80 mm. Results for trapezoidal web arch (Figure 3) are compared with those obtained for
a circular arch with planar web.
Two load cases are considered in our buckling analysis, with the load applied on the
upper flange. The first case corresponds to a uniform gravitational load (Figure 4). The
arch is supposed to have clamped end points. Figures 5–7 show internal forces distribu-
tions for an applied load equal to 1 kN/m. It can be seen that compression forces are the
most significant internal forces, reaching a value of 6.61 kN at supports. Nevertheless,
shear forces and bending moments are not negligible and the arch is prone to lateral tor-
sional buckling.
The second load case corresponds to a uniform pressure load (Figure 8). Internal forces for
a pressure load equal to 1 kN/m are shown in Figures 9–11. It can be seen that bending effects
are almost negligible. Consequently, out-of-plane buckling for this load case would relate to
either torsional buckling or flexural buckling, but no to lateral torsional buckling.

Figure 1. Circular medium line. Figure 2. I-section and trapezoidal wave shape.

516
Figure 3. I-section circular arch with trapezoidal Figure 4. Gravitational (G) load case.
corrugated web.

Figure 5. Compression forces for G load. Figure 6. Shear forces for G load.

Figure 7. Bending moments for G load. Figure 8. Pressure (P) load case.

Figure 9. Compression forces for P load. Figure 10. Shear forces for P load.

Figure 11. Bending moments for P load.

3 BASIC BUCKLING MODES

The first arch buckling mode depends on the web and flange thicknesses, the I-section geometry,
the type of load and the lateral constraint. Six basic buckling modes may develop. The two first
correspond to web local buckling (Figure 12) and flange local buckling (Figures 13). These local
buckling modes may be present in different arch locations depending on the inner forces distribu-
tion. In the transition from web buckling to flange buckling a combined mode may be present.
Non-laterally constraint arches may develop three global buckling modes: torsional buck-
ling (Figure 14), lateral torsional buckling (Figure 15) and out-of-plan flexural buckling
(Figure 16). On the other hand, if the arch is laterally constraint it may globally buckle

517
Figure 12. Web local buckling. Figure 13. Flange local buckling.

Figure 14. Torsional buckling. Figure 15. Lateral torsional buckling.

Figure 17. In-plane flexural buckling.


Figure 16. Out-of-plane flexural buckling.

following an in-plane flexural buckling mode (Figure 17). As with local buckling modes,
global buckling may have transition zones where a mode combination appears.

4 ARCH BUCKLING RESULTS

More than 800 elastic buckling analysis have been performed using ANSYS Academic Mech-
anical finite element software. Half of these analysis corresponds to non-laterally constraint
arches (NLC) and half to laterally constraint arches (LC). Figures 18–29 summarize buckling
results. In figure captions, gravitational load and pressure load are referred to as G and
P respectively. As indicate above, three web morphologies are considered: planar web (PW),
40 mm depth corrugated web (CW40), and 80 mm depth corrugated web (CW80). First buck-
ling mode is indicated on the figures using the following codes: WB (web buckling), FB
(flange buckling), TB (torsional buckling), LTB (lateral torsional buckling), OPFB (out-of-
plane flexural buckling), and IPFB (in-plane flexural buckling).

4.1 Non laterally constraint arch


Figures 18 and 19 show results for non-laterally constraint arches with a flange thickness of
12 mm subjected to gravitational load (NLC-GL-12) and pressure load (NLC-PL-12),

518
respectively. As expected, plate corrugation significantly increases web buckling resistance,
which is the first buckling mode for small web thickness. It can be seen that gravitational load
leads to lateral torsional buckling due to the presence of significant bending moments. On the
other hand, pressure load, which implies significant arch compression, produces out-of-plane
flexural buckling. Moreover, for specific values of planar web thickness, the arch develops tor-
sional buckling, something not present in arches with corrugated webs.
Similar results are obtained for I-sections with flange thickness of 24 mm (Figures 20 and 21)
and 36 mm (Figures 22 and 23). In all cases, buckling loads have been obtained for aches with
a range of web thickness reaching the flange thickness value. Each curve mark on figures corres-
ponds to a FEM analysis.
In general, buckling resistance increases with corrugation depth. Arches with trapezoidal
corrugated web show more buckling resistance than arches with planar web. However, the dif-
ference diminishes with web thickness and may even be in favor of planar webs in some par-
ticular cases for high values of web thickness.
Web buckling significantly increases with corrugation. The relation between buckling loads
for corrugated webs and buckling loads for planar webs is in the order of h2/a12 as can be
expected from the general expression of buckling stress for compressed plates. More corruga-
tion depth slightly increases buckling loads due to a little improvement in boundary conditions.
Bending moment, which are present when gravitational loads are applied, locally increases
compression stresses. Consequently, corrugated webs, which develop small size web buckling
waves, are negatively affected. On the other hand, the presence of bending moment have
a positive effect on planar webs.
Regarding lateral torsional buckling and out-of-plane flexural buckling, the benefit of cor-
rugation is related to the improvement on cross section properties such as torsional constant

Figure 18. Buckling loads for NLC-G-12. Figure 19. Buckling loads for NLC-P-12.

Figure 20. Buckling loads for NLC-G-24. Figure 21. Buckling loads for NLC-P-24.

519
Figure 22. Buckling loads for NLC-G-36. Figure 23. Buckling loads for NLC-P-36.

and week axis moment of inertia. The relative improvement diminishes with flange thickness.
A quantification of the corrugation effect needs to be determined.

4.2 Laterally constraint arch


When arches have their lateral displacement prevented, no longer torsional buckling, lateral-
torsional buckling, and out-of-plane flexural buckling are present. Figures 24 and 25 show
buckling load results for arches with a flange thickness of 12 mm. As with non-laterally con-
straint arches, the first buckling mode corresponds to web buckling (WB), with significantly
higher buckling loads for trapezoidal corrugated webs. Then, arches with corrugated webs
develop flange buckling (FB), something which is not present in arches with planar web. For
laterally constraint arches, flange buckling loads diminishes with the increase of corrugation
depth. This is due to the fact that flange buckling length increases with corrugation depth.
As with the previous case, when flange thickness is equal to 24 mm, arches with planar web
only develop web buckling (Figures 26 and 27). On the other hand, as web thickness increases,
arches with corrugated web first develop local web buckling, then local flange buckling and,
finally, global in-plane flexural buckling (IPFB).
Finally, Figures 28 and 29 show the buckling loads for arches with a flange thickness equal to
36 mm. It can be seen that arches with thick planar web develop global in-plane flexural buckling
with a buckling load bigger than that obtained for corrugated webs. This result comes from the
fact that planar web have more compression capacity than corrugated webs. Additionally, arches
with corrugated web under pressure load do not develop flange buckling (Figure 29).
For almost the entire range of web thickness, web buckling is the buckling mode present in
laterally constraint planar web arches. Consequently, the benefit of using corrugated webs is
very significant for laterally constraint arches.

Figure 24. Buckling loads for LC-G-12. Figure 25. Buckling loads for LC-P-12.

520
Figure 26. Buckling loads for LC-GL-24. Figure 27. Buckling loads for LC-PL-24.

Figure 28. Buckling loads for LC-GL-36. Figure 29. Buckling loads for LC-PL-36.

5 CONCLUSIONS

Buckling loads and buckling modes for circular I-section arches with trapezoidal corrugated web
have been studied. The analysis has considered two load cases: gravitational load and pressure
load. Buckling loads have been obtained for two trapezoidal depth: 40 mm and 80 mm. The results
for corrugated webs have been compared with those obtained for I-section arch with planar web.
The use of trapezoidal corrugated plates have shown to be beneficial in three main cases.
First, when web buckling is the first buckling mode, which happens to be the case for all arches
with relatively thin webs, buckling resistance significantly increases with corrugation. Of course,
this increase comes primarily from the fact that buckling loads of plates are inversely propor-
tional to the square of plate width. Corrugation of webs forces buckling mode to be localized
on each of the planes that integrate the trapezoidal wave. Additionally, corrugation introduces
more rigid boundary conditions. With the I-section dimensions used in the paper, corrugated
web buckling loads are 20 to 30 times higher that planar web buckling loads.
The second benefit of using corrugated webs is related to lateral torsional buckling of
non-laterally constraint arches under gravitational loads. This benefit is related to the improve-
ment of I-section torsional properties with corrugation. However, this improvement diminishes
with the increase of web thickness and may even be negative in some cases for thick flanges.
For the particular case of a web thickness of 6 mm, the increase of lateral torsional buckling
varies from 25% to 39% for a corrugation depth of 80 mm, and from 16% to 26% for 40 mm.
For pressure load, where bending is not present and the arch is subjected to compression,
the main global buckling mode is the out-of-plane flexural buckling. The study has shown
that corrugated webs also improve the arch resistance to this buckling mode. As with the lat-
eral torsional buckling, the improvement produced by corrugation diminishes with flange

521
thickness. Nevertheless, for a web thickness of 6 mm the increase is superior to 50% for
a corrugation depth of 80 mm, and to 40% for 40 mm.
For small enough plate thicknesses, laterally constraint arches develop web local buckling, as
in non-laterally constraint arches, and flange local buckling. Again, corrugated webs produce
a significant increase in buckling loads. However, when local buckling is not present, in-plane
flexural buckling is the first buckling mode. Contrary to lateral torsional buckling and out-of-
plane flexural buckling, in-plane flexural buckling is better resisted by arches with plane webs,
and corrugation has a negative effect. Moreover, the studied cases have shown that corrugation
depth of 40 mm leads to a higher buckling load than corrugation depth of 80 mm.
In general, the use of trapezoidal corrugated webs produces a significant improvement in
arch buckling resistance for non-laterally constraint arches, particularly in what refers to web
buckling. Only for in-plane flexural buckling planar webs have shown better results than cor-
rugated webs. However, in-plane flexural buckling is only present with relatively thick web
plates, which have questionable practical use.

REFERENCES

Abbas, H.H., Sause, R. & Driver, R.G. 2006. Behavior of corrugated web I-girders under in-plane loads.
Journal of Engineering Mechanics 132 (8): 806–814.
Abbas, H.H., Sause, R. & Driver, R.G. 2007a. Analysis of flange transverse bending of corrugated Web
I-girders under in-plane loads. Journal of Structural Engineering 133 (3): 347–355.
Abbas, H.H., Sause, R. & Driver, R.G. 2007b. Simplified analysis of flange transverse bending of corru-
gated web I-girders under in-plane moment and shear. Engineering Structures 29 (11): 2816–2824.
Bradford, M.A. & Pi, Y.-L. 2004. Design of steel arches against in-plane instability. Int. J. Appl. Mech.
Eng. 9 (1): 37–45.
Chen, H., Guo, Y.-L., Bradford, M.A., Pi, Y.-L. & Yuan, X. 2017. Strength and Design of Pin-Ended
Circular Arches with Sinusoidal Corrugated Web under Combined In-Plane Loads. Journal of Struc-
tural Engineering 143 (2): art. no. 04016158, 1–14.
Elgaaly, M., Seshadri, A. & Hamilton, R.W. 1997. Bending strength of steel beams with corrugated
webs. Journal of Structural Engineering 123 (6): 772–782.
Elgaaly, M. & Seshadri, A. 1998. Steel built-up girders with trapezoidally corrugated webs. Engineering
Journal 35 (1): 1–10.
Hasan, Q.A., Wan Badaruzzaman, W.H., Al-Zand, A.W. & Mutalib, A.A. 2017. The state of the art of
steel and steel concrete composite straight plate girder bridges. Thin-Walled Structures 119: 988–1020.
Hassanein, M.F. & Kharoob, O.F. 2013a. Behavior of bridge girders with corrugated webs: (I) real
boundary conditions at the juncture of the web and flanges. Engineering Structures 57: 554–564.
Hassanein, M.F. & Kharoob, O.F. 2013b. Behavior of bridge girders with corrugated webs: (II) shear
strength and design. Engineering Structures 57: 544–553.
Guo, Y.-L., Chen, H., Pi, Y.-L. & Bradford, M.A. 2016a. In-plane strength of steel arches with
a sinusoidal corrugated web under a full-span uniform vertical load: Experimental and numerical
investigations. Engineering Structures 110: 105–115.
Guo, Y.-L., Chen, H., Pi, Y.-L., Dou, C. & Bradford, M.A. 2016b. In-plane failure mechanism and
strength of pin-ended steel I-section circular arches with sinusoidal corrugated web. Journal of Struc-
tural Engineering 142 (2): art. no. 04015121, 1–15.
Jáger, B., Dunai, L. & Kövesdi, B. 2015. Girders with trapezoidally corrugated webs subjected by com-
bination of bending, shear and path loading. Thin-Walled Structures 96 (art. no. 4603): 227–239.
Jáger, B., Dunai, L. & Kövesdi, B. 2017a. Flange buckling behavior of girders with corrugated web Part
I: Experimental study. Thin-Walled Structures 118: 181–195.
Jáger, B., Dunai, L. & Kövesdi, B. 2017b. Flange buckling behavior of girders with corrugated web Part
II: Numerical study and design method development. Thin-Walled Structures 118: 238–252.
Jáger, B., Dunai, L. & Kövesdi, B. 2017c. Experimental investigation of the M-V-F interaction behavior
of girders with trapezoidally corrugated web. Engineering Structures 133: 49–58.
Leblouba, M., Barakat, S., Altoubat, S., Junaid, T.M. & Maalej, M. 2017a. Normalized shear strength
of trapezoidal corrugated steel webs. Journal of Constructional Steel Research 136: 75–90.
Leblouba, M., Junaid, M.T., Barakat, S., Altoubat, S. & Maalej, M. 2017b. Shear buckling and stress
distribution in trapezoidal web corrugated steel beams. Thin-Walled Structures 113: 13–26.

522
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Effect of stiffener position on buckling behavior of H-shaped steel


beam with upper flange restraint

N. Igawa & K. Ikarashi


Department of Architecture and Building Engineering, Tokyo Institute of Technology, Tokyo, Japan

ABSTRACT: When the displacement and rotation of the upper-flange are restrained by
a slab, etc. the beam cross-sectional web deforms with lateral buckling. Stiffener-plates posi-
tioned vertical to the material axis are effective not only for local buckling but also for lateral
buckling when the upper-flange is restrained. In this paper, a finite element method analysis is
used for H-shaped section beams with various cross-sectional shapes and lengths, and the
effective stiffening position for each buckling form was studied. It has been found that an
effective stiffening position is determined by checking the deformed shape of the beam without
the stiffener-plates. Further, a formula to evaluate the position was derived.

1 INTRODUCTION

So far, there have been many studies on lateral buckling. In Japan, design guidelines for steel
structure are used to set the spacing of the lateral stiffening, the stiffening rigidity, and the
stiffening force of the stiffeners. It is also widely known that the translation and rotation of
the upper part of the beam cross-section is continuously and intermittently restrained by the
floor slab or purlin members, and this restraint acts effectively against lateral buckling. In
recent years, the elastic lateral buckling strength and plastic deformation capacity of the
beams-restrained upper-flange were clarified using an energy method or static load test. More-
over, continuous restraint on the beam has been studied in various restraint forms: some stud-
ies clarified buckling behavior using the energy method; for example, a case of restraint on the
lower-flange of the beam receiving a lift force by the wind, or a case of the beam where the
cross-section is uniaxially symmetric I-shaped.
Further, using stiffener-plates is one of the effective methods of stiffening against the buck-
ling of members. By providing a stiffener in an appropriate direction according to the buck-
ling form of the member to be stiffened, it is possible to obtain stiffening effects such as an
increase in the buckling resistance of the member and an improvement in the plastic deform-
ation performance. Especially, when stiffener-plates are provided vertical to the material axis,
they act on web local buckling. This is because the rigidity of the web at the cross section of
the stiffening portion is increased. Therefore, these stiffener-plates provided in this direction
are ineffective for lateral buckling; the members twist without deformation of the cross-
sectional shape.
In the case of an upper-flange restraint, the lateral buckling mode changes the one with the
deformation of the cross-sectional web because only the lower-flange sways and rotates. For
this reason, stiffener-plates work effectively not only for plate local buckling but also for lat-
eral buckling. In order to obtain the stiffening effect, it is necessary to provide plates at appro-
priate places; however, few studies have investigated the relation between the stiffening
position and the stiffening effect, and this method of stiffening has not been established yet.
This study intends to clarify the influence of the stiffener-plates on the buckling behavior of
the restrained beams and to determine the stiffening position where the stiffening effect can be
obtained accurately.

523
2 FINITE ELEMENT METHOD ANALYSIS

Figure 1 shows the analytical model. In this study, a finite element method (FEM) analysis
program ABAQUS is employed. This model is a cantilevered H-shaped cross-section beam
and composed of 4-node shell elements. This model assumes the time of the horizontal
load action due to the earthquake. The boundary conditions of this model are listed in
Table 1. The right end is simply supported and a connector element at the junction of the
web and the lower-flange at this end connects two rigid bodies. Therefore, the web and
lowerflange rotate separately. In order to reproduce the upperflange restraint, all nodes on
the joining line between the web and the upper-flange are restrained. Further, when consid-
ering plate local buckling, the translation of the lower junctions of the web and flange are
constrained every other 3H, where H is the height of the beam’s cross-section. Only one
pair of stiffener-plates are set such they sandwich that the web from both sides and are
slightly thicker than one of the web. The important index in this study is Ls, which is the
parameter indicating the stiffener position and represents the distance from the fixed-end
to the stiffener-plates.
A large-deformation analysis is carried out using this model, and the case where lateral
buckling occurs and the case where local buckling occurs are examined separately. By this, the
influence of the stiffener-plates on plastic deformation capacity after elastic buckling and the
effective stiffener position are considered. In addition, Table 2 lists parameters employed in
this study.

Figure 1. Analytical model.

Table 1. Boundary restraint conditions of model.


Displacement Rotation

Restraint conditions ux uy uz θx θy θz

Fixed-end 1 1 1 1 1 1
Pin support-end 0 0 1 1 0 0
Upper-flange
0 0 1 1 1 0
restraint
Lateral stiffening 0 0 1 0 0 0

1 = (complete) restraint, 0 = free.

524
Table 2. Parameters used in this study.
bw Width of web
H Width of H-shaped beam (H = bw + tf)
bf Half the width of flange
tw Thickness of web
tf Thickness of flange
Ls Distance between fixed end and stiffener-plate
L Length of H-shaped beam
E Young’s module (N/mm2)
ν Poisson’s ratio (ν = 0.3)
Aw Section area of web ( = bwtw)
Af Section area of flange ( = 2bftf)
GJ Torsional rigidity of H-shaped beam
EIω Bending-torsional rigidity of H-shaped beam
Mp Full plastic moment of H-shaped beam (Nmm)
θp Rotation angle of H-shaped beam at full plastic moment (rad)
WF New width thickness ratio indicator
λb Slenderness ratio under upper-flange restraint (λb = (Mp/McrR)1/2)
McrR Elastic lateral buckling strength of restrained H-shaped beam (Nmm)
Mcre Elastic lateral buckling strength of normal H-shaped beam (Nmm)
TS The indicator for increase rate of elastic buckling strength by upper-flange restraint

3 EFFECT OF STIFFENER ON LATERAL BUCKLING OF H-SHAPED STEEL


BEAM WITH UPPER FLANGE RESTRAINT

3.1 Decision of the effective stiffener position


In this section, lateral buckling is considered. First, beams that undergo lateral buckling are
selected using (1), which classifies the buckling form of the restrained H-shaped beams. If
WF/λb is less than 1.4, this beam causes lateral buckling, and if it is larger than 1.4, it causes
plate local buckling. Coupled buckling is not considered in this study; therefore, the beams
representing WF/λb ≤ 1.30 are selected.

WF 51:4    Lateral buckling
ð1Þ
λb ≥ 1:4    Plate local buckling

The stiffening effect caused by the stiffenerplates installed at the lateral stiffening position as per
the Japanese guideline (Architectural Institute of Japan, 2010) is considered. In equation (2)
represents the interval of lateral buckling stiffening. In equation (2.a) is used because SS400
steel material is assumed in this study.
8  
> lb H lb M
>
< A  250 and i  65
> 1:0 
Mp
 0:5
f y
SS400   ð2:aÞ
>
> lb H lb M
>
:  375 and  95 0:5   1:0
Af iy Mp
8  
> lb H lb 
M
>
>  200 and  50 1:0   0:5
< Af iy Mp
SS490   ð2:bÞ
>
> lb H lb M
>
:  300 and  75 0:5   1:0
Af iy Mp

lb Interval of lateral buckling stiffening (mm)


iy Radius of gyration (mm)
M The smaller end bending moment of the stiffening section (Nmm)

525
Figure 2 shows the result of the analysis. This represents the comparison of the load dis-
placement relationship of the restrained beams due to the stiffening position. Position A is the
location obtained by calculating in Equation (2.a), and the effect of the stiffener-plates at that
position is small. Position B is the place where the largest deformation of the web-section of
the restraint beam (not stiffened) occurs, as shown in Figure 3(a). These pictures show the
state of web deformation when the strength of the restrained beam reaches 0.8Mmax,
after Mmax; directing arrows shown in Figure 2(a) presents the timing at 0.8Mmax. The state at
this time is defined as the “ultimate state” in this study.
From these results, in this scenario, the lateral stiffening position in the Japanese regulation
is not appropriate, and the position obtained by checking the deformed shape in the “ultimate
state” of the beam not stiffened by stiffener-plates is better for the stiffening position. This is
because the guideline assumes the lateral buckling stiffener, such as small beam sticks, and
stiffener-plates are not assumed. Further, this effective stiffening position B is defined by Lse,
representing the distance between the fixed end and the position.

3.2 Evaluating equation of the effective stiffener position for lateral buckling
In the previous section, the effective stiffening position is defined as the position where the web
plate deformation of the restrained, not stiffened beam, occurs. Therefore, in this section, the
evaluation equation for the effective stiffening position Lse is considered.
Indicator TS is explained first. equation (3) shows a part of the procedure for calculating
the elastic buckling strength of the restrained H-shaped beam. Mcre is the strength of the non-
restrained H-shaped beams, Mcrm’ is a minimum convergence value of the elastic lateral buck-
ling resistance considering web plate. Further, A1 and A2 are numerical values determined by
the boundary condition, but the details are omitted here. TS is the index that evaluates the
increase in the ratio of lateral buckling strength by continuous, complete restraint on the

Figure 2. Load displacement relationship.

Figure 3. The “Ultimate state” of the beams in Figure 2(a).

526
upper-flange, derived from approximating Mcrm’/Mcre (eq (4)). kβ is a numerical value based
on the boundary conditions on both ends, and Kw is the out-of-plane flexural rigidity of the
web; the large TS means that the beam is long (large L) and has a large Kw, that is, a beam
having a cross-section with a thick web and thin small flanges. Therefore, from (4), if TS is
a large value, the strength increase rate McrR/Mcre is also large. TS has a value of 1 to 20 for
a general-purpose cross-section and length.

McrR Mcrm 0
¼ A1 þ A 2 ð3Þ
Mcre Mcre
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Mcrm 0 Kw L2
≈  2 ¼ TS ð4Þ
Mcre GJ þ EIω π=kβ L

Figure 4 shows the Lse from a variety of beams obtained using the procedure mentioned previ-
ously. The vertical axis represents the numerical value obtained by dividing Lse by L and the
horizontal axis represents the index TS. It can be confirmed that the points showing the effective
stiffening position are good arranged well, and they are evaluated by TS.
The following eq (5) is the approximation of the plotted points, the plotted values, and the
approximated values show good agreement, as shown in (5). Further, the range on WF/λb
available for (5) is also set because the beams whose WF/λb values are small (0.7 or less) are
plotted apart from the approximate curve.
 
Lse 0:351 WF
¼ pffiffiffiffiffiffi 0:75   1:30 ð5Þ
L TS λb

3.3 Stiffening range


The method for calculating the effective stiffener position is suggested. In practice, however, it
may not be possible to set the stiffener in the position to be derived, for the convenience of
construction. Therefore, in this section, the stiffening range for stiffener-plates is examined to
demonstrate an adequate stiffening effect. Based on the distribution of the points in Figure 4,
four values, TS = 2, 4, 10, and 20 are selected and the corresponding beams are set. Table 3

Figure 4. Arrangement of Lse.

527
Table 3. Beam size in consideration.
Beam size

TS Section L (mm) Lse (mm)

2 H - 700×250×12×25 6000 1462


4 H - 850×350×19×36 10000 1769
10 H - 550×250×16×32 13000 1410
20 H - 400×200× 9×12 13000 1019

lists the size of the beams which have respective TS. Furthermore, five stiffening positions are
set. The positions are a position where Ls = Lse, and positions that shifted back and forth by
3% and 5% of L around the position.
Figure 5 shows the result of the competitions of the load-displacement relationship. (a) is
a case where the plates stiffened at Ls = Lse and Ls = Lse ± 3% are compared, (b) is a case at
Ls = Lse and Ls = Lse ± 5%. In the case where TS has a small value, even if the stiffening pos-
ition deviates from Ls = Lse, the same plastic deformation capacity is exhibited. However, in
the case of a large value, even if there is a small deviation, a large difference occurs in the load
displacement relationship, and the obtained stiffening effect becomes very small.
Thus, the equation defining the stiffening range for stiffener-plates is shown as (6).
The upper and lower limit of Eq (6) is also shown in Figure 4, with the dashed line. In
the small range of TS, a deviation of about 5% is permitted; however, in a large range,
slight deviation is not recognized. It can also be confirmed that both plots are within
this stiffening range.

0:28 Lse 0:42


pffiffiffiffiffiffi   pffiffiffiffiffiffi ð6Þ
TS L TS

Figure 5. Effect of a gap of stiffening position on load displacement relationship.

528
4 EFFECTIVE STIFFENER POSITION OF THE WEB PLATE LOCAL BUCKLING
OF AN H-SHAPED STEEL BEAM WITH THE UPPERFLANGE RESTRAINT

A case where web local buckling occurs is considered in this chapter. Equation (1) is initially
employed to select beams that will undergo plate local buckling, and the effective stiffening
position is considered following the same procedure as in the case of lateral buckling. Note
that a case of receiving pure shear force that causes the deformation of the entire web is not
considered.
As a result of checking the “ultimate state,” the local buckling deformation pattern is
roughly classified as shown in Figure 6(a) and (b) is bent and one or two waveforms are gener-
ated, and (c) is also affected by shear at the same time. Further, A, B, and C shown in the
figure are the set stiffening positions, and they capture the large deformation position and the
front and back thereof.
Figure 7 shows the load displacement relationship of beams shown in Figure 6. In either
case, it is possible to obtain a generally high stiffening effect by providing a stiffener at the stiff-
ening position B. Therefore, other beams are similarly classified by buckling waveforms, and
the position Ls = Lse corresponding to position B are obtained for each case and sorted out.
The results are shown in Figure 8. The vertical axis is the numerical value obtained by div-
iding Lse by H, and the horizontal axis is the web width thickness ratio. The classified buckling
waveform (a) is plotted by ●, (b) by 〇, and (b) by ▲. In the case where the web width thick-
ness ratio is small, (a) becomes dominant, and the smaller the web width thickness ratio, the
more the effective stiffening position is away from the fixed end. This is considered to be due
to lateral buckling between the lateral buckling stiffening positions. On the other hand, (b)
and (c) are seen when the width to thickness ratio is large, and the effective stiffening position
in this case is Lse = 0.5H.
An approximate eq (7) of the plotted points is obtained from Figure 8. In this study, since
lateral buckling stiffening is provided every 3H, it becomes such an equation. Therefore, by
making the stiffening interval shorter, the effective stiffening position is presumed to be
a fixed value at 0.5. In addition, this value 0.5 is consistent with the value obtained from
experiments for a usual unrestrained H-shaped beam, and it is confirmed that the upper-
flange restraint does not affect web local buckling.

Figure 6. Classification of buckling wave at the “ultimate state” and setting stiffening position (not
stiffened).

Figure 7. Comparison of load-displacement relationship.

529
Figure 8. Arrangement of stiffener position Lse.

8    
>
> 1
13  0:1
bw bw
575
Lse <
11 tw  tw 
¼ ð7Þ
H >
> bw
: 0:5  75
tw

5 CONCLUSIONS

The conclusions obtained based on the previously mentioned considerations are shown below.
1. Stiffener-plates provided at the position that obtained the “ultimate state” of the restrained
unstiffened beam are better than that provided at the position decided by Japanese guide-
line on lateral buckling stiffening, if lateral buckling occurs in restrained beams.
2. If web local buckling occurs in restrained beams, the effective stiffener positions can be
decided from the deformation shape at the “ultimate state.”
3. Equations for calculating effective stiffener positions were shown as eq (5) and (7).

REFERENCES

Architectural Institute of Japan. 2010. Recommendation for plastic design of steel structures (2nd).
Architectural Institute of Japan. 2018. Recommendation for stability design of steel structures (4th).
Bradford, M.A. 1988. Buckling of elastically restrained beams with web distortions. Thin-walled struc-
tures 6: 287–304.
Bradford, M.A. 1998. Inelastic buckling of I-beams with continuous elastic tension flange restraint. Jour-
nal of construction steel research 48: 63–77.
Ikarashi, K. & Ohnishi, Y. 2014. Elastic Buckling Strength of H-Shaped Beams with Continuous Com-
plete Restraint on Upper Flange. Journal of structural and construction engineering, Architectural Insti-
tute. of Japan 706: 1899–1908.
Ikarashi, K. et al. 2018. Collapse Mode and Plastic Deformation Capacity of H-Shaped Beams with Con-
tinuous Restraint Upper Flange. Journal of structural and construction engineering, Architectural Insti-
tute. of Japan 745: 491–501.
Ikarashi, K. & Sano, T. 2018. Influence of Upper Flange Restraint Condition on Lateral Buckling Behav-
ior of H-Shaped Beam. Journal of structural and construction engineering, Architectural Institute. of
Japan 749: 1063–1073.
Suzuki, T. et al. 1982. Experimental study on plastic deformation shape of stiffener. Academic lecture
summary collection, Architectural Institute. of Japan: 1951–1952.

530
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Local buckling strength of vertical haunch H-shaped beam under


shear bending

W. Ishida & K. Ikarashi


Department of Architecture and Building Engineering, Tokyo Institute of Technology, Tokyo, Japan

ABSTRACT: Vertical haunch H-shaped beams have been widely employed in steel struc-
tures, as they are cost-effective, provide the required strength for the joints, and can connect
beams of different heights. However, there have only been few studies on the mechanism and
strength of the elastic local buckling of the vertical haunch H-shaped steel beams. Therefore,
the purpose of this study is to demonstrate the influence of various factors on the elastic local
buckling and to propose an elastic local buckling strength evaluation method for vertical
haunch H-shaped beams.

1 INTRODUCTION

Vertical haunch beams have been widely used in steel structures for the purposes of reducing
the amount of steel used, strengthening the joints of columns and beams, and connecting
beams of different heights. Several experimental and analytical studies related to vertical
haunch beams have been conducted so far. However, many of these studies have investigated
beams with a limited cross section. When the beam reaches the maximum yield strength due
to buckling, the buckling behavior may be different due to the influence of the cross sectional
shape of the beams. A consistent method for designing a vertical haunch beam without any
limitations has not yet been proposed.
Therefore, the purpose of this research is to propose a method for designing a vertical
haunch beam that is applicable in various situations. In this report, the relationship between
the elastic local buckling property, buckling strength, and the haunch shapes are clarified
using a theoretical analysis and the finite element method analysis, and an evaluation method
for the elastic local buckling was proposed.

2 STRESS DISTRIBUTION OF THE BEAM UNDER SHEAR BENDING IN THE


ELASTIC RANGE

2.1 Derivation of the stress distribution considering the taper gradient


In this section, a theoretical analysis was used to derive the stress distribution of a vertical
haunch beam that was subjected to bending shear. Figure 1 shows an example of a model of
a vertical haunch beam. This beam comprises a cantilever with a haunch on the fixed end side.
The target vertical haunch beam has a maximum cross section beam with a height Ds,
a haunch length Ls, and a haunch angle θ. Different symbols and parameters are shown in
Tables 1 and 2. A vertical load was applied to the free end side of this beam, and a bending
moment of M = QL was induced on the maximum section side.
In this section, the degree of stress generated in each component plate element, when the
vertical haunch beam was subjected a bending shear force, was derived in consideration of the
geometrical characteristics due to the taper gradient. For regular uniform cross beams, it was
assumed that the shear force on the flange was negligible, because the thickness of the flange

531
Figure 1. Vertical haunch beam.

Table 1. Parameter definition.

Q : external shear force M ðxÞ ¼ M ð1  x=LÞ : acting bending moment


B ¼ 2b : beam width D : beam height d ¼ D  tf : web height
tw : web thickness tf : flange thickness L : member length Ls : haunch length
β ¼ Ls =L : haunch ratio γ ¼ 1  d=ds : taper ratio θ : haunch angle
Aw ¼ dtw : minimum web area Af ¼ b=tf : flange area
A
η ¼ 16 þ Awf : shear bending ratio Z ¼ 16 tw d 2 þ Btf d ¼ Aw dη : section modulus

Ds ¼ D=ð1  γÞ : max beam height ds ¼ Ds  tf : max web height


e ¼ Ds  D : widening height
A
Aws ¼ ds tw : max web area ηs ¼ 16 þ Awsf : max shear bending ratio
Zs ¼ 16 tw ds 2 þ Btf ds ¼ Aws ds ηs : max section modulus

Table 2. Stress level.

σw ¼ ZMs ¼ η AMws Ds : web bending stress at x ¼ 0; y ¼ Ds


s
τw ¼ AQws ¼ DQs tw : web shear stress at x ¼ 0; y ¼ Ds
σfx ¼ η AMws Ds : x component of flange axial directional stress at x ¼ 0
s
σfy ¼ σfx tanθ : y component of flange axial directional stress at x ¼ 0

plate was sufficiently larger that of the web. This assumption can also be applied in the verti-
cal haunch beams. However, as the stress degree in the flange axial direction was produced
along the inclination, as shown in Figure 2, the y-axis direction component varied according
to the flange angle θ. It was considered that this component bore a part of the shear force
generated in the web.
If the component in the axial direction of the flange in the x-axis is equal to the degree of
bending stress at the edge of the web, as shown in Figure 2, the magnitudes of both the upper
and lower flanges are equal and are represented by the Equation (1).

M ðxÞ ηe 1  x=L
σf 1 ðx; yÞ ¼ σfx2 ðx; yÞ ¼ Z ðxÞ ¼ σfx ηðxÞ ð1  γx=Ls Þ2 ð0  x  Ls Þ ð1Þ

In contrast, the y-direction component of the axial stress generated only in the lower flange
varied according to the haunch angle θ. This is expressed by Equation (2) as follows.

e ηe 1  x=L
σfy2 ðx ; yÞ ¼ σfy2 ðx; yÞ  tan θ ¼ σfx ð0  x  Ls Þ ð2Þ
Ls ηðxÞ ð1  γx=Ls Þ2

532
Figure 2. Shearing force load on lower flange.

Further, the magnitude of the shearing force load on the lower flange in the arbitrary cross
section of the haunch part is expressed by Equation (3).
 
1 Lx
Qf ðxÞ ¼ σfy2 ðxÞ  Af ¼ Qγ 1  ð0  x  Ls Þ ð3Þ
6ηðxÞ Ls  γx

However, it is necessary to observe that the lower flange was buried as a direct stress and not
as a shear stress.
The shear force load on the web at an arbitrary cross section of the haunch part was obtained
by subtracting the force represented by the Equation (3) from the shearing force acting on the
beam, and is expressed by the Equation (4).
 
1 1  x=L
Qw ðxÞ ¼ Q  Qf ðxÞ ¼ Q 1  γ 1  ð0  x  Ls Þ ð4Þ
6ηðxÞ β  γx=L

The shearing force load on the web was determined from Equation (4) using the taper ratio
γ, the haunch ratio β, and the shear bending ratio η, independently of the haunch angle θ.
Considering the shearing force burden due to the flange, the shear stress intensity distribution
of the web is expressed as follows:
8 n o
ηe 1  x=L
>
< σw ðx; yÞ ¼ σw  ηðxÞ   2y
 1 ð0  x  Ls Þ
ð1  γx=Ls Þ2 dðxÞ
 ð5Þ
>
: σ ðx; yÞ ¼ σ  ηe 1  x=L 2y
w w η  ð1  γÞ2 d  1 ðLs  x  LÞ
8 h n o i
< τ ðx; yÞ ¼ τ  ds  1  γ 1  1 1  x=L
w w 6ηðxÞ  1  γx=Ls ð0  x  Ls Þ
dðxÞ
ð6Þ
: τw ðx; yÞ ¼ Q ¼ τw ðLs  x  LÞ
Aw 1γ

The bending stress distribution is a triangle distribution, and the shear stress distribution is
a rectangular distribution, as shown in Figure 1.
Figure 3 shows an example of the stress distribution. In this example, the beam H-400 ×
200 × 9 × 22 was considered with a member length of 1600 mm and a haunch length Ls of 400,
and the widened beam height Ds was changed. As observed from this figure, the shear stress
acting on the web was discontinuous before and after the haunch. Further, the bending stress
generated in the haunch part was alleviated by the enlargement of the cross section and
approached the equally distributed load in the vicinity of the fixed end.

2.2 Verification of stress distribution using the finite element analysis


In this section, the stress distribution theoretically derived in the previous section is analyzed
using the finite element analysis. A general-purpose finite element analysis program Abaqus
2017 was employed for the analysis. Figure 4 shows a finite element analysis model.

533
Figure 3. Stress distribution by theoretical analysis.

Figure 4. FEM analysis model.

The element used was a 4-node shell element. The boundary conditions in the cross
section with x = 0 were fixed for all the displacements and rotations. The displacement in the
z-direction and the rotation around the x-axis were fixed in the cross section with x = L. In the
research, the lateral buckling braces were set at the position shown in the Figure 4 to eliminate
the influence of the lateral buckling. Similar kinds of steel were used in the web and the flange.
Their material properties were defined using Young’s modulus E = 2.05 × 105 [N/mm2], and the
Poisson’s ratio was defined as ν = 0.3, The stress distribution in the elastic region, where the
beams were not plasticized, was investigated.
Figure 5 shows the stress intensity distribution of the analysis. The axis of ordinates repre-
sents the distance from upper flange edge. The horizontal axis represents the stress intensity
distribution. These graphs show plots of the shear stress, the bending stress, and the direct
stress in the y-axis direction of the web, obtained by numerical analysis, and the stress distri-
bution obtained by the Equations (5) to (8) was indicated by broken lines and dashed dotted
lines. Haunch beam could be divided into (a) Haunch part, (b) Starting point, (c) Equal cross
section, and had different correspondences in stress distribution at each part.
In (c), the bending stress and the shear stress corresponded well. The bending stress in (a) and
(b) corresponded relatively well, but in the vicinity of the lower flange, numerical values were
higher than calculation due to the stress concentration. In (b), the y-axis direction direct stress by
the numerical analysis is greatly acting. Figure 6 shows the distribution of the y-axis direction
direct stress on the web. The y-axis compressive stress was caused by the transmission of in-plane
stress of the lower flange. The reduction of shear stress by numerical analysis in (a) and (b) was
considered to be caused by the concentration of compressive stress at the starting point.
2.3 Stress transmission mechanism around the haunch
In the haunch part, the shearing force of the web was applied on the lower flange as a direct
stress, while the compressive stress concentrated on the web. In this section, consider the
micro elements near the haunch are considered and the stress transmission is discussed.

534
Figure 5. Comparison of stress by calculation and analysis (H-400 x 200 x 9 x 22-1600, Ds = 500, Ls = 400).

Assuming that the stress of the web was in a plane stress state, the stress state of the uniform
cross section developed into the state shown in Figure 7(a). However, when the stress state of
an element was transmitted to the haunch part, it was tilted by the haunch angle θ, as shown
in Figure 7(b). When the stress was transmitted from the state shown in Figure 7(a) to the
state shown in Figure 7(b), the coordinates of the point B and B’ rotated by 2θ from the point
A and A’ on the stress circle shown in Figure 8(a) indicates the stress of the state shown in
Figure 7(b). The angle formed by the principal stress plane was set to φ. Using the addition
theorem, σx0 , σy0 , and τxy0 could be represented by σx, σy, and τxy, respectively, as expressed in
Equations (7), (8), and (9), respectively.
0 σx þ σy σx  σy
σx ¼ þ cos 2θ þ τxy sin 2θ ð7Þ
2 2
0 σx þ σy σx  σy
σy ¼  cos 2θ  τxy sin 2θ ð8Þ
2 2
0 σx  σy
τxy ¼ τxy cos 2θ  sin 2θ ð9Þ
2

It can be observed from Equation (8), that a part of the shear stress was transmitted as
a component of the y-direction straight stress to the lower haunch flange. Moreover, as men-
tioned in the previous section, it can be observed that the lower flange bore part of the web
shear stress as a direct stress in the tapered portion.
When the stress was transmitted from the state shown in Figure 7(b) to the state shown in
Figure 7(a), the coordinates of the point D and D’ rotated by 2θ from the point C and C’ on
the stress circle shown in Figure 8(b) indicates the stress of the state shown in Figure 7(a). The
angle formed by the principal stress plane was set to ψ. Using the addition theorem, σx, σy,
and τxy can

535
Figure 6. y-axis compressive stress acting on the web (H-400 x 200 x 9 x 22-1600, Ds = 500, Ls = 400).

Figure 7. Planar element near the haunch starting point.

be represented by σx0 , σy0 , and τxy0 , respectively, as expressed in Equations (10), (11), (12),
respectively.

σx 0 þ σy 0 σx 0  σy 0
σx ¼ þ cos 2θ  τxy 0 sin 2θ ð10Þ
2 2
σx 0 þ σy 0 σx 0  σy 0
σy ¼  cos 2θ þ τxy 0 sin 2θ ð11Þ
2 2
σx 0  σy 0
τxy ¼ τ0xy cos 2θ  sin 2θ _ ð12Þ
2

Let us consider the condition of Equation (11) becoming zero. Considering the case of realistic
vertical haunch beams, let θ was defined as 0 < θ ≦ 30. Provided the haunch length was not
similar to the material length, σy0 had the same sign as σx0 , and its absolute value was much
lower than σx0 (so that Equation (11) became negative), it is evident that the compressive
stress was generated in the web at the starting point of the haunch.

3 EVALUATION METHOD OF ELASTIC LOCAL BUCKLING STRENGTH OF


VERTICAL HAUNCH BEAMS

3.1 Elastic local buckling resistance using eigenvalue analysis


In this chapter, the elastic local buckling resistance of a vertical haunch beam was obtained
using the finite element method and an evaluation curve was proposed.

536
Figure 8. Mohr’s stress circle.

Eigenvalue analysis of the vertical haunch beam was carried out using the finite element
method. The analysis model is similar to the one shown in Figure 4. The parametric analysis
was carried out using the cross sectional shape, maximum beam height, haunch length, and
the beam length. The cross section that is the object of this time often possesses a relatively
large web width-thickness ratio.
The cross section beams corresponding to the haunch beams possess the same beam length,
with a cross section that equals the smallest cross section of the haunch beam. Eigenvalue ana-
lysis was carried out to evaluate the buckling durability.

3.2 Elastic local buckling durability evaluation formula


As shown in Figure 9, the influence of the web compressive stress was dominant in the haunch
beam, and the buckling waveform indicated by the web crippling was remarkable. When the
beam buckles and web crippling occurred, the elastic buckling resistance was greatly reduced
compared to that of the uniform section beam. The reduction in the width of the buckling
proof stress increased, as the haunch angle increased. Therefore, the influence of the compres-
sive stress was dominant in buckling.
Equations (13) and Figure 10 show the elastic local buckling proof stresses. The vertical
axis shows the reduction rate of the buckling durability compared to the uniform section
beams. The horizontal axis indicates a coefficient ρ, as shown in Equation (14). This coeffi-
cient corresponds to the magnitude of the shear force acting on the web at the starting point
of the haunch. Therefore, it can be expressed using the shear bending ratio η, the haunch ratio
β, and the taper ratio γ similarly to Equation (4). Based on the shapes of the commonly used
beams and haunches, ρ was set to 8.60 or less.
Using this formula, the elastic local buckling resistance was evaluated with good accuracy.

2:00
h Qcr =Qcr ¼ ð13Þ
ρ þ 2:00
 
1 1β γ
ρ ¼ 1  8:60 ð14Þ
6η β 1γ

Figure 9. Example of web crippling (H-400 × 200 × 9 × 22-1600).

537
Figure 10. Elastic local buckling durability evaluation formula.

4 CONCLUSION

In this study, the stress distribution, stress transmission mechanism, buckling behavior including
web crippling, and the elastic local buckling resistance of a vertical haunch H-shaped steel beam
that was subjected to bending shear were discussed. The following results were obtained from this
study.
1. The stress distribution was theoretically derived considering the tapered shape of the verti-
cal haunch. Finite element analysis was also conducted to verify the stress distribution. The
results of this analysis revealed that the compressive stress acts on the web at the starting
point of the haunch. We also explained the mechanism of the stress transmission near the
starting point of haunch.
2. Eigenvalue analysis was performed on various haunch beams to obtain the elastic local
buckling resistance. We proposed the factor ρ, which significantly influenced the buckling
resistance and the elastic local buckling resistance evaluation formula.
In the future, we will consider further investigation of the elastic local buckling resistance such
as examination of the section that buckles at the flange and examination during the compres-
sion of the upper flange We are currently developing specific experiments to study the mech-
anical behavior, buckling behavior, and the deformation capacity in the plastic range.

REFERENCES

Petricone, R. & Salerno, V.L. 1970. Plastic Analysis and Design of A Haunched Beam, Transactions
New York Academy of Science: 657–674.
Uang, C. & Yu, Q.K. & Gross, J. 2000. Seismic Rehabilitation Design of Steel Moment Connection with
Welded Haunch, Journal of Structural Engineering: No.126, 57–68.
Uang, C. & Yu, Q.K. & Noel, S. & Gross, J. 2000. Cyclic Testing of Steel Moment Connections Rehabili-
tated with RBS or Welded Haunch, Journal of Structural Engineering: No.126, 57–68.
Yu, W.W. 2000. Cold-Formed Steel Design Third Edition. New York. John Wiley & Sons, Inc.
Kikuo, I. 2003. Buckling Strength of Simply Supported Web Plate under the Action of Bending Shear
Stress, Architectural Institute of Japan Structural Papers Report Collection: No.656, 135–141.

538
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Imperfection sensitivity of corrugated web girders subjected


to lateral-torsional buckling

B. Jáger, M. Kachichian & L. Dunai


Department of Structural Engineering, Faculty of Civil Engineering, Budapest University of Technology
and Economics, Budapest, Hungary

C. Égető
Department of Geodesy and Surveying, Faculty of Civil Engineering, Budapest University of Technology
and Economics, Budapest, Hungary

ABSTRACT: Current standards and specifications do not provide prescriptions how to deter-
mine the lateral-torsional buckling (LTB) strength of trapezoidally corrugated web girders. In
the literature some research work can be found on the investigation of the LTB behavior mainly
by FE analysis, and on the derivation of possible analytical solutions for the resistance. The
number of available experimental results is limited in the literature therefore the authors per-
formed an extensive test program on eleven full-scale test specimens in laboratory conditions.
Based on the experimental test results an advanced FE model is developed. Standards provide
equivalent geometric imperfections including initial geometric imperfections and residual stres-
ses for structural members. These proposals have not been, however, investigated for the appli-
cation to trapezoidally corrugated web girders. Thus in this study an imperfection sensitivity
analysis is performed and a proposal is developed for the equivalent geometric imperfections to
be applied in FEM based design.

1 INTRODUCTION

Corrugated web girders are increasingly used in the structural engineering praxis, thus the
local and global stability phenomena have a greater importance in the design. In the literature
the effect of local stability on cross-sectional resistances, namely the bending moment, shear
buckling and transverse force strengths have been experimentally and numerically studied by
several researchers and there are quite accepted modelling issues on how to determine the
FEM based strength of corrugated web girders.
The global member stability of this type of plate girders, namely the lateral-torsional
buckling strength of corrugated web girders was investigated first by Lindner (1990).
Since then a few researches studied the elastic critical moment and ultimate strength of
corrugated web girders. There is a conclusion of these researches that the elastic critical
moment of corrugated web girders is greater than those of with flat web, however, the
increment has been considered in different ways by attributing to different cross-sectional
properties. Furthermore, the ultimate lateral-torsional buckling strength has been investi-
gated mostly by nonlinear FE analysis; the number of available test result is limited. In
the nonlinear FE analysis the imperfections, namely the initial geometric imperfection
and residual stresses have a key role, thus the choice in the imperfection shape and mag-
nitude and residual stress distribution have great importance to recover the ultimate
strength and to determine the required buckling curves to be used for the prediction of
the ultimate strength by reduction factor based design methodology. The EN1993-1-1
standard provides equivalent geometric imperfection magnitudes for member stability
analysis, but the recommendation is rather conservative compared to magnitudes

539
proposed in the literature. Therefore the aim of this study is to determine the required
magnitudes using commonly accepted equivalent geometric imperfection shapes on the
basis of the test results of 11 large-scale test specimens. Based on the test results, an
advanced FE model is developed and an imperfection sensitivity analysis is performed.

2 PREVIOUS RESEARCHES

The elastic critical moment of a simply supported conventional flat web girder subjected to
uniform bending moment can be calculated by Equation 1 according to EN1993-1-1 standard
and Timoshenko & Gere (1961).
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 ffi
π π2 EIw
Mcr ¼ EIz GIt 1 þ 2 ð1Þ
L L GIt

In the above equation the L is the buckling length (span of the girder), E is the elastic modu-
lus, G is the shear modulus, Iz is the moment of inertia about the minor axis, It is the torsional
constant and Iw is the warping constant. In the literature several researchers studied the elastic
critical moment of trapezoidally corrugated web girders by FE analysis and concluded that
the elastic critical moment and the ultimate strength is greater than that of girders with con-
ventional flat webs (Sayed-Ahmed 2005, Moon et al. 2009, Nguyen et al. 2010, Zhang et al.
2011, Larsson & Persson 2013, Ilanovsky 2015). The increase in the elastic critical moment
has been, however, attributed to different sectional properties by different authors. Firstly,
Lindner (1990) stated that the greater performance may be attributed to the warping constant.
On the other hand Larsson & Persson (2013), Lopes et al. (2017) and Guo & Papangelis
(2018) performed notable FE studies and found that the torsional constant of corrugated web
girders is significantly greater than those of with flat web, however, there is just a minor differ-
ence between the warping constants.
The lateral-torsional buckling strength was experimentally investigated under three-point-
bending by Kubo & Watanabe (2007) on nine trapezoidally corrugated web specimens and
Hannebauer (2008) and Pimenta et al. (2015) on three and four sinusoidally corrugated web spe-
cimens, respectively. In addition Zhang et al. (2017) tested four sinusoidally corrugated web gir-
ders by cantilever arrangement with concentrated load introduction at the end of the cantilever.
The imperfections have key role in the determination of the FEM based strength by the non-
linear analysis. EN1993-1-1 suggests to use L/300 as equivalent geometric imperfection magnitude
on welded members subjected to lateral-torsional buckling if second-order plastic analysis is
applied. On the other hand, Lindner (1990) measured and found that the initial out-of-plate
imperfection of the compression flanges slightly exceeds L/1000. In the nonlinear FE analysis
Moon et al. (2009, 2013) used L/1000 and L/500, while Ibrahim (2014) and Elkawas et al. (2018)
applied L/1000 as equivalent geometric imperfection magnitudes. In addition Elkawas et al.
(2018) studied and found that the residual stresses have no influence in the case of stocky flanges.

3 EXPERIMENTAL BACKGROUND

3.1 Test specimens


The experimental research is performed at the Budapest University of Technology and Econom-
ics, Department of Structural Engineering in 2018. In the frame of the program 11 large-scale
girders are tested by four-point-bending arrangement. Six different girder geometries having dif-
ferent flange sizes and the same trapezoidal profile are investigated. The measured geometrical
and material properties of the tested girders and the notations are given in Figures 1-2 and in
Table 1. For all the specimens the widths of the parallel and inclined folds of the web are equal
(a1 = a2 = 98 mm) as mostly used in bridges with a corrugation angle of 45°.

540
Figure 1. Applied trapezoidal profile geometry.

Figure 2. Cross-sectional layout of specimens.

Table 1. Properties of specimens.


tf bf tw h hw ts fyf fuf fyw fuw
Specimen [mm] [mm] [mm] [mm] [mm] [mm] [MPa] [MPa] [MPa] [MPa]

1/1 13.8 139 6 547.0 519.3 10 357 528 289 379


1/2 14.1 140 6 545.5 517.3 10 357 528 289 379
2/1 14.0 158 6 548.5 520.6 10 357 528 289 379
2/2 14.0 162 6 546.5 518.4 10 357 528 289 379
3/1 14.0 181 6 546.5 518.5 14 357 528 289 379
3/1 14.0 179 6 548.0 520.0 14 357 528 289 379
4/1 16.7 219 6 553.0 519.6 14 379 534 289 379
5/1 16.5 250 6 552.0 519.0 14 379 534 289 379
5/2 16.6 250 6 550.0 516.8 14 379 534 289 379
6/1 15.9 300 6 551.0 519.1 16 372 520 289 379
6/2 16.1 300 6 550.0 517.9 16 372 520 289 379

Figure 2 presents the scaled drawings of the six specimen types having different flange sizes;
the numbering of the girders is set accordingly in the order of increasing flange sizes. The nom-
inal flange thickness (tf) of type #1, #2, and #3 specimens is 14 mm, while it is 16 mm for type
#4, #5, and #6 tabulated in the second column of Table 1. The nominal flange widths (bf) are
140, 160, 180, 220, 250, and 300 mm, respectively. The thickness of the web (tw) is 6 mm for all
specimens with a nominal web depth (hw) of 520 mm. At the load introduction and support loca-
tions vertical stiffeners are placed with a nominal thickness (ts) of 10 to 16 mm. The steel material
has a prescribed steel grade of S355 for flanges and S235 for the web with a nominal yield
strength of 355 MPa and 235 MPa, respectively. The material properties are evaluated by cou-
pons according to ISO 6892-1; the results can be found in the last columns of Table 1 regarding
the flange yield (fyf) and ultimate strength (fuf) and web yield (fyw) and ultimate strength (fuw).

3.2 Test setup


The applied test arrangements are presented in Figure 3. The general purpose of the test
layout is to apply large spans (6000 mm), ensuring large buckling length for the develop-
ment of lateral-torsional buckling. The total length of the tested girders is 8.2 m; the spe-
cimens are vertically supported at the cross-sections 1100 mm from both ends and the

541
Figure 3. Test setup.

concentrated vertical forces are introduced at the cross-sections 100 mm from both ends
creating a four-point-bending loading condition with lever arms of 1000 mm. The load is
produced by hydraulic jacks with maximum loading capability of 1000 kN. The concen-
trated transverse force is introduced through rigid steel load transfer elements placed on
the upper flange. The girders are laterally supported at the vertical supports and load
introduction places in order to restrain the out-of-plane rotation, thus to anchor the
transverse bending moment. In addition, the specimens are strengthened by vertical stiff-
eners at the locations of vertical supports and load introduction places in order to avoid
local failure. By this setup the twist and warping are restrained, so thus the torsional
moment is anchored; meaning that bimoment can develop. Furthermore, the two
1000 mm long cantilevers are strengthened by braces to improve the shear strength and
avoid shear buckling failure of specimens when it would have been the governing failure.
The deflection, lateral displacement and rotation of the middle cross-section and the sup-
port motions are also captured by linear displacement transducers. The longitudinal strains in
the flanges are also measured by strain gauges placed on the flanges at the middle and support
cross-sections. Further details can be found in (Jáger et al. 2019).

3.3 Test results


The experimentally observed ultimate loads and failure modes are presented in Table 2. The
flange classes of specimens are 2, 3 and barely 4 according to EN1993-1-1 which is valid for
trapezoidally corrugated web girders (Jáger et al. 2017a, 2017b). The cross-sectional bending
moment strength of compact flanges can be calculated by neglecting the web contribution
according to Equation 2 (EN1993-1-5).
 
My ¼ bf tf fyf hw þ tf ð2Þ

The comparison of the measured and theoretical strength of specimens is shown by the sixth
column of Table 2. The failure modes were rigid cross-sectional movement and no distorsional
failure of the web appeared. For specimen types #1, #2, #3, and #4 pure lateral-torsional

542
Figure 4. Lateral-torsional buckling failure (LTB) of specimen 1/1.

Table 2. Test results.


Specimen cf/tf flange class Mtest [kNm] My [kNm] Mtest/My failure mode

1/1 7.5 2 297.4 366.2 0.81 LTB


1/2 7.4 2 321.8 374.7 0.86 LTB
2/1 8.1 2 366.0 421.2 0.87 LTB
2/2 8.2 3 351.9 421.4 0.84 LTB
3/1 8.9 3 449.2 481.4 0.93 LTB
3/1 8.8 3 419.9 478.4 0.88 LTB
4/1 8.6 3 720.0 743.8 0.97 LTB
5/1 9.7 3 795.3 837.7 0.95 LTB → LFB
5/2 9.6 3 844.5 841.5 1.00 LTB → LFB
6/1 11.6 4 972.7 951.2 1.02 LFB + LTB
6/2 11.5 4 977.0 960.6 1.02 LFB + LTB

buckling failure (LTB) occurred, while in the case of specimen type #5 the lateral-torsional
buckling failure is followed by local flange buckling (LTB→LFB) in the post-buckling range.
In the case of specimen type #6 the lateral-torsional buckling and local flange buckling are
coupled at ultimate (LFB+LTB). The detailed description and evaluation of the experimental
results can be found in (Jáger et al. 2019).

4 FE MODEL DEVELOPMENT

Advanced numerical model is developed by ANSYS 15.0 finite element program. It is a full
shell model using eight-node-thin shell elements (SHELL281). The ultimate resistances are
determined by geometrical and material nonlinear analysis using equivalent geometric imperfec-
tion (GMNIA). For the imperfections the scaled first eigenmode shapes are applied. Figure 5
presents the developed geometrical model with the boundary and loading conditions applied in
the tests. The numerical model is simply supported at the location of the vertical supports.

Figure 5. Developed FE model.

543
A linear elastic-hardening plastic material model with von Mises yield criterion is used. The
material model behaves linear elastic up to the yield stress (fy) by obeying Hook’s law with
Young’s modulus equal to 200000 MPa. The yield plateau is modelled up to 1% strains with
a small increase in the stresses. By exceeding the yield strength the material model has an
isotropic hardening behaviour with a reduced modulus until it reaches the ultimate strength (fu).
In the numerical model the measured values of the yield and ultimate strengths are implemented
for all the tested girders. Convergence study is also performed by using different elements and
mesh sizes. The previous FE studies are confirmed that by applying the eight-node shell elem-
ents using parabolic serendipity base functions, and 4 elements along the fold lengths could be
acceptable (Jáger et al. 2017b).

5 IMPERFECTION SENSITIVITY

It was proved previously (Jáger et al. 2016) that the first eigenmode shape is the most practical
and relevant in terms of the imperfection sensitivity and recovering the actual failure mode
and safe side load carrying resistances of the test specimens. In this section the first eigenmode
shape is applied as equivalent geometric imperfection and its necessary magnitude is calibrated
to ensure adequate solutions provided by the FE simulations. A typical first eigenmode shape
is presented in Figure 6. Regarding all the test specimens the results of the imperfection sensi-
tivity analysis are shown in Figure 7a-d. The different curves represent the numerically com-
puted load carrying capacities divided by the measured resistances. The vertical axis
represents the MFE/Mtest ratio, the horizontal axis represents the imperfection magnitudes.
The horizontal red lines demonstrate the equality of the computed and measured load carry-
ing capacities (MFE = Mtest). The abscissas of the intersection points between the sensitivity
curves and the horizontal red lines give the required magnitudes of the equivalent geometric
imperfections.
It is shown in Figure 7 that the more slender is the girder to lateral-torsional buckling the
imperfection sensitivity is greater. However, the differences in the necessary imperfection mag-
nitudes within one specimen type are significant; meaning that the uncertainties in the imper-
fections are notable shown by the tabulated data in Table 3 as well. It can be observed that
the required imperfection magnitudes range from 0.0 to 6.3 mm and the minimum value for
the scaling factor on the span length L is obtained to 954.
Figure 8 presents the comparison of the GMNIA based and test results by using L/1000, L/500
and L/300 magnitudes on the first eigenmode shapes as equivalent geometric imperfections. It is to
be noted that almost all of the results are on the safe side if L/1000 is applied, however, more con-
servative GMNIA based strengths are provided if L/500 or L/300 is applied. The statistical evalu-
ation is shown in Table 4. It can be seen that by using L/1000 the average is obtained to 0.957 on
the safe side with maximum deviation on the unsafe side of 1.005. By using the EN1993-1-1 pro-
posal (L/300) the results are very conservative; the average is obtained to 0.816 on the safe side
which means around 20% safety margin in the FE analysis in the case of slender girders. If L/500
is applied then a moderately conservative FEM based strength is obtained.

Figure 6. First eigenmode shape.

544
Figure 7. Imperfection sensitivity analysis – magnitude determination.

Table 3. Test results.


Specimen e0 [mm] L/e0 [-]

1/1 3.0 2002


1/2 1.8 3401
2/1 3.5 1705
2/2 6.3 954
3/1 2.6 2301
3/2 5.7 1048
4/1 0.8 7232
5/1 2.5 2369
5/2 0.0 -
6/1 0.6 9719
6/2 1.0 6109
Ave 3685
Min 954

Table 4. Statistics of using different imperfection magnitudes.


MFE,L/1000/Mtest MFE,L/500/Mtest MFE,L/300/Mtest

Average 0.957 0.890 0.816


Std. Dev. 0.031 0.043 0.060
CoV 0.032 0.048 0.074
Min 0.889 0.802 0.716
Max 1.005 0.945 0.905

545
Figure 8. Comparison of the results using different imperfection magnitudes.

6 CONCLUSIONS

In the paper an experimental research program and the test results on the lateral-torsional
buckling strength of trapezoidally corrugated web girders are presented. Based on the experi-
mental study an advanced FE model is developed and calibrated considering the first eigen-
mode shape as equivalent geometric imperfection shapes. On the basis of the experimental
and numerical studies the following conclusions can be done:
– for specimens having compact flanges pure lateral-torsional buckling failure occurred, while
in the case of slender flanges the flange buckling coupled with lateral-torsional buckling;
– the imperfection sensitivity analysis shows that the more slender is the girder to lateral-
torsional buckling the imperfection sensitivity is greater;
– the results of the imperfection sensitivity analysis confirmed the previous proposals that the
first eigenmode shape with a scaling factor of L/1000 could be applied as equivalent geo-
metric imperfection in the range of the studied geometry;
– the recommendation of the EN1993-1-1 (L/300) provides very conservative solutions for
the LTB resistance.
In the future the restraining effects of the support conditions of the current test arrangement is
to be investigated by FE analysis and then a parametric FE study can be performed in order
to determine the applicable buckling curves for the prediction of the lateral-torsional buckling
strength of trapezoidally corrugated web girders.

ACKNOWLEDGEMENT

The presented research program is part of the “BridgeBeam” R&D project No. GINOP-
2.1.1-15-2015-00659; the financial support is gratefully acknowledged. Through the first author
the paper was also supported by the ÚNKP-19-3-IV. New National Excellence Program of the
Ministry of Human Capacities; the financial support is gratefully acknowledged.

REFERENCES

ANSYS® v15.0, Canonsburg, Pennsylvania, USA.


Elkawas, A.A., Hassanein, M.F. & Elchalakani, M. 2018. Lateral-torsional buckling strength and behaviour
of high-strength steel corrugated web girders for bridge construction. Thin-Walled Structures 122: 112–123.
EN 1993-1-1:2005, EUROCODE 3: Design of steel structures, Part 1-1: General rules and rules for
buildings.
EN 1993-1-5:2005, EUROCODE 3: Design of steel structures. Part 1-5: Plated structural elements.
Guo, C. & Papangelis, J. 2018. Torsion of beams with corrugated webs. Proceedings of the Ninth Inter-
national Conference on Advances in Steel Structures 373-382, Hong Kong, China, 5-7 December.

546
Hannebauer, D. 2008. Zur Querschnitts- und Stabtragfähigkeit von Trägern mit profilierten Stegen. PhD
dissertation, Brandenburgischen Technischen Universität, Cottbus, Germany.
Ibrahim, S.A. 2014. Lateral torsional buckling strength of unsymmetrical plate girders with corrugated
webs. Engineering Structures 81: 123–134.
Ilanovsky, V. 2015. Assessment of bending moment resistance of girders with corrugated web. Pollack
Periodica 10(2): 35–44.
ISO 6892-1:2016, Metallic materials – Tensile testing – Part 1: Method of test at room temperature.
Jáger, B., Dunai, L. & Kövesdi, B. 2016. Experimental based numerical modelling of girders with trape-
zoidally corrugated web subjected to combined loading. Proceedings of the 7th International Confer-
ence on Coupled Instabilities in Metal Structures, Baltimore, MD, USA, 7-8 November.
Jáger, B., Dunai, L. & Kövesdi, B. 2017a. Flange buckling behavior of girders with corrugated web Part
I: Experimental study. Thin-Walled Structures 118: 181–195.
Jáger, B., Dunai, L. & Kövesdi, B. 2017b. Flange buckling behavior of girders with corrugated web Part
II: Numerical study and design method development. Thin-Walled Structures 118: 238–252.
Jáger, B., Kövesdi, B. & Dunai, L. 2019. Experimental study on the lateral-torsional buckling strength of
trapezoidally corrugated web girders. Proceedings of the SSRC Annual Stability Conference 2019,
St. Louis, MO, USA, 2-5 April.
Kubo, M. & Watanabe, K. 2007. Lateral-torsional buckling capacity of steel girders with corrugated web
plates. Doboku Gakkai Ronbunshuu A 63(1): 179–193.
Larsson, M. & Persson, J. 2013. Lateral-torsional buckling of steel girders with trapezoidally corrugated
webs. MSc thesis 57. Gothenburg, Sweden.
Lindner, J. 1990. Lateral torsional buckling of beams with trapezoidally corrugated webs. Proceedings of
the 4th International Colloquium on Stability of Steel Structures 79-82. Budapest, Hungary.
Lopes, G.C., Couto, C., Real, P.V. & Lopes, N. 2017. Elastic critical moment of beams with sinusoidally
corrugated webs. Journal of Constructional Steel Research 129: 185–194.
Moon, J., Yi, J., Choi, B.H. & Lee, H.E. 2009. Lateral-torsional buckling of I-girder with corrugated
webs under uniform bending. Thin-Walled Structures 47: 21–30.
Moon, J., Lim, N.H. & Lee, H.E. 2013. Moment gradient correction factor and inelastic
flexural-torsional buckling of I-girders with corrugated steel webs. Thin-Walled Structures 62: 18–27.
Nguyen, N.D., Kim, S.N., Han, S.R. & Kang, Y.J. 2010. Elastic lateral-torsional buckling strength of
I-girder with trapezoidal web corrugations using a new warping constant under uniform moment.
Engineering Structures 32: 2157–2165.
Pimenta, R.J., Queiroz, G. & Diniz, S.M.C. 2015. Reliability-based design recommendations for
sinusoidal-web beams subjected to lateral-torsional buckling. Engineering Structures 84: 195–206.
Sayed-Ahmed, E.Y. 2005. Lateral torsion-flexure buckling of corrugated web steel girders. Proceedings of
the Institution of Civil Engineers, Structures and Buildings 158(1): 53–69.
Timoshenko, S.P. & Gere, J.M. 1961. Theory of Elastic Stability. 2nd edition, London: McGraw-Hill.
Zhang, Z., Li, G. & Sun, F. 2011. Flexural-torsional buckling of H-beams with corrugated webs.
Advanced Materials Research 163-167: 351–357.
Zhang, Z., Pei, S. & Qu, B. 2017. Cantilever welded wide-flange beams with sinusoidal corrugations in
webs: Full-scale test and design implications. Engineering Structures 144: 163–173.

547
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Enhanced buckling capacity of axially compressed stiffened plates


taking into account the shear-lag effect

Andreas Jäger-Cañás
Bauwesen GmbH, Lohfelden, Germany

ABSTRACT: Compressed plates are typical elements of bridges and other built-up mem-
bers. Optimized material distribution requires wide and thin plates, which need sufficient stiff-
ening. Due to the shear-lag effect, the compressive stress decreases from the web to the center
line or the side of the plate. The stress of the plate may be substantially smaller away from the
web. This allows for optimized design of compressed plates. However, this is not widely
applied.
For this paper a finite element parametric study has been conducted on stiffened
plates of different aspect ratios with three stiffeners of variable geometry subject to axial
compression, which varies over the plate width depending on the shear lag, and a fixed
shear stress.
The outcome of the study shows that more efficient designs are possible, when the shear lag
effect is taken into account. Since the standard Eurocode procedure showed to be uneco-
nomic, hand calculation procedures are developed and compared. A recommendation is pro-
vided on which design method is preferable for the dimensioning of stiffened plates subject to
axial compression and small shear loads.

1 INTRODUCTION

For typical engineering structures such as bridges or light-weight built-up, welded members
thin plates are reinforced by open or closed stiffeners. The beam theory is usually employed
for the structural analysis although in the vicinity of the bearings the member forces alone are
not satisfactory to determine the stress state of the members. To overcome the necessity to
apply higher theories such as shell theory, the cross-sectional stress distribution is calculated
taking into account the shear-lag effect. Formulae have been developed, which are successfully
used for reinforced concrete structures for many years. For steel structures EN 1993-1-5
(2007b) provides equations that allow to take the shear-lag into account as well. The evalu-
ation is printed in Figure 1 for an exemplary bridge bottom plate with 5000 mm width, 3 stiff-
eners of 30 mm thickness and 342 mm width. The yield strength fy was chosen to 355 N/mm².
It is clear that the shear lag plays a role for a huge range of member lengths. In the presented
case, up to 138 m of cantilever length are necessary to reach a geometry parameter κ less than
0.02, which is required to be able to consider the full cross-section for the calculation of
normal stresses.
Especially in regions of high bending moment gradients, the shear lag reduces the effective
cross-section, which is particularly important for the analysis of members or their parts in the
vicinity of bearings. For simplification, in EN 1993-1-5 (2007b) it is proposed to reduce the
sectional area for the determination of design stresses in the beam analysis and apply the max-
imum stress to a reduced area. This approach is safe-sided for the beam analysis since the
stresses are usually clearly overestimated in the effective cross-section.
However, for the design of parts of the member, e.g. for the bottom plate of a bridge girder,
the stress distribution used for the beam analysis is not sufficient since it does not represent the
actual stress state of the plate. Adopting such stress distribution underestimates the stress state

548
Figure 1. Dependence of the shear-lag effect on the geometry of the beam and its cross-section.

Figure 2. Normal stress distribution over the cross-section depending on β (r.), engineers’ approaches (l.).

in the middle as well as it overestimates the stress closer to the boundary of the plate (Figure 2,
σc,beam). Adopting the stress at the first stiffener (σc,1st) as a constant load for the plate design
may result in an overestimation of the resistance since yielding at the edges is preliminary
assumed when the check is carried out against the yield strength. Usually, the maximum stress is
then adopted for the analysis of the plate, e.g. in (Timmers, et al., 2015) or (Johannson, et al.,
2007). However, depending on the value of β, this assumption may lead to a quite safe design as
can easily be shown with the stress distribution given in Figure 2 (r.).
As a contribution to achieve lighter, more economical structures, a numerical study was car-
ried out concerning stiffened axially compressed plate with a fixed shear stress of 50 N/mm².
The geometry was optimized to not allow local buckling to govern the resistance. Therefore,
the plate thickness t was varied between 15 up to 40 mm in 5 mm steps. Three stiffeners of
thicknesses of tst 10 up to 40 mm and a width bst of 11.4 tst at a distance ast of 34.2 t were
employed. Plate lengths of 2000, 3000, 4000 and 5000 mm were studied. The simulations were
compared with modifications of the approach of EN 1993-1-5 (2007b) to achieve a more eco-
nomical design. A recommendation is presented, which allows to use the enhanced bearing
capacity via the inclusion of the ψlag factor.

2 ANALYSIS METHOD

Numerical calculations are carried out using the commercial software package Sofistik (Sofis-
tik, 2013). Shell elements (termed QUAD in the software) with four nodes and linear deform-
ation approximation are used. The mesh size was primarily studied and fixed for the
parametric study to one percent of the plate length with a ratio of width to length of 2. The
stiffeners were modelled with four elements in the height direction.
As recommended in (Braun, 2010), the material model with strain hardening according to
BSK (2007a) was chosen. The parameters are given in Table 1.

549
Table 1. Material properties.

ϵ fy E ϵ fy E
[‰] [N/mm²] [N/mm²] [‰] [N/mm²] [N/mm²]

999.00 532.5 0 ‒999.00 ‒532.5 0


62.26 532.5 0 ‒62.26 ‒532.5 0
16.55 355.0 3883 ‒16.55 ‒355.0 3883
1.69 355.0 0 ‒1.69 ‒355.0 0
0.00 0.0 210000 0.00 0.0 210000

The boundary conditions were applied to allow almost free movement in the x-y plane
(Figure 3). Since the loads, which were applied as line loads according to Figure 2 r. on the
four edges (compression and shear) respectively compression only on the stiffeners, were in
equilibrium, no large in-plane deflection occurred so that the influence on the results is negli-
gible. For the determination of the buckling shapes the stiffness of boundary springs (termed
cx, cy and cz in Figure 3) had to be adjusted to prevent rigid body rotation.
Imperfections were generated from different buckling modes given in Figure 4. Local buckling
modes were scaled to 1/267 of the length between two stiffeners (Figure 4 m.) respectively the
width of the longitudinal stiffener (Figure 4 l.). The final pre-deformed shape was then determined
by combining the local with the global shape scaled to 1/400 of the plate length (Figure 5).
The geometrically and materially nonlinear analysis of the imperfect structure (GMNIA)
was carried out employing the modified Newton-Raphson iteration scheme. To overcome
numerical instabilities, a subsequent dynamically stabilized calculation was made, which
started at the maximum load level derived using Newton-Raphson. The quasi-static analysis
was carried out with time increments of 100 s and load steps of 5% of the yield strength fy.
When the nodal residual forces at the end of a load step were larger than 0.5 kN, the analysis
was terminated and the load factor rGMNIA was denoted.
The procedure was calibrated with the experiments R1, R3 and T3 published in (Byung,
et al., 2009). A good agreement was found with an overestimation of the test results by max-
imum 8%. Hence, a calibration factor of 1.10 is chosen to account for the overestimation of
the tests.

Figure 3. Model for the parametric study (spring constants in kN/m).

550
Figure 4. Imperfection shapes (l. stiffener, m. plate, r. column like buckling).

Figure 5. Final imperfection shapes (l. upward column buckling, r. downward column buckling).

Table 2. Calibration of the numerical calculation procedure.


Type fy,test Diff.
[‰] N/mm² σtest/f,test σFE/f,test %

R1 357.4 0.553 0.585 5.8


R3 357.4 0.691 0.745 7.8
T3 357.4 0.934 0.985 5.5

3 DESIGN PROPOSAL

The design procedure was developed in accordance with current codified provisions. The
geometry was chosen to exclude the possibility of local buckling of the stiffeners or the plate
in most cases so that at the edge with highest compression the yield limit could be reached.
It is necessary to compute the critical stress σcrc for column like buckling following the guid-
ance given by EN 1993-1-5 (2007b). The hand calculation procedure is recommended since
using FE software to analyze the plate without vertically restrained longitudinal edges, usually
many buckling modes have to be reviewed to find the correct column like buckling shape. No
extrapolation to the most compressed edge shall be made to prevent over-conservatism.
For the evaluation of the numerical calculations, the first critical buckling mode was used
to derive the plate buckling bifurcation stress σcrp by multiplying it with the acting design
stress σ1 = fy. It was disregarded whether global or local buckling was shown by the shape of
the first eigenmode. This approach is safe, since in tendency the more detrimental column like
buckling gets decisive for the design as the critical plate buckling stress is reduced. In many
cases the approach leads to the correct result since local buckling was not probable due to
geometries used.
The values of the bifurcation stress at which column like and plate buckling occur, increase
with a decreasing value of β (Figure 1). This is due to force, which the stiffener needs to carry,
is smaller at low β. It could be observed from the numerical calculations that the plate like
behavior more dominantly altered the resistance of the stiffened plate compared to column
like buckling. Hence, a simple and effective approach is to modify the critical stress of plate
like buckling so that the interaction in tendency more likely predicts plate like behavior. To

551
do so, the factor ψ (to emphasize its use in this paper termed ψlag) shall be determined as given
in Equation 1.

σ2
ψlag ¼ ð1Þ
σ1

where σ1 = edge stress and σ2 = stress in mid-plate (minimum stress due to shear-lag effect)
To alter the plate buckling critical stress, the formula for σE proposed in EN 1993-1-5
(2007b), which allows to take into account variable stress depending on the distance to
the most compressed edge, was used and modified as given with Equation 2 for the
“ψ-approach”.

!k
8:2=4
σcrp;ψlag ¼ αcr σ1 ð2Þ
1:05 þ ψlag

where αcr = buckling eigenvalue; σ1 = edge stress and k = exponent to alter interaction.
Another possible enhancement can be achieved by modifying the critical plate buckling stress
according to Equation 3 (“σmean-approach”). The critical plate buckling stress is higher at low
values of β. In tendency more often plate like behavior is anticipated in the design check.

σ þ σ 1 2σ1
1 2
σcrp;σmean ¼ αcr σ21 ¼ σcrp ð3Þ
2 σ1 þ σ2

where αcr = buckling eigenvalue at β = 1; σ1 = edge stress; σ2 = stress in the mid-plate.


The simplified engineer’s approach is to reduce the complex plate buckling problem
with possible column like behavior into column like buckling alone. However, this nor-
mally yields quite conservative results, especially when thicker plates are combined with
quite slender stiffeners and the interaction parameter ξ = σcr,p/σcrc –1 reaching values
close to one or higher. It is possible to achieve economically reasonable results at low
values of ξ. Due to its conservative nature in a huge range of the parameters and the
impossibility to account for shear buckling, the “column-buckling-approach” is not
scope of further detailed studies.
Shear buckling plays a role in all approaches and shall be analyzed according to EN
1993-1-5 (2007b) for the stiffened plate with a proper interaction formulation between
shear and axial compression to be used. The decrease in overall capacity due to the
shear load as considered for the parametric study is negligibly small with up
three percent reduction.

Figure 6. Impact of k on the approximation of the numerical calculations depending on stiffener


stiffness.

552
Figure 7. Impact of k on the approximation of the numerical calculations.

Figure 8. Approximation of the numerical calculations using the “σmean-approach“.

4 RESULTS

4.1 Results determined using the “ψ-approach”


At the example of the longest plate length of 5000 mm, it can be shown, that neglecting the
shear-lag effect (k = 0) in the design may lead to quite conservative approximations of the
finite element calculations as can be seen in Figure 6, where the utilization of the numerical
results ηnum is compared to the codified calculations ηEC. When weak stiffeners are employed
(relative stiffener stiffness γsl ≤ 100), the capacity is underestimated by up to 190%. Heavier
stiffeners typically lead to quite high buckling reduction factors. Hence, when column like
buckling is decisive, the predicted capacities are quite similar for all β, regardless of the chosen
value for k. Analyzing plates with very heavy stiffeners may yield unsafe results of up to 35%
compared to the FE calculations. This is due to local buckling of the stiffener and the plate in
combination with numerical problems and a model factor, which at 1.10 is too high in some
cases.

553
Typically, higher exponents yield a closer approximation of the FE calculations. As ψlag
grows, the value of k loses its influence on the resistance. Since at ψlag = 1 the exponent k does
not alter the capacity prediction at all, the proposed design procedure is fully compatible with
the current philosophy of EN 1993-1-5 (2007b).
The prediction of the proposed design procedure rprop,, which expresses the relation of the
calculated resistance as the outcome of Equation 2 divided by σ1, is plotted in Figure 7
together with the results determined using EN 1993-1-5 (2007b) termed rEC. Both are shown
in relation to the resistance derived by numerical calculations (rGMNIA). In Figure 8 the rela-
tive resistance rσmean = σR,mean/σ1 is depicted against rGMNIA. Values below the diagonal line
are unsafe.
It may easily be concluded that especially capacities determined for low values of β are
much closer to the numerical results when a high value for k is chosen. When k is not modi-
fied, the difference between numerical and predicted resistances for β = 1 is about 35% to the
unsafe side and 170% to the safe side. Since at this particular value of β the results do not
depend on the chosen magnitude for k, the design approach reveals that the codified proced-
ure may need some adjustments to achieve a closer match to numerical results.

4.2 Results determined using the “σmean-approach”


Employing this design approach (results are depicted in Figure 8) yields closer predictions of
finite element calculations than EN 1993-1-5 (2007b). However, the “ψ-approach” allows for
a better approximation, especially in the range between 0.40 < r < 0.70. Again, for β = 1, the
outcome is the same as the prediction of EN 1993-1-5 (2007b), so that no enhancement is
possible.

4.3 Comparison of both approaches


In Figure 9 a comparison is made using histograms. For each approach the amount of the
occurrence of certain factors calculated by relating the predictions of the numerical results to
the design approach is counted and plotted as bar plots.
All three methods are unsafe in a couple of cases. This is usually the case when β = 1 or
very heavy stiffeners are combined with very thin plates. Not all unsafe results are captured by
the safety factor γM1 = 1.10, which is represented with the dashed line in Figure 9.
The least conservative results are predicted by the “ψ-approach” with k equal to 4, for which
a maximum of about 160% underestimated numerical calculations is achieved. The “σmean-
approach” predicts the results with the same maximum underestimation but is more conserva-
tive in the range of 25% up to 150% of the numerical determined capacities. Furthermore
a wider spread of the predictions between 0% and 25% of the FE calculations is achieved com-
pared to the “ψ-approach”, which makes the latter the preferable choice for the design method.
The standard Eurocode procedure is up to 200% uneconomic, especially for smaller values
of β, while it is unsafe in a couple of cases. Hence, it is not the best choice for the design of
stiffened plates, which are influenced by the shear-lag effect.

Figure 9. Comparison of different approaches (Eurocode l., “ψ-approach” (k=4) m., “σmean-approach r.).

554
5 SUMMARY

Typical light-weight structures under axial compression consist of thin plates, which are stiff-
ened. Often, due to the shear-lag effect the stress away from the boundary of the plate is
reduced. In consequence, plate and stiffeners have to carry less load than anticipated when
a constant stress state is assumed.
A parametrical study considering the shear-lag effect has been carried out to explore the
effect of the shear-lag on the design of stiffened plates.
It was observed that a constant load is a mostly safe but uneconomical choice for the
design. Hence, proposals were derived, which allow to take into account the beneficial stress
reduction away from the boundaries. Most promising is the “ψ-approach” with k = 4, for
which the critical plate buckling stress is increased to rise its importance in the interaction
between column like and plate buckling. The “σmean-approach” achieves the same effect while
the peak and the minimum stress are averaged and the reduction factor is calculated.
It could be shown that the “ψ-approach” with k = 4 is the superior choice since only few
numerical calculations are seriously underestimated. Most calculated results are predicted in
the range of 0.90 up to 1.35 times the numerical resistance.

REFERENCES

Braun, B., 2010. Stability of steel plates under combined loading. Dissertation. Stuttgart: University of
Stuttgart.
BSK, 2007a. Boverkets Handbok om Stålkonstruktioner. Karlskrona: Boverket.
Byung, H., Hwang, M., Yoon, T. & Yoo, C., 2009. Experimental study of inelastic buckling strength and
stiffness requirements for longitudinally stiffened panels. Engineering Structures, Volume 31, pp.
1141–1153.
EN1993-1-5, 2007b. EN 1993-1-5:2007 - Eurocode 3 - Design of steel structures - Part 1-5: Plated structural
elements. Brussels: CEN.
Johannson, B. et al., 2007. Commentary and worked examples to EN 1993- 1-5“Plated structural elements,
Brussels: JRC scientific and technical reports.
Sofistik, 2013. VERiFiCATiON MANUAL. HQ Oberschleissheim: Sofistik AG.
Timmers, R. et al., 2015. Stabilitätsnachweise nach EN 1993-1-5 – Theorie und Beispiele. In Stahlbau
Kalender 2015: U. Kuhlmann (Ed.).

555
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Axial buckling behavior of welded ring-stiffened shells

Andreas Jäger-Cañás
Bauwesen GmbH, Lohfelden, Germany
Chair of Steel and Timber Structures, BTU, Cottbus, Germany

Zheng Li & Hartmut Pasternak


Chair of Steel and Timber Structures, BTU, Cottbus, Germany

ABSTRACT: Ring-stiffened cylindrical shells are important structures employed for the
storage of liquids and solids. The stiffeners are usually required to provide sufficient stability
against wind and external pressure. In this paper it is shown that the ring beams also enhance
the resistance against axial compression. The influence of the radius to thickness ratio, the
stiffener spacing, the placement of the weld and the ring beams’ cross-sectional area is studied.
It is approved that under certain circumstances a much higher resistance may be achieved if
the ring stiffeners are taken into account for the axial buckling check.

1 INTRODUCTION

Ring-stiffened shells are widely used as tanks or silos to store a variety of goods, such as water, oil
or grain and other solids. The stiffeners are typically intended to strengthen the structure against
wind and external pressure. However, it has been recently shown (Jäger-Cañás & Pasternak,
2017) that the ring stiffeners allow for a higher axial buckling
pffiffiffiffiresistance, especially when they are
closely spaced at maximum of one buckling wave (l = 3:46 rt). One recent approach to take the
strengthening effect of arbitrarily spaced stiffeners into account was presented in (Jäger-Cañás &
Pasternak, 2018a) and refined in (Jäger-Cañás & Pasternak, 2018b).
An axis-symmetric eigenform-affine imperfection was used to develop the design procedure.
This kind of imperfection is known as being deleterious but not practically relevant. While for
unstiffened shells a weld type A shape deviation according to (Rotter & Teng, 1989) leads to
similar results as the axis-symmetric eigenform-affine imperfection, for ring-stiffened shells no
specific studies are available on ring-stiffened cylinders that allow a comparison.
For this paper several numerical analysis were carried out to close the present gap in know-
ledge. Practical radius to plate thickness (r/t) ratios have been analyzed, which had variably
placed weak to very strong stiffeners. The placement of the weld joint was studied for a medium
strength ring beam. Furthermore, the role of the stiffener spacing was studied. A wide range of
imperfection depths was taken into account to allow a broad view on the subject.

2 UNSTIFFENED SHELLS

2.1 Analysis procedure


The analysis was carried out using the commercial software Sofistik, which is widely used in
structural engineering practice and has been approved for many finite element calculations
(Sofistik, 2013). The “QUAD” element was employed, which is a four node shell element
using linear functions to calculate the deformations. A linear-elastic, ideal-plastic material law
with a yield strength of 355 MPa was chosen for the calculations. A mesh convergence study

556
pffiffiffiffi
was carried out, which led to the conclusion that a meridional lengthpffiffiffiffi of 0.24 rt yields satis-
factory results when the circumferential length does not exceed 0.96 rt.
The length of the shell was kept constant at a length to radius (l/r) ratio of two if not stated
otherwise. Only a quarter of the whole shell was modelled due to geometrical and load sym-
metry. Appropriate boundary conditions were applied on the vertical edges. The horizontal
edges were chosen to represent continuity with very light radial springs of 1 kN/m stiffness to
ensure numerical stability. The edge was clamped around the tangential direction. The bottom
had restrained vertical movement while the top was free to move vertically. The load was
applied as a uniform line load at the top edge.
The geometrically and materially nonlinear analysis (GMNA; GMNIA when imperfections
are included) was carried out employing the Newton-Raphson iteration scheme. Typically,
Sofistik uses the modified approach in combination with a line-search. However, in some
cases the full Newton-Raphson algorithm allowed for better convergence.
For comparison, a dynamically stabilized quasi-static analysis was used. The increments
were chosen as one percent of the critical load. The resistance was determined by carrying out
an accompanying eigenvalue check along the load deflection path. When the critical load
factor switched from above one to below one between two load increments, the limit load was
determined by linear interpolation between the adjacent load increments.
A weld type A imperfection according to (Rotter & Teng, 1989) was used for the analysis.

2.2 Results
Many authors used weld type imperfections to study the behavior of the axially compressed
cylinder. Amongst them are Al lawati (Al lawati, 2016) and Rotter (Rotter, 2004). Eigenform-
affine imperfections have been used in (Jäger-Cañás & Pasternak, 2017) to determine the
resistance. In Figure 1 these results have been plotted for comparison. Al lawati found out,
that for a specific length range, minimal capacities are determined (Min), while a different set
of parameters was proposed for adoption into the code framework (Prop). The curve deter-
mined using an eigenform (EF) is plotted with half of the amplitude compared to the other
graphs due to the inward and outward deviation of the imperfection shape. The dashed line
represents the calculated imperfection sensitivity curve when no stiffeners are present. As
a further comparison the codified elastic imperfection buckling reduction factor according to
(EN1993-1-6, 2017) is drawn for fabrication tolerance class C (TC C).
All sensitivity curves comply well. While Rotter’s calculated curve and the proposal of Al
lawati are more optimistic. The other sensitivity curves agree very well with the codified rela-
tionship between amplitude of imperfection and bearing capacity.
Therefore, the chosen model and analysis procedure may be concluded as a satisfactory
choice for the parametric study including ring stiffeners.

Figure 1. Analysis results of unstiffened shells in comparison with results from literature.

557
3 RING-STIFFENED SHELLS

3.1 Analysis procedure


The analysis procedure is principally in accordance with Section 2.1. The stiffeners have been
modelled using QUAD elements. Their mesh was chosen to three elements in the width direc-
tion of the stiffener. The dimensions were chosen with a width to thickness (bst/tst) ratio of 4:1.
The cross-sectional area is expressed using the ring parameter kst (Eq. 1), which was kept con-
stant at a value of 0.4 if not
pstated
ffiffiffiffi otherwise. The spacing was varied in multiples of the buck-
ling half-wave length (1.73 rt, BuHW) for the eigenform-affine imperfection with a minimum
distance of one half-wave.pWeldffiffiffiffi depressions have beenpvaried
ffiffiffiffi as multiples of the linear bend-
ing half-wave length (2.44 rt, BeHW) starting at 3.66 rt to prevent a major influence of two
adjacent imperfections. Both cases may be compared by evaluating the curves but it is not
possible to compare FEM results directly due to different stiffener spacings.

Ast
kst ¼ ð1Þ
ast tsh

where Ast = cross sectional area of the stiffener, ast = stiffener spacing and tsh = shell thickness
The elastic imperfection reduction factor α relates the resistance rGMNIA of the shell to the
bifurcation load factor rcr, which is determined employing a linear buckling analysis (LBA).
Often, this parameter alone is not sufficient to describe the effect of the imperfection for ring-
stiffened cylindrical shells since plasticity may influence the capacity. It was shown in (Jäger-Ca
ñás, 2019) that strong, closely spaced stiffeners let cylinders with a r/t ratio of up to 5000 exhibit
yielding before buckling occurs. To determine satisfactory buckling curves for the evaluation
according to (EN1993-1-6, 2017), the curve parameters α, β and η have to be extracted from the
modified capacity curve. The relation of the buckling capacity rGMNIA relative to its plastic limit
load factor rpl, which may be determined employing a materially nonlinear analysis (MNA), is
termed buckling reduction factor χ and may be calculated with eqs. 2-6.
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffi
λ̅ x;rst ¼ rMNA =rLBA ¼ rpl =rcr ð2Þ
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
αx;rst
λ̅ xp;rst ¼ ð3Þ
1  βx;rst

λ̅ x;rst  λ̅ x0;rst 1 ð4Þ


 ηx;rst
λ̅ x;rst  λ̅ x0;rst
λ̅ x0;rst < λ̅ x;rst  λ̅ xpl 1  βx;rst ð5Þ
λ̅ xpl;rst  λ̅ x0;rst

λ̅ xpl < λ̅ x;rst αx;rst = λ̅ 2x;rst ð6Þ

with λ̅ x0:rst as the squash limit relative slenderness equal to 0.20 and the indices x for axial
buckling and rst for ring-stiffened shell

3.2 Variation of the r/t ratio


The variation of strength depending on the r/t ratio is shown in Figures 2 and 3 for different
placements of the weld depression. The parameter a/d = 0 means that the weld is in the line of
the stiffener while a/d = 1/2 represents the weld seam between two plates being half way
between two stiffeners. The spacing for this evaluation is chosen to four BeHWs.
With an increasing value of r/t or the imperfection amplitude the capacity decreases. This is
shown for both placements of the weld. Describing the capacity using the elastic imperfection
reduction factor α as printed in Figure 3 shows that almost no difference exists when the weld
is the line of the stiffener. Due to yielding, for r/t = 500 a slightly lower value is calculated.

558
Figure 2. Imperfection sensitivity of ring-stiffened shells depending on r/t ratio in terms of χ.

Figure 3. Imperfection sensitivity of ring-stiffened shells depending on r/t ratio in terms of α.

The difference between the two evaluated stiffener placements is huge. Choosing a/d = 0
allows for about 150% more capacity than a/d = 1/2 would.
With the weld in the stiffener line, the analysis type is rather a GMNA than a GMNIA
because the effect of the imperfection such as circumferential compression and tension is
beared by the strong stiffener so that no circumferential buckling is probable. Hence, the che-
quer board pattern cannot develop and the resistance rises. When the weld is placed half way
between both adjacent stiffeners, the shell buckles as a medium length cylinder with the stiff-
eners acting as the boundary conditions since they prevent radial expansion. A chequer board
pattern may develop, which reduces the capacity severely. Comparing with Figure 1, it may be
concluded that the unstiffened shell reaches almost the same capacity as the ring-stiffend shell
with a/d = 1/2.

3.3 Variation of the placement of the circumferential weld


As shown in Figure 4, the placement of the weld has a huge impact on the bearing capacity.
The curves are presented in terms ofpthe
ffiffiffiffi relative distance of the weld to one stiffener with, e.g.
1/8 meaning that the stiffener is 0.61 rt away from one stiffener.

559
The further the weld is moved away from the ring beam, the less pronounced becomes the
strengthening effect of the stiffener. Circumferential buckling as a result of the inward and out-
ward deformation in the proximity of the weld is more likely to influence the capacity when the
weld is further away from the ring beam. Even a very small distance of 1/8 of the stiffener spacing
leadsptoffiffiffiffi a drop in resistance of about 15%. With the weld being situated at a/d = 1/4 (equal to
1.22 rt) the capacity drops by 50% to 90%. When the weld is placed half way between two adja-
cent stiffeners, the cylinder behaves like a medium cylinder with the ring stiffeners forming the
boundaries. The capacity then drops down to the resistance of an unstiffened shell.

3.4 Variation of the ring parameter


The distance of the stiffeners is kept constant for this analysis while the cross-section of the
ring beam is varied to achieve different values of ring parameters.
To allow for an easy comparison with Figure 1, in Figure 5 the curves are plotted in relation
to the elastic imperfection reduction factor α. It is shown that light ring-stiffeners allow for
a huge strength gain. For the case of the weld being in the stiffener line, a ring parameter of
0.1 is sufficient to achieve about 240% in strength gain (at Δw/t = 1.5). Up to an imperfection
depth of 2.5 times the wall thickness, the capacity is almost independent on the cross-sectional
area of the ring beam. A further increased imperfection amplitude leads to less loss in capacity
for more heavy stiffeners, which allow for about 260% more strength at Δw/t = 5.
When the weld is far away from the ring beam, the stiffener is less effective. A maximum
gain in resistance of about 90% may be achieved for very heavy ring stiffeners (kst = 1) at
a high imperfection amplitude of Δw/t = 5. However, no large difference in capacity may be
concluded for smaller imperfection amplitudes. Compared to the unstiffened shell, the
strength gain gets larger as the cross-sectional area of the ring stiffener is increased.

3.5 Variation of the stiffener spacing


As the stiffeners are placed further away from each other, the bearing capacity decreases. Due
to the reduction of circumferential stiffness, chequer board patterns are more likely to develop
so that the shell behaves more as if it was unstiffened when the spacing is large. It is shown in
Figure 6 r. that the transition is quite smooth between very narrowpffiffiffistiffening
ffi to huge stiffener
distances
pffiffiffiffi while the
pffiffiffiffi lowest capacity is reached at about 7.32 rt . A huge decrease from
3.66 rt to 4.88 rt is observed, which means that the stiffeners only allow for a reasonable
strength gain in a limited spacing range when the weld is placed half way between the ring
beams.

Figure 4. Imperfection sensitivity of ring-stiffened shells depending on weld placement.

560
Figure 5. Imperfection sensitivity of ring-stiffened shells depending on kst for a/d = 0 and 1/2.

Figure 6. Resistance of ring-stiffened shells depending on the stiffener spacing, determined using a weld
type imperfection.

pffiffiffiffi
It is worth to mention that at a distance of 3.66 rt the weld imperfections almost meet,
which produces an eigenform-affine like imperfection shape. If the stiffeners are considered
even closer together, the imperfections overlay each other, making this imperfection configur-
ation a less preferable choice for a buckling analysis.

3.6 Comparison with FEM calculations employing eigenform-affine substitute imperfections


The calculations are taken from (Jäger-Cañás, 2019). The curve shapes are quite similar in
Figures 6 and 7 while for the latter results were determined
pffiffiffiffi using
pffiffiffiffi an eigenform-affine shape.
It is immediately observable that from 3.66 rt to 6.92 rt stiffener spacing the bearing
capacity decreases rapidly until the resistance of the unstiffened shell is reached.
In comparison with the analysis employing a weld depression as imperfection, the results
comply quite well. For the most narrowly stiffened shell, the capacities are almost the same
independently of the imperfection. As the stiffeners are moved away from each other, pffiffiffithe
ffi results
determined with a weld imperfection are more conservative (e.g. Figure 6: 4.88 rt: α ≈ 0.3

561
Figure 7. Resistance of ring-stiffened shells depending on the stiffener spacing, determined using an
eigenform-affine imperfection shape.

for Δw/t = 1; Figure 7: α ≈ 0.37 for Δw/t = 1). The reduction of the resistance is more empha-
sized when the weld type imperfection is used. Contrary, deep imperfections lead to higher
capacities compared to the calculations employing an eigenform-affine shape deviation pattern.
While for Δw/t = 5 the maximum achievable elastic imperfection reduction factor is 0.20 when
the eigenform is used, a resistance of α = 0.28 is possible employing the weld imperfection. The
lowest capacities for both imperfection types are around α = 0.15 and hence are independent
from the chosen shape deviation.
The evaluation of the FEM results allows the conclusion that the weld type imperfection
can be employed to analyze
pffiffiffiffi ring-stiffened structures subject to axial compression if the stiff-
eners are at least 3.66 rt away from each other. The bearing capacities from the numerical
calculations are in good agreement with the results determined using eigenform-affine
imperfections.

4 CONCLUSION AND FUTURE WORKS

The structural behavior of ring-stiffened shells with weld imperfections, which are subjected to
axial compression, has been studied. It could be shown that the stiffeners alter the bearing
capacity of tanks and silos and may greatly enhance the resistance.
If the shell is sufficiently thin, the r/t ratio is of minor importance. A radius 500 times larger
than the plate thickness proved to be a sufficient lower value from which on the elastic imper-
fection reduction factor is only dependent on the imperfection amplitude.
It could be shown that the placement of the weld in the proximity of the stiffener is most
advantageous, since the circumferential force arising from the radial shape deviation is beared
by the stiffener. Since this ring beam is quite stiff against buckling perpendicular to the cylin-
drical wall, no chequer board pattern is likely to develop, which effectively rises the bearing
capacity of the shell. As the weld is moved further away from the stiffener, the capacity is
reduced until the weld is half way between two adjacent ring beams. In this case, the shell
buckles as a medium length cylinder with no effect from the stiffener but them acting as the
boundary.
Even weak stiffeners of ten percent the connected shell cross-sectional area allow a huge
strength gain of about 250%, if the weld is placed in the line of the ring beam. The further
away the weld is placed from the stiffener, the lesser is the influence, while very heavy stiff-
eners enhance the resistance by up to 90% for deep imperfections.

562
Especially narrowly stiffened shells achieve a very high bearing capacity. From the compari-
son with FEM calculations employing eigenform affine imperfection shapes the value for
a narrowly stiffened shell can be concluded to be one buckling wave-length, which is approxi-
mately 1.5 times the linear bending half-wave length of a cylindrical shell. When this spacing
is surpassed, the effect of stiffeners is rapidly decreased as they are moved further away from
each other. The capacity is limited by the resistance of the unstiffened shells when the stiff-
eners are very far away from each other.
Much more research is necessary to develop appropriate buckling curves in accordance
with (EN1993-1-6, 2017) that explicitly deal with ring-stiffened shells with weld imperfections.
Multiple parameter combinations have to be examined and buckling parameters should be
extracted from the modified capacity curve. Especially for narrowly ring-stiffened shells the
elastic imperfection reduction factor is not sufficient to describe the resistance correctly, since
even very thin shells may exhibit yielding before buckling occurs if the stiffeners are heavy.
For arbitrarily ring-stiffened shells, a design procedure was developed using eigenform-
affine imperfections. The formulae may be taken from (Jäger-Cañás, 2019).

REFERENCES

Al lawati, H. A. R. M., 2016. Buckling of axially compressed cylindrical shells under different conditions.
Dissertation. Edinburgh: University of Edinburgh.
EN1993-1-6, 2017. EN 1993-1-6:2017 Eurocode 3: Design of steel structures - Part 1-6: Strength and sta-
bility of shell structures. Brussels
Jäger, A., 2016. Grundlegende Studien zum Beulverhalten von eng ringversteift en Kreiszylinderschalen
unter Axialdruck. DASt-Kolloqium 2016, Essen
Jäger, A. & Pasternak, H., 2016. Studien zum Beulverhalten von eng ringversteiften Kreiszylinderschalen
unter Axialdruck. Bauingenieur Vol. 91, pp. 401–409.
Jäger-Cañás, A., 2019. Beitrag zur Traglastermittlung ringversteifter Kreiszylinderschalen unter
Axialdruck. Dissertation. Cottbus: Brandenburgische Technische Universität
Jäger-Cañás, A. & Pasternak, H., 2017. Influence of closely spaced ring-stiffeners on the axial buckling
behavior of cylindrical shells. Eurosteel 2017, Copenhagen, ce/papers.
Jäger-Cañás, A. & Pasternak, H., 2017. On the axial buckling of very thin-walled cylindrical shells. Kopen-
hagen: ce/papers.
Jäger-Cañás, A. & Pasternak, H., 2018a. Über das Tragverhalten ringversteifter Kreiszylinderschalen unter
Axialdruck. DASt-Kolloqium 2018, Kaiserslautern
Jäger-Cañás, A. & Pasternak, H., 2018b. On the buckling behavior of ring-stiffened shells under axial
compression. USB Proceedings of Eighth International Conference on Thin-Walled Structures
(ICTWS 2018), paper 27, Lisbon.
Rotter, J., 2004. Cylindrical shells under axial compression. In: Buckling of Thin Metal Shells. London:
Spon Press.
Rotter, J. & Teng, J., 1989. Elastic Stability of Cylindrical Shells with Weld Depressions. Journal of
Structural Engineering, American Society of Civil Engineers, vol. 115(5), pp. 1244–1263.
Sofistik, 2013. VERiFiCATiON MANUAL. HQ Oberschleissheim: Sofistik AG.

563
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Ductility assessment of structural steel and composite joints

J.-P. Jaspart, A. Corman & J.-F. Demonceau


Liege University, Liège, Belgium

ABSTRACT: Eurocode 3, in its Part 1-8 on the design of structural joints, and Eurocode 4
provide designers with assessment procedures for the initial rotational stiffness and the resist-
ance of steel and composite joints respectively. These design procedures refer to the so-called
component approach and have been validated through numerous comparisons with test
results and numerical non-linear simulations. For beam and column members, the resistance
level considered by the code is the one which could not be exceeded at ULS and it depends on
the cross-section class (Class 1 to Class 4). For Class 1 cross-sections, the plastic resistance
may be considered and internal rotations may take place and develop in the cross-section in
the case ductility criteria are met. If plastic rotation capacity is available, a plastic global ana-
lysis of the structure may be contemplated. For connections and joints, a similar concept is to
be applied, but unfortunately very few information is provided in the Eurocodes which would
enable the designer to check whether enough plastic rotational capacity is locally available. In
this paper, a procedure to estimate the rotation capacity of joints is presented. As for the
evaluation of the stiffness and resistance properties, it refers to the component method
approach. Its validity is demonstrated through comparisons with experimental data.

1 INTRODUCTION TO THE COMPONENT METHOD

1.1 Principles of the method


In order to account for the influence of the structural joints on the actual response of building
frames, their mechanical properties, in terms of rotational stiffness, resistance and possibly
deformation capacity has to be evaluated. To achieve this goal, reference is nowadays widely
made (Jaspart and Weynand, 2016) to the so-called “component method” which considers
any joint as a set of macroscopic “individual basic components”. In the particular case of
Figure 1, the relevant components are the following ones:
– compression zone:
• column web in compression
• beam flange in compression
– tension zone:
• column web in tension
• column flange in bending
• bolts in tension
• end-plate in bending
• beam web in tension
Figure 1. Joint in bending with an extended
– in shear zone:
end-plate connection.
• column web panel in shear
Each of these basic components possesses its own level of strength and stiffness in tension
compression or shear. The coexistence of several components within the same joint element -
for instance the column web which is simultaneously subjected to compression (or tension)
and shear - can obviously lead to stress interactions that are likely to decrease the strength
and the stiffness of each individual basic component; this interaction affects the shape of the

564
deformability curve of the related components but does not call the principles of the compo-
nent method in question again.
The application of the component method includes three successive steps:
a. listing of the “activated” or “active” components for the studied joint;
b. evaluation of the stiffness and/or strength properties of each individual basic component
(specific properties - initial stiffness, design strength,. . . or whole deformability curve);
c. assembly of the components in view of the evaluation of the stiffness and/or strength char-
acteristics of the whole joint (specific properties - initial stiffness, design resistance,. . . or
whole deformability curve).

1.2 Field of application and levels of refinement


The combination of all the components for which characterization rules are presently available
(CEN, 2004; CEN,2005; Jaspart et al, 2005) allows covering a wide range of joint configura-
tions, what should largely be sufficient to satisfy the needs of practitioners as far as beam-to-
column joints, column bases and beam splices are concerned, whatever is the loading situation.
Besides that, the framework of the component method is such that it allows the use of vari-
ous techniques for the component characterization and the joint assembly: in particular, the
stiffness and strength characteristics of the components may result from experimentations in
laboratory, numerical simulations by means of finite element programs or analytical models
based on theory. Similar levels of sophistication exist also, as those presented in (Jaspart and
Weynand, 2016) for what regards the joint assembly.
In Eurocode 3 Part 1-8 “Design of joints” (CEN, 2005), practical application rules are pro-
vided which allow characterizing steel joints under static loading. Complementary rules for
steel-concrete composite joints are available in Eurocode 4 (CEN, 2004). And in (Jaspart and
Weynand, 2016), extensions of the component method are proposed to accommodate fire or
earthquake conditions and to mitigate the risk of progressive collapse.

1.3 Present limitation to the application of the component method


Most of research efforts in the last decades have focused on the characterization of the stiff-
ness and resistance properties of the components, while few investigations have been devoted
to the prediction of their deformation capacity. This one is however a key parameter to
master in different design situations as the three following ones:
design of a structure with partial-strength joints based on a plastic global analysis, so
requiring from the joints a sufficient plastic rotational capacity;
– design of a structure to mitigate the risk of progressive collapse under exceptional loading
through the use of the alternative load-path method;
– design of a structure with rather rigid but partial-strength joints under a severe earthquake,
so requiring energy dissipation in the joints.
In the present paper, a general approach for the determination of the rotational capacity of
joint is presented and validated through comparisons with results of experimental tests.

2 PREDICTION OF THE JOINT ROTATION CAPACITY

2.1 General model


The rotational response of a joint is presented in the form of an M-φ moment-rotation curve
where M and φ represent respectively the bending moment to which the joint is subjected and
the resultant relative rotation between the connected members. This curve may be drawn as
well for joints in bending than for joints subjected to more complex loading, including add-
itional axial forces.

565
Figure 2. Main joint properties characterising actual M-φ curves a) Well marked bi-linear response
b) Less marked bi-linear response.

For classical steel or composite joints made of welded and bolted connections, the shape of
the M-φ curve is approximately bi-linear and may therefore be characterized by four key
parameters:
– an initial stiffness Sj,ini;
– a plastic bending resistance MRpl;
– a strain hardening (more generally post-plastic) stiffness Sj,st;
– an ultimate bending resistance MRu.
When no instability or early brittle failure occurs in the joint at ultimate state, MRu differs
significantly from MRpl, and the bi-linear shape of the M-φ curve is well marked (Figure 2.a);
when instability or early brittle failure occurs - for instance in the column web in compression
or in bolts in tension – MRu comes closer to MRpl, what tends to give a more or less round
final shape to the M-φ curve (Figure 2.b). Whatever the case, the ultimate rotation capacity φu
may be derived at the intersection of the M-φ curve with the MRu horizontal line.
Several mathematical expressions integrating these four parameters may be used so as to
closely approximate the non-linear character of actual M-φ curves. In (Jaspart, 1991), reference
is made to an exponential expression and its adequacy is shown on the basis of many compari-
sons with experimental test results in which the four key properties were predicted through
rather sophisticated analytical models. In the present paper, the approach is focusing on less
complex prediction approaches, as the one proposed in Eurocode Part 1-8 (CEN, 2005).
In this European normative document, expressions are provided for the characterization
and the assembly of components, but assembly procedures are only suggested for the evalu-
ation of the initial stiffness Sj,ini and the bending resistance MRpl of the joints. So nothing is
said in terms of strain-hardening or post-plastic stiffness Sj,st and ultimate bending
resistance MRu. On the basis of the two obtained values, a simplified M-φ curve is built as
shown in Figure 3.a. In these ones, no limit is provided to the yield plateau.
In the next paragraphs, procedures for the evaluation of the Sj,st and MRu values on the
basis of the component approach are proposed, enabling the designer to approximate the
actual M-φ curve as shown in Figure 3.b and, in line with the objective of the present paper,
to derive an estimation of the total joint rotational capacity φu and the plastic one (φu - φpl).

2.2 Derivation of Sj,st and MRu

2.2.1 Strain-hardening (post-plastic) stiffness


In Eurocode 3 Part 1-8, the following expression is proposed for the evaluation of the initial
stiffness of a joint in bending:

Ez 2
Sj; ini ¼ P 1 ð1Þ
 ki; m
m
where:
– E is the modulus of elasticity of steel
– z is the internal level arm in the joint (see Figure 1)
– ki is the elastic stiffness coefficient characterising each of the joint components

566
Figure 3. Simplified modelling of the joint M-φ curves, a) According to Eurocode 3 Part 1-8 b) Accord-
ing to the present paper.

A similar approach is followed to derive the strain-hardening stiffness Sj,st.


This requires, in a first step, to evaluate the strain-hardening stiffness coefficient for each
basic component. Studies of numerous test results on components (Jaspart, 1991) allow pro-
posing the following expressions:

Est
kst ¼ ki ð2:aÞ
E

for: column webs in compression/tension and column flanges and end plates in bending;

2ð1 þ vÞ Est
kst ¼ ki ð2:bÞ
3 E

for: column web panels in shear


where:
kst is the strain-hardening stiffness coefficient;
ki is the initial stiffness coefficient;
vEst is the strain-hardening modulus of elasticity,
v is the steel Poisson coefficient.

In a second step, the assembly procedure has to be contemplated. This one is highly dependent
on the relative importance of the design moment resistance MRpl,comp i of each individual basic
component, MRpl,comp i being calculated by considering temporally the component as the only
active component in the joint, when compared to joint design moment resistance MRpl.
For instance, let assume a joint in which one of the components is much weaker than the
others. Sj,st will result, in such a case, from the combination of the strain-hardening stiffness
of the weak component and the initial stiffness of the others; as a matter of fact, the latter
remain in the elastic range of behaviour for applied moments higher than MRpl, while the
former is in its strain-hardening range of behaviour.
In more usual joints, the successive apparition of yielding in the different components
during the joint loading beyond MRpl leads to a progressive decrease of the actual strain-
hardening stiffness in comparison with the previous case. The complexity of the problem has
been overcome in (Jaspart, 1991) as explained here below.
Each component which possesses a high design moment resistance in comparison with MRpl
will contribute in an elastic way to Sj,st, which, in fact, should probably be better called “post-
limit” stiffness. In the contrary, a component, the design resistance of which is closer to MRpl
will experience strain-hardening and will affect more significantly Sj,st. The simplified evalu-
ation of Sj,st consists therefore in the classification of the components according to their
design resistance MRpl,comp i in order to distinguish those which will contribute to Sj,st by
means of their initial elastic stiffness coefficient ki from those which will contribute by means
of the strain-hardening coefficient kst. A deep study of experimental tests on joints with end-
plates has allowed the determination of a boundary value of the moment capacity:

MRpl; limit ¼ 1; 65MRpl ð3Þ

567
which allows to classify the components (elastic contribution to Sj,st if MRpl,comp i > MRpl,limit;
strain-hardening contribution if MRpl,comp i  MRpl,limit).
The strain-hardening stiffness of the joint Sj,st may therefore be evaluated as follows:

Ez 2
Sj; st ¼ P 1 ð4Þ
k

where:

X1 X 1  X 1 
¼ þ ð5Þ
k ki;m kst; k MRpl; comp; k MRpl; limit
m MRpl; comp;j 4MRpl; limit k

k and m are component indices.

2.2.2 Ultimate moment resistance


A good estimation of the ultimate moment resistance MRu of the joint may simply be obtained
by substituting:
– the yield stress of the steel material fy by the ultimate stress fu;
– the design resistance of the bolt in tension by the ultimate resistance of the bolt in tension
(stress area times ultimate yield strength);
in the formulae proposed in (CEN, 2005) for the evaluation of the joint design moment
resistance MRpl. The risks of instability of the column web in compression and of the beam
flange in compression have however not to be forgotten. As for MRpl, the ultimate moment
resistance MRu is associated to the ultimate resistance of the weakest component.

2.3 Comparisons of full M-φ curves with the simplified proposed model
Numerous comparisons are presented in (CEN, 2005), for various connection types, and the
good agreement between the prediction of the four key values characterizing the joint
response, coupled to the use of an exponential M-φ joint model, and the test results have
allowed to validate the analytical procedure. The fact that series of tests performed by differ-
ent persons in different laboratories have been considered in this comparative study is likely
to increase the confidence in the model. Examples of such comparisons are given in Figure 4,
but this time through the use of the simplified M-φ joint model presented in Figure 3.b, refer-
ring to various failure modes (endplate in bending, column web in compression and concrete
reinforcement bars in tension). A quite good agreement is generally obtained, except in cases
of thin endplates (or column flanges) in which membrane effects result from the high ultimate
component displacements. The predicted rotation capacity is there a bit underestimated but
still on the safe side.

2.4 Evaluation formula for the joint rotational capacity


As a result, finally, the ultimate rotation capacity of the joints may be evaluated as equal to:
 
’u ¼ MRu  MRpl Sj;st ð6Þ

and the plastic rotation capacity (see Figure 3) as:

  .
’u  ’u ¼ MRu  MRpl Sj;st  3MRpl Sj;ini ð7Þ

568
Figure 4. Comparisons between predicted and M-φ curves, a) Test 07 on a single-sided joint with an
extended endplate connection (Jaspart, 1991) b) Test 013 on a single-sided joint with an extended end-
plate connection (Jaspart, 1991) c) Test T9 on a single-sided joint with an extended endplate connection
(Zoetemeijer, 1974) d) Test T1 on a composite joint with an flush endplate steel connection (Demonceau
& Jaspart, 2004).

3 CONCLUSIONS

Present paper proposes an analytical procedure for the evaluation of the ultimate and plastic
rotational capacities of joints based on the application of the component approach. It comple-
ments provisions provided by the Structural Eurocodes for the design of steel and composite
steel-concrete joints. Its application should be of particular interest for the plastic analysis of
structures with partial-strength joints, for the design of appropriate joints in structures under
seismic actions or even for the justification of the sufficient structural robustness of a building
subjected to an extreme exceptional event.

REFERENCES

CEN, 2004. Eurocode 4: Design of composite steel and concrete structures - Part 1-1: General rules and
rules for buildings, Brussels, European Committee for Normalisation.
CEN, 2005. Eurocode 3: Design of steel structures - Part 1-8: Design of joints, Brussels, European Com-
mittee for Normalisation.
Demonceau, J.-F. & Jaspart J.-P, 2004. Experimental and analytical investigations on single-sided compos-
ite joint configuration, 5th International PhD Symp. in Civil Engineering, Balkema, pp. 341–349.
Jaspart, J.P., 1991. Etude de la semi-rigidité des nœuds poutre-colonne et son influence sur la résistance et la
stabilité des ossatures en acier, PhD Thesis, Liège University.
Jaspart, J.P. & al, 2005. Development of a full consistent design approach for bolted and welded joints in
building frames and trusses between steel members made of hollow and/or open sections, CIDECT report
5BP, Comité International pour l’Etude et de Développement de la Construction Tubulaire.
Jaspart, J.P. & Weynand, K., 2016. Design of joints in steel and composite structures, ECCS Eurocode
Design Manual, Wiley, Ernst & Sohn.
Zoetemeijer, P., 1974. A design method for the tension side of statically loaded, bolted beam-to-column con-
nections. Heron, Netherlands, Vol. 20, N°1.

569
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Study on the influence of reduced beam sections on the seismic


behaviour of a moment resisting frame

A. Jiménez, E. Mirambell & E. Real


Department of Civil and Environmental Engineering, Universitat Politècnica de Catalunya-BarcelonaTech,
Barcelona, Spain

ABSTRACT: This paper presents the results of a series of nonlinear time history analyses
performed in Abaqus on a single-span two-storey moment resisting frame. The aim of this
paper is to evaluate the base shear when the structure is provided with Reduced Beam Sec-
tions (RBS) close to the beam-to-column connections and to confirm that stress concentra-
tions in the beam-to-column welds are avoided. Also, interstorey drifts and the total
dissipated plastic work are evaluated. The structure is modelled with shell elements in order
to reproduce with accuracy the formation of plastic hinges in the critical sections of the
structure with a Chaboche cyclic hardening model for the S275 steel. The dynamic analyses
are performed by introducing recorded accelerograms scaled to match a target seismic
demand by means of the software SeismoMatch.

1 INTRODUCTION

Findings after the earthquakes of Northridge and Kobe (in 1994 and 1995, respectively) have
come to show that steel structures are very much sensitive to cyclic actions. A large amount of
detailing of steel joints have been found to develop cracks in the beam-to-column welds, which
leads to a brittle failure of the joint. New strategies have since been developed, such as the
Reduced Beam Section (RBS) (Plumier, 1990), which consists on weakening the beam member
of the connection to concentrate plastic strains away from the connection and into the beam.
The reduced beam section (RBS) connection (or dog-bone) is a good alternative to the
reinforcement of the connection due to its ease of execution and its reduced cost. Owing to
their convenience, RBS connections are becoming increasingly popular in seismic areas and
are being object of study of many researchers all around the world (Giugliano, Longo, Mon-
tuori, Piluso, 2011; Montuori, 2015; Montuori, Sagarese, 2018).
In this paper, a numerical model is developed using the software Abaqus (Dassault Sys-
tèmes, 2016) to reproduce the behaviour of a two-storey single-span moment resisting
frame when subjected to a ground acceleration by means of a nonlinear time-history ana-
lysis. A series of seismic records have been scaled up by means of the software SeismoMatch
(SeismoSoft Ltd, 2012) in order to match a specific level of seismic demand. The structure
of study is analysed with and without dog-bones in order to identify the differences in the
behaviour of the structures when subjected to a seismic action. Base shear, interstorey drifts
and dissipated plastic work are evaluated.

2 DEFINITION OF THE SEISMIC ACTION

In order to impose the scaled seismic records to the structure matching the same level of seis-
mic severity an elastic response spectrum needs to be defined to scale all the natural records
to. In this paper, the software SeismoMatch (SeismoSoft Ltd, 2012) is used to that effect.

570
A total of three recorded accelerograms provided by the software were used. These acceler-
ograms correspond with the earthquakes of Kobe (Japan, 1994), Imperial Valley (United
States and Mexico, 1940) and Friuli (Italy, 1976).

2.1 Elastic response spectrum and matching of the records


The seismic action considered in this study is represented by a peak ground acceleration
pga = 0.55g, 5% damping, spectrum type 1, soil type C and importance class II. The correspond-
ing elastic spectrum is shown in Figure 1 according to EN1998-1, clause 3.2.2 (CEN, 2009).
The recordings provided in SeismoMatch (SeismoSoft Ltd, 2012) did not match the desired
seismic severity (see Figure 2a). After the proper scaling of the records, very good agreement
has been achieved between the target spectrum in EN1998-1 (CEN, 2009) and the spectrum of
each of the earthquakes considered in this study (Figure 2b). The matched recordings accurately
represent the seismic demand of study and have been used in the analyses performed herein.

3 NUMERICAL MODEL

3.1 Description of the frame


The structure of study is a moment resisting frame with a single span and two storeys. It is fixed
at the column bases and is working only in-plane. The storey height is 5m and the beam span is

Figure 1. Elastic response spectrum used for the scaling of the seismic records.

Figure 2. Target response spectrum and spectra derived from the seismic recordings (a) and derived
from the matched recordings (b).

571
11m (measured between the centrelines of the columns). It has been designed under dead and
live loads and checked against ultimate limit state as well as serviceability requirements according
to EN1993-1-1 (CEN, 2011). Also, it has been checked that the structure can resist the imposed
seismic demand under seismic combination by means of an equivalent lateral load approach
obtained from the design spectrum (using the adequate behaviour factor q = 6). The sections
used are IPE450 for the beams and HEB340 for the columns. All members are made of steel
with S275 steel grade. Figure 3 shows the geometry of the structure considered in this study.

3.2 Geometry of the reduced beam section


The radius cut to be performed on the beams has been designed according to the prescriptions
given in EN1998-3, annex B clause B.5.3.4 (3) (CEN, 2011) (see Figure 4).

3.3 Characterisation of the material


The material used in the numerical model is a Chaboche nonlinear isotropic-kinematic harden-
ing model with three backstresses (Chaboche, 1986) for carbon steel grade S275. The param-
eters introduced in Abaqus (Dassault Systèmes, 2016) have been obtained from cyclic test data
in the literature (Krolo, Grandić, Smolčić, 2016). The values used are reported in Table 1.

3.4 Numerical model and analysis


The numerical model has been developed using only S4R shell elements for the whole struc-
ture. All shells were located at the middle plane of the plates. The radius root between web
and flanges of the sections was not considered in the model. The columns were modelled as
fixed at their bases by means of a reference point and a constraint in order to make it easier to
obtain base shear results. The top flanges of the beams have their movement restrained in the
out-of-plane direction and cannot rotate due to torsion effects, simulating fixity due to the
presence of a floor slab. Also, three cross-sections of the beams have their movements con-
strained to avoid any lateral-torsional buckling. Adequate stiffening of the panel zone and
continuity plates were modelled. The welds were not taken into account in the model but the
beams and columns were connected by sharing mesh nodes. An elastic simulation was per-
formed with Abaqus and SAP2000 to validate the model with the design loads corresponding
to the seismic combination. As it can be observed in Table 2, very good agreement was
achieved. The compared magnitudes are the maximum lateral displacement at the roof level
and the reactions at the right support.

Figure 3. Structure of study with dimensions in mm (a) and view of the beam-to-column connection (b).

572
Figure 4. Scheme of the geometry of the Reduced Beam Section of study. Dimensions in mm.

Table 1. Parameters for the definition of the properties of the material.


E σ|0 C1 C2 C3 Q∞
Material (MPa) ν (MPa) (MPa) γ1 (MPa) γ2 (MPa) γ3 (MPa) b

S275 207000 0.3 285 13921 765 4240 52 1573 14 25.6 4.4

Table 2. Comparison of the results for the validation of the model.


ux,max Rx Rz My
(mm) (kN) (kN) (kNm)

Abaqus 21.90 –83.43 554.43 188.29


SAP2000 21.82 –84.59 569.39 187.67

The analysis conducted in Abaqus is a dynamic analysis with implicit integration. Masses due
to self-weight and the loads acting in seismic combination have been introduced in the model.
Vertical loads in seismic combination were introduced slowly to avoid these loads causing any
relevant dynamic response of the structure. The target value of vertical loads was achieved after 5
seconds and the target value was maintained during the rest of the dynamic analysis. After these
loads were slowly applied (at the time instant t = 5s) an acceleration boundary condition was
applied to reproduce the ground movement caused by the matched seismic records.

4 RESULTS

After performing the six nonlinear time-history analyses the base shear, the interstorey drifts
and the dissipated plastic work have been evaluated for each of the three matched seismic
records. In the subsequent sections the obtained results are compared.

4.1 Base shear


The following graphs have been obtained after the analyses evaluating the base shear of the
structure at every instant during the simulated seismic event.
As it can be observed in Figure 5, the differences in base shear are not very apparent
between the cases with and without RBS. Indeed, it has been checked that the fundamental
period of the structure does not change in a noteworthy manner due to the weakening of the
beams and therefore, the forces acting on the structure remain the same.

573
Both graphs tend to follow a very similar pattern, indicating similar responses from both
structures. This is especially so during the first seconds of the simulation, when little damage
has occurred and there is very small influence on the stiffness by the RBS. After some damage
is developed, differences start to be more evident.

4.2 Interstorey drift


The interstorey drift was measured for both storeys throughout the whole duration of the seis-
mic event simulation. The following figures illustrate the results obtained for the first storey.
The following charts illustrate the evolution of the interstorey drift for the top storey. It can
be observed that all the graphs associated to the same seismic event are very recognisable.
As it can be observed in Figures 6 and 7, the maximum displacements achieved by the struc-
ture are always larger in the case of the frame that is provided with RBS. This seems to indi-
cate that, although base shear seems to be the same (see Figure 5), the structure is more
flexible and can reach larger target displacements when plastic fuses are adopted.
This increase in lateral flexibility needs to be checked for damage limitation requirements
with a lower intensity seismic action (EN1998-1, clause 4.4.3.2 (CEN, 2009)), which causes
lower plastic strains in the frame members. As a result, differences in interstorey drifts in
damage limitation combinations between frames with and without RBS will be smaller than
those in ultimate limit state combinations.

4.3 Dissipated plastic work


The total dissipated plastic work in the structure was also evaluated during the time-history
analyses. For a single Gauss point, the dissipated plastic work wpl is defined as shown in
Equation 1 (Lemaitre, Chaboche, 2009).
ðt
wpl ¼ σij ðτÞ_εpl
ij ðτÞdτ ð1Þ
0

Figure 5. Evolution of the base shear of the structure during the analysis. From top to bottom: Kobe
matched earthquake, Imperial Valley matched earthquake, Friuli matched earthquake.

574
Figure 6. Evolution of the interstorey drift for the first storey. From top to bottom: Kobe matched
earthquake, Imperial Valley matched earthquake, Friuli matched earthquake.

Figure 7. Evolution of the interstorey drift for the second storey. From top to bottom: Kobe matched
earthquake, Imperial Valley matched earthquake, Friuli matched earthquake.

575
Figure 8. Evolution of the total dissipated plastic work for the whole structure. From left to right:
Kobe matched earthquake, Imperial Valley matched earthquake, Friuli matched earthquake.

Figure 9. Equivalent plastic strain at the first-storey right beam end. Imperial Valley Matched earth-
quake. Comparison between the intact section (left) and Reduced Beam Section (right).

The total dissipated plastic work for each matched accelerogram is displayed in Figure 8.
In the three cases analysed, the total dissipated plastic work wpl is at least twice as much
when the structure is provided with dog-bones. Since the structure is exactly the same in both
cases with the single difference of the presence of RBS, this great difference in dissipated
energy clearly is due to the higher dissipation in the dog-bones when experiencing plastic rota-
tions. This fact can be observed in Figure 9, where equivalent plastic strains are compared for
the Imperial Valley matched earthquake. For the structure with RBS the strains correspond
with t = 21.63s and for the structure without RBS t = 21.73s. The times are slightly different
to ensure a similar level of base shear for the structure. The maximum equivalent plastic
strains for the RBS (0.209) are almost three times those of the intact section (0.077).

5 CONCLUSIONS

In this paper a numerical model in Abaqus (Dassault Systèmes, 2016) has been developed to
reproduce the behaviour of a single-span two-storey moment resisting frame when equipped
with dog-bone joints and to compare it with an identical structure without dog-bones. The
material model has been taken from experimental cyclic data for steel S275 found in the litera-
ture (Krolo, Grandić, Smolčić, 2016). The seismic action is defined by a series of three seismic
records that have been scaled up to meet a desired seismic demand with the software Seismo-
Match (SeismoSoft Ltd, 2012). Based upon the time-history analyses performed, the following
conclusions can be drawn:

576
The time history records of total base shear of the structure during the simulated seismic
event show very similar values for maximum shear for the case of the moment-resisting frame
equipped with RBS and without it.
The structure when provided with RBS at the beam ends can reach larger displacements
than when it is not, making it more ductile when subjected to large displacement demands
during a severe seismic event.
The total dissipated plastic work for the whole structure is largely increased with the presence of
the RBS. Indeed, for the performed analyses in this study the dissipation with dog-bones ranges
between 2.23 and 3.70 times that of the structure with conventional welded connections. This is
due to the achievement of higher rotations in the dog-bones, promoting more plastic dissipation.

ACKNOWLEDGEMENTS

This study was developed in the frame of the Project 2017 DI 061 “Efficient design of steel
joints subjected to cyclic loads” funded by Generalitat de Catalunya.

REFERENCES

CEN 2009. Eurocode 8: Design structures for earthquake resistance – part 1: General rules, seismic actions
and rules for buildings, European Comitee for Standardisation, Brussels.
CEN 2011. Eurocode 3: Design of steel structures – part 1-1: General rules and rules for buildings, Euro-
pean Comitee for Standardisation, Brussels.
CEN 2011. Eurocode 8: Design of structures for earthquake resistance – part 3: Assessment and retrofitting
of buildings, European Comitee for Standardisation, Brussels.
Chaboche, J.L. 1986. Time-independent constitutive theories for cyclic plasticity. International Journal of
Plasticity, vol. 2, no. 2, pp. 149–188.
Computers and Structures, INC. July 2016. CSI Analysis Reference Manual.
Dassault Systèmes, 2016. Abaqus 2016 Online Documentation.
Giugliano, M.T., Longo, A., Montuori, R., & Piluso, V. 2011. Seismic reliability of traditional and innova-
tive concentrically braced frames. Earthquake Engineering and Structural Dynamics, 40(13),1455–1474.
doi:10.1002/eqe.1098.
Krolo, P.; Grandić, D. & Smolčić, Ž. 2016. Experimental and numerical study of mild steel behaviour
under cyclic loading with variable strain ranges. Advances in materials science and engineering. Volume
2016, Article ID 7863010.
Lemaitre, J. & Chaboche, J.L. 2009. Mechanics of solid materials. New York: Cambridge University Press.
Montuori, R. 2015. Less is more: the reduction of beam section for the seismic behaviour improvement of
existing steel MRFs. 6th International Conference on Theoretical and Applied Mechanics, Salerno, Italy.
Montuori, R., & Sagarese, V. 2018. The use of steel RBS to increase ductility of wooden beams. Engin-
eering Structures, 169, 154–161.
Plumier A. 1990. New idea for safe structure in seismic zone. Proceedings of IABSE symposium on mixed
structures including new materials. p. 431–436.
SeismoSoft Ltd. 2012. SeismoMatch: User’s manual.

577
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Stability of double-symmetric sections subjected to axial force,


bending moments and torsion

F. Jörg
Researcher at the Institute of Structural Design, University of Stuttgart, Germany

U. Kuhlmann
Timber and Composite Construction, Institute of Structural Design, University of Stuttgart, Germany

ABSTRACT: To investigate the stability behavior of double-symmetric sections subjected to


axial force, biaxial bending and torsion, 12 tests were performed at the laboratory of the Univer-
sity of Stuttgart. The test specimens consisted of three different double-symmetric sections loaded
with varying axial forces and different load eccentricities. These tests were numerically recalcu-
lated so that the numerical model has been validated. This model was used for further investiga-
tions in a large parameter study to examine the influence of additional design parameters not
covered by the experimental testing program, e.g. different moment diagrams, steel grades and
various internal force combinations. Starting from the solution for N-M-T presented in Annex
C of the new draft prEN 1993-1-1 (which could be confirmed) a slightly modified suitable analytic
design model has been developed. This equation determines the member resistance as well as the
cross-sectional resistance taking into account axial compression, biaxial bending and torsion in
one equation. A criterion for neglecting the warping moment has also been developed.

1 INTRODUCTION

Double-symmetric steel rolled sections are the most important elements used in steel buildings
such as industrial halls. Nowadays, they are used for columns and trusses as well as for purlins
and crane girders. This is due to their high load-bearing capacity and flexibility as a result of its
large product selection. Consequently, the slenderness makes them vulnerable to buckling. Con-
fronted with various external influences, this results in a combination of different internal forces,
see Figure 1.

Figure 1. Practical application scenarios.

578
This can be illustrated by the example of a crane girder. The crane girder is subjected to
biaxial bending stresses with an additional torsional moment due to horizontal load applica-
tion at the top flange. Moreover, axial forces in longitudinal direction of the girder may occur
e.g. through acceleration and braking of the crane. Currently, various alternative design rules
exist to prove buckling resistance, however, these are often too complicated for practical
applications, e.g. finite element calculations or do not include all internal forces, as the current
equivalent member method in EN 1993-1-1.
Within the framework of the research project “Interactions of axial force, bending moments
and torsion: harmonization and stability check of rolled sections”, different work packages were
planned to expand and improve the existing design rules. This new approach should allow an easy
calculation for interactions like axial force, biaxial bending and torsion. To investigate the stability
behavior, 12 tests on specimens with different double symmetric sections with varying load eccen-
tricities were performed at the material testing institute of the University of Stuttgart (MPA).

2 EXPERIMENTAL INVESTIGATIONS

2.1 General
The experimental part of the research project involved tests on the resistance of beams subjected
to axial compression, bending and additional torsion. Therefore, three series of profiles com-
monly used in practice, namely IPE (purlins), HEB (columns) and HEA (light crane girders) were
investigated. For each series two tests focused on the stability behavior of the strong axis (y-y)
with different eccentric loading, also an additional test on the weak axis (z-z) was performed. The
axial compression load was applied according to the buckling resistance Nb,Rk of each profile.
For the IPE series, two different axial compression forces were tested, and for the other two series
only one level. The parameters of the testing program are summarized in Table 1.

2.2 Test girders


In order to apply the axial compression at the supports, end plates had to be welded at both
ends of the test specimens. Since the specimens could not be fork-supported directly at these
end plates, an excess length of 100 mm was allowed. The fork supports were spaced at
a distance of exactly 4000 mm, as the test girders had a total length of 4200 mm. The test
specimens differed from each other only by one parameter in order to analyze its influence on

Table 1. Testing program.


Load eccentricitiesa

Profile Identification Axial compression y-y z-z Fult/F*-151b

I-151 b/6 – 1.00


I-152 0.15 · Nb,Rk b/3 – 0.92
IPE 300 I-153 – h/2 0.53
I-301 b/6 – 0.95
0.30 · Nb,Rk
I-302 b/3 – 0.82
I-303 – h/2 0.40
II-151 b/6 – 1.00
HEB 200 0.15 · Nb,Rk
II-152 b/3 – 0.86
II-153 – h/2 0.73
III-151 b/6 – 1.00
HEA 240 III-152 0.15 · Nb,Rk b/3 – 0.79
III-153 – h/2 0.65
a
y-y – strong axis, z-z – weak axis
b
maximum load of test 151 corresponding to the relevant series

579
the stability behavior and to have a database for the validation of the FE-model used. Add-
itional rolled sections, slenderness values, internal force combinations as well as moment dia-
grams were investigated in numerical studies to examine design parameters that were not
covered by the testing program.
Investigations of Sedlacek et al. 2004 and preliminary numerical calculations have shown
local buckling of the compression flange caused by the eccentric load application of the test
girder. To apply the eccentric loading through a loading frame while simultaneously preventing
local buckling, stiffeners were welded in the middle of the girders. To avoid additional residual
stresses in the compression flange, the stiffeners were only welded to the web of the profile.
For the testing program, a uniform steel grade of S235 J2 was used. Material parameters,
i.e. yield stress, tensile strength and E-modulus were determined by additional tensile tests on
the flange and web following the recommendations given in EN 10025-1.

2.3 Test setup and procedure


The test setup of the experiments corresponded with a three-point bending test, including axial
compression at the supports and torsion through an eccentric loading at mid span of the test
girder. Due to the torsion and the lateral torsional buckling of the test girders, a fork restraint
had to be arranged. To achieve a fork restraint for the supports, a frame at each end was used.
Here the flanges of the test specimens were supported by threaded rods with a teflon coating.
Axial compression was applied on the end plates through two horizontal compression cylin-
ders which were connected to establish uniform compression. The bending was applied
through a concentrated load at mid span of the specimens. By a specific load eccentricity, an
additional torsional moment was introduced in the specimens. The loading construction con-
tained a stiff loading frame consisting of HEM profiles that was connected with the stiffeners
of the specimens. The load was applied with a hydraulic cylinder, which was constructed as
a pendulum rod. To ensure a constant load, the hydraulic cylinder was moved horizontally on
a roller rail system during loading. The whole test setup can be seen in Figure 2.
The test procedure contained two loading steps. During the first loading step, axial com-
pression – magnitude depending on the buckling resistance – was applied through the two
compression cylinders. The load was increased through path control and after reaching the
planned magnitude of axial compression, the load was kept constant by using force control.
Subsequently, during the second loading step, bending and torsion were applied. To ensure
a constant load, the rotation of the hydraulic cylinder was observed through an inclination
sensor. In the event of small rotations, the vertical cylinder was moved horizontally during load-
ing on a roller rail system. Through force control, the load was increased, until approximately
60% of the previously calculated numerical loading capacity had been reached. Afterwards, the
load was increased further in a path-controlled manner up to the maximum loading capacity.

Figure 2. Test setup with fork restraints, loading frame and hydraulic cylinder on a roller rail system.

580
2.4 Measurements
Before testing the specimens the geometric imperfections were measured with a laser scanner.
Afterwards, the measured values of the point cloud were transformed into a sinusoidal pre-
forming which was used for the numerical recalculations of the tests, see Figure 4. Several dif-
ferent measuring devices were used to measure the behavior of the test specimens. The
horizontal and vertical displacements were recorded using several displacement sensors at the
supports and through integrated sensors in the hydraulic cylinders. The rotations were meas-
ured with several inclination sensors in the middle and at both fork supports. To observe the
yielding, strain measurements were added in the middle of the specimens.
Apart from the mechanical measurements, the three-dimensional deformations of the speci-
mens were measured in the middle of the span by photogrammetric means, see Figure 3. With
the photogrammetric method, photos of the decisive panel could be taken with four special
cameras from different positions. The data were generated by an intersection in space. The
surface of the measuring panel was marked with coded targets to clearly follow the deform-
ations of a specific point during testing, see Figure 5.

2.5 Comparison of ultimate load capacity


Table 1 includes the load capacity of each specimen in relation to the maximum load of the
respective test series (I-151, II-151 and III-151), thereby revealing the influence of each param-
eter. The highest load capacity of all three profiles can be found when subjected to the strong
axis with a small eccentricity of b/6 and an axial force of 15% of the buckling resistance Nb,Rd.
The highest load capacity for all specimens can be found in the HEA 240 profile.
For the test specimens subjected to bending on the strong cross-section axis (y-y), mainly
a lateral torsional buckling failure occurred in the tests. For the tests in series I with an axial
force of 30% of the buckling resistance a flexural buckling could be observed first, which
changed into a lateral torsional buckling failure with increasing bending. On the other hand,
the test girders subjected to bending on the weak cross-section axis (z-z) were more likely to
show a cross-section failure.

Figure 3. Setup for photogrammetric measurements.

581
Figure 4. Measured values and sinusoidal preforming for numerical recalculations.

Figure 5. Photogrammetric measurements.

3 NUMERICAL INVESTIGATIONS

3.1 Numerical simulations of tested girders


In collaboration with Prof. Taras from the University of the Federal Armed Forces Munich,
a numerical model was developed using the software package Abaqus. This numerical model
was validated by comparing loading capacity, load-displacement behavior and local deform-
ations between the experiments and the numerical recalculations.
The numerical model is a shell model with an ideal fork restraint as boundary conditions and
free warping at the ends. For recalculation of the tests, the end plate and overlap were modeled
in order to take restraint warping into account. Four-node linear shell elements with reduced
integration (S4R) were chosen to model the studied sections. The mesh contained 16 elements
per flange and web, where the residual stresses according to Figure 7 were applied. The single
plates were connected to each other by means of rigid coupling beams. Through kinematic
couplings at the supports a stiff load introduction was achieved that allowed the cross-section
to twist and rotate freely. By using additional beam elements, which were placed in the center
of the flanges the fillets of rolled sections were taken into account. The cross-sections of these
beam elements corresponded to equivalent box cross sections which allowed the modelled
member to match the torsional stiffness given by the production standards for rolled sections.
For the numerical simulation of the tests, the material characteristics of the flange and web
were gathered from the tensile tests. These engineering material curves were then translated
into true strain-stress curves. The elastic modulus E used in the analysis was achieved by
means of tensile material tests, furthermore the Poisson’s ratio was assumed to be ν = 0.3.

582
3.2 Load-carrying capacity
The overall results have been compared with the load-displacement curves collected on the
basis of tests and numerical simulations. At the same time, the loading capacity and stiffness
acquired in the numerical simulations are compared with the test results.
Figure 6 shows the results for specimen II-151 subjected to the strong axis y-y with an axial
compression of approx. 15% of the buckling resistance and a load eccentricity of b/6. The
load-displacement curves show a good agreement.

3.3 Parameter study


This validated model was used for further numerical investigations in a large parameter study
in order to analyze an appropriate design approach. The numerical model used showed free
warping at the ends, geometric imperfections according to the lateral-torsional buckling eigen-
mode, residual stresses according to the ECCS Publication No. 33 and a material approach
agreeing to the Swedish Steel Standard with additional strain hardening, see Figure 7.
The parameters investigated within the numerical study consisted of various steel grades from
S235 up to S460, nine profiles, related slenderness values from 0.4 up to 2.0, different moment
diagrams (constant, linear, parabolic) and numerous varying internal force combinations.

4 DESIGN RESISTANCE

4.1 Interaction according to prEN 1993-1-1


An initial approach to calculate the design resistance for members subjected to axial compres-
sion, biaxial bending and torsion was developed during the research project P554 in 2004, car-
ried out in collaboration of the Universities of Aachen, Berlin and Bochum leading to the
equation in EN 1993-6, Annex A. A new approach, which can be found in Annex C of the

Figure 6. Load-displacement curves for test girder II-151 obtained by numerical recalculation and test.

Figure 7. Numerical model for the parameter study.

583
new draft prEN 1993-1-1, combined the equation of current EN 1993-1-1, 6.3.3 for axial com-
pression and biaxial bending with an additional term for a warping moment, see Knobloch
et al. 2017. Using the approach for stability verification, an additional cross-sectional resist-
ance check has to be provided. This new design approach evaluated by a numerical database
that was created by researchers from the University of Bochum. Independently, the tests and
the numerical database developed within this running project confirmed that the approach
given in the new draft shows a safe side solution, in case aside of the member stability check
also the cross section has been verified separately.

4.2 Modified design approach


The approach of the new draft was modified according to the results of the numerical database.
The modified version, which can be seen in (1) showing the equation for the strong axis, does not
only cover the member resistance but also cross-sectional verification when shear stresses are of
negligible size. These rules fully fit into the rules as they are now given in Annex C of the new
draft.

NEd kyω  My;Ed Mz;Ed BEd


þ kyy   þ kyz   þ kn  kmz  kα  ð1Þ
χy  NRk=γ χLT  My;Rk γ Mz;Rk
γM1
BRk=
γM1
M1 M1

with

kyω ¼ 1  0:25  mω

kn ¼ 1  nz  mω

kmz ¼ 1  mz
 
My;Ed 1
kα ¼ 1 2
Mcr

The factor kα in the interaction equation covers the effects of the 2nd theory order, whereas kmz
considers the dissipation of the warping moment due to the bending moment Mz. The function kn
is evaluated and adopted for the numerical results. The factor kyω leads to more economic results
and is a function dependent on the warping moment. kyy and kyz are taken from the main text in
EN 1993-1-1. In case the warping moment and torsion effects tend to zero the equation fully rep-
resent the original equation of axial compression force and biaxial bending in the main text of
EN 1993-1-1. The modified approach can be seen in Figure 8, where the numerical results re,FEM
are compared with the analytic results rt. In (a) the results for the constant moment curve and
steel grade S235 are visible, whereas in (b) a linear moment curve and steel grade S355 can be
found. Most of the results lie within the +/– 10% deviation. For beams that are almost exclusively
subjected to Mz and B, the load capacities are partly slightly overestimated with the analytical
model. However, at the same time these cases are not of great practical importance.
The safety level has also been checked according to the procedure in EN 1990 Annex D which
leads to a γM1 value of 1.05 for the modified design approach, comparable with a γM1 value of
1.01 for the current equation in the new draft. For the evaluation, the moment diagrams constant,
linear and parabolic as well as the steel grades S235 up to S460 were considered.

5 CRITERION OF NEGLECTION

In collaboration with the University of Bochum a criterion has been developed indicating if
the warping moment may be neglected. When the criterion (2) is satisfied the interactions
equation simplifies to the common N-My-Mz verification of EN 1993-1-1. A limit of 0.07 lead

584
Figure 8. Comparison between analytical rt and numerical re load factors.

Figure 9. Diagrams to check the criterion for disregarding the warping moment.

to a maximum deviation of 5% of the load resistance factor. This criterion was also taken into
account in the analytic results in Figure 8 and could be presented in the form of diagrams, for
the most commonly used rolled sections.

kα  mω  0:07 ð2Þ

The warping moment BEd was calculated for different rolled sections in dependence on the
torsional moment mx,d and Mx,d. Considering the cross-sectional values, for example the follow-
ing diagrams in Figure 9 result. Since the diagrams are shown as a function of the torsional
moment, the warping moment does not have to be calculated separately. Thus, the user can dir-
ectly read the maximum permissible torsional moment and if necessary, calculate back the max-
imum permissible load eccentricity up to which a simple N-My-Mz verification is sufficient.

6 CONCLUSIONS

Experimental investigations have been conducted on 12 rolled sections within this research
project carried out by the University of Stuttgart. The tests focused explicitly on the influence
of additional torsional moments on the interaction of axial compression and bending
moments. During the testing program, the axial force and the load eccentricity were varied. It
can be seen in the tests that a load placed on the strong axis tends to lead to stability failure
and a load placed on the weak axis tends to lead to a predominant cross-sectional failure.

585
For further numerical investigations, the experimental tests have been recalculated to verify
the FE model. In order to verify, the load-displacement curves including maximum loads and
deformations of the girders were compared between tests and numerical recalculations result-
ing in a very good accordance. This numerical model was used to create a large numerical
database with different steel grades, slenderness values, moment diagrams and various internal
force combinations. Meanwhile, investigations also for mono-symmetric sections have started.
The rules for the design resistance were originally developed based on investigations carried out
during a research project at the Universities of Aachen, Berlin and Bochum. Additional research
has recently been done by the University of Bochum in the frame of development of the second
generation of Eurocodes. In this paper a modified design approach based on the numerical results
has been presented, which considers the cross-sectional resistance as well as the resistance of the
member and allows the application of a simple criterion for neglecting the warping moment.

ACKNOWLEDGMENTS

The work presented has been carried out as part of the research project “Interactions of axial
force, bending moments and torsion: harmonization and stability check of rolled sections”.
This project of the research association DASt is financed over the AiF within the development
program for industrial community research and development (IGF) from the Federal Ministry
of Economic Affairs and Energy (BMWi) based on a decision of the German Bundestag.
Special thanks goes to “Stahlwerk Thüringen GmbH” for the donation of test specimens
and “Haller GmbH” for their fabrication. The valuable contributions of the Institute for
Photogrammetry in Stuttgart, especially Dipl.-Ing. Alessandro Cefalu and Stefan
Schmohl, M.Sc., for the photogrammetric measurements are greatly appreciated, including
the material testing institute at the University of Stuttgart (MPA) for carrying out the tests.

REFERENCES

Abaqus, v. 6.13 2013, Dassault Systems/ Simulia, Providence, RI, USA.


ECCS: 1984, Ultimate Limit State Calculations of Sway Frames with Rigid Joints. European Convention
for Constructional Steelwork – TC 8, ECCS Publication No. 33, Brussels, Belgium.
EN 10025-1: 2005, Hot rolled products of structural steels – Part 1: General technical delivery conditions.
EN 1993-1-1: 2005, Eurocode 3: Design of steel structures – Part 1-1: General rules and rules for
buildings.
EN 1993-6: 2007, Eurocode 3: Design of steel structures – Part 6: Crane supporting structures.
Knobloch, M. & Winkler, R. & Walter, A. 2017, Lateral-torsional buckling resistance of steel members
subjected to axial compression, biaxial bending and torsion. Presentation at ECCS TC 8 Structural
Stability, Coimbra, Portugal.
prEN 1993-1-1: 2018, Eurocode 3: Design of steel structures – Part 1-1: General rules and rules for build-
ings. Final Draft, Document TC250/SC3/N 2661.
Sedlacek, G. & Stangenberg, H. & Lindner, J. & Glitsch, T. & Kindmann, R. & Wolf, C. 2004, Investiga-
tions on the influence of torsion effects on the plastic load-bearing capacity and the member resistance
of steel profiles. Research for practical use P554. Forschungsvereinigung Stahlanwendungen e.V.
(FOSTA), Düsseldorf, Germany.

586
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Distortional buckling of stiffeners in stainless steel profiled sheeting

J. Jůza & M. Jandera


Czech Technical University in Prague, Prague, Czech Republic

ABSTRACT: The research presents a numerical study of compressed stainless steel cold-formed
thin-walled plates with an intermediate stiffener as is common for flanges of profiled sheeting. The
numerical model was validated on tests from literature. These test results were substantially
extended by a parametric study to cover the whole slenderness range and common geometries of
stiffeners. As stainless steel material exhibits more nonlinear stress-strain behaviour than carbon
steels, the results of the numerical study also shown significant deviation of the resistance which is
contributed to the material non-linearity. The distortional buckling curve for stainless steel was
lower for the whole slenderness range and a modified design procedure was therefore proposed.

1 INTRODUCTION

1.1 Stainless steel material


As it is known, stainless steel material exhibits more non-linear stress-strain diagram without
a clearly visible yielding point. This difference in behaviour led to development of supplemen-
tary design rules for stainless steel structures, e.g. in Europe EN 1993-1-4 (2006), which pro-
vides modifications to the carbon steel rules where stainless steel behaviour differs
significantly. These modifications are mainly covering behaviour in compression affected by
instability such as global buckling, local buckling or shear buckling.

1.2 Distortional buckling in the Eurocodes


There is no additional rule suggested for stiffeners of cold-formed stainless steel sections in the
code EN 1993-1-4 (2006) and therefore the rules for carbon steel applies (1):

χd ¼ 1:0 for λd  0:65


χd ¼ 1:47  0:723λd for 0:655λd 51:38
ð1Þ
0:66
χd ¼  for 1:38  λd
λd

where the reduction factor χd for distortional buckling resistance is obtained from the non-
dimensional slenderness λd .
The Eurocode 9 for design of aluminium structures (EN 1999-1-4 2007) contains a different
formula for the distortional buckling (2). This reduction factor is significantly lower than for
carbon steel over the whole slenderness range.

χd ¼ 1:0 for λd  0:25


χd ¼ 1:115  0:62λd for 0:255λd 51:04
ð2Þ
0:53
χd ¼  for 1:04  λd
λd

587
As aluminium alloys exhibit similar degree of non-linearity as stainless steels, it may indicate
a lower reduction factor χd may also apply to stainless steels.
There was only little investigation done so far by Blaß & Ummenhofer (2010), with
a distortional buckling curve proposal published later in Misiek et al. (2010). As expected,
a significantly lower distortional buckling curve for stainless steel stiffeners was suggested.

2 NUMERICAL MODEL OF INTERMEDIATE STIFFENERS

2.1 Basic characteristics of the numerical model


The numerical model of a flange with an intermediate stiffener was created in Abaqus/CAE 6.14-
1. The experiments carried out by Blaß & Ummenhofer (2010) were numerically simulated at first.
The finite element model was made using four-node shell elements. Boundary conditions of
the model were implemented so that the model corresponds with the conditions of the above-
mentioned research. Linear supports were defined at the edges of the model. At all the edges,
the displacement perpendicular to the flange surface was prevented. At the bottom edge, the
displacement parallel with the stiffener (loading direction) was prevented. Additionally, the
displacements in all three directions were prevented in one node. Rotations of all edges
remained free. Boundary conditions are shown in Figure 1a.
The numerical model was loaded by a uniaxial displacement of the top edge, so that the flange
was compressed (Figure 1b). Resulting load-bearing capacity of the model was determined as
a maximum summation of the reactions of the top edge nodes during the process of loading.

2.2 Material model


Considering the material behaviour of stainless steel, the stress strain diagram curve is not pos-
sible to be accurately simplified by a bilinear function, as usual for carbon steels. The stress-strain
diagram curve was therefore determined by formula (3) taking account of material nonlinearity:
8  n
< σ
þ 0:002 fσy for σ  fy
E
ε ¼ σf   n ð3Þ
: y þ εu  fu fy  ε0:2 σfy þε0:2 for fy 5σ  fu
Ey Ey fu fy

where the meaning of the parameters of the model and their most recent recommendation can
be found in in Arrayago et al. (2015).

Figure 1. Numerical model (a) boundary conditions, (b) loading by displacement.

588
Table 1. Parameters of the stress strain diagram curve for models of tests.
Thickness t [mm] fy [MPa] fu [MPa] εu [-] n [-] E [MPa]

0.50 295 673 48.4 12 200 000


0.60 295 667 47.1 12 200 000
0.80 327 673 47.4 12 200 000

For the stress-strain diagram of models of the tests, experimentally measured values (Blaß
& Ummenhofer, 2010) of steel grade 1.4301 were used. The measured values of yield strength
(fy), ultimate tensile strength (fy) and ductility are presented for every nominal sheet thickness.
Considering that 3 different nominal sheet thicknesses were used in the experimental research,
3 different stress strain diagrams were created here. Also, the hardening exponent n was taken
from the experimental research, considering value n = 12 for all three thicknesses. The value
of the Young’s modulus was taken over from the code EN 1993-1-4 (2006) being 200
000 MPa. The values are summarised in Table 1.

2.3 Imperfections
Considering the fact, that compressed thin-walled plates are slender elements, their load-
bearing capacity is highly influenced by the imperfections. As membrane residual stress are
generally very low for cold-formed sections and the influence of the residual stress bending
component in the flats (being also relatively low) is present in the stress-strain diagram, only
geometric imperfections were considered.
These were considered by the shape of the elastic buckling eigenmodes. The eigenmodes
were determined by a linear buckling analysis, method Lanczos, also in Abaqus software.
According to EN 1993-1-5, the imperfection amplitude can be taken as 80% of the fabrication
tolerances given in EN 1090-2 for the governing imperfection (not necessary the lowest buck-
ling curve but the one with higher influence on the resistance). Additional imperfections are
further reduce to 70%. Two buckling eigenmodes were taken into account for imperfections of
the model of the stiffened flange. One eigenmode for distortional buckling and one eigenmode
for local buckling of the flat parts of the flange. Elastic buckling eigenmodes for local resp.
distortional buckling are shown in Figure 2.

Figure 2. Elastic buckling eigenmodes for local (left) resp. distortional buckling (right).

589
2.4 Numerical model validation
In part, the numerical model results are compared to the experimental ones. For all of the 48
specimens, which were previously tested experimentally at KIT the average prediction was
12.9% lower (safe side) with quite significant standard deviation (15.5%). There were only pre-
dictions of three models on the unsafe side (by 0.2; 1.2 and 6.6%), but some model predictions
were significantly conservative.
The cause of the lower value of the capacity according to the numerical simulation is prob-
ably in the way of implementing the imperfections. The size of the imperfections was con-
sidered in the model based on conservative fabrication tolerances. However, despite the real
sizes of imperfections of specimens tested experimentally were not measured, they were prob-
ably significantly lower. As some specimens were very sensitive to the imperfection magnitude,
it may explain the lower resistance prediction of the model.
Finally, it can be said, that the accuracy of the model is acceptable and the above presented
modelling procedure is possible to be used for a parametric study.

3 PARAMETRIC STUDY

3.1 Determination of the specimen length for numerical simulations


As the length of the experimentally tested specimens was limited by the dimensions of the testing
machine, but the typical length of trapezoidal sheeting is significantly higher, it was necessary to
model the flanges much longer in comparison to the experimentally tested specimens. The para-
metric study should represent a real flange, where distortional buckling is not affected by the
length of the flange and where the results could be used for a safe structural design procedure
development.
A series of numerical simulations with the gradually increasing length of the model was car-
ried out. It was needed to determine the length of the flange, for which a further increase of
length does not lead to a significant change of the resistance. The length of the model had
been changed in the multiples of the width of the flange. As it is shown in Figure 3 for two of
the models with different distortional buckling slenderness, considering the length by ten
times the width is a safe value and further extension in the length would just need higher com-
putational time but no difference in the resulting resistance.

3.2 The geometrical and material properties


Numerical models of flanges with an intermediate stiffener were created for the parametric
study. The models differed from each other in the geometry of the cross section and in the
sheet thickness. The length of the model was always equal to ten times its width. Models of
the flanges with an intermediate stiffener were divided in two groups according to the geom-
etry of the cross section, to flanges with a narrow groove and flanges with a wide groove
(Figure 4). In total, there were 60 different models according to the geometry. All of them
were created in four different material variants.

Figure 3. Load-bearing capacity of the flange depending on its length.

590
Figure 4. Flange with a narrow (above) and wide (below) groove.

In the Table 2 are presented material model parameters of the four considered stress-strain
diagrams. The first one represents the average from the tests, the other three are representative
values for cold-rolled stainless steel sections as recommended by Afshan et al. (2019).
All numerical models were loaded by displacement and the load-bearing capacities were
determined. Elastic critical stress and relative slenderness were calculated for every model.
Since the load-bearing capacity was known from the numerical simulation and considering
the reduction factor for local buckling in EN 1993-1-4 is established accurately (Gardner &
Theofanous, 2008), the reduction factor χd for distortional buckling resistance could have
been iteratively calculated. The resulting values of the non-dimensional slenderness for distor-
tional buckling and the reduction factor χd for particular models were plotted to the chart so
as a comparison could be assessed.

3.3 Comparison of the numerical results with existing distortional buckling curves.
The reduction factors for particular models are plotted in the charts (Figures 5 to 7) depend-
ing on their slenderness. The comparison is done with the codified distortional buckling
curves for carbon steel and for aluminium, and with the buckling curve proposed by Misiek
et al. (2010). There are three separate charts, each one for a different material type.
From the comparison of the above mentioned buckling curves it is clear, that the resulting
points from the parametric study are corresponding the best to the curve proposed by Misiek
et al. (2010), where the main deviation of the results is mainly for lower slendernesses of the
stiffener. Concerning the codified distortional buckling curves, the buckling curve for alumin-
ium (EN 1999-1-4) is closer to the results for stainless steel. The buckling curve for carbon
steel (EN 1993-1-3) appears to be the least suitable. Almost all the results are under the curve,
which means that the curve significantly unsafe for stainless steel.

3.4 Proposal of distortional buckling curve for stainless steel


A proposal of the distortional buckling curve for stainless steel was the final step of the paramet-
ric study. The formula with its parameters was determined with the least-squares method as (4):

Table 2. Material model parameters.


Material fy [MPa] fu [MPa] εu [-] n [-] m [-] E [MPa]

Austenitic 295 670 0.48 12.0 2.2 200 000


Austenitic 280 580 0.50 9.1 2.3 200 000
Duplex 530 770 0.30 9.3 3.6 200 000
Ferritic 320 480 0.16 17.2 2.8 200 000

591
Figure 5. Reduction factor χd for austenitic steel.

Figure 6. Reduction factor χd for duplex steel.

χd ¼ 1:0 for λd  0:43


χ ¼ 1:43  λd for 0; 435λd 50:82
d ð4Þ
0; 5
χd ¼  for 0:82  λd
λd

592
Figure 7. Reduction factor χd for ferritic steel.

Table 3. Comparison of reduction factors χd for distortional buckling resistance according to different
distortional buckling curves.
Formula/FE Misiek et al. (2010) EN 1993-1-3 EN 1999-1-4 proposal

Mean value 1.04 1.28 0.999 0.965


Standard deviation 0.058 0.146 0.061 0.055

3.5 Statistical evaluation


Apart from the graphical evaluation, all the mentioned distortional buckling curves were
evaluated statistically. For every numerical model of the study, there was calculated
a reduction factor χd. Further, the ratio of the reduction factor according to the buckling
curve to the reduction factor determined iteratively according to the FEM analysis was calcu-
lated (Formula/FE) for all buckling curves in the comparison. It is given together with the
standard deviation of the ratio in Table 3. As concluded form the previous graphical compari-
son, the proposal of Misiek et al. (2010) is fitting the results best and the new curve (4) is just
a small improvement of this proposal.

3.6 Determination of the partial safety factor γM


In connection with the proposed distortional buckling curve it was also necessary to determine
the value of the partial safety factor γM. The procedure was based on the comparison of the
theoretical value of the resistance calculated using the buckling curve and the resistance deter-
mined by FEM analysis. Error propagation term Vrt was determined in a simplified way using
the maxima of the material and geometry related coefficients of variation Vmat and Vgeom.
Vmat was adopted from Afshan et al. (2015) whereas the value of Vgeom was taken from the
final draft of the code prEN 1993-1-1:2015 as published in 2017 resp. its Annex E. During the
procedure of calculating the design value of the resistance, mean values of all basic input vari-
ables were needed. Again, material parameters were acquired from Afshan et al. (2015) and
geometrical parameters were acquired from the Annex E of prEN 1993-1-1:2015 – Final
Draft. Finally, partial safety factor γM was calculated for all specimens. As shown in Table 4,

593
Table 4. Statistical evaluation procedure according to EN 1990 Annex D.
Correction
Vδ Vgeom Vmat Vr,t fy,mean/fy,min γM,mean*
Material factor b

Austenitic 0.9893 0.0484 0.025 0.06 0.0650 1.41 1.070


Duplex 1.0671 0.0618 0.025 0.03 0.0391 1.20 1.081
Ferritic 1.0179 0.0529 0.025 0.045 0.0515 1.38 1.033

the mean value of γM for each stainless steel grade is below 1.1. Therefore, the codified value
γM0 = 1.1 can be accepted.

4 CONCLUSION

The paper describes development and validation of a numerical model of compressed stainless
steel flange with an intermediate stiffener. The validation was done using previously published
tests (Blaß & Ummenhofer 2010).
Later, the model was modified for its use in a parametric study. The study considered various
geometries of the flange, shape and size of the stiffener and three stainless steel grades (austenitic,
duplex, ferritic). The calculated results were compared to the existing procedures for distortional
buckling reduction factor χd. The currently used buckling curve, originally determined for carbon
steel, according to the EN 1993-1-3, was shown to be significantly unsafe for stainless steel mem-
bers. On the other hand, better correspondence with the obtained results was shown for the buck-
ling curve for aluminium alloy elements given by EN 1999-1-4. The last evaluated buckling curve
was the one proposed by Misiek et al. (2010). This curve shown a good correspondence with the
calculated results with only some deviation for smaller stiffener slendernesses.
Finally, a new distortional buckling curve, which correspond to the parametric study results
accurately was determined and its safety proved by a reliability study.

ACKNOWLEDGEMENT

The authors would like to thank to Thomas Misiek for his support and consultations. The
support of the GAČR 17 - 24769S “Nonlinear stability and strength of slender structures with
nonlinear material properties” is gratefully acknowledged.

REFERENCES

Afshan, S., Francis, P., Baddoo, N.R. & Gardner. L. 2015. Reliability analysis of structural stainless
steel design provisions. Journal of Constructional Steel Research 114: 293–304.
Afshan, S., Zhao, O. & Gardner, L. 2019. Standardised material properties for numerical parametric
studies of stainless steel structures and buckling curves for tubular columns. Journal of Constructional
Steel Research 152: 2–11.
Arrayago, I., Real, E. & Gardner, L. 2015. Description of stress-strain curves for stainless steel alloys.
Materials and Design 87: 540–552.
Blaß, H. J. & Ummenhofer, T. 2010. Stahltrapezprofile aus nichtrostenden Stählen. Bericht Nr.: 091503
Karlsruher Institut für Technologie (KIT).
EN 1090-2. 2008. Execution of steel structures and aluminium structures – Part 2: Technical requirements
for steel structures. Brussels: European Committee for Standardization (CEN).
EN 1993-1-4. 2006. Eurocode 3 - Design of steel structures, Part 1-4: General rules - Supplementary rules
for stainless steels. Brussels: European Committee for Standardization (CEN).
EN 1993-1-5. 2006. Eurocode 3 - Design of steel structures, Part 1-5: Plated structural elements. Brussels:
European Committee for Standardization (CEN).
EN 1999-1-4. 2007. Eurocode 9 - Design of aluminium structures, Part 1-4: Cold-formed structural sheeting.
Brussels: European Committee for Standardization (CEN).

594
Gardner, L. & Nethercot, D.A. 2004. Experiments on stainless steel hollow sections – Part 1: Material
and cross-sectional behaviour. Journal of Constructional Steel Research 60(9): 1291–1318.
Gardner, L. & Theofanous, M. (2008). Discrete and continuous treatment of local buckling in stainless
steel elements. Journal of Constructional Steel Research 64(11),1207–1216.
Misiek, T., Krüger, H., Kathage, K. & Ummenhofer, T. 2010. Buckling of stiffeners for stainless steel
trapezoidal sheeting. Steel Construction 3(4): 225–230.

595
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Appropriate spring stiffness models for the end support of bolted


single steel angle members in compression

M. Kettler, H. Unterweger & T. Harringer


Institute of Steel Structures, Graz University of Technology, Austria

ABSTRACT: Preliminary numerical and experimental investigations by the authors have


highlighted the fact that the stiffness of the rotational restraints at the gusset plates near the
member’s ends are crucial for the prediction of the compression member capacity of bolted
angles. Consequently, analytical models for the estimation of appropriate spring stiffness
values have been developed for several practical applications in buildings and two-bolt con-
nections at both member’s ends. Within this paper the developed formulae are presented
alongside with a detailed background of their derivation. The formulated stiffness values can
in future be used to improve the accuracy of the prediction of the compression member cap-
acity of single steel angles with bolted end-connections.

1 INTRODUCTION

Single steel angles with bolted end-connections are often used as bracing members in trusses and
frames for buildings or in lattice transmission towers, because they can be fabricated and erected
very easily and very quickly. The eccentric connection on only one leg induces additional bending
moments leading to a complex load carrying behaviour, especially for compression members.
In (Kettler et al. 2017 & 2019), the authors highlighted the fact that the stiffness of the rotational
restraints at the gusset plates near the member’s ends are crucial for the prediction of the compres-
sion member capacity of bolted angles. Laboratory tests as well as numerical finite element calcu-
lations have been carried out. Three different boundary conditions on both ends have been
investigated: Boundary condition 1 (BC1) is a clamped support. BC2 is a knife edge support that
allows only for rotations about the axis parallel to the connected leg. BC3 is a fully hinged
support.
Figure 1 shows a comparison between the conducted laboratory tests, a numerical parametric
study and relevant European design standards for L80 x 8 sections with two-bolt connections at
both members ends. In the diagram, the compression member capacity NR is related to the plas-
tic capacity of the section Npl and plotted over the slenderness λv about the minor axis of the
section. The experimental tests and the numerical calculations comprise all three different
boundary conditions. Obviously, the difference in resistance between BC1 (clamped support)
and BC2 (knife edge support) is much higher than between BC2 and BC3 (hinged support). The
ultimate compressive strength for members with BC1 is also significantly higher than current
code provisions. Generally, the high influence of different boundary conditions is evident.
Figure 2 shows the results of an additional numerical parametric study. The transition from
BC2 to BC1 is analysed by varying the rotational stiffness cφ out of plane of the gusset plate
at both ends. Three different angle sections (L60 × 6, L80 × 8 and L120 × 12) with two differ-
ent member lengths (λv ≈ 1:0 and λv ≈ 2:8) are investigated. The thickness of the gusset plate is
assumed very high (t = 25 mm), to prevent bending deformations out of plane. The compres-
sion member capacities NR are again related to the plastic capacity of the section Npl and plot-
ted over the logarithmic scaled rotation stiffness cφ. The results show a lower shelf (cφ < 107
Nmm/rad) and an upper shelf (cφ > 109 Nmm/rad) behaviour, corresponding to the resistances
for BC2 and BC1 in Figure 1, respectively.

596
Figure 1. Comparison of L80x8 test results for two-bolt connections with design standards and numer-
ical calculations for different end support conditions (BC1, BC2 and BC3).

Figure 2. Influence of variation of stiffness cφ out of the gusset plate plane on the compression capacity
of the member - transition from BC2 to BC1.

597
Further numerical investigations have shown that common connections for practical appli-
cations usually show stiffness values in the transition zone (Figure 2) between the two border-
line cases. This points out the importance of the consideration of the actual rotational
restraints at the member’s ends for the calculation of realistic compression member capacities.
Therefore, this paper presents appropriate spring stiffness models for three different connec-
tion details that are often used in practical applications. The comprehensive parametric study
for validation of these analytical formulae are given in (Harringer, 2017).

2 SPRING STIFFNESS MODELS FOR CONNECTIONS IN PRACTICE

2.1 Detail 1 – Rigid end plate connection


The first investigated constructional detail is a simple gusset plate that connects the attached
bolted steel angle with a fixed support (comparable to BC1 of the laboratory tests). This detail
represents a hybrid joint to a concrete wall with a steel end plate. This steel end plate gives
a clamped support to the end of the gusset plate. Figure 3 shows two configurations with dif-
ferent orientations of the steel angle. Detail 1a (Figure 3 left) with a horizontal angle section
and detail 1b (Figure 3 right) with an inclination of 45°. The Figure also illustrates the
moment distribution out of plane of the gusset plate within the theoretical model of
a cantilever beam, that is the basis for the rotational stiffness cφ1. The shown moment distribu-
tions have been validated with a numerical parametric study, using finite element models with
solid elements. Interestingly, detail 1 shows a linear increasing moment distribution from the
outermost bolt (and not from the second bolt, as it was expected) to the end of the angle sec-
tion. From that point on to the end of the gusset plate the moment is constant. The model

Figure 3. Theoretic model for detail 1a (left) and detail 1b (right) for spring stiffness cφ1.

598
represents those rotation direction (orientation of moment M), which results in the smaller
stiffness cφ1.
It is worth noting that additional calculations also verified that the rotational stiffness in
plane of the gusset plate can be taken as infinite.
Equation 1 shows the derivation of the rotational stiffness cφ1 for detail 1:

M 1 EI1 3EI1
c’1 ¼ ¼ ¼Ð  ¼ ð1Þ
’ ’ M  M dx 3x þ y

where E = Young’s modulus; x, y = geometric dimensions, see Figure 3.


The flexural rigidity of the gusset plate I1 can be calculated with Equation 2.

heff  t3
I1 ¼ ð2Þ
12

where t = thickness of the gusset plate; heff = effective plate height, see Figure 3.
The therein used effective plate height heff, has been calibrated with finite element calcula-
tions for detail 1a and detail 1b. The resulting formulae are given in Figure 3. For detail 1a,
heff is calculated based on the geometrical parameters x and y and a load distribution angle of
2:1. For practical reasons it is limited to the height of the gusset plate h. For detail 1b, heff can
be calculated with the distances a, b and c, assuming a plastic yield line that is no longer
straight, but shows one or two kinks.

2.2 Detail 2 – joint to girder flange


The second investigated constructional detail consists of a single steel angle that is bolted to
the upper flange of an I-shaped girder of length L, see Figure 4. Compared to detail 1, detail 2
is more complex, because now two deformation modes (a local and a global one) have to be
considered for the rotational stiffness out of plane of the upper flange. The local mode repre-
sents the local deformation of the upper flange and the restraining effect of the girder’s web,
while the global mode covers the rotation of the whole I-section about its longitudinal axis.
Additional finite element calculations have shown that the rotational stiffness in plane of
the upper flange can be taken as infinite for practical design configurations.

Figure 4. Detail 2 with relevant moment M (left) and investigated local and global modes (right).

599
Equation 3 shows the derivation of the local rotational stiffness cφ2,local out of plane for
detail 2. The moment M = 1 is acting at the upper edge of the web, that is assumed to be
a simply supported beam, supported by the two flanges. This results in a triangular moment
distribution in the web.

M 1 EI2 3EI2 3Et3w  leff ;2


c2;local ¼ ¼ ¼R  dx ¼ ¼ ð3Þ
  MM hw 12hw

where tw = web thickness of girder; hw = web height of girder.


The parameter leff,2 represents the effective width of the web (in direction of the girder axis)
that is contributing to the rotational stiffness. Additional finite element calculations have been
carried out to calibrate this parameter for practical configurations. The derived formula is pre-
sented in Equation 4.

2:05  h0:5
w  tf
1:25
 b0:5
f
leff ;2 ¼ ð4Þ
t1:25
w

where tf = flange thickness of girder; bf = flange width of girder.


In addition to the local mode, also the global mode (i.e. the rotation of the whole I-section
girder) has to be considered for detail 2. A numerical sensitivity analysis has confirmed the fact
that the rotational restraint from I-section girders is usually not high enough to assume the
girder as rigid for the verification of the angle member. The global rotational stiffness cφ2,global
can either be calculated by means of a sophisticated beam model or by Equation 5. The latter is
based on a simply supported girder with end fork conditions and a single torsional moment M
= 1 for the girder in the middle between the two supports.
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
M 1 1 IT L
c’2;global ¼ ¼ ¼  with εT ¼  ð5Þ
’ ’ L
 1  tanh εT 2:6  Iω 2
4GIT εT

where L = girder length; G = shear modulus; IT = torsional stiffness; Iω = warping stiffness.


The total rotational stiffness of detail 2 can then be calculated with Equation 6 assuming
a series connection of the two spring stiffness parameters:

1
c’2 ¼ ð6Þ
1
c’2;local þ c’2;global
1

2.3 Detail 3 – joint to girder web


The third investigated constructional detail consists of a gusset plate that connects the bolted
angle member to the web of an I-shaped girder of length L, see Figure 5. The distance of the
gusset plate from the upper flange u can vary between 0 ≤ u ≤ hw. As illustrated in Figure 5,
the stiffness values in and out of plane of the gusset plate have to be considered for detail 3,
because of the flexibility of the web plate in both directions. The stiffness parameter cφ,3,out
(Figure 5 left) has to take into account the flexibility of the gusset plate (as for detail 1), the
local bending stiffness of the girder’s web and the global stiffness of the girder (as for detail 2,
global effect), while cφ,3,in (Figure 5 right) has to cover the local bending stiffness of the gir-
der’s web and the global stiffness of the girder. The corresponding formulae are described in
the two following chapters.

600
Figure 5. Detail 3 with moment out of plane Mout (left) and moment in plane of gusset plate Min (right).

2.3.1 Rotational stiffness out of plane of gusset plate


The rotational stiffness cφ,3,out of the constructional detail 3 comprises the local stiffness
cφ,3,out,local and the global stiffness cφ,3,out,global. The local stiffness cφ,3,out,local can be estimated
with Equation 7 and consists of the following two parts: i) the flexibility of the gusset plate cφ1
that can be calculated in accordance to the constructional detail 1 in chapter 2.1, ii) the local
bending stiffness of the girder’s web cφ,3,out,web, see Equation 8.

1
c’3;out;local ¼ ð7Þ
1
c’1 þ c’3;out;web
1

The local bending stiffness of the girder’s web is derived by assuming a moment Mout = 1 at
a position u/hw from the upper flange. The web is thereby modelled as a simply supported
beam with supports at the web-flange connections and an effective width leff,3 in longitudinal
direction of the girder. Equation 8 also shows the result for the specific case of u/hw = 0.5.

3EI3 3E  leff ;3  t3w forhuw ¼ 0:5 E  leff ;3  t3w


c3;out;web ¼  2 ¼  2 ¼ hw
hw  1  3 huw þ 3 u
hw 12hw  1  3 huw þ 3 huw

ð8Þ

where u = distance from gusset plate to upper flange.


The geometric parameter leff,3 has been calibrated on a numerical parametric study for prac-
tical configurations of detail 3. The resulting formula is given in Equation 9.

70  h0:75
leff ;3 ¼ w
ð9Þ
t1:4
w

The global rotational stiffness value cφ,3,out,global can be estimated in accordance to the con-
structional detail 2 in chapter 2.2 (Equation 5).

c’3;out;global ¼ c’2;global ð10Þ

601
The total out of plane rotational stiffness of detail 3 can then be calculated with Equation
11 assuming a series connection of the local and global spring stiffness parameters:

1
c’3;out ¼ ð11Þ
1
c’3;out;local þ c’3;out;global
1

2.3.2 Rotational stiffness in plane of gusset plate


For detail 3, also the rotational stiffness in plane of the gusset plate has to be taken into
account. This was confirmed by additional finite element calculations. Again, the total stiff-
ness covers the local and the global stiffness parameters. The model for the local rotational
stiffness cφ,3,in,local is based on the assumption that the acting moment Min = 1 in Figure 5
(right) can be interpreted as a pair of forces F with a distance hplate in longitudinal direction of
the I-shaped girder, where hplate is the width of the gusset plate. The out of plane deformation
w of the girder’s web due to a single load F (position of loading at x = 0) can be formulated
based on plate theory (Girkmann 1963):
 
Fh2w X 1 nπ  x nπ : x nπ  u nπ  z Et3w
w¼ 1 þ  e hw sin sin with K ¼ ð12Þ
3
2Kπ n n 3 hw hw hw 12ð1  v2 Þ

where x = coordinate in longitudinal direction of the I-shaped girder; z = coordinate from the
upper flange in vertical direction; ν = Poisson’s ratio for steel (ν = 0.3).
Since only the deformation at the height z = u is of interest for the following investigations,
Equation 12 can be simplified to Equation 13:
 
Fh2w X 1 nπ  x nπ  u
 e hw sin2
nπx
w¼ 1þ ð13Þ
2Kπ3 n n3 hw hw

Inserting F = Min/hplate = 1/hplate in Equation 13 and stopping the summation after the first
term, the out of plane deformation w due to a single load F can be written in the following
form:
 
h2w πx πx πu
w¼ 1þ  e hw sin2 ð14Þ
2Kπ3 hplate hw hw

where hplate = width of gusset plate.


The difference in out of plane deformation Δw between the two load introduction points
(i.e. the endpoints of the gusset plate) due to both loads F can be calculated by Equation 15:
 

  h2w 2πu π  hplate πh
 hplate
Δw ¼ 2  wðx ¼ 0Þ  w x ¼ hplate ¼ sin 1  1 þ  e w ð15Þ
Kπ3 hplate hw hw

For hplate/hw < 0.75, the term in parenthesis of Equation 15 can be approximated by hplate/hw
(see Figure 6).
The approximated difference in out of plane deformation Δw* is therefore given by:

hw πu
Δw ¼ sin2 ð16Þ
Kπ3 hw

The resulting local rotational stiffness in plane of the gusset plane is presented in Equation 17:

602
Figure 6. Linearisation of the parenthesis-term in Equation 15.

Figure 7. Compression member with rotational restraints at member ends and eccentric loading N.

603
1 hplate Kπ3 hplate Et3w
c’3;in;local ¼ ¼ 
¼ with K ¼ ð17Þ
’ Δw 2 πu
hw sin hw 12ð1   2 Þ

The global stiffness in plane of the gusset plate can be calculated with Equation 18 under the
assumption that the moment Min is acting on the girder in the middle between two supports
(leading to minimum spring stiffness).

12EIz
c’3;in;global ¼ ð18Þ
L

where Iz = flexural stiffness of girder about the minor axis.


The total in plane rotational stiffness of detail 3 can then be calculated with Equation 19:

1
c’3;in ¼ ð19Þ
1
c’3;in;local þ c’3;in;global
1

3 CONCLUSIONS

In this paper, analytical models for the estimation of appropriate spring stiffness values have
been developed for several practical applications in buildings and two-bolt connections at
both member’s ends. The formulated stiffness values can in future be used to improve the
accuracy of the prediction of the compression member capacity of single steel angles with
bolted end-connections. Figure 7 illustrates the theoretical model of a bolted angle member
with rotational restraints at both member’s ends and an eccentric compression loading N. For
slender angle members, prone to flexural buckling, a second order beam analysis (1D-model)
with equivalent geometric imperfections and with the newly developed end restraints (cφ,3,in
and cφ,3,out) should provide realistic results of the compression member capacity. Therefore,
current activities of the authors focus on a comparison between the discussed beam model
with realistic end restraints and laboratory tests as well as finite element calculations to end
up with accurate equivalent geometric imperfections.

REFERENCES

Girkmann, K. 1963. Flächentragwerke, 6th edition. Vienna: Springer.


Harringer, T. 2017. Ermittlung baupraktischer Anschlusssteifigkeiten von Winkelprofilen als maßgebende
Einflussgrößen der Knicktragfähigkeit. Master thesis. TU Graz.
Kettler, M., Taras, A. & Unterweger H. 2017. Member capacity of bolted steel angles in compression:
Influence of realistic end supports. Journal of constructional steel research 130: 22–35.
Kettler, M., Lichtl, G. & Unterweger, H. 2019. Experimental tests on bolted steel angles in compression
with varying end support conditions. Journal of constructional steel research 155: 301–315.

604
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Global buckling strength of built-up cold-formed steel column


under compression

T. Kobashi & N. Shimizu


Nippon Steel Corporation, Chiba, Japan

ABSTRACT: This paper reports the numerical analysis results regarding the flexural buckling
and maximum strengths of built-up cold-formed steel members under compression. Four different
section types were investigated in terms of their flexural buckling and maximum strengths. In add-
ition, we proposed a design formula about the flexural buckling strength of built-up members.
Through a comparison between the numerical analysis results and the proposed evaluation results,
it was found that the proposed method agreed well the numerical analysis results.

1 INTRODUCTION

Cold-formed steel structures are widely used as primary structural components of wall panel
structures in Japan. These wall panel structures are resistant toward lateral shear force caused
by earthquakes, and the shear force is translated to the foundation through stud members
that are framed inside the wall panels. Thus, increasing the loading capacity of the stud
member is an important design issue.
Using a built-up member is a common method to increase the maximum strength of the
stud member. These built-up members typically exhibit higher global buckling strength than
the sum of the individual member capacities because the members comprising the built-up
column section restrained their flexural buckling each other. This capacity increase is well-
known and current design standards and specifications (for examples: AISI. 2016 and AIJ.
2002) provide design equations to modify the slenderness ratio of built-up members. In add-
ition, many investigations regarding the buck-to-buck built-up lipped C section members have
been conducted and original design equations and methods have been developed (for
examples: Fratamico et al., 2018 and Sato et al. 2015).
Recently, the wall panels with a large shear loading capacity have been developed using
a steel sheeting (Kawai et al., 2016). In addition, 4 story buildings which applied this high
strength wall panel have been constructed in Japan, a country with serious seismic issues. This
shear loading capacity increase requires the maximum strength increase of the stud members,
thus, complicated built-up section shapes are employed. On the other hand, the current design
method assumes only buck-to-buck built-up sections, therefore, we could not estimate the
maximum loading capacity of other types built-up section members without further experi-
ments or finite element analysis.
This study was performed to provide a design formula that can evaluate the flexural
buckling strength of built-up section members not only for back-to-back C section mem-
bers but also for other types of built-up section members. In this paper, four different
section types were investigated. We conducted eigenvalue and incremental elasto-plastic
analyses to investigate the flexural buckling strength and the maximum strengths of the
built-up section members. In addition, the design formula for elastic flexural buckling
strength is proposed.

605
2 ELASTIC FLEXURAL BUCKLING STRENGTH ON BUILT-UP SECTION
MEMBERS

2.1 Design equations for the flexural buckling strength of two layers built-up members
To evaluate the flexural buckling strength of built-up section members, we assumed a two
layers model (Newmark 1951) as shown in Figure 1. This model is composed of two members
with the same section shape. The members are connected by an elastic spring that can trans-
late only shear force when a relative displacement in the longitudinal direction occurs. From
Figure 1(b), we could obtain the equilibrium of force and moment as follows:

Forces in the xdirection Na ¼ Nc þ Nt ð1aÞ

Forces in the ydirection Qa ¼ Qc þ Qt ð1bÞ

Moment Ma ¼ Mc þ Mt þ Nc xc þ Nt xt ð1cÞ

When we consider the small element in Figure 1(c), we also obtained the equilibrium of force
in the z direction where v is the deflection in the orthogonal direction as follows:

d2v d2v
Tensile side dQT þ NT dx þ qdx dx ¼ 0 ð2aÞ
dx2 dx2
d2v d2v
Compression side dQC þ NC 2
dx  qdx 2 dx ¼ 0 ð2bÞ
dx dx

Both tensile side and compression side members had the same section shapes, thus, Mc = Mt
and Mc+dMc = Mt+dMt. Therefore, the relationship dQT = dQC holds true and we could
obtain the following equation by substituting it into Eqs. (2a) and (2b).

d2v d2v
ðNT  NC Þ 2
dx þ 2qdx 2 dx ¼ 0 ð3Þ
dx dx

When δ is defined as the relative displacement between the individual members in the longitu-
dinal direction, the shear force q can be expressed as kδ, where k is the shear stiffness. In add-
ition, δ can be expressed as uC-uT-2xt(dv/dx), where uC is the displacement at the centroid of the
compression side member in the x direction, uT is the displacement at the centroid of the tensile
side member in the x direction, and xt is the length between the centroidal axis of the built-up
member and that of the individual member. Thus, by substituting Eq.(1a), (1b), and (1c) into

Figure 1. Outline of the two layers model.

606
Eq.(3), the second-order differentiate for x and general solution of the differential equation was
obtained as follows. It should be noted that Nc/EA = duc/dx and Nt/EA = dut/dx.
 
d4v Na 2kx2t 2k d 2 v kNa
 2 þ þ þ v¼0 ð4Þ
dx4 4EI EI EA dx2 EAEI

v ¼ C1 coshðα1 xÞ þ C2 cosðα2 xÞ ð5Þ


vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u   2
u N
t a kx 2 k Na kx2 k kNa
α1 ¼ þ tþ þ þ tþ  ð6aÞ
4EI EI EA 4EI EI EA EAEI
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u    2
u
t Na kxt 2 k Na kx2t k kNa
α2 ¼  þ þ þ þ þ  ð6bÞ
4EI EI EA 4EI EI EA EAEI

The boundary condition assumed in this paper is v = 0 when x = L/2 and d2v/dx2 = 0 when
x = L/2. After substituting these conditions into Eq. (5), the following equation was obtained.
2     3
L L
cosh α cos α2
6 1 7
6  2   2  7 C1 ¼ 0 ð7Þ
4 L L 5 C2
α2 cosh α1 β2 cos α2
2 2

From Eq. (7), the buckling load was obtained when the C1 and C2 are non-zero values simul-
taneously. It should be noted that the determinant of the 2 × 2 square matrix on the left side is
equal to zero. Based on the boundary condition, we obtained the following equation describ-
ing the flexural buckling strength on a built-up member.
  
2EIπ2 π2 =L2 þ 2k x2t =EI þ 1=EA
Ncre n ¼ ð8Þ
L2 ð2K=EA þ π2 =L2 Þ

2.2 Shear stiffness at the screw connection in the proposed equation


The screw was modeled as a continuous shear spring with a stiffness of k in the preceding sec-
tion. We evaluated the stiffness k by assuming that the elastic potential energy stored in the
shear spring is equal to the elastic strain energy of the plate element.
Figure 2(a) shows a diagram of the screw connection where the centroidal axis of the fas-
tened plate elements is eccentric so that the bending moment, M, occurs. This bending
moment causes rotational deformation of the screw itself, resulting in relative deformation in
the longitudinal direction. For the built-up member where the web plates are connected at the
pitch l0 (Figure 2b), if the relative displacement occurs so that the shear force q is equal to all
the screws, the bending moment is distributed as an anti-symmetric shape. Therefore, assum-
ing the plate element loads the bending moment at the edges, the moment distribution at the
edges of the plate elements is approximated as M = M(y). Using this assumption, the

Figure 2. Rotation and moment around the screws.

607
relationship between M(y) and q was obtained in Eq.(9a). It should be noted we assumed that
the screws were provided symmetrically about the center of the web plate elements, resulting
in the moment distribution M(y) becoming an even function with respect to the y direction.
ð
4 M ð yÞdy ¼ eq ð9aÞ

X
∞ nπ
M ð yÞ ¼ Mn cos y ð9bÞ
n¼1
b

When we define qn and δn as the shear force and relative displacement when Mncos(nπy/b)
appears, the elastic potential energy of plate element Uns can be described as follows:
UnS ¼ qn δn =2 ð10Þ

The elastic strain energy of the plate element can be obtained from the equilibrium of the plate
element described by Eq. (11) (Timoshenko 1961) as follows:

∂4 w ∂4 w ∂4 w
þ2 2 2þ 4 ¼0 ð11Þ
∂x 4 ∂x ∂y ∂y

From the edge moment equal to Mncos(nπy/b), the deflection of the plate element can be
expressed as wn(x, y) = f(x)cos(nπy/b). Therefore, substituting wn(x, y) = f(x)cos(nπy/b) into
Eq. (11) allows us to solve it as an ordinary differential equation about x based on the bound-
ary condition. The obtained results are as follows, where the boundary condition was assumed
to be f(x) = 0 and d2f(x)/dx2 = −Mn/D in the case of x = l0/2.
 nπ 2 sinhðnπl =2bÞ nπ  nπ
Mn bl0 coshðnπl0 =2bÞ 0
wðx; yÞ ¼ sinh x  x cosh x cos y
4Dnπ sinh2 ðnπl0 =2bÞ b l0 coshðnπl0 =2bÞ b b
ð12Þ

The elastic strain energy of the plate element by bending moment is equal to the total work
by Mn cos(nπy/b) loaded at the plate edge. Therefore, the elastic strain energy Unε can be
expressed as follows:
( )
Mn2 l0 b 1 2b coshðnπl0 =2bÞ
Unε ¼  ð13Þ
4D sinh2 ðnπl0 =2bÞ nπl0 sinhðnπl0 =2bÞ

Herein, we assumed that UnS is equal to Unε. Thus, from Eqs. (10) and (13) we can obtain the
relationship between the deformation δn and shear force qn. It should be noted that the eccen-
tric bending moment qne should be equal to the ∫Mncos(nπy/b)dy.
16E t
nqn ¼ βn δn ð14aÞ
3πð1   Þ l0
2

!1
nπl0 1 coshðnπl0 =2bÞ
βn ¼  þ ð14bÞ
2b sinh2 ðnπl0 =2bÞ sinhðnπl0 =2bÞ

From Eq. (14b), the βn converges to unity with increasing nπl/2b. In addition, βn increased
with decreasing nπl/2b when nπl/2b > 0. This indicates that the value of βn should be larger
than unity. In addition, this large/small relationship is true for any value of n. Therefore, we
could obtain the following inequality:

608
X

16E tX ∞
nqn  δn ð15Þ
n¼1
3πð1   Þ l0 n¼1
2

Based on the principle of superposition, ∑δn is equal to the relative shear deformation at the
connection δ. Furthermore, the eccentric bending moment qne is equal to 4∫Mncos(nπy/b)dy
and nqn equals −8Mnsin(nπ/2)/πt. Using these relationships, it was found that ∑(nqn) was
a convergent function, so that using the temporary constant value Jq, the following inequality
about the ratio of shear spring stiffness k to Young’s modulus E could be obtained.

k 16 t
 ð16Þ
E Jq 3πð1  2 Þ l0

Eq. (16) suggests that the minimum ratio of k to E could evaluate as a value that is a constant
multiple of t/l. From these considerations, we conducted eigenvalue analyses of built-up mem-
bers and approximated the relationship between k and t/l in Section 2.3.

2.3 Evaluation of the minimum value on k/E


To evaluate the minimum value of the k/E, the eigenvalue analyses of the buck-to-buck built-
up members were conducted using the numerical analysis program MARC. Figure 3 shows an
outline of the analysis model. In this analysis, half of the built-up members were modeled con-
sidering symmetricity. The z-direction displacement and Rx, Ry rotation were fixed at the
bottom edges and axial loads were loaded at the centroid of each individual member where
the x and y displacement at the top edges were restrained. To avoid unexpected local deform-
ation at the top edges, the nodes at the top side edge of the lipped-C channel were connected
to the centroidal axis using a rigid rink for each member. The model was comprised of shell
elements and the screw connections were modeled by rigid links. The models have assumed
the use of steel materials with Young’s modulus of E = 205,000 N/mm2 and Poisson’s ratio of
υ = 0.3. The analysis variables are listed in Table 1.
Figure 4 shows examples of the buckling mode obtained by eigenvalue analyses. The flex-
ural buckling mode (Figure 4(a)) was obtained mostly, but the local buckling mode where the
plate elements deformed inward and outward alternately (Figure 4(b)) was observed when
L = 1800 and bf was 50 or 60 mm. The objective of this analysis was to evaluate the flexural
buckling strength so we used data only from flexural buckling mode.
From the numerical analysis results, we evaluated k/E using Eq. (8). The relationship between
k/E and t/l0 is shown in Figure 5 where the numerical analysis results were plotted. Although fluc-
tuations were observed, the k/E value generally increased with increasing t/l. From the numerical

Table 1. List of analysis variables.


Variables Values (mm)

t 1.6, 3.0
L 1,800, 3,600
bf 20, 30, 40, 50, 60
l 100, 300, 900

Figure 3. Outline of the FE model.

609
Figure 4. Buckling modes obtained by eigenvalue analyses.

Figure 5. k/E versus t/l0. Figure 6. Evaluation results.

analysis results, we approximated the minimum value of k/E using the following equation
(broken line in Figure 5). It should be noted that this approximation provided conservative
results.

K=E ¼ 0:22t=l ð17Þ

Using Eq. (17) we evaluated the flexural buckling strength on the built-up members and
Figure 6 shows the evaluation results. The vertical axis is Ncr_FEA/Ncre_n (ratio of the elastic
buckling strength Ncr_FEA obtained by eigenvalue analysis to the evaluated strength Ncre_n
obtained from the design equation)
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ffi and the horizontal axis is the slenderness ratio λ (where λ
is equal to 2π2 E=Ncr FEA ). The white plots are the results obtained by the proposed method
and the grey plots were obtained from the evaluation of flexural buckling strength by AISI
S100 (NSA. 2016). The proposed method agreed well with the eigenvalue analysis results and
the coefficient of variation (C.V. in Figure 6) obtained using the proposed method was smaller
than that obtained using the AISI method.

2.4 Elastic flexural buckling strength of the four different section types of built-up members
We investigated the flexural buckling strength of built-up members, as shown in Figure 7. Type-
A indicates sections comprised of the two lipped C channels and one rectangular section. This
model assumes that the lipped channels do not connect with each other and that their web plate
is connected via another structural steel component. Type-B describes sections comprised of
two lipped C channels and two channels. This model assumes that the flange plate is connected
via another structural component. Type-C is comprised of two lipped-C channels where the
flange width at the upper and lower sides differs. It assumes the built-up members are connected
to each other through the flange plates. All analysis models adopted identical boundary condi-
tions as described showed above (Figure 3), where the axial loads were distributed depending on
the ratio of each individual cross-sectional area. The analysis variables include the member
length L, pitch of the screw connections l0, the width of the flange plate, and thickness of the
plate elements. The complete list of variables is provided in Table 2.

610
Figure 7. Section examples.

Table 2. List of analysis variables.


Section L(mm) w(mm) l0(mm)

Type-A 1800, 3600 60, 80 100, 300, 900 Figure 8. Ratio of the buckling strength obtained
Type-B 1800,3600 40, 80 100, 300, 900 by FEA to that obtained by the proposed
Type-C 1800,3600 20, 40 100, 300, 900 equations.

For Types-A and B, we assumed that the flexural buckling strength could be evaluated by
linear summation of the built-up lipped-C section and the other individual steel members. For
these two sections, the shear force between the lipped-C members was translated through the
two screws connections. Therefore, we used the series spring model and k/E equal to 0.11t/l.
Figure 8 shows the evaluation results. The vertical axis is the ratio of Ncr_FEA/Ncre_n (ratio of
the eigenvalue analysis results from Ncr_FEA to the buckling strengthpobtained by the ffi proposed
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
method Ncre_n) and the horizontal axis is the slenderness ratio λ (¼ 2π2 E=Ncr FEA ). The pro-
posed equation agreed well with the analysis results, showing an evaluation error of ˂10%.
The applicable range of the proposed design formula will be determined in future studies.

3 MAXIMUM STRENGTH OF THE BUILT-UP MEMBERS UNDER


COMPRESSION

3.1 Outline of the analysis model


The finite element analyses of the axially compressed built-up members were conducted and their
maximum strength was evaluated using the proposed design equation. Figure 9 shows an outline
of the FE analysis model. Half of the built-up members were modeled herein. The model was
composed of shell elements and the screw connections were modeled using a rigid link. The nodes
at the edges of the top side were connected to the centroidal axis of a built-up member by rigid
links. The boundary condition is shown in Figure 9(a). The nodes at the bottom side provided the
symmetric boundary condition (where the displacement in the z direction and rotation around
the x and y axes were restrained) and the enforced displacement was provided on the centroid of
the built-up member at the top side. The section shapes employed in this analysis are shown in
Figure 9(b). A total of 4 section shapes were used in the analyses. The thickness of the plate elem-
ents was 3.0 mm. The analysis variables are listed in Table 3. All models assumed that steel mater-
ial was used. The stress and strain relationships used in the models are shown in Figure 9(c). The
analysis models have a geometrical imperfection that was given based on the flexural buckling
mode determined by eigenvalue analysis. We set the maximum imperfection deformation as
1.8 mm (0.1% of the member length). The residual stress and the yield stress increase at the
corner portion by cold forming were ignored since we assumed that they offset one another.

3.2 Evaluation of maximum strength of the built-up members under compression


The maximum strength of the built-up members obtained from the FEA and the nominal
strength obtained from AISI S100 (NSA. 2016) were compared. The equations used were as
follows. Nne is the nominal axial strength, σy is the yield stress of the steel material, Ag is the

611
Table 3. List of incremental analyses.
Section Type L(mm) Width (mm) Screws pitch (mm)

buck-to-buck Lipped C 1800 40, 80 100, 300, 500, 900


Type-A 1800 60 100, 300, 500, 900
Type-B 1800 40 100, 300, 500, 900
Type-C 1800 40 100, 300, 500, 900

Figure 9. Outline of elasto-plastic analysis model.

Figure 10. Comparison of the maximum strength of a built-up member under compression.

gross section area of the built-up member, Ncre_b is the flexuralp


buckling strength obtained
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
from the proposed method, and λc is the slenderness ratio (λc ¼ Ag σy =Pcre b ). It should be
noted that, considering the point where all members have a slenderness factor of local buck-
ling of ˃0.776, we ignored the effect of the local buckling.

Pne ¼ 0:658λc Ag σy
2
For λc  1:5 ð18aÞ
!
0:877
For λc 41:5 Pne ¼ Ag σy ð18bÞ
λ2c

Figure 10 shows the relationship between Pmax/Agσy (ratio of the maximum strength Pmax to yield
strength of the built-up member Agσy) and λc. The plots represent the finite element analysis

612
results and the solid line represents the evaluation results obtained from Eqs. (18a) and (18b). As
shown in Figure 10, the plots and solid lines agreed well. Furthermore, the solid line provided
a slightly conservative evaluation result. As mentioned above, the proposed method approxi-
mated the minimum shear stiffness value. Thus, the maximum strength obtained by finite element
analyses was conservatively evaluated.

4 CONCLUSIONS

The flexural buckling strength of built-up members was investigated and design equations
were proposed based on the two layers model. We conducted elastic and elasto-plastic FE
analyses and compared the numerical analysis results with those obtained using the proposed
equation. The conclusions can be summarized as follows:
– The design equations regarding the flexural buckling strength on the built-up members
were proposed by modeling the screws as elastic shear springs. The proposed equations pro-
vided good evaluation results. Through eigenvalue analyses, the proposed method was
adequate for elastic flexural buckling strength evaluation of the buck-to-buck built-up
lipped-C channels and for investigation of other types of built-up members.
– The ratio of the screw connection shear stiffness to Young’s modulus could be used to
evaluate the lower limit as a constant multiple of t/l0 (ratio of the plate thickness t to
screw pitch l0). We conducted numerical analyses and proposed design equations
describing the relationship between k/E and t/l0. Although the developed equations
were approximations of the eigenvalue analysis of the buck-to-buck built-up members,
the proposed equation can be used to evaluate flexural buckling strength on other
types of built-up members.
– Incremental elasto-plastic analyses of axially compressed built-up members were conducted
based on the finite element method and their maximum strength was determined using the
proposed equations. The numerical analysis results mostly agreed well with the evaluation
results.

REFERENCES

AISI. 2016. North America Specification for the Design of Cold Formed Steel Structural Members 2016
edition. Washington DC: American Iron and Steel Institute.
AIJ. 2002. Recommendations for the Design and Fabrication of Light Weight Steel Structures. Tokyo:
Architectural Institute of Japan.
Fratamico, D.C. Torabian, S. Zhao, X. Rasmussen, K.J. Schafer, B.W. 2018. Experiments on the global
buckling and collapse of built-up cold-formed steel columns, Journal of Constructional Steel ResearchI.
Vol. 144: pp.65–80
Sato, A. Mitsui, K. Okada, H. Ono, T. 2015. Flexural elastic buckling strength of batten type light gauge
built-up compression member Part 1, Journal of Structural and Construction Engineering. Vol.80:
no.714 (in Japanese)
Kawai, Y. Ono, T. Sato, A. Tohnai, S. & Kondo, M. 2016. Allowable Design Formula for Steel Sheet
Shear Walls with Burring Holes. Proceedings of CIMS2016
Newmark, N.M. 1951. Test and analysis of composite beams with incomplete interaction. Proceedings of
the Society for Experimental Stress Analysis. vol.9: no.1
Timoshenko S. & Gere J. 1961. Theory of elastic stability, NewYork: McGraw-Hill

613
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Coupled buckling of steel LC-beams under bending

Z. Kołakowski, T. Kubiak & M. Kamocka


Department of Strength of Materials, Lodz University of Technology, Lodz, Poland

ABSTRACT: The interactive buckling and load carrying capacity of thin-walled steel lip
channel section beams subjected to bending loading is discussed in this paper. The beam was
subjected to pure bending in the web plane. A special attention on the influence of primary
and secondary global-distorsional modes was paid on. A method of the modal solutions to the
interactive buckling problem within Koiter’s asymptotic theory, using the semi-analytical
method (SAM) and the transition matrix method, was applied. Results obtained from this
approach were validated by FEM outcomes.

1 INTRODUCTION

Thin-walled cold-formed steel (CFS) C-section and LC-section (i.e., lip channel) beams are basic
structural elements in the construction industry that are primarily subject to bending. A load
carrying capacity to resist loads in thin-walled beams is limited first of all due to stability.
Only the latest and most important publications related to this subject scope are presented
here. The numerical methods often applied in nonlinear analysis of stability and load carrying
capacity are as follows: Finite Strip Method (FSM), Finite Element Method (FEM), the con-
strained Finite Element Method (cFEM) (Adany et al. 2018, Adany 2019), Generalised Beam
Theory (GBT) (Martins et al. 2018), Direct Strength Method (DSM) (Martins et al. 2017) and
a semi-analytic method based on Koiter’s (Garcea et al. 2017). In papers (Niu et al. 2014 a, b)
attention was drawn to the distortional-global interaction buckling during in wide experimen-
tal and numerical investigations of steel C-beams. The development in the theory of inter-
active buckling of thin-walled structures is discussed in (Hancock 2018).
An influence of the distortional-lateral buckling mode on the interactive buckling of thin-
walled short channels with imperfections subjected to the major-axis bending moment was
analysed in (Kolakowski et al. 2016). In (Kolakowski et al. 2018) an effect of various lengths
of C-beams subjected to the bending moment, with linearly variable stresses in the web plane,
on the interactive buckling and load carrying capacity was dealt with. Attention was paid in
particular to an effect of the secondary global buckling mode on an interaction between
modes, including distortional buckling modes. In papers (Kolakowski et al. 2016, 2018) the
results were obtained using semi-analytical method (SAM).
The aim of the presented work is numerical verification using FEM of the phenomena which
were observed in (Kolakowski et al. 2016, 2018). In those works it was found that for LC-beams
the highest influence of secondary global modes on interactive buckling and load carrying cap-
acity is for medium-long beams. In the presented paper the only the LC-beam analysis of a given
length for this range was performed. The above considerations decided that the results obtained
from the semi-analytical method (SAM) (Kolakowski et al. 2018) were compared with FEM.

2 PROBLEM FORMULATION

Prismatic thin-walled steel LC-beams built of plates connected along longitudinal edges and
under the uniform major-axis bending moment were considered. The beams were simply

614
supported at their ends (Kolakowski et al. 2016, 2018). It was assumed that the steel material
the structure was made of obeyed Hooke’s law. The problem was solved by two methods
SAM and FEM.

2.1 Semi-Analitycal Method(SAM)


In semi-analitycal method (SAM) a plate model (i.e., 2D) of thin-walled structures was
applied in order to account for all modes of global, local and coupled buckling. For each plate
component, precise geometrical relationships (i.e., full Green’s strain tensor) were assumed in
order to consider both out-of-plane and in-plane bending of the i-th plate (Kolakowski et al.
2016, 2018). The shear lag phenomenon, as well as an effect of cross-sectional distortions are
included.
The employed semi-analytical method (SAM) (Kolakowski et al. 2018) is based on Koiter’s
asymptotic perturbation theory with Byskov-Hutchinson’s formulation. The displacement
fields U and the sectional force fields N were expanded into power series (Koiter’s type expan-
sion for the buckling problem) with respect to the dimensionless amplitude of the r-th mode
 the total
deflection ζ r . For thin-walled structures with the linear geometric imperfections U,
potential energy has the form (Kolakowski et al. 2018):

 
1 2 1X J
M 1XJ X
J X
J
1XJ

X
J
M
 ¼  M a0 þ 
ar ζ r 1 
2
þ 
apqr ζ p ζ q ζ r þ brrrr ζ 3r  ar ζ r ζ r

2 2 r¼1 Mr 3 p q r 4 r r
Mr

ð1Þ

where: M is a magnitude of the applied bending moment, Mr , ζ r , ζ r – the buckling moment of


the r-th buckling mode, the dimensionless amplitude of the r-th buckling mode and the dimen-
sionless amplitude of the initial deflection corresponding to the r-th buckling mode, respect-
ively, and r = 1,2,..J (where J are all the relevant buckling modes that are believed to be
important in the structural response). In the semi-analytical method (SAM), one postulates to
determine approximated values of the second order coefficients (1)  brrrr on the basis of the
linear buckling problem. This approach allows the values of the first order coefficients (1)  apqr
to be precisely determined, according to the applied nonlinear Byskov and Hutchinson theory
(Kolakowski et al. 2016, 2018). On the basis of equation (1) one can determine the postbuck-
ling equilibrium paths for coupled buckling.

2.2 FEM model


Numerical simulation on the basis of finite element method (FEM) was performed in ANSYS
18.2® software (User’s Guide ANSYS® 18.2, Ansys Inc.). Numerical model was created using
structural element SHELL 181 - the four node element with six degrees of freedom at each
node. Boundary conditions were applied to the model in this way to ensure pure bending: con-
centrated forces acting on the nodes in both ends of cross-sections were applied in this way to
correspond to the bending stress distribution (Figure 1). In the nodes representing the support
of the beam (z = 0 and z = L), the displacement in the transverse direction (UX = 0, UY = 0)
was removed.
Bending moment Mb was calculated as the sum of forces acting on the beam and the dis-
tance from the neutral axis and angle of rotation α was determined based on the displacement
of the flange in the place of beam support. The analysis of the displacement reveal the rotation
of beam around two axis during the loading: the rotation around X axis and Y axis. Thus the
angle of rotation around X axis denoted as α was determined based on the displacement of
node localized on the neutral axis. Angle of rotation α was determined from the following
formula:

615
Figure 1. Discretized FEM model with applied boundary conditions.

α ¼ arctg Uy z n
ð2Þ
max

where Uz n - displacement of the point localized on the neutral axis in y direction;


ymax - maximum distance from neutral axis to the outer layer.

3 ANALYSIS OF THE CALCULATION RESULTS

Thin-walled lip channel section steel beam was taken into consideration. The beam cross section
is presented in Figure 2 with the following dimensions: b1 = 80 mm, b2 = 40 mm, b3 = 10 mm,
t = 1.0 mm. Mechanical properties of the beam were assumed: Young modulus E = 210 GPa,
Poisson ratio ν = 0.3. In the prebuckling state beams are subjected to linearly variable stresses
caused by a bending moment in the web plane (i.e., the upper flange is subject to tension,
whereas the lower one is compressed).

Figure 2. Cross section of the considered lip-channel section.

616
Firstly, linear analysis was performed to determine the buckling stresses and corresponding
to them buckling modes. Figure 3 presents the relationship between maximum critical stresses
σb and the half-wavelength Lb for a broad range of length variation. The lower curve (denoted
as curve I mode) corresponds to the lowest values of buckling loads, often referred to as pri-
mary buckling loads. The upper curve (marked as curve II Mode) refers to higher critical
values, which can be called secondary buckling loads. By dotted lines results obtained by
Finite Strip Method from CUFSM program (www.ce.jhu.edu/bschafer/cufsm) are depicted
while by solid one those obtained from SAM. Very good compliance of SAM and CUFSM
results can be observed.
According to the conclusions from (Kolakowski et al. 2018), further analysis was limited to
medium-long LC-beams with length L = 340 mm. In further analysis for SAM, two- and
three- mode approaches were adopted.
Three buckling modes were desired:
• m = 1 – primary global distortional-lateral buckling mode (m corresponds to the magni-
tude of halfwaves in the longitudinal direction),
• m > 1 – local buckling mode,
• m = 1s – secondary global distortional-lateral buckling mode.
Linear buckling analysis was performed to determine the buckling moment and buckling
stress for each of three considered buckling modes. In Table 1 results obtained from SAM
were confronted with those obtained from FEM and FSM. The following index markings
have been adopted: 1 - the primary global mode, 2 - the lowest local buckling mode, 3 - the
secondary global mode. Differences between the results are up to 6%. Table 2 shows the

Figure 3. Buckling stress as a function of the half-wavelength for considered LC-beam.

Table 1. The buckling stresses σr (r = 1,2,3) with the corresponding number of halfwaves m along the
longitudinal direction of the LC-beam.
σ1 σ2 σ3

MPa MPa MPa

SAM 398 (1) 533 (8) 4098 (1)


FEM 374 (1) 542 (8) 4194 (1)
CUFSM 410 (1) 528 (8) 4198 (1)

617
Table 2. Comparison of buckling stresses obtained from SAM, FEM and FSM.
Method SAM FEM FSM

Primary global modes

Secondary global modes

Local modes

buckling modes obtained for the three methods of solution. Very good character compatibility
was obtained.
In the next step, nonlinear analysis was conducted with implementing the small initial geomet-
rical imperfection corresponding to the buckling modes. For local buckling mode w0 = 0.1t and
for global distortional w0 = t were assumed. The influence of each buckling modes separately
and their interaction was investigated. The sign of imperfection was chosen to obtain the lowest
magnitude of buckling moment.
Detailed numerical computations of the influence of geometric imperfection and its inter-
action were performed. The M/Mcr equilibrium paths in the function of α/αcr for

618
Figure 4. The influence of initial imperfection on postbuckling behaviour.

imperfection ζ1*= │1.0│, ζ2*= │0.1│, ζ3*= │1.0│were determined, where Mcr and σcr
were obtained for the primary global buckling mode. The influence of each of the considered
buckling mode (m = 1; m = 8; m = 1s) was investigated. The worst case of sign (for which
the lowest buckling moment was obtained) was chosen.
The influence of each of the considered buckling mode (m = 1; m = 8; m = 1s) was investi-
gated. The worst case of sign (for which the lowest buckling moment was obtained) was chosen.
In Figure 4 the influence of each geometrical imperfection for FEM are presented. In the charts
0 refers to the model without initial geometric imperfection, and next m = 1 to initial imperfec-
tion ζ1* corresponded to first global mode, m = 8 for ζ2* according first local buckling mode
and m = 1s for ζ3* according secondary global buckling mode. It could be observed that initial
imperfection according to the local buckling mode has negligible effect on postbuckling behav-
iour. Little higher effect is observed in case of applying imperfection corresponded to the second-
ary distortional-lateral buckling mode. The highest influence on the buckling and postbuckling

Figure 5. Equilibrium path for 2 modes (m = 1; m=8).

619
Figure 6. Equilibrium path for 2 modes (m = 1; m =1s).

response has first distortional-global buckling mode. For global imperfections the load carrying
capacity is achieved with a lower value for the primary global mode.
In the second part the interaction of 3 modes (m = 1; m = 8; m = 1s), 2 modes for two cases
(m = 1; m = 8) and (m = 1; m = 1s) using SAM was analyzed. This analysis was intended to
determine the buckling mode which determine load carrying capacity of the bending LC-beam.
Results of SAM were compared with FEM calculations. Figure 5 presents buckling and post-
buckling response of considered beam with implemented initial geometric imperfection corres-
pond to m = 1; m = 8. When postbuckling structure response is analyzed, the curves obtained
from two approaches differs significantly. Based on the results of FEM, non-dimensional load
carrying capacity M/Mcr = 0.87 was obtained while for SAM postbuckling equilibrium path
increases monotonically.
The effect of primary and secondary global modes on structure response in presented in
Figure 6. High agreement both in the range before buckling and after loss of stability is

Figure 7. Equilibrium path for 3 modes (m = 1;m=8; m=1s).

620
observed. Results of SAM reveal significant influence of this interaction on postbuckling
response.
The influence of interactive buckling for three mode approaches on structure response is
depicted in [5] and is presented in Figure 7. Significantly good agreement between SAM and
FEM is obtained. Slight difference in beam stiffness (the slope of straight line in prebuckling
range) was obtained. SAM results obtained with taking into account interactive buckling of
primary and secondary global modes (Figures 6 and 7) indicates on the significant effect of
distortional-lateral global mode on the load carrying capacity of bending LC-beam in com-
parison to the case with neglecting the secondary global one (Figure 5). Therefore, application
of SAM allowed to indicate which mode of buckling determines the LC-beam load carrying
capacity.

4 CONCLUSIONS

In this work, the influence of secondary global distortional-lateral buckling on load carrying
capacity of bending LC-beams with medium-long was presented. Results obtained from semi-
analytical method (SAM) were verified by FEM and additionally in case of critical loading by
finite strip elastic buckling analysis (CUFSM). A very good agreement of results was
obtained.

ACKNOWLEDGEMENT

The studies have been conducted as a part of the research projects financed by the National
Science Centre, Poland - decision numbers: UMO-2017/25/B/ST8/00007.

REFERENCES

Adany, S. 2018. Constrained shell Finite Element Method for thin-walled members, Part 1: constrains
for a single band of finite elements. Thin-Walled Structures 128: 43–55.
Adany, S., Visy, D. & Nagy, R. 2018. Constrained shell Finite Element Method, Part 2: application to
linear buckling of thin-walled members. Thin-Walled Structures 128: 56–70.
Garcea, G., Leonetti, L., Magisano, D., Goncalves, R. & Camotim, D. 2017. Deformation modes for the
post-critical analysis of thin-walled compressed members by a Koiter semi-analytic approach. Inter-
national Journal of Solids and Structures 110–111: 367–384.
Hancock, G.J. 2018. Coupled Instabilities in Metal Structures (CIMS) – What have we learned and are
we going? Thin-Walled Structures 128: 2–11.
Kolakowski, Z. & Urbaniak, M. 2016. Influence of the distortional-lateral buckling mode on the inter-
active buckling of short channels. Thin-Walled Structures 109: 296–303.
Kolakowski, Z. & Jankowski, J. 2018. Interactive buckling of steel C-beams with different lengths –
From short to long beams. Thin-Walled Structures 125: 203–210.
Martins, A.D., Landesmann, A., Camotim, D. & Dinis, P.B. 2017. Distortional failure of cold-formed
steel beams under uniform bending: Behaviour, strength and DSM design. Thin-Walled Structures,
118: 196–213.
Martins, A.D., Camotim, D., Goncalves, R. & Dinis P.B. 2018. GBT-based assessment of the mechan-
ics of distortional-global interaction in thin-walled lipped channel beams. Thin-Walled Structures
124: 32–47.
Niu, S., Rasmussen. K.J.R. & Fan F. 2014. Distortional-global interaction buckling of stainless steel
C-beams: Part I – Experimental investigation. Journal of Constructional Steel Research 96: 127–139.
Niu, S., Rasmussen K.J.R. & Fan F. 2014. Distortional-global interaction buckling of stainless steel
C-beams: Part II – Numerical study and design. Journal of Constructional Steel Research 96: 40–53.

621
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Crashworthiness performance of tubular energy absorbing


structures with triggers

M. Kotełko
Department of Strength of Materials, Lodz University of Technology, Poland

M. Ferdynus & K. Okoń


Department of Machine Construction & Mechatronics, Lublin University of Technology, Poland

ABSTRACT: The paper presents results of the parametric study into energy absorption cap-
ability of thin-walled tubular columns of prismatic cross-section with different types of triggers:
cylindrical indentations (redrawing dents) and spherical concave or/and convex indentations
(“bumps”) of different shape and location. Columns are subjected to axial impact compressive
load. Extensive parametric analysis is performed using Finite Element numerical code. Several
crashworthiness indicators are examined to asses energy absorption effectiveness of structures
under investigation, among them the peak crushing force, which decreases due to the trigger
and this reduction is mostly desirable from biomechanical reasons. The FE numerical models
and numerical results are validated by means of experimental impact tests performed on the
drop hammer rig experimental stand. Conclusions concerning optimal design of examined
structures from energy absorption effectiveness point of view are derived.

1 INTRODUCTION

In the early sixties of the 20th century automotive safety regulations stimulated the develop-
ment of the new concept of a crashworthy (safe) vehicle that had to fulfil integrity and impact
energy management requirements (AISI Automotive Appl. Com.-2004). A designer of any
impact attenuation device must meet two main, sometimes contrary, requirements: The initial
collapse load has to be not too high in order to avoid unacceptably high impact velocities of
the vehicle. On the other extreme, the main requirement is a possibly highest energy dissipa-
tion capacity, which may not be achieved if the collapse load of the impact device is too low.
The latter may result in dangerously high occupant “ridedown” decelerations. Thus, maximiz-
ing energy absorption and minimizing peak to mean force ratio by seeking for the optimal
design of these components are of great significance.
There are numerous types of metallic thin-walled energy absorbers (converting energyof
impact into an energy of plastic deformation) that are cited in the literature (Baroutaji et al.
2017). Namely, there are steel drums, thin tubes or multi-corner columns subject to compres-
sion, compressed frusta (truncated circular cones), simple struts under compression, sandwich
plates or beams (particularly honeycomb cells) and many others. Among all those design solu-
tions, mentioned above, thin-walled metal tubes are widely used as energy absorption systems
in automotive industry due to their high energy absorption capability, easy to fabricate, rela-
tively low price and sustainability at collapse.
In the case of tubular thin-walled members subjected to axial compression acting as energy
absorbers, it is essential to come up with a design that would promote a progressive buckling
mechanism to stimulate the highest energy absorption capacity. One of the possible design solu-
tions is the application of a trigger (notch or dent) to release the desired crushing mechanism.

622
A trigger may induce the desired crushing (collapse) mode, leading not only to higher energy
absorption and a higher mean-to-peak crushing force ratio, but also to lower peak crushing
force. Flaws or dents acting as triggers can be periodically or non-periodically situated.
Relatively few publications are devoted to tubular structures with dents, dimples or other
flaws. An overview of tubular sections with different kinds of imperfections and fillers subjected
to axial impact loads was published by Yuen & Nurick (2008), who analysed crashworthiness
performance of tubular structures with different types of imperfections (flaws), including pre-
buckle, parallel and dished indentations, cutouts, stiffeners, fillers, and wrapping.
An interesting solution has been proposed by Yang et al. (2017), who investigated the crush-
ing behavior of a thin-walled circular tube with internal gradient grooves. The authors fabri-
cated a stainless steel thin-walled tube with preset internal circumferential rectangular groove
defects using the SLM 3D printing method. They observed a double buckling-splitting crush-
ing mode. Empty and foam-filled circumferentially grooved thick-walled circular tubes under
axial low velocity impact were investigated theoretically and experimentally by Darvizeh et al.
(2017). The effect of discontinuity size on the energy absorption performance of square profiles
was reported by Estrada et al. (2017). Lancaster & Palmer (2016) studied the behavior of tubu-
lar columns with dents, but these flaws were treated as imperfections resulting from damage.
An attempt to use columns with dents as energy absorbers, i.e., a static analysis of axially
compressed square section tubes with dents in the corners, is presented by Ferdynus (2013).
The subject of the investigation was a square-section column with periodically (along its
height) situated dimples (dents produced by redrawing) in the corners. FE simulations of the
crushing behavior were performed. Recently, Yang et al. (2017) investigated circular tubes
with periodically situated (both along the circumference and the height) ellipsoidal dimples,
subjected to axial crushing force. FE simulations of the crushing behavior were performed
and experimentally validated by quasi-static tests on 3D printed brass tubes.

2 SUBJECT AND OBJECTIVES OF THE STUDY

The state of art review, presented in the introduction, shows a gap in research undertaken so
far into crushing behavior and crashworthiness of thin-walled columns of prismatic cross-
section with non-periodically situated dents.
Thus, the subject of the present investigation were thin-walled tubular aluminum columns of
prismatic square cross-section with different new types of triggers, situated non-periodically:
cylindrical indentations (redrawing dents) and spherical concave and convex indentations
(“bumps”) of different shape and location (Figure 1). Namely, columns with the following trig-
gers were under investigation (Figure 1):

Figure 1. Columns with different triggers.

623
– Columns with cylindrical indentations (redrawing dents) –column D (Figure 1a),
– Columns with convex spherical “bumps”– column B1 (Figure 1b,
– Columns with concave spherical “bumps” –column B2 (Figure 1c),
– Columns with both convex and concave ‘bumps” on opposite walls, respectively column B3
(Figure 1d).
All columns were subjected to axial impact compressive load.
The objective of the study was the investigation into an influence of a type of trigger and
several geometrical parameters of triggers, e.g. diameters and depth of redrawing and position
a trigger upon the energy absorption capacity of the absorber. The study is a continuation of
the previous one, focused on columns type D (Ferdynus et al. 2018).

3 METHODOLOGY

3.1 FE numerical models


Crushing behavior analysis of investigated columns was performed on the basis of FE simula-
tions. FE explicit analysis was carried out using ABAQUS 6.14 code. The column was situated
between two rigid plates (Figure 2). A FE model of the column was created using 4-node
deformable shell elements with reduced integration. The rigid plates were modelled using
4-node bilinear, quadrilateral rigid elements.
It was assumed the columns to be made from aluminum alloy EN AW6063- T6
(σY = 175 MPa, σult = 250 MPa, ν = 0.33). Since aluminum alloys do not display a significant
sensitivity to the strain rate, a bi-linear material model was applied, neglecting an influence of
the strain rate, but taking into account the strain hardening (E = 70000MPa, Et = 937,5MPa).
In order to connect the rigid plates, with the column (shell) tie links were applied. In the
rigid plates reference points were created, at which impact force (at the bottom point) and
acceleration, velocity, displacement (at the top point) were recorded. All degrees of freedom
were taken (switched off) at the bottom rigid plate, while at the top plate only one degree of
freedom, in axial vertical direction, was possible. In order to create a uniformly divided mesh,
the technique of partition was used. Global size of elements both in the rigid plates and in the
columns was 2 mm. Two step FE analysis was performed: buckling and, subsequently, explicit
dynamic analysis. In the second step the first buckling mode was applied as the geometrical
imperfection of the amplitude equal to 1/10 of wall thickness.
Dimensions of columns of all types were the same (square 40×40 [mm], wall thickness
t = 1,2 mm, length L = 180 mm). Columns D were subjected to impact load of the kinetic
energy Ek = 1,47 kJ, which corresponds to the mass m = 60 kg dropping with the initial
velocity V0 = 7m/s. Columns B were subjected to impact load of the kinetic energy Ek = 1,715
kJ, which corresponds to the mass m = 70 kg, dropping with the initial velocity V0 = 7m/s.

Figure 2. Exemplary FE model of column B2: a) geometrical model, b) FE model with generated mesh,
c) first buckling mode (I. step analysis), d) – failure mode (dynamic explicit analysis).

624
3.2 Experimental tests
Impact tests were performed on the drop hammer rig (Kotełko et al. 2018). Impact tests were
limited to columns type D only. Specimens were made from standard aluminum tubes of
square section 40x40 and wall thickness 1,2 mm. Redrawing dents were made using plastic
working (cold burnishing) on special device. Material parameters and impact test parameters
were exactly the same, as assumed in FE calculations for columns type D. Impact tests were
recorded using a high-speed camera, mounted in front of the drop hammer stand. It enabled
one to register the crushing behaviour of the column and the real-time shortening of the
object, measured by means of Aramis Digital Image Correlation system.

4 PARAMETRIC STUDY INTO ENERGY ABSORPTION EFFECTIVENESS

4.1 Crashworthiness indicators


The parametric study into optimal triggers shape, position and geometry was performed using
several crashworthiness indicators, commonly applicable in crashworthiness analysis (Jones –
2003). In the present study four following indicators were taken into account:
– an initial Peak Crushing Force (PCF) (mostly desirable from biomechanical reasons),
– Crash Load Efficiency CLE,defined as a Mean Crushing Force (MCF) to peak crushing
force PCF ratio:

MCF
CLE ¼ 100% ð1Þ
PCF
EAðdx Þ
MCF ¼ ð1aÞ
dx

where: EA(dz) is energy absorbed corresponding to the deformation dz,


- Stroke Efficiency SE (representing the deformation capacity of an absorber):

U
SE ¼ ð2aÞ
L0

Alternatively, the stroke efficiency is expressed as:

LU
STE ¼ ð2bÞ
L

where: L is the initial length of the column and U is the shortening.


- Total efficiency TE (assessing the whole performance of an energy absorber - Kotełko et al
(2018), Jones (2010).
TE ½%
¼ CLE x SE ð3Þ

4.2 Experimental results and results of parametric study


Parametric study was performed for all types of columns with triggers shown in Figure 1,
taking into account the following varying parameters:
– for columns type D (with dents - Figure 1a), the varying parameters were the dent position
and relative depth; the columns are denoted as DYY_X (YY stands for position measured
from the bottom and X stands for relative depth of the dent),

625
– for the columns type B1 (Figure 1b), the varying parameters were the redrawing diameter
and redrawing depth and the columns are denoted as B1/d/t (d stands for the diameter and
t for the redrawing depth,
– for columns type B2 (Figure 1c) the varying parameters were as above and are denoted as
B2/d/t
– for columns type B3 (Figure 1d) the varying parameters were also as above, and the col-
umns are denoted as B3/d/t.
Previous extensive parametric study of the columns type D (Ferdynus et al. 2018) indicated,
that the main factor influencing a crushing mode and, subsequently, energy absorption capabil-
ity, is a dent depth. The dent distance from the base is of less importance. Thus, the position of
“bumps” in columns B1, B2 and B3 was kept constant (33 mm above the bottom). Diameter of
redrawing bumps varied from 28 -36 [mm] and redrawing depth from 1,2 to 4,8 [mm].
Exemplary load-shortening diagram obtained from the impact test of specimen D33_30 is
shown in Figure 3, together with stages of the crushing patterns for selected time instants. Num-
bers in frames in Figure 3a correspond to the numbers shown in Figure 3b. Figure 4 presents the
comparison of numerical (FE) and experimental results for column D33_30. Numbers of crushing
patterns correspond to the numbers in Figure 3. After reaching the first initial peak (PCF), the
next peaks (3, 6,8) are caused by the development of subsequent folds (wrinkles). It is a typical
progressive buckling behavior (Jones – 2011). The agreement of FE and experimental results is
satisfactory. Also crushing patterns obtained from FE simulations and observed during the test
are in good agreement. It confirms the correctness of applied numerical FE models.
For all types of triggers and nearly all redrawing depths of triggers, crushing was initiated by
the local plastic mechanism situated exactly at the trigger, which appeared to be the optimal
crushing mode, from the energy absorption effectiveness point of view. For all types of triggers
a significant decrease of peak crushing force was observed (Ferdynus et al. 2018, Okoń 2018).
Exemplary diagrams of three crashworthiness indicators for columns with triggers type
D (dents) are shown in Figure 5 (SM stands for column without trigger – smooth). The bigger is
the dent depth, the higher are CLE and TE indicators, while stroke efficiency displays minimum
and then increases again. CLE indicator is higher than for SM columns, for dent relative depth
greater than 10%. However, TE indicator is higher than for SM column already for relative dent
depth greater than 5%.

Figure 3. Exemplary results of impact test of column D33_30: a) load-shortening diagram, b) -stages of
crushing process.

626
Figure 4. Comparison of experimental and numerical results for column D33_30: load – shortening dia-
grams and crushing patterns.

Figure 5. Crashworthiness indicators CLE, SE and TE for D-33 columns (with dents) versus relative
dent depth.

Figure 6. Crashorthiness indicators CLE, and STE for B columns (with concave and concave/convex
bumps) – Okoń (2018).

Exemplary diagrams of two crashworthiness indicators CLE and STE (given by 2b) for col-
umns with triggers type B (bumps) are shown in Figure 6. More effective appeared to be trig-
gers with both concave and convex bumps (higher CLE). The least effective appeared to be
columns with bumps of big redrawing diameters and small redrawing depth, regardless a type
of bump (concave, convex or both), e.g. B1/36/1/(Okoń 2018).
Assembly of three crashworthiness indicators for the most effective types of triggers is given
in Table 1. Among all examined triggers the most effective were those of type B3 9 with both

627
Table 1. Crashworthiness indicators for the most effective triggers.
Triggertype Column model CLE [%] SE[-] TE [%]

D (dent) D_33_40 45,32 0,723 32,76


B1 (bump) B1/36/4 42,81 0,724 30,99
B2(bump) B2/36/4 42,01 0,745 31,29
B3 (bump) B3/36/4 52,77% 0,734 38,73

convex and concave bumps). Triggers with only convex or only concave bumps were less
effective than triggers type D (dents).

5 CONCLUSIONS

For all types of examined triggers the predominant factor influencing the energy absorption
effectiveness is the relative dent depth. As mentioned above, the most effective are triggers
with both concave and convex bumps but triggers with only convex or only concave bumps
were less effective then those with redrawing cylindrical dents. It was also observed that for
only convex or only concave bumps the crashworthiness indicators were nearly the same.
In columns of all types of triggers crushing was initiated by the local plastic mechanism situ-
ated exactly at the trigger, except triggers with the smallest depth. In the latter case crash-
worthiness effectiveness was much less and the crushing behavior was significantly different.

REFERENCES

Baroutaji A., Sajjia M., Olabi A.G. 2017. On the crashworthiness performance of thin-walled energy
absorbers: recent advances and future developments, Thin-Walled Structures 118: 137–163.
Darvizeh A., Meshkinzar A., Alitavoli M., Rajabiehfard R. 2017. Low velocity impact of empty and
foam filled circumferentially grooved thick-walled circular tubes, Thin-Walled Structures 110: 97–105.
Estrada Q., Szwedowicz D., Majewski T., Martinez E., Rodriguez A. Effect of quadrilateral discontinuity
size on the energy absorption of structural steel profiles, Eksploatacja i Niezawodność 18.
Ferdynus M. 2013. An energy absorber in the form of a thin-walled column with square cross-section
and dimples, Maintenance and Reliability 15: 253–258.
Ferdynus M., Kotelko M., Kral J. 2018. Energy absorption capability numerical analysis of thin-walled
prismatic tubes with corner dents under axial impact, Maintenance and Reliability 20: 252–259.
Ferdynus M. et al. 2018. Crashworthiness performance of thin-walled Prismatic tubes with corner dents
under axial impact – numerical and experimental study. Eighth International Conference on Thin-
Walled Structures (ICTWS2018) Lisbon, Portugal, July 24–27, 2018.
Jones N. 2003. Structural impact.
Jones N. 2010. Energy-absorbing effectiveness factor, International Journal of Impact Engineering 37: 754–765.
Kotelko M., Ferdynus M., Jankowski J. 2018. Energy absorbing effectiveness– different approaches,
Acta Mechanica et Automatica 12: 54–59.
Lancaster E., Palmer S. Model testing of mechanically damaged pipes containing dents and gouges,
ASME Pressure Vessels & Piping Conference New York, vol. 235: 143–148.
Okoń K. 2018. Numerical analysis of energy absorbers as thin – walled square section tubes with round
ribbing. Master Thesis, Lublin University of Technology. Lublin.
Prasad P., Belwafa J.E. (eds.). 2004. Vehicle crashworthiness and occupant protection. AISI Automotive
Appl. Committee, Michigan.
Yang Z., Yu Y., Wei Y., Huang C. 2017. Crushing behavior of a thin-walled circular tube with internal
gradient grooves fabricated by slm 3d printing, Thin-Walled Structures 111: 1–8.
Yang K., Xu S., Zhou S., Shen J., Xie Y.M. 2017. Design of dimpled tubular structures for energy
absorption, Thin-Walled Structures 112: 31–40.
Yuen S.C., Nurick G. 2008. The energy-absorbing characteristics of tubular structures with geometric
and material modifications: an overview, Applied Mechanics Reviews 61.

628
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Failure plastic mechanisms in TWCFS columns under eccentric


compression

M. Kotełko
Department of Strength of Materials, Łódź University of Technology, Poland

V. Ungureanu & D. Dubina


Department of Steel Structures and Structural Mechanics, Politehnica University of Timisoara, Romania
Laboratory of Steel Structures, Romanian Academy - Timisoara Branch, CCTFA, Romania

ABSTRACT: The paper presents a database of local plastic mechanisms of failure in lipped
channel section columns subjected to eccentric compression about minor axis. Selected results
of both numerical and experimental validation of developed mechanisms models are pre-
sented. Comparative diagrams of equilibrium paths obtained experimentally and from FE
simulations together with post-failure curves obtained via yield line analysis, based on devel-
oped mechanisms models, are shown. Conclusions concerning the applicability of presented
mechanisms models for determination of load-carrying capacity of structures under investiga-
tion are derived. Further research perspectives are presented.

1 INTRODUCTION

The EN 1993-1-1:2005 code defines four classes of cross-sections of steel structural elements
as shown in Figure 1. The role of cross-section classification is to identify the extent to which
the resistance and rotation capacity of cross-sections is limited by its local buckling resistance.
Class 1 cross-sections are those which can form a plastic hinge with the rotation capacity
required from the plastic analysis without reduction of the resistance. Class 2 cross-sections
are those which can develop their plastic moment resistance but have limited rotation capacity
because of local buckling, while Class 3 cross-sections are those in which the stress in the
extreme compression fibre of the steel member assuming an elastic distribution of stresses can
reach the yield strength, but local buckling is liable to prevent the development of the plastic
moment resistance. Class 4 cross-sections are those in which local buckling will occur before
the attainment of yield stress in one or more parts of the cross-section. Consequently, for
Class 4 cross-sections, effective widths may be used to make the necessary allowances for
reductions in resistance because of local buckling.
Thin-walled cold-formed steel structures are usually made of thin-walled members of class 4
sections. Since these sections are prematurely prone to local or distortional buckling and they
do not have a real post-elastic capacity, a failure of such members is initialized by the local-
global interactive buckling of plastic-elastic type, not an elastic-elastic one. Thus, the failure at
the ultimate stage of those members, either in compression or bending, always occurs by form-
ing a local plastic mechanism. This fact suggests the possibility to use the local plastic mechan-
ism analysis to characterize the ultimate strength (load carrying capacity) of such members,
since a proper identification of the geometry of a local plastic mechanism is crucial for correct
evaluation of the post-buckling rigid-plastic curve (post-ultimate curve), which subsequently
results in a correct estimation of the load carrying capacity of the member, as well as of the
energy absorption at collapse (Kotełko 2004, 2008).
An appropriate upper-bound estimation of the load-carrying capacity of those members is
an intersection point of the rigid-plastic (post-ultimate) curve and imperfect elastic one (point

629
Figure 1. Structural behaviour of a TWCFS member: (a) cross-section behaviour classes, (b) equilib-
rium paths for members subjected to compression.

F in Figure 1(b)). This approach is particularly useful in the concept of Erosion of Critical
Buckling Load (ECBL) method (Ungureanu & Dubina 2004a, b). This method enables to
determine in a more realistic way the ultimate strength in local/distortional - global interactive
buckling. Based on the concept of Erosion of Critical Bifurcation Load (ECBL), Dubina
(2001) proposed an approach to evaluate the ultimate strength in local/distortional-global
interactive buckling. This approach enables to use of the Ayrton-Perry format of European
buckling curves (Ungureanu & Dubina 2002, 2004a) to calibrate appropriate buckling curves
for any kind of interactive local/distortional-global buckling.
In the case of in the plastic-elastic interactive buckling, the imperfection factor used in the
Ayrton-Perry formula takes the form:
pffiffiffiffiffiffiffi
ψ2 Qpl
α¼  pffiffiffiffiffiffiffi ð1Þ
1  ψ 1  0:2 Qpl

N
where: Qpl ¼ Aσpl;m
y
and Npl,m is the local plastic mechanism strength (point F in Figure 1b).
Thus, the main problem of this concept is to evaluate properly the plastic strength of a thin-
walled member, via the yield line analysis (YLA), based on the local plastic mechanism model.
Plastic mechanism analysis (YLA) is widely used in solving problems of the load capacity,
ductility and energy absorption of thin-walled structures (Murray 1986, Ungureanu et al.
2010, Kotelko et al. 2011). However, as mentioned above, the determination of the load-
carrying capacity of TWCFS members subjected to combined load, particularly eccentric
compression, is still an open question.
On the other extreme, development of adequate models of plastic mechanisms of failure in
TWCFS members of steel framing subjected to combined load is necessary to perform
a proper advance analysis based on performance criteria.
Stage of failure of a thin-walled structure, when plastic mechanisms are well developed, can
be analyzed based on the rigid-plastic theory and yield line analysis (Murray 1986). Significant
progress in this theory related to thin-walled members took place in 80’s of the XXth century
(Murray & Khoo 1981).

630
2 SUBJECT AND METHODOLOGY OF THE STUDY

The aim of the present study is to establish a database of plastic mechanisms of failure for
stub columns with lipped channel section subjected to eccentric compression about the minor
axis, based on previous works (Ungureanu et al. 2016, 2018). Positive and negative eccentrici-
ties along the symmetry axis are investigated, as shown in Figure 2. The range of investigated
eccentricities was from -100 mm to 100 mm. Dimensions of the typology of cross- sections
under investigation are given in Table 1. The length of columns was assumed to be a multiple
of buckling half-wave and was determined based on buckling analysis.
The yield line theory (Yield Line Analysis - YLA) applied to thin-walled steel structures
allows one to perform an analysis of structural behaviour in the vicinity of the ultimate load
and in the post-failure stage. The basic assumption is that the plastic mechanism is fully devel-
oped, and the plastic zones developed in the walls of thin-walled steel member are concen-
trated at yield lines, either stationary or travelling. At the level of yield lines, the material is
considered fully plastic. Initially, the strain hardening was neglected in the analysis. Later, the
modified yield line theory (Kotełko 2004) considers this phenomenon.
The plastic mechanism approach is based on two basic methods, namely the energy method
(work method) and the equilibrium strip method (Murray 1986). Using the energy method, the
Principle of Virtual Velocities is applied, having the following form:
ð
P  δ_ ¼ σij ε_ pij ðβ; χÞdV ð2Þ
V

where P is generalized load, δ is the global generalized displacement, δ_ is the rate of change of
the global generalized displacement, β is the vector of kinematical parameters of the plastic
mechanisms (kinematical admissible displacements), χ is a vector of geometrical parameters of
the plastic mechanisms, σij is the stress tensor and ε_ pij is the strain rate tensor.
In open-section columns subjected to eccentric compression the energy of plastic deformation is
a sum of bending strain and membrane strain energy and generally may be expressed as follows:
Xk
W ¼ Wb þ j¼1
Wmj ð3Þ

Figure 2. The subject of the investigation; dimensions of the lipped channel section.

Table 1. Typology dimensions.


No. a (mm) b (mm) c (mm) t (mm)

1 150 60 20 1; 2
2 150 47 16 1; 1.5; 2
3 250 100 25 1.5; 2; 2.5; 3

631
If a rigid-plastic material model is assumed, the bending strain energy is given as follows:

Xn σy  t2
W ¼ mp l  βi ; mp ¼
i¼1 i
ð4Þ
4

where σy is the initial yield stress.


If strain-hardening of the material is considered, the bending strain energy is expressed as
(Kotełko 2004):

Xn  2
p βi Et  t2
Wb ¼ li  mp βi þ H ; H p ¼ ð5Þ
i¼1 2 12n

where li is the length of the yield line and βi is the angle of relative rotation of two walls of the
plastic mechanism along with that line, mp is the fully plastic moment in the wall cross-section,
Et is a material tangent modulus and n is multiple of wall thickness.
The membrane strain energy Wmj is the energy absorbed in tension fields, either in the web
or in the flange (the first component in Eqn. (3)):
ð
 
Wmj ¼ N0 εp dAi ð6Þ
Ai

3 FAILURE MECHANISMS DATABASE

The collapse plastic mechanisms identified based on FE simulations by Ungureanu et al. (2018)
for lipped channel columns under eccentric compression about the minor axis are listed in Table 2.
For small positive eccentricities the true flange mechanism CF1, initiated by distortional buckling,
was observed (see Figure 3a). For larger positive eccentricities (e = 30 – 100 mm) the CF quasi-
mechanism (CFQM) was developed (Figure 3c). Figures 3b and 3d present the corresponding real
failure modes.
For small and medium negative eccentricities, a modified pitched-roof mechanism, named
pitched-roof(I) (PR I), has been identified, as shown in Figure 4b. It consists of local
pitched-roof mechanism in the web and two local mechanisms in the flanges. Thus, the total
energy of plastic deformation is a sum of bending strain energy and membrane strain energy
in tension fields (shadowed in Figure 4a). For very small negative eccentricities the mechan-
ism pitched roof in the web (PR - Figure 4a), neglecting tension fields in the flanges was
a good approximation of the real mechanism. For large negative eccentricities, the roof
mechanism (Figure 4c) was identified.

Table 2. Failure mechanisms for lipped channel


columns under eccentric compression.
Failure mechanism
Eccentricity e [mm] (theoretical model)

5 – 20 (positive) CF1
30 – 40 (positive) CF1/CFQM
50 – 100 (positive) CFQM
-5 – -60 (negative) PR/PR I
-60 – -100 (negative) PR I/roof

632
Figure 3. Failure mechanisms for positive eccentricities: (a) theoretical model CF1, (b) corresponding real
failure mode (e = + 20 mm), (c) theoretical model CFQM, (d) corresponding real failure mode (e=+60 mm).

Figure 4. Failure mechanisms for negative eccentricities: (a) theoretical model PR; (b) theoretical model
PR I; (c) theoretical model “roof”; (d) real failure mode (e = -15 mm); (e) real failure mode (e = -60 mm).

4 COMPARATIVE NUMERICAL AND EXPERIMENTAL ANALYSIS

In order to evaluate theoretical models of failure mechanisms, both numerical and experimen-
tal validation was performed. Numerical FE analysis (using ABAQUS code) was carried out
for the wide range of columns dimensions and applied eccentricities, as shown in Table 1.
Experiments were carried out on stub steel columns, situated in the special grip, as shown in
Figure 5. The short columns were installed in the bottom and top plates with grooves (indi-
cated by the arrow in Figure 5), each groove corresponding to certain eccentricity. The plates
were fastened to cradles, which enabled hinge support on both ends of the column. The grip
with the specimen was installed on the testing machine Instron. The applied load and shorten-
ing of the specimens were measured using an integrated measurement system of the testing
machine. In order to obtain a more detailed (field) information about deformations of speci-
mens’ walls, the ARAMIS Digital Image Correlation system (DICS) was used. Experiments
were performed on columns of dimensions a = 150 mm, b = 60 mm, c = 20 mm, t = 1 mm.

633
Figure 5. Experimental stand: (a) general view; (b) exemplary deflection map in the flange (ARAMIS,
e = +20 mm).

Figure 6 shows the ultimate loads versus the eccentricity (numerical and experimental
results). Ultimate loads decrease significantly with an increase of positive eccentricity and an
absolute value of negative eccentricity. However, for very small positive eccentricities an
increase is observed (with the maximum for e = +5 mm).
In the case of positive eccentricities, two types of plastic mechanisms were recorded experi-
mentally. For relatively small eccentricities the theoretical model CF1 was confirmed mainly
(Figure 3a). For larger positive eccentricities CFQM quasi-mechanism was confirmed, as
shown in Figure 3b. These two mechanisms were developed in the specimens, in which local
buckling and initial yielding were observed in the flange.
Plastic mechanisms recorded for negative eccentricities experimentally were different from
those, derived based on FE simulations (Ungureanu et al. 2018). However, it should be

Figure 6. Ultimate loads versus eccentricity (numerical and experimental results for column dimensions
a = 150 mm, b = 60 mm, c = 20 mm, t = 1 mm.

634
emphasized, that FE calculations presented by Ungureanu et al. (2018) concerned columns
with different a/t and a/b ratios.
The mechanisms recorded experimentally and confirmed by FE simulations were
a combination of pitched roof (PR) or roof mechanism in the web and tension fields in the
flanges (PRI) or a local mechanism similar to the local mechanism in flanges of a box-section
beam subject to pure bending (Kecman 1983).
Plastic mechanisms recorded in FE simulations for a wider range of a/t and a/b ratios were
(in some cases) similar to PR I mechanisms presented in Figure 4b or roof mechanism pre-
sented in Figure 4c.
Figure 7 shows the comparative load - shortening diagrams of numerical and experi-
mental results. In Figure 7a the comparison of experimental and numerical FE structural
curves for two small eccentricities (e = +10, 15 [mm]) with theoretical YLA post-failure
curve, obtained for the mechanism CF1 is presented. Continuous curve CF1 stands for
rigid-perfectly plastic material characteristics, while the dotted curve CF1 for rigid-plastic
characteristics with linear strain-hardening. Figure 7b shows the theoretical CFQM

Figure 7. Load - shortening diagram for the positive eccentricities; comparison of FE, experimental
results and YLA analysis (wall thickness t = 1 mm, a = 150 mm, b = 60 mm, c = 20 mm, fy = 199 MPa):
(a) small positive eccentricities, (b) medium and large positive eccentricities.

635
Figure 8. Load - shortening diagram for the negative eccentricities; comparison of FE results and YLA
analysis (wall thickness t = 2 mm, a = 150 mm, b = 60 mm, c = 20 mm, fy = 355 MPa): (a) comparison
with PR I mechanism post-failure curve, (b) comparison with roof mechanism post-failure curve.

post – failure curve compared with experimental and numerical FE curves for eccentrici-
ties e = + 40 mm and e = +60 mm.
Figure 8 shows load-shortening diagrams for negative eccentricities obtained from FE simu-
lations compared with post-failure curves obtained for PR, PR I and roof mechanism (numer-
ical results only).

5 FINAL REMARKS

The obtained results confirm the possibility to use the local plastic mechanism analysis to
characterize the ultimate strength of short columns under eccentric compression. The database
of those mechanisms for different eccentricities will be useful to create a map of plastic mech-
anisms for a wide range of eccentricities (from positive to negative values). However, for

636
a certain eccentricity, an appropriate model of plastic mechanism should be calibrated, based
on FE simulations and, simultaneously, experimental results.
The developed models of local plastic mechanisms for columns subjected to compressive
load with positive eccentricities are in good agreement with both numerical and experimental
results. For columns subjected to eccentric loads with negative eccentricities (particularly
large), models of plastic mechanisms should be improved.
Thus, further research into the derivation of appropriate theoretical models of local plastic
mechanism for a wider range of columns’ dimensions (particularly for negative eccentricities)
as well as their experimental verification will be continued.

ACKNOWLEDGEMENT

The research was supported by ROMANIAN - POLISH JOINT RESEARCH PROJECT


“Yield line theory for load-capacity estimation of thin-walled cold-formed steel members
under combined loading” (2019-2021) under the agreement on scientific cooperation between
the Romanian Academy and the Polish Academy of Sciences and by a grant of the Romanian
Ministry of Research and Innovation, project number 10PFE/16.10.2018, “PERFORM-
TECH-UPT - The increasing of the institutional performance of the Politehnica University of
Timișoara by strengthening the research, development and technological transfer capacity” in
the field of “Energy, Environment and Climate Change”, within Program 1 - Development of
the national system of Research and Development, Subprogram 1.2 - Institutional Perform-
ance - Institutional Development Projects - Excellence Funding Projects in RDI, PNCDI III”.

REFERENCES

Dubina, D. 2001. The ECBL approach for interactive buckling of thin-walled steel members. Steel &
Composite Structures 1(1): 75–96.
Dubina, D. & Ungureanu, V. 2002. Plastic strength of thin-walled members. Proc. of the Sixteenth Int.
Specialty Conference on Cold-Formed Steel Structures, Orlando, Florida, 17-18 October 2002.
EN 1993-1-1:2005. Eurocode 3: Design of steel structures - Part 1-1: General rules and rules for
buildings (including EN 1993-1-1:2005/AC, 2009). European Committee for Standardization,
Brussels, Belgium.
Kecman, D. 1983. Bending collapse of rectangular and square section tubes. International Journal of
Mechanical Sciences 25(9-10): 623–636.
Kotełko, M. 2004. Load-capacity estimation and collapse analysis of thin-walled beams and columns –
Recent advances. Thin-Walled Structures 42(2): 153–175.
Kotełko, M. 2010. Load-carrying capacity and mechanisms of failure of thin-walled structures. WNT
Warszawa (in Polish).
Kotełko, M., Ungureanu, V., Dubina, D. & Macdonald, M. 2011. Plastic strength of thin-walled plated
members - Alternative solutions review. Thin-Walled Structures 49(5): 636–644.
Murray, N.W. & Khoo, P.S. 1981. Some basic plastic mechanisms in thin-walled steel structures. Int.
J. Mech. Sci. 23(12): 703–713.
Murray, N.W. 1986. Introduction to the theory of thin-walled structures. Clarendon Press, Oxford.
Ungureanu, V. & Dubina, D. 2004a. Recent research advances on ECBL approach. Part I: Plastic-elastic
interactive buckling of cold-formed steel sections. Thin-Walled Structures 42(2): 177−194.
Ungureanu, V. & Dubina, D. (2004b). Post-elastic strength and ductility of cold-formed steel sections. Proc.
of the Fourth International Conference on Thin-Walled Structures, Loughborough, UK, 22–24 June 2004.
Ungureanu, V., Kotełko, M., Mania, R.J. & Dubina, D. 2010. Plastic mechanisms database for thin-walled
cold-formed steel members in compression and bending. Thin-Walled Structures 48(10-11): 818-826.
Ungureanu, V., Kotełko, M. & Grudziecki, J. 2016. Plastic mechanisms for thin-walled cold-formed steel
members in eccentric compression. Acta Mechanica et Automatica 10(1): 33–37.
Ungureanu, V., Kotełko, M., Karmazyn, A. & Dubina, D. 2018. Plastic mechanisms of thin-walled
cold-formed steel members in eccentric compression. Thin-Walled Structures 128: 184−192.

637
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Experimental study of beam-to-column connection with bolted joints

Y. Koyama, A. Sato & H. Idota


Nagoya Institute of Technology, Nagoya, Aichi, Japan

Y. Sato & S. Yagi


IIJIMA Structural Design Office, Japan

S. Takaki & M. Kamada


Sumikin System Buildings Co. LTD, Japan

ABSTRACT: It is well known that in-elastic behaviour in the panel zone of beam-to-
column connections can provide stable performance; it can be used to absorb the energy to
avoid the collapse of the structure. Although the contribution of plastic deformation at the
panel zone is limited up to 30% in Eurocode 8, there is no limit in the Japanese code when the
structural performance is verified. In this paper, full-scale testing of the beam-to-column con-
nections where the beam is connected to the column though T-stub by high-strength bolts are
tested. In the tested specimens, the shape similar to T-stub is used for the continuity plate, and
partial reinforcement at the panel is expected. From the testing, the maximum strength, and
the deformation capacity of the beam-to-column connection will be evaluated.

1 INTRODUCTION

In Japan, it is common to use a steel hollow structural section (HSS) for the columns in the
building structures. On the other hand, a wide-flange section is often used for the columns in
the industrial buildings. When the wide-flange section is used for the columns, the dissipative
zone might be assumed in the panel zone when the doubler plate is not installed. It is also well
known that in-elastic behaviour in the panel zone can provide stable performance; it can be
used to absorb the energy to avoid the collapse of the structure. Although the contribution of
the plastic deformation at the panel zone is limited up to 30% in Eurocode 8 (2009), there is
no limit in the Japanese code when the structural perfo`rmance is verified.
In the Kobe earthquake, fractures at the beam ends, where the Complete Joint Penetration
(CJP) is done, were observed. After the decades, numerous researches and the improvement in
the configurations and materials provided the solution to avoid the fractures at the joints. The
CJP components where the axial stress is subjected should be welded by the qualified welder, and
the Ultrasonic Test (UT) inspection is necessary by the qualified inspector to guaranty the quality
(AIJ 2018); therefore, time and cost will be the burden for the project budget. High-strength bolts
(HSB) are also commonly used at the joints. HSB is an industrial product, and the quality can be
managed easily. Moreover, installation of the HSB can be done by anybody and well-trained skill
is not required. HSB can be a potential to be used in the beam-to-column connections instead of
welding, and it might solve the problems that are considered in welded connection.
In this study, beam-to-column connections where the components are connected by HSB
are proposed and tested in full-scale. The behaviour zone is assumed in the panel zone, and
the capacity design principle is used to determine the other components (AIJ 2012). The wide-
flange beam is connected to the wide-flange column through T-stubs by HSB. Continuity
plates, the configuration is similar to T-stub, are also installed in the panel zone by HSB.
From the full-scale testing of the beam-to-column connections, the maximum strength, the
deformation capacity, and the failure mode at the Ultimate Limit State (ULS) are evaluated.

638
Moreover, the effect of bolted continuity plate to the structural performance of the beam-to-
column connections is also evaluated in detail.

2 EXPERIMENTAL PLAN OF SUBASSEMBLAGE FRAME

2.1 Test specimens


The configuration of the test specimens is shown in Figure 1. The beam is connected to the column
though T-stub by HSBs. Figure 2 shows the continuity plates that are installed in the panel zone to
stiffen the column flange, and a smooth load pass from the beam to the column is obtained. The
continuity plates are installed to coincide the centreline of the T-stub and the rib-plate.
The dissipative zone is designed to be in panel zone (except specimen No.4); therefore, the
size of the T-stub is determined from the capacity design, and the thickness of the flange is
larger than the beam and column. For the comparison study purpose, specimen No.4 was
designed to have the dissipative zone in the joints at the beam flange and T-stub, and other
components were designed to fulfil the capacity design. Table 1 summarises the applied load
(see Figure 5) corresponding to the strength of each component. To compute the correspond-
ing load, rigid plastic analysis was used except the considering structural fuze component.
Geometrical configuration was considered in the calculation; following formulas were used to
compute the moment at the beam-to-column working point (Kuwahara & Inoue 1996).

Figure 1. Configuration of the test specimens.

639
Figure 2. Configuration of the continuity plates.

Table 1. Computed load corresponding to the mechanism.


Yield load cal Qy Collapse load cal Qp , cal Qu

column beam panel joint column beam panel joint


No. Name cal c Qy cal b Qy cal p Qy cal j Qy cal c Qp cal b Qp cal p Qp cal j Qu

1 HY-600-p 420 322 282 307 470 477 302 578


2 H588-p 432 340 298 301 483 490 322 462
3 H346-p 69 53 51 58 77 78 55 170
4 BH350-j 217 183 225 59 265 269 250 167

cal Qp : full plastic load, cal Qu : ultimate load

1
Mc  ¼  Mc ð1Þ
1db =2H
1
Mb  ¼  Mb ð2Þ
1 ðdc þ aÞ=2L

1
Mp  ¼  Mp ð3Þ
1db =2H  dc =2L
1
Mj  ¼  Mj ð4Þ
1dc =2L

MX: yeild moment and ultimate moment of X X = c (column), b (beam), p (panel), j (joint)
a: distance between outside surface of column and beam cross-section exceedingly concen-
trated stress
dc: width of panel zone
db: height of panel zone
L, H: refer to the fig.5 (L=H=3000 mm)

Table 2 summarizes the material properties of the steel that are used in the specimens. The
properties shown in Table 2 were obtained from the coupon tests per Japanese Industrial
Standard (JIS). The values shown in Table.1 are computed based on the measured thickness
and strength summarised in Table 2. Table 3 summarizes the HSBs that is used to connect the
components. Bolt grade “S10T” is the twist off high-strength bolt that is used for slip-critical
joint; specified pre-load is introduced during the assembly.

640
Table 2. Material properties from tensile tests.
Member Spec. No. t [mm] E [N/mm2] σy [N/mm2] σu [N/mm2] Y.R. [%] Elong [%]

1 flange 22.0 212000 244.5 407.6 60.0 34.7


web 12.2 214900 286.5 421.0 68.1 30.3
column & beam SS400 2 flange 19.2 213900 294.0 425.9 69.0 30.8
web 11.7 210400 330.2 448.8 73.5 26.4
3 flange 8.5 210200 306.3 441.4 69.4 27.9
web 5.6 206600 368.6 475.1 77.6 18.6
SM490A 4 flange
27.9 205600 342.0 491.9 69.5 21.9
web
T-stub SN490B all flange 34.9 212700 354.8 519.7 68.3 32.8
web 17.9 201700 393.3 538.9 73.0 26.4
continuity plate SN490B all - 12.1 210300 392.5 536.0 73.2 23.2

t: Measured plate thickness, E: Young’s modulus, σy: Yield stress, σu: Ultimate stress, Y.R.: yield ratio (= σy/σu×100),
Elong: Elongation

Table 3. The bolts used in test specimens


Joint Type Grade Size

beam—T-stub M20
T-stub — column — flange-plate Twist-off type high-strength bolt S10T M24
web-plate — column M20

2.2 Test program


Figure 3 shows the elevation of the test setup. The beam end was supported as a pin, and the
load was applied to the column end which is supported as a roller. To conduct the testing in
in-plate, two lateral supports were attached to the column and beam, respectively.

Figure 3. Elevation of the test setup.

641
Figure 4. Test setup.

Figure 4 is the photograph shows also the test setup before loading (with specimen No.1).
Testing was controlled by story drift ratio R of the subassemblage (See Figure 5(a)). Figure
5(b) shows the load sequence subjected to the specimen.

3 TEST RESULTS

3.1 Load vs. story drift ratio


Hysteresis behaviour of the specimens are shown in Figure 6. The vertical axis is the applied load
and the horizontal axis is the story drift ratio or the component of the story drift ratio. The circle
legend shown in the figures corresponds when the first slippage occurred in the bolted joint. More-
over, the triangle legends (open and solid) shown in the figures corresponds to the Maximum load
in the positive and negative loading, respectively. The symbol Rp represents the component of
panel deformation to the story drift; the symbol Rj represents the component of joint deformation

Figure 5. Loading program.

642
Figure 6. Load vs. story drift ratio.

643
to the story drift. As can be recognized from the figures, specimens where the panel zone is
designed to be the behaviour zone (i.e., Specimen No.1 to No.3) had a large plastic deformation in
the panel zone. The specimen where the behaviour zone is designed to be in the T-stub joint had a
large plastic deformation in the joint and significant slippage and hardening due to bearing were
observed.
Structural characteristics measured from the testing are summarized the Table 4. The elastic
stiffness in the positive and negative loading showed different values; negative loading showed
slightly higher value in all specimens. The elastic stiffnesses were obtained from the regression
analysis of the data in the elastic range. The maximum load measured in the negative loading
also showed larger value than the value measured in positive loading. The strength degrad-
ation started after the story drift ratio go beyond 3.0%; specimen No.4 where the joint is
designed to be the behaviour zone did not have the strength degradation.

3.2 Strain characteristics in panel zone


Figure 7 shows the principal strains computed from the tri-axis strain gauge attached in panel
zone. The strain level shown in the figure corresponds to the maximum load in the positive load-
ing. The arrow facing outside represents tension strain; the arrow facing inside represents com-
pression strain. Be aware that specimen No.4 have a different scale due to the small plastification
in the panel.
As shown in Figure 7, the direction of the principal strain can be defined as the region of
the panel zone that is indicated by the broken lines. The photos in Figure 8 shows the
deformed shape at the panel zone when the story drift ratio reached 10%. It can be confirmed
that the deformed shape shown in Figure 8 corresponds well with the direction of principal
stress shown in Figure 7. Specimen No.4 did not have the closeup photo of the panel zone.

Table 4. Test result (p: positive loading, n: negative loading).


No. Name pK [kN/rad] pQmax [kN] pRmax [rad] nK [kN/rad] nQmax [kN] nRmax [rad]

1 HY600-p 30600 323.6 0.0503 31600 439.7 0.0602


2 H588-p 28200 239.9 0.0447 30000 427.8 0.0472
3 H346-p 3200 71.4 0.0411 3800 72.7 0.0348
4 BH350-j 9000 138.9 0.0340 11100 140.9 0.0304

K: Stiffness of the partial flame, Qmax: Maximum load, Rmax: Story drift ratio at the maximum load

Figure 7. Strain characteristics in panel zone.

644
Figure 8. (a) Whole view at R=1/10 (No.1), (b-d) Deformation of panel zone at R=1/10 (No.1-3).

4 CONCLUSIONS

In this paper, the full-scale testing were conducted to clarify the maximum strength, the
deformation capacity, and the failure mode of the bolted beam-to-column connection. The
effect of bolted continuity plate to the structural performance of the beam-to-column connec-
tions is also evaluated in detail.
Following results were found:
1) The structural performance of the beam-to-column connection with bolted joints was
shown. The test results clearly showed that the specimens where the panel zone shall be the
dissipative zone can provide stable in-elastic behaviour, and strength degradation occurred
after the story drift goes beyond 3.0%.
2) Bolt slippage was not found in the bolt joints where it was adequately designed based on
the capacity design principal.
3) Panel zone of the bolted continuity plates can be defined between the area where the HSB
are used to connect the continuity plates to the column web.
4) It was confirmed that the proposed beam-to-column connection can be used as the dissipative
zone.

REFERENCES

Architectural Institute of Japan. Recommendation for Design of Connections in Steel Struc- tures, Japan,
2012.3
Architectural Institute of Japan. Technical Recommendation for Steel Construction for Buildings, Part1
Guide to Steel-rib Fabrications, Japan, 2018.1
European Committee for Standardization: Eurocode 8: Design of structures for earthquake resistance,
Part1: General rules, seismic actions and rules for buildings, EN 1998-1-1, 2009.7
Kuwahara S, Inoue K, “An approximate calculation for the plastic collapse load of the rigid frame con-
sidering joint panels”, Journal of Structural and Construction Engineering, Japan, 441-449, 1996.

645
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Material strength statistics and reliability aspects for the


reassessment of end-of-service-life steel bridges

R. Kroyer & A. Taras


Institute of Structural Engineering, Bundeswehr University Munich, Germany

ABSTRACT: The determination of the reliability of engineering structures within


a continuously ageing infrastructure is a topic of increasing international importance. This
paper focuses on probabilistic aspects in the reassessment of the structural capacity of existing
steel bridges. A consistent and comprehensive collection of statistical material strength data
for existing steel structures is presented. Beginning with puddle steel, the collected data of
former steel products covers the era since about 1900 until today. The subsequent discussion
refers to the implementation of a semi-probabilistic approach in compliance with Eurocode
1990 for determination of adequate partial safety factors. The paper concludes with a concept
to enable a more accurate reappraisal of specific in-service structures. In this context, an
extended probabilistic approach using Bayesian statistics is presented.

1 INTRODUCTION

Structural reliability in civil engineering design is verified by the adoption of an overall safety
concept. This approach allows to consider statistical uncertainties related to scattering of param-
eters as well as stochastic uncertainties due to simplifications in structural modeling. However,
structures are exposed inevitably to structural deterioration during their lifetime. The statistical
characteristics of the relevant load cases such as the load magnitude and the return period might
also evolve with time. This is especially valid for the statistical characteristics of the predominant
load case from traffic on bridge structures. Consequently, a periodical reassessment of bridge
structures is required. Focusing on the structural capacity, reliability aspects in context with load
capacity predictions of engineering structures within a continuously ageing infrastructure are
a topic of internationally increasing importance. Acknowledging this requirement, the proposed
paper is devoted to the reassessment of existing steel bridge structures, particularly with regards
to the ultimate limit state capacity of individual members and sections. It comprises a consistent
and comprehensive collection of statistical material strength data for existing steel structures
taken from the literature with the intention of comparative appraisal of material strength vari-
ability. Beginning with puddle steel, the collected data of former steel products covers the era
since about 1900 until today. Using this statistical database on material strength characteristics
of steel products available in Central Europe since 1900, a semi-probabilistic methodology in
conjunction with Monte-Carlo simulations for determination of partial safety factors is pre-
sented and implemented for illustration purposes. The proposed paper furthermore includes an
extended probabilistic approach for the exploitation of material strength data available from
material tests for a specific in-service structure. Using Bayesian statistics, this approach is sup-
posed to enable a more accurate reappraisal of the structural reliability.

2 HISTORICAL REVIEW ON STEEL PRODUCTION

Structural reassessment of existing structures according to state-of-the art design standards


requires usually a retrospective appraisal of former construction material properties. Focusing

646
on steel, the material properties were figured out to evolve typically with further development of
steelmaking processes. The distinction of different statistical populations of material strength
characteristics according to the manufacturing technique is thus reasonable and adopted in this
study. Hence, this section recovers briefly the historical development of steelmaking processes.
The development of the proportionate quantities by manufacturing process over time is illus-
trated in Figure 1 and 2. Existing processes were typically repressed by new ones. The historical
development is described more comprehensively in Kroyer & Taras (2018).
Focusing on products with industrial importance, the puddling process can be considered
as the first industrial procedure in steel production (Langenberg, 1995). In the time range
from 1785 until about the year 1860, a large part of the available steel products was provided
by the puddling process. However, the demand for steel increased tremendously in that time
and could not be satisfied by the puddling process. Attempts towards more effective proced-
ures resulted in various techniques for production of mild steel. The production technique for
mild steel which came up first was the Bessemer process (Kuhlmann, 2017). In contrast to the
puddling process, the outcome of the Bessemer process showed basically isotropic mechanical
properties. For the first time, this procedure satisfied the demands for a high-quality steel
product with reproducible characteristics suitable for mass production (Lüddecke, 2006).
Unlike the Bessemer process, the Thomas process didn’t require the import of appropriate
ore. This aspect allowed for rapid establishment of the more flexible Thomas process for mild
steel production, which was used successfully for many years (Kuhlmann, 2017). Industrial
development continued and for reasons of improved technological material properties, from
1910 on the Siemens-Martin process was the leading technique for mild steel production (Lüd-
decke, 2006). Besides high purity and the increased suitability for welding and cold forming,

Figure 1. Global steel production (1860-1990) by process according to (Kuhlmann, 2017).

Figure 2. German steel production (1950-2008) by process according to (Kuhlmann, 2017).

647
a decisive advantage of the Siemens-Martin process may be seen also in the composition of
the raw material. The process allows the use of scrap metal and benefited therefore from
ongoing increase of those amounts beginning around the year 1900. Nowadays the oxygen
steelmaking process and the electrical steel production are the predominant techniques in steel
production. With the oxygen steelmaking process, a significant improvement of mild steel
material properties was achieved. The process was developed by VÖEST in Linz, Austria. In
1950, it was applied for the first time for industrial purposes (Kuhlmann, 2017). Subsequently,
the process gained rapidly industrial importance. In 1979, only 9.9 % of the produced steel
was still provided by Siemens-Martin facilities in Germany. To date, there are no longer any
Siemens-Martin facilities in Germany. Considering statistical data from 2005 as up to date,
the oxygen steelmaking process provided 68,9 % of the produced steel in this year, whereas
31,1 % were produced as electrical steel (Pasternak et al. 2010).

3 STATISTICAL DATA ON MATERIAL STRENGTH

This section contains a consistent and comprehensive collection of statistical material strength
data for existing steel structures. Beginning with puddle steel, the collected data of former
steel products covers the era since about 1900 until today. The presented statistical data is
taken from appropriate literature, where results of material testing are documented. As con-
sistency of the data is a rather important aspect for comparative appraisal of the material
strength variability of former steel products and subsequent reliability analysis, appropriate
references are rare. Besides the extent of documented data, the sampling location from the
cross section is important. Additional information how to reveal the manufacturing process,
which was figured out as decisive aspect for distinction of different material strength popula-
tions, is described in Bucak & Mang (1998). The influence of the product shape on the mater-
ial properties is outlined in Reiche (2000). For the purpose of consistency, the data presented
in this study was therefore limited to samples taken from the flange of rolled sections. Kroyer
& Taras (2018) contains a more comprehensive recapitulation of the overall issue.
In the first instance, Tables 1 and 2 show some selected statistics for the material properties
of puddle steel and mild steel. Both data sets were collected by the accredited material testing
laboratory MFPA Leipzig. The available data is thus considered highly reliable. Comparing
the statistical material properties of puddle steel and mild steel, the mean of the yield stress
ReH as well as of the ultimate stress Rm is basically higher for mild steel than for puddle steel.
At the same time, the standard deviations of both material properties show higher values for
mild steel than for puddle steel. Although the minimum values of the discussed properties
show higher values for mild steel, which might be considered as a meaningful indication for
improved material properties, there is no striking indication so far for a significant improve-
ment of material properties by the introduction of mild steel fabrication techniques.
The collected data on mild steel as presented in Table 2 wasn’t further distinguished with
regard to the different mild steel production techniques in this paper. The material testing
data available for statistical evaluation was thus assigned to different mild steel production
techniques by chemical analysis of the microstructure. Table 3 shows correspondingly

Table 1. Statistical material properties of puddle steel from period 1850-1920


(Reiche, 2000).
Statistical Characteristic ReH [N/mm2] Rm [N/mm2] Rm/ReH

x̅ 235 371 1.59


σ 24.6 21.8 0.15
CoV 0.105 0.058 0.094
Min 177 305 1.28
Max 307 426 2.05

648
Table 2. Statistical material properties of mild steel from period 1896-1936
(Reiche, 2000).
Statistical Characteristic ReH [N/mm2] Rm [N/mm2] Rm/ReH

x̅ 268 393 1.48


σ 41 37 0.14
CoV 0.152 0.094 0.094
Min 192 326 1.03
Max 422 513 1.86

Sample size: n = 88 respectively 128, x̅ : mean, σ: standard deviation, CoV: coefficient of


variation.

Table 3. Statistical material properties of mild steel by manufacturing technique


(Reiche, 2000).
Statistical Parameter

Process Characteristic n x̅ σ CoV min max

Bessemer ReH [N/mm2] 12 282 49 0.17 234 357


Rm [N/mm2] 12 396 25 0.06 351 433
Thomas ReH [N/mm2] 133 268 39 0.15 208 422
Rm [N/mm2] 133 390 37 0.09 326 512
Siemens-Martin ReH [N/mm2] 92 265 43 0.16 192 405
Rm [N/mm2] 92 397 40 0.10 327 513

Sample size: n = 128, x̅ : mean, σ: standard deviation, CoV: coefficient of variation.

a statistical evaluation of material properties by the manufacturing technique. Comparing the


documented statistical material properties for Bessemer steel, Thomas steel and Siemens-
Martin steel as presented in Table 3, the statistical characteristics of Thomas steel and Sie-
mens-Martin steel are very similar. This is valid for both the yield stress and the ultimate
stress. Although the Bessemer process was the first process for mild steel production, the
mean of the yield stress is higher than for the other processes. As the scattering of the yield
stress, represented by the standard deviation, is also higher than for Thomas and Siemens-
Martin steel, the reliability assessment of those materials is reserved for reliability analysis.
According to the statistical data documented in Table 3, the mild steel manufacturing tech-
niques don’t differ significantly in the mean value of the ultimate stress. The corresponding
standard deviation of the ultimate stress of Bessemer steel is significantly lower in comparison
to Thomas steel and Siemens-Martin steel.
Although appropriate data for comparative purposes is rare, Reiche (2000) contains an add-
itional reference for mild steel strength statistics in relation to a specific structure. More pre-
cisely, when the crossing over the river Weichsel was built close to the former town of Fordon,
Poland in 1893, extensive material testing was performed. The bridge itself, called Fordoner
Weichselbrücke, was a truss bridge made of steel. At the time of construction, it was one of
the longest railway bridges in Europe. Mild steel from Thomas and Siemens-Martin manufac-
turing sites was furthermore also used for the first time to a greater extent. This was in prin-
ciple the reason for the extensive material testing campaign. To the authors’ knowledge for
the first time, the results from material testing were evaluated statistically and documented in
the literature. Figure 3 shows some selected statistics for the yield stress respectively for the
ultimate stress (Reiche, 2000). According to the data documented from Fordoner Weichsel-
brücke, the mean values of the yield stress and the ultimate stress are very similar for Thomas
steel and Siemens-Martin steel. However, Siemens-Martin steel shows nearly twice the stand-
ard deviation of Thomas steel with regard to both the yield stress and the ultimate stress. In

649
Figure 3. Material strength statistics documented from Weichselbrücke near Fordon (Reiche, 2000).

comparison to the statistical data documented by MFPA Leipzig, the mean values of the
material properties are slightly higher. Moreover, the standard deviations are significantly
lower than those ones indicated by MFPA Leipzig for Thomas steel respectively for Siemens-
Martin steel.
The historical revision of manufacturing techniques for steel revealed the predominance of
the oxygen steelmaking process as well as the electrical steel production from 1950 on. There
was no further main progress in manufacturing techniques afterwards. However, the literature
research carried out in this study gave rise to the assumption of basically consistent statistical
distributions of material properties in the period between 1950 and 2000. However, there were
naturally minor changes in manufacturing techniques. With Table 4 a statistical documentation
of extensive material testing for mild steel grades St37 and St52 is available. The material testing
was performed in the time period from 1949 until 1952 respectively from 1967 until 1971.

4 CALIBRATION OF ADEQUATE SAFETY FACTORS

Reassessment of existing structures according to state-of-the-art design rules applies a load


and design factor method (LRFD) using prescribed nominal strength values and partial safety
factors. This section describes a semi-probabilistic methodology in conjunction with Monte-
Carlo simulations for determination of adequate partial safety factors. Beyond the advantage
of avoiding fully probabilistic analyses, this approach is particularly useful when no stochastic

Table 4. Yield stress material test statistics for St37/St52 mild steel in different
periods (Petersen, 1987).
a. Time period 1949 and 1952
Steel x̅ σ CoV
Grade n [N/mm2] [N/mm2] [-]

St 37 4232 277 21.5 0.0776


St 52 2950 388 14.2 0.0366
b. Time period 1967 and 1971
St 37 1320 278 17.8 0.0640
St 52 2750 397 20.4 0.0514

where n = sample size; x̅ : mean; σ: standard deviation; CoV: coefficient of variation.

650
information on the distribution of the load effects is available. More precisely, the approach
allows for separate reliability analysis of the structural capacity. This is achieved in compli-
ance with Eurocode 1990 by a separation of the overall reliability into proportionate ones of
the basis variables. The concept is recovered briefly for convenience by means of Figure 4.
This schematic illustration shows the joint probability distribution of the very basic two-
parameter reliability problem in a standardized space from top view. For simplicity, the joint
probability distribution of the structural capacity and the load effect, denoted by the basis
variables R and E, is represented by concentric circles. The corresponding marginal distribu-
tions of R and E are shown adjacently to the axes. Applying a First Order-Reliability Method
(FORM) as a simplified probabilistic analysis, the procedure yields the minimal distance β
from the center of the joint probability distribution to the limit state. The corresponding point
on the limit state is termed design point. The projection of the vector from the center of the
joint probability distribution to the design point in the 2D-plane of R and E results in the
quantiles of the marginal distributions defining the design point. The overall reliability β is
thus separated into proportionate safety margins of the basis variables R and E. The semi-
probabilistic approach intends to simplify reliability analysis de-facto by prescribing the mar-
ginal safety margins. This is achieved by introduction of the so called weighting factors αR
and αE as fraction of the overall safety margin. The corresponding values are obtained by trig-
onometric expressions. Following this approach, there is no need to carry out a FORM ana-
lysis. The quantiles of the marginal distributions fulfilling adequate reliability can be
determined instead using the weighting factors. Further theoretical background is explained
comprehensively in Melchers & Beck (2018).
The separation of the overall reliability margin to its basis variables enables also usage of less
complicated methodologies for reliability analysis as Monte-Carlo simulations. Monte-Carlo
simulations were applied consequently as a suitable approach for the determination of adequate
partial safety factors for former steel products in this study. This is a simple and robust
approach for evaluation of mathematical models with stochastic parameters independent of its
complexity (Papaioannou et al., 2013). Following this approach, the various input parameters
are sampled individually according to the corresponding stochastic models from the literature
review. The mathematical model is evaluated repeatedly in the sequence using input parameter
sets from the samples. This results in a sample for the mathematical model of the structural
resistance with stochastic input parameters. This sample is supposed to include all statistical
uncertainties of the input parameters. As illustrated in Kroyer & Taras (2018), this is also valid
for the statistical uncertainty with regard to geometrical characteristics such as the cross-
sectional area and the moment of inertia. Once the sample for the structural capacity is
obtained, the procedure requires to determine the quantile of adequate marginal safety. This

Figure 4. Methodology of the semi-probabilistic approach.

651
quantile value is denoted by Rd,MCS, wherein the index “d” specifies a design value according to
the terminology in the Eurocode design standards and the index “MCS” highlights the gener-
ation by Monte Carlo simulations. The quantile Rd,MCS is determined by use of the weighting
factor αR and the inverse of the standard normal distribution Φ-1 in conjunction with the statis-
tical moments of the sample for the structural capacity μR and σR as follows:

σE
for 0:165 57:6 : αR ¼ 0:8; αE ¼ 0:7 ð1Þ
σR
βR ¼ αR  β ¼ 0:8  3:8 ¼ 3:04 ð2Þ

Rd;MCS ¼ 1 ð3:04Þ  σR þ μR ð3Þ

The calibration of adequate partial safety factors for former steel products is carried out
finally by comparison of a nominal resistance Rk with the design quantile Rd,MCS taken from
the Monte-Carlo sample for the structural resistance. As the studies presented herein were car-
ried out for revision of the guideline RiL 805 for reassessment of steel railway bridges, the
nominal resistance value was determined corresponding to this guideline.

Rk
γM ¼ ð4Þ
Rd;MCS

5 PRESENTATION OF SELECTED RESULTS

The design concept implemented in the Eurocode design codes for steel structures makes use of
three different material partial safety factors according to the limit state to be verified. For
assessment of the structural capacity of a member in tension with respectively without holes, as
well as for members under predominant compression, different material partial safety factors
are proposed. An overall study for calibration of those partial safety factors for former steel
products of different periods was carried out by evaluation of three limit states. The limit states
considered in this context and the corresponding results are presented more extensively in
Kroyer & Taras (2018). For illustration of the calibration methodology some specific results are
depicted in this paper however. For that purpose, the limit state of a structural member in ten-
sion without holes was selected. This limit state is illustrated schematically in Figure 7. The ratio
of the nominal capacity Rk and the sampled capacity Rd,MCS is shown appropriately as histo-
gram in Figure 5. The partial safety factor γM* verifying satisfactory reliability is determined

Figure 5. Exemplary results from probabilistic evaluation of tensile stressed brutto steel section limit state.

652
according to Equation 4 and marked in the figure within the range of the sample indicated by
red limiters.

6 EXTENDED PROBABILISTIC APPROACH FOR INDIVIDUAL ASSESSMENT

A methodology for determination of appropriate partial safety factors from overall statistical
distributions was shown in the previous sections. This section proposes a concept for consider-
ation of building-specific material data provided such data is available. The approach makes
use of Bayesian statistics for updating of stochastic models. According to this concept, an
a-priori stochastic model is transformed to an a-posteriori stochastic model using a sample for
the parameter subjected to statistical uncertainty. The mathematical framework is provided
by the Bayes-Theorem yielding a revised probability density for the event A when additional
information is included from a conditioning event E. The mathematical relation is typically
formulated according to Melchers & Beck (2017) as follows:

PðA \ E Þ
PðAjE Þ ¼ ð5Þ
PðEÞ

When the approach described above is applied to a stochastic model of material strength with
distribution parameters θ and an available sample x ¼ fx1 ; x2 ; . . . ; xn g of material tests with
size n, an appropriate formulation is intended using the probability density distribution fR’ of
the a-priori model as well as fR’’ of the a-posteriori model. An exemplary outcome of this
approach is illustrated in Figure 6 for the stochastic model of structural resistance. In this
case, the a-posteriori stochastic model shows a higher mean value and a lower standard devi-
ation than the a-priori stochastic model. The required material partial safety factor is there-
fore supposed to be lower when the a-posteriori stochastic model is used instead of the

Figure 6. Bayesian updating of stochastic models.

Figure 7. Tensile stressed member without holes.

653
a-priori stochastic model. However, as the sample is taken from a certain structure, the ana-
lysis result is basically a building-specific outcome and must not be transformed directly to
similar structures.

7 CONCLUSION

The content of this paper is devoted to material strength statistics of former steel products
and the determination of appropriate material partial safety factors for structural reassess-
ment. An extensive literature research was carried out for statistical data collection of former
steel products. The paper comprises pertinent statistics for material strength properties of
puddle steel and mild steel produced by different manufacturing techniques in the era since
about 1900 until today. For the purpose of determination of adequate partial safety factors,
a semi-probabilistic approach in conjunction with Monte-Carlo simulations and in compli-
ance with Eurocode 1990 was proposed. In order to illustrate the procedure, the methodical
outline concludes with an exemplary probabilistic evaluation of a specific limit state for
puddle steel and general mild steel statistical data. The results documented herein were
obtained in the course of methodological work for reappraisal of steel railway bridges. The
methodology and the documented statistics are certainly not limited to this case of applica-
tion. The paper concludes with an extended probabilistic concept for inclusion of structure-
specific statistical material data. Following this concept, Bayesian statistics are applied for
probabilistic updating of existing stochastic models in case an adequate sample of material
testing data is available.

REFERENCES

Bucak, Ö. & Mang, F. 1998. Erfahrungen mit alten Stahlkonstruktionen. Stahlbau 67(1): 46–60.
Kroyer, R. & Taras, A. 2018. Appraisal of the partial factors of safety for Eurocode-LRFD-conform
assessment of existing steel bridges. 12th Japanese German Bridge Symposium; Proc. Intern. Symp.,
Munich, 4-7 September 2018.
Kuhlmann, U. 2017. Stahlbau-Kalender 2017. Ernst & Sohn. Berlin.
Langenberg, P. 1995. Bruchmechanische Sicherheitsanalysen anrissgefährdeter Bauteile im Stahlbau.
RWTH Aachen. Aachen.
Lüddecke, F. 2006. Ein Beitrag zur Ertüchtigung bestehender Stahltragwerke unter besonderer Berück-
sichtigung des Fügeverfahrens Schweißen. BAM. Berlin.
Melchers, R. E. & Beck A. T. 2017. Structural Reliability Analysis and Prediction. John Wiley & Sons.
Hoboken.
Papaioannou et al. 2013. Reliability Sensitivity Analysis with Monte Carlo Methods. ICOSSAR 2013.
Proc. Intern. Symp. New York, 16-20 June 2013.
Pasternak, H. et al. 2010. Stahltragwerke im Industriebau. Ernst & Sohn. Berlin.
Petersen, C. 1987. Stahlbau. Friedr. Vieweg & Sohn. Braunschweig.
Reiche, A. 2000. Zustandsbewertung von metallischen Tragwerkskomponenten. Fraunhofer IRB Verlag.
Stuttgart.

654
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

The influence of stiffeners width on buckling modes of steel


LC-beams subjected to bending

T. Kubiak, Z. Kolakowski & F. Kazmierczyk


Lodz University of Technology, Lodz, Poland

ABSTRACT: The paper presents the influence of the stiffener’s width of the lip-channel sec-
tion beams’ wall on buckling modes. During calculation all type of buckling mode were deter-
mined i.e. global, local, distortional, lateral, torsional and coupling forms. The beams under
consideration were simply supported at the ends and subjected to bending in the web plane.
A plate model (2D) was adopted. The shear lag phenomenon and distortional deformations
were considered. A method of the modal solution using the analytical-numerical method
(ANM) and the transition matrix method, was applied. The special attention has been paid
for differences in buckling modes primary and secondary for m=1 as well as local mode (m>1)
and they interactions as well as influence on postbuckling behaviour. The results obtained by
developed analytical-numerical method based on asymptotic Koiter theory have been verified
by results obtained by FEM calculations.

1 INTRODUCTION

As it is well known the cold formed steel beams are widely used in many different branches of
industry. It is well known that the behavior of thin-walled structures is determined by many
factors, e.g. geometrical dimensions, type of load, initial imperfections (shape inaccuracies or
residual stresses). Due to the mentioned above fact, in world wise literature a lots of books
and papers devoted to stability of thin-walled structures can be easily found. In spite of this,
that issue is still of interest to many scientists.
The problem of stability and postbuckling behavior is solved using well-known and still
developed numerical methods, which are as follow: Finite Strip Method (FSM), Finite Elem-
ent Method (FEM) and Generalised Beam Theory (GBT).
Focusing only on the publications of the last few years, regarding the interactive buckling
of thin-walled open-section beams, one could mention the work of teams gathered around the
following scientists: Ádány and Schafer (Schafer & Adany 2006, Adany 2018), Hancock (Han-
cock 2018), Rhodes and Loughlan (Loughlan & Rhodes 1979, Rhodes & Loughlan 1980),
Rasmussen (Niu et. al. 2014a, 2014b), Silvestri and Camotim (Abambres et. al. 2014, Bebiano
et. al. 2018)) and many others.
To the best knowledge of the authors, there are not enough papers showing the influence of
the stiffener width in lipped channel beam subjected to pure bending on buckling modes, their
interaction and postbuckling behavior. The problem has been solved using analytical-
numerical method (ANM) based on Koiter asymptotic theory (Kolakowski & Krolak 2006).

2 ANALITYCAL-NUMERICAL METHOD

The employed ANM analytical-numerical method (Kolakowski & Krolak 2006) is based on
Koiter’s asymptotic perturbation theory with Byskov-Hutchinson’s formulation (Byskov &
Hutchinson 1977) and it has been developed by Kolakowski et. al. (Kolakowski & Krolak
2006, Kolakowski & Mania 2013) for more than 30 years. The software based on mentioned

655
methods allow to solve non-linear problem of interaction buckling of thin walled structures.
The structures under consideration could be the prismatic thin-walled beam-columns built of
plates connected along longitudinal edges and subjected to the longitudinal load i.e. uniform
compression, pure bending or they combination (eccentric compression). The shear lag phe-
nomenon and distortional deformations are considered. In order to account for all possible
buckling modes from global through local to coupled, a plate model (i.e. 2D) of thin-walled
structures was applied. The beams were simply supported at their ends. It was assumed that
the beam material from which the structure was made obeyed Hooke’s law.
For each plate component, precise geometrical relationships (i.e., full Green’s strain tensor)
were assumed (Kolakowski & Krolak 2006, Kolakowski & Mania 2013). According to the
Koiter’s theory the displacement fields and the sectional force fields were expanded into
power series with respect to the dimensionless amplitude of the r-th mode deflection (Byskov
& Hutchinson 1977, Kolakowski & Krolak 2006, Kolakowski & Mania 2013):

U ≡ ðu; v; wÞ ¼ λU0 þ ζ r Ur þ ζ 2r Urr þ …


ð1Þ
N ≡ ðNx ; Ny ; Nxy Þ ¼ λN0 þ ζ r Nr þ ζ 2r Nrr þ …

where: load factor is λ, the pre-buckling (i.e., unbending) fields are U0, N0, the first-order non-
linear fields are Ur, Nr and the second-order non-linear fields – Urr, Nrr, respectively. The range
of indices is [1, J], where J is the number of interacting modes. The summation is supposed on
the repeated indices. The two buckling modes UI and UK are mutually orthogonal in the follow-
ing sense σ0  l11 ðUI ; UK Þ ¼ 0; ðI; K Þ ¼ ½1; J
; I≠K where J are all the relevant buckling modes
that are believed to be important in the structural response. For more detailed analysis, see
(Byskov & Hutchinson 1977 Kolakowski & Krolak 2006).
For thin-walled beams under bending with the geometric imperfections U  (only the linear
initial imperfections, i.e., U  ¼ ζ  Ur ), the total potential energy has the form (Kolakowski &
r
Krolak 2006, Kolakowski & Mania 2013):

 
1 1X J
M 1XJ X
J X
J
1 XJ

X
J
M
 ¼  M 2 a0 þ 
ar ζ 2r 1  þ 
apqr ζ p ζ q ζ r þ brrrr ζ 3r  ar ζ r ζ r

2 2 r¼1 Mr 3 p q r gh4 r r
Mr

ð2Þ

and the equilibrium equations are as follows:


 
M M 
1 ξ þ apqr ξ p ξ q þ brrrr ξ 3r ¼ ξ r ¼ 1; . . . ; J ð3Þ
Mr r Mr r

where: M is a magnitude of the applied bending moment, Mr, ζ r , ζ r – the buckling moment of
the r-th buckling mode, the dimensionless amplitude of the r-th buckling mode and the dimen-
sionless amplitude of the initial deflection corresponding to the r-th buckling mode, respect-
ively. The constant coefficients a0 ;  apqr and brrrr can be determined with the equations
ar ; 
described in the literature (Byskov & Hutchinson 1977, Kolakowski & Krolak 2006, Kola-
kowski & Mania 2013).
It is worth mentioned that the analytical-numerical method (Kolakowski & Krolak 2006) as
well as semi-analytical method (Kolakowski & Mania 2013) allow a much simpler analysis of
observed phenomena and their interpretation than based on results obtained by FEM.

3 FEM MODEL

Numerical simulation on the basis of finite element method (FEM) was performed in ANSYS
18.2® software. Numerical model was created using structural element SHELL 181 with four

656
Figure 1. Discretized FEM model with applied boundary conditions.

node element and six degrees of freedom at each node. Boundary conditions were applied to
the model in this way to ensure pure bending: concentrated forces acting on the nodes in both
ends of cross-sections were applied in this way to correspond to the bending stress distribution
(Figure 1). In the nodes representing the support of the beam (z = 0 and z = L), the displace-
ment in the transverse direction (UX = 0, UY = 0) was removed.
Bending moment Mb was calculated as the sum of forces acting on the beam and the dis-
tance from the neutral axis and angle of rotation α was determined based on the displacement
of the flange in the place of beam support. The analysis of the displacement reveals the rota-
tion of beam around two axes during the loading: the rotation around X axis and Y axis.

4 RESULTS OF CALCULATIONS

The calculations have been performed employing analytical-numerical method shortly


described in Section 2 and commercial software ANSYS based on finite elements method. The
lipped channel section beams made of steel (E = 210 GPa, v = 0.3) with cross-section dimen-
sion b1 = 80 mm, b2= 40 mm, b3= 4, 10 or 15 mm and thickness t = 1mm presented in Figure 2
have been considered.
The buckling modes with buckling stress σcr [MPa] for different length of considered beams
are presented in Tables 1–3 and by points in signature buckling curves presented in Figures 3–5.
The course of the curves for primary modes (lower curves) and secondary modes (upper curves)
depends on the width of stiffener. For the stiffener width b3 = 4 mm the one maximum and two
minima can be distinguished (Figure 3) for both signature buckling curves. Additional it was
observed that the type of buckling mode change when the length increases and the maximum is
passed. For example, for the beam length lower or equal to 720 mm the primary buckling
modes correspond to distortional mod and for higher length the torsional modes are observed
(see first row denoted as GP – “global primary” in Table 1). Similar relation between secondary
global (GS – Table 1) buckling modes and the course of corresponding signature buckling
curve (Figure 3) is observed – for L ≤ 260mm the flexural-distortional buckling modes exist, for
260mm < L ≤ 720 mm – distortional – lateral modes, for 720 mm < L ≤ 2000 mm – distortional
and torsional modes and for L > 2000 mm global flexural. The minimal local modes (i.e. for
number of half waves in longitudinal direction m >1) also change from local flexural for very

657
Figure 2. Cross-section dimension of considered lipped channel beam.

Table 1. Buckling modes and stress σcr [MPa] for different length of lipped channel section beam with
stiffeners b3 = 4 mm.
Length L [mm]

200 263 400 720 1000 2000

GP

188 221 363 631 451 136

GS

3731 4763 2795 1420 1735 2592

m=5 m=2 m=2 m=4 m=6 m=11


470 209 188 185 187 185

GP – primary global buckling mode, GS – secondary global, L – local mode

658
Table 2. Buckling modes and stress σcr [MPa] for different length of lipped channel section beam with
stiffeners b3 = 10mm.
Length L [mm]

80 130 200 270 340/450 600 830 2000

GP

695 862 571 430 398/444 570 647 170

GS

1530 2205 3593 5418 4098/2627 1689 1304 2172

LD - - - - -

m=2 m=2 m=6


407 422 398

LF

m=2 m=3 m=5 m=7 m=8/11 m=15 m=21 m=49


532 534 532 533 533/532 532 532 532

GP – primary global, GS – secondary global, LD – local distortional, LF – local flexural

short beams (L = 200mm) to local distortional for L > 200 mm. Similar relations is observed for
beams with higher width of stiffener.
Analyzing the course of signature buckling curves presented in Figure 4 and corresponding
buckling modes presented in Table 2 the same relations between curves maxima and change of
modes can be easily found. Additionally, it was found that for short beam (L ≤ 450 mm) the
lowest local buckling mode correspond to flexural one (row denoted as LF – “local flexural”

659
Table 3. Buckling modes and stress σcr [MPa] for different length of lipped channel section beam with
stiffeners b3 = 15mm.
Length L [mm]

140 170 300 450 850 2000

GP

1213 1292 687 550 718 203

GS

2717 2597 4718 2845 1215 1941

m=3 m=4 m=7 m=11 m=20 m=48


526 522 523 523 522 522

GP – primary global, GS – secondary global, L – local mode

Figure 3. Signature buckling curves of lip channel beam with stiffeners b3 = 4 mm.

660
Figure 4. Signature buckling curves of lip channel beam with stiffeners b3 = 10 mm.

Figure 5. Signature buckling curves of lip channel beam with stiffeners b3 = 15 mm.

Figure 6. Exemplary local modes: flexural (a) for L = 170mm, b3 = 15 mm and distortional (b) for L =
263 mm, b3 = 4 mm.

in Table 2 or in Figure 6a) but for longest beam i.e. L > 450 mm the lowest local mode is local
distortional mode (row denoted as LD – “local distortional” in Table 2 or in Figure 6b).
In Figure 5 the signature primary and secondary curves for lipped channel section beam
with the highest considered width of stiffener b3 = 15 mm are presented. The similar relation
can be found. Additionally, it was found that the maxima of signature primary buckling curve

661
Figure 7. Exemplary global modes: distortional (a) for L = 340 mm, b3 = 10 mm and lateral-distortional
(b) for L = 600 mm, b3 = 10 mm.

Figure 8. Exemplary global torsional mode for L = 2000 mm, b3 = 15 mm.

correspond to minima of signature secondary buckling curve. It means that the primary and
secondary buckling modes change crossing the same value (see buckling modes in Table 3)
Looking at values of buckling stress, especially corresponding to secondary modes, it is very
easy noticed that they are extremely high (higher not only then yield limit but also ultimate
stress). But corresponding modes interact with the lowest local modes and with primary global
leading to decreasing buckling stress in case of interactive buckling, they have also influence on
postbuckling behavior as well as on load-carrying capacity (Kolakowski & Jankowski 2018).
The calculations allowing to find buckling modes for beams with different stiffener
width were also performed employing finite element software ANSYS. Very similar
results have been obtained to those from analytical numerical method. Very good agree-
ment has been obtained for buckling modes (the same were found employing both
methods) as well as for buckling stress (the highest differences between ANM and FEM
was lower than 10%).
The exemplary buckling modes are presented in Figures 6–8. The modes presented in
Figure 6 allow to conclude that for very narrow stiffeners the distortional modes (Figure 6b)
appear instead of flexural like for beam with wide stiffener (Figure 6a).

5 CONCLUSIONS

The influence analysis of stiffeners width of lipped channel section beam subjected to pure
bending on buckling load have been performed. It was found that not only buckling load
(buckling stress) but also buckling load strictly depend on width of stiffener. Due to previous
analysis performed not only by Authors’ of this paper (Kolakowski & Jankowski 2018) but
also by other Scientists (eg. Martins et. al. 2017) dealing with distortional buckling mode
influence on postbuckling behavior the following can be conclude:

662
– the interactive buckling depends not only on the buckling stress but on buckling mode
which interact;
– the signature buckling curve, especially the maximums allow to predict how many different
global primary and secondary modes can appears with increasing beam length (length of
a half-wave);
– the local buckling mode also depend on stiffener’s width of lipped channel beam – the local
distortional mode can appear which correspond to even two times lower buckling stress
than local flexural buckling mode (cf. row L in Table 1 and rows LD and LF in Table 2).

ACKNOWLEDGEMENT

The studies have been conducted as a part of the research projects financed by the National
Science Centre, Poland - decision numbers: UMO-2017/25/B/ST8/00007.

REFERENCES

Abambres M., Camotim D. & Silvestre N. 2014. Modal Decomposition of Thin-Walled Member Col-
lapse Mechanisms, Thin-Walled Structures, 74: 269–291.
Adany S. 2018. Constrained shell Finite Element Method for thin-walled members, Part 1: constrains for
a single band of finite elements. Thin-Walled Structures, 128: 43–55.
Bebiano R., Camotim D. & Gonçalves R. 2018. GBTUL 2.0 − A second-generation code for the GBT-
based buckling and vibration analysis of thin-walled members, Thin-Walled Structures, 124: 235–257.
Byskov E. & Hutchinson J.W. 1977. Mode interaction in axially stiffened cylindrical shells. AIAA J., 15
(7): 941–948.
Hancock G.J. 2018. Coupled Instabilities in Metal Structures (CIMS) – What have we learned and are
we going? Thin-Walled Structures, 128: 2–11.
Kolakowski Z. & Krolak M. 2006. Modal coupled instabilities of thin-walled composite plate and shell
structures. Composite Structures, 76: 303–313.
Kolakowski Z. & Mania J.R. 2013. Semi-analytical method versus the FEM for analysis of the local
post-buckling. Composite Structures 97: 99–106.
Kolakowski Z. & Jankowski J. 2018. Interactive buckling of steel C-beams with different lengths – From
short to long beams. Thin-Walled Structures, 125: 203–210.
Loughlan J. & Rhodes J. 1979. The interactive buckling of lipped channel columns under concentric or
eccentric loading, in Rhodes J., Walker A.C. (Eds.), Proc. of Int. Conf. on Thin-walled Struct., Gran-
ada, London, University of Strathclyde, Glasgow, pp. 40–55.
Martins A., Landesmann A., Camotim D., & Dinis P. B. 2017. Distortional failure of cold-formed steel
beams under uniform bending: Behaviour, strength and DSM design. Thin-Walled Structures, 118:
196–213.
Niu S., Rasmussen K.J.R. & Fan F. 2014a. Distortional-global interaction buckling of stainless steel
C-beams: Part I – Experimental investigation, J. Constr. Steel Res., 96: 127–139.
Niu S., Rasmussen K.J.R. & Fan F. 2014b. Distortional-global interaction buckling of stainless steel
C-beams: Part II – Numerical study and design, J. Constr. Steel Res., 96: 40–53.
Rhodes J. & Loughlan J. 1980. Simple design analysis of lipped channel columns, Proc., Fifth Int. Spe-
cialty Conf. on Cold-formed Steel Struct., St. Louis, Missouri, pp. 241–262.
Schafer B.W. & Adany S. 2006. Buckling analysis of cold-formed steel members using CUFSM: conven-
tional and constrained finite strip method, 18th International Specialty Conference on Cold-Formed
Steel Structures, October 26–27, 2006, Orlando, Florida.

663
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Design of slender compressed plates in structural steel joints by


component based finite element method

M. Kuříková & F. Wald


Czech Technical University in Prague, Praha, Czech Republic

J. Kabeláč
Brno University of Technology, Brno, Czech Republic

ABSTRACT: This contribution shows the current trends in advanced modelling of connec-
tions and differences of the research oriented and design oriented models. The principles of
Component Based Finite Element Modelling and the system response quantity and applica-
tion features for design of slender compressed elements was developed. The elements are ana-
lysed by material non-linear and by buckling analyses. The proper behaviour of components,
e.g. of bolts, anchor bolts, welds etc., is treated by introducing components representing its
behaviour in term of initial stiffness, ultimate resistance and deformation capacity in 3D envir-
onment. To show this process a contribution is prepared, which summarises the history of
achievements in prediction of behaviour of structural connections. The validation on an
example of an unstiffened haunch in a welded portal frame joint and verification on the
column web panel and a column stiffener is shown.

1 INTRODUCTION

Slender plates in compression, their geometrical imperfections, and residual stresses are
taken into account in geometric and material nonlinear analysis by research oriented
finite element model (ROFEM), see EN1993-1-5:2005. This should be précised according
to the different plate/joint configuration. The design oriented finite element model
(DOFEM) procedure naturally offers the prediction of the elastic and inelastic buckling
load of the joint. The design procedure for class 4 cross-sections according to reduced
stress method is described in Annex B. It allows to predict the post buckling resistance
of the joints. Critical buckling modes are determined by material linear and geometric
nonlinear analysis. In the first step the minimum load amplifier for the design loads to
reach the characteristic value of the resistance of the most critical point coefficient αult,k
is obtained by material nonlinear analysis. Ultimate limit state is reached by 5 % plastic
strain. The critical buckling factor αcr is determined by eigenvalue analysis and stands
for the load amplifier to reach the elastic critical load under complex stress field. The
load amplifiers are related to the non-dimensional plate slenderness, which is determined
as follows:
rffiffiffiffiffiffiffi
αult
λ¼ ð1Þ
αcr

Reduction buckling factor ρ is calculated according to Annex B of EN1993-1-5:2005. Conser-


vatively, the lowest value from longitudinal, transverse and shear stress is taken. Figure 1 visu-
alize the relation between plate slenderness and reduction buckling factor.

664
Figure 1. Buckling reduction factor ρ according to Annex B of EN1993-1-5:2005.

The verification of the plate is based on the von Mises yield criterion and reduced stress
method. Buckling resistance is assessed as:
αult :ρ
1 ð2Þ
γM1

where γM1 is the partial safety factor.


The component column web panel in shear is described in Cl. 6.2.6.1 EN1993-1-8:2006. The
design method is limited to column web slenderness d/tw ≤ 69 ε, where d is the plate width, tw
the plate thickness, ε = √(235/fy), and fy the yield strength. Webs with higher slenderness are
designed according to EN1993-1-5:2005 Cl. 5 and Annex A. The shear resistance is made of
shear buckling resistance of the web panel and resistance of the frame made of the flanges and
stiffeners surrounding the panel. The buckling resistance of the web panel is based on the
shear critical stress

τcr ¼ kτ σE ð3Þ

where σE is the Euler critical stress of the plate


 2
π2 E tw
σE ¼ ð4Þ
12ð1   2 Þ hw

The buckling coefficient kτ is obtained in EN 1993-1-5 Annex A.3. The slenderness of the web
panel is
sffiffiffiffiffiffi
λw ¼ 0; 76 fyw ð5Þ
τcr

The reduction factor χw may be obtained in EN 1993-1-5 Cl. 5.3. The shear buckling resistance
of the web panel is

χ fyw hw tw
Vbw;Rd ¼ wpffiffiffi ð6Þ
3γM1

The resistance of the frame may be designed according to Cl. 6.2.6.1 EN 1993-1-8:2006.
General rules for the design of slender plates in structural steel joints are specified in EN
1993-1-8:2006, based on research done by Zoetemeijer (a, b 1981). For haunches and column

665
stiffeners, using the same or higher plate thickness and steel grade for the haunch web, flange
and column stiffeners as for the beam flange is recommended. The typical plate is designed as
a first or second class cross-section which allows the plastic design. Fourth class haunches
with thin web and reduced flange stiffness are not covered in the design code. The previous
research of the authors proved that the design resistance of a joint with the fourth class
haunch is significantly increased as is proved in (Kurejková and Wald, 2014). Practice is
asking for a design method for a general plate, which is not limited to plastic design only. The
design method should be fully adapted to advanced FEM analysis.
Haunches are widely used in seismic moment connections. Brittle fractures in steel moment
connections caused by earthquakes motivated engineers to design a connection with plastic
zones away from the connection. Welding a triangular haunch has been shown to be very effect-
ive in seismic steel connections (Yu et al. 2000). To minimize the construction cost, new arrange-
ments are proposed. The study by Lee et al. (2003) is focused on a straight haunch with one free
end and a seismic design procedure based on a strut model is described in (Lee and Uang,
2001). In the past 30 years, studies have focused on the design of haunches without a flange.
Design methods are published in (Laustsen et al. 2012), (Martin 1979) and (Shakya and Vinna-
kota, 2008). A large experimental investigation is published in (Robinson 1983) and (Salmon
et al. 1964). All methods are not ready to be adopted directly in the component method or in
numerical calculations, because the estimation of the failure mechanism is required.
Each type of the above mentioned stiffeners has a different design procedure and limits the
stiffener size by the recommended plate thickness or dimensions. The design procedure for
general plate dimensions is not available and joints for which the component method is not
applicable are analysed by the finite element method (FEM). The research oriented FEM
(ROFEM) model, which is based on materially and geometrically nonlinear analysis with
imperfections (GMNIA), is not suitable for the design of common joints, especially for time-
consuming modelling, difficulties in entering the imperfections and the definition of the ultim-
ate limit state. On the other hand, the component based FEM model (CBFEM) with material
nonlinear analysis without imperfections and with the assessment of the components of bolts,
welds and plates becomes an alternative approach to the design of joints as is described in
(Martin 1979) and (Shakya and Vinnakota, 2008). The design of fourth class stiffeners in the
joints of steel structures takes into account local buckling of a slender compressed plate, shear
buckling of a slender web panel and local buckling of a compressed plate between the bolts,
see (Kurejková et al. 2015).
The optimisation of stiffener thickness is limited by buckling of slender plates. The design
of stiffeners, which was so far limited to the 1st and 2nd class cross-sections, is replaced by
FEM design procedure including buckling analysis. The general procedure for FEM models is
proposed and verified for slender stiffeners, for local web panels in shear and for slender
haunches.
Stiffeners are primarily provided to increase the buckling resistance of various plated struc-
tures. Experimental results reviewed in (Shi et al. 2007) and (Abidelah et al. 2012) indicate
clearly that the presence of stiffeners in the extended end-plate connections has a great influ-
ence on the behaviour of the joints. However, buckling and post-buckling of thin-walled stiff-
ened plates are nonlinear phenomena and no design formulas for the design of stiffeners in
structural steel joints have been suggested. Nevertheless, if optimum stiffener dimensions are
provided, the stiffened plate should gain maximum buckling and post-buckling resistance.
In the first part of the study, haunches with and without flanges are studied by laboratory
tests with the detailed measurement of deflections and strains. In the second part, the resist-
ance is calculated and validated against experimental data. For the ROFEM validation pur-
poses, material and geometrical nonlinear finite element analysis with imperfections is
applied. The validity is demonstrated on the comparison of load-deflection curves.
Hence, this paper uses finite element analysis to describe the effect of the plate thickness on
the buckling resistance to avoid local buckling. Subsequently, curves and formulas for the
design of slender plates in design oriented FEM models are presented. The proposed design
procedure is verified for the column web panel and column stiffener joint (Kurejková and
Wald, 2014).

666
2 EXPERIMENTAL INVESTIGATION

The experiment of haunches covered six specimens with different types of edge reinforcement.
Three specimens, A to C, are without flanges, three specimens, D to F, have flanges with dif-
ferent stiffness. Unstiffened specimens differ in the web thickness tw and the web width bw.
Reinforced specimens differ in the web thickness tw, the flange thickness tf and the flange
width bf. The dimensions of specimens are summarized in Table 1. Vertical and horizontal
plates with the thickness of 20 mm are welded to each other. In the inner corner of the triangu-
lar web, there is a cut-out for a better fit and weld to the vertical and horizontal plates. The
flange is welded to the triangular web and to both plates. The material characteristics of the
steel plates are summed up in Table 2. Young’s Modulus, the yield strength and the ultimate
tensile strength are measured for the plate thicknesses of 4, 6 and 12 mm, while the nominal
values are taken for the plate thickness of 20 mm.
The test set-up for the specimen without a flange is shown in Figure 2. Vertical load at the
top of the specimen was applied by hydraulic press. Lateral torsional buckling is prevented by
lateral supports. Hinges are used at the bottom. Loading is carried out in predetermined steps
until the collapse of the specimen. Every test starts with loading to approximately one tenth of
the load capacity determined on numerical models before testing followed by complete
unloading. Then, the specimens are reloaded to the collapse. The measured quantities are
load, vertical and horizontal displacements and strains in predefined places.

3 RESEARCH ORIENTED FEM MODEL

The ROFEM was modelled in the advanced FEM software Dlubal RFEM. In the numerical
model, 4-node quadrilateral shell elements with nodes at its corners are applied with a maximum
side length of 10 mm. Six degrees of freedom are in every node: 3 translations (ux, uy, uz) and 3
rotations (φx, φy, φz). Material and geometric nonlinear analysis with imperfections (GMNIA) is
applied. A true stress-strain material model is chosen and the von Mises yield criterion is applied.
Equivalent geometric imperfections are derived from the first buckling mode and the amplitude is
set according to Annex C in EN1993-1-5:2005. Large deformation analysis is used and the
Newton-Raphson method for solving systems of equations is chosen. The number of loading

Table 1. Specimens geometry.


Triangular web Flange

Specimen bw [mm] hw [mm] tw [mm] bf [mm] tf [mm]

A 200 400 6 - -
B 400 400 6 - -
C 400 400 4 - -
D 400 400 6 60 6
E 400 400 6 120 12
F 400 400 4 120 12

Table 2. Material characteristics used in numerical models.


Plate thickness [mm]

Material characteristics 4 6 12 20

Young’s Modulus E [GPa] 163,0 158,7 159,8 160,0


Yield strength fy [MPa] 417,5 323,5 395,4 355,0
Ultimate tensile strength fu [MPa] 499,3 467,0 529,6 510,0

667
Figure 2. Test set-up of specimens without flange.

steps is set to 50, the convergence criteria for tolerance to 1,0 % and the maximum number of
iterations to 50. Loading is applied through displacement increments, which reflect the experiment
conditions better. The analysis stops at a certain limit of displacement. The numerical model is
shown in Figure 3.
An example of the comparison of ROFEM and experimental test on the load-deflection
behaviour is shown in Figure 4a. The comparison of ultimate resistances measured in
experiment and obtained from ROFEM is shown in Figure 4b. The resistance calculated
in the numerical model is displayed on the horizontal axis. The resistance measured in the
experimental study is displayed on the vertical axis. It can be seen that good agreement
exists. It can be found that good agreements between numerical models and experimental
results of the haunches exist in the failure mode. For more details, see (Kurejková and
Wald 2017).

Figure 3. ROFEM model of specimens without flange.

668
Figure 4. a) Load deflection curve of a haunch without a flange b) Experiments’ ultimate resistances
compared against ROFEMs’.

4 DESIGN ORIENTED FEM MODEL

Design procedure for class 4 cross-sections is described in Introduction of this paper. The design
procedure is verified on the comparison of design oriented FEM (DOFEM) and research
oriented FEM (ROFEM) models. The buckling analysis is implemented in the software. The
calculation of the design resistances is done according to design procedure. FCBFEM is interpol-
ated until ρ ∙ αult,k/γM1 is equal to 1,0. A beam-column joint with a slender column web is stud-
ied. The height of the beam web and width of the column web panel are changing. The
geometry of the examples is described in Table 3. The joint is loaded by bending moment only.
The global behaviour of a beam to column joint with slender column web described by
moment-rotation diagram in CBFEM model is shown in Figure 5 and Table 4. Attention is

Table 3. Specimens geometry for a sensitivity study of the width of the column web panel.
Column flange Column web Beam

bf tf hw tw
Sample [mm] [mm] [mm] [mm] IPE Material

IPE400 250 10 820 4 400 S235


IPE500 250 10 820 4 500 S235
IPE600 250 10 820 4 600 S235

Figure 5. Moment-rotation curve of example IPE600.

669
Table 4. Verification of the critical load Mcr and the resistance MR predicted by CBFEM to CM and
ROFEM.
Mcr αcr MR αult,k Difference

ROF- CBF- CBF- ROF- CBF- CBF- MCBFEM MCBFEM


CM CM
EM EM EM EM EM EM /MROFEM /MCM
Sample [kNm] [kNm] [kNm] [-] [kNm] [kNm] [kNm] [-] [%] [%]

IPE400 256 275 303 1,75 170 177 186 1 9 5


IPE500 216 234 236 1,31 177 194 180 1,29 2 8
IPE600 195 210 210 1,13 200 205 186 1,52 8 10

focused on the main characteristics: the design resistance MCBFEM and critical load Mcr. The
diagram is completed with a point where yielding starts Myield and resistance by 5 % plastic
strain MR.
The design resistance calculated by CBFEM is compared with ROFEM and CM. The com-
parison is focused on the design resistance and critical load. The results are ordered in Table 4.
The diagram in Figure 5c shows the influence of the width of the column web on the resistances
and critical loads in the examined examples. The results show good agreement in critical load and
design resistance. The CBFEM model of the joint with a beam IPE600 is shown in Figure 6a.
The first buckling mode of the joint is shown in Figure 6b.
A beam-to-column joint with a slender column web stiffener is studied. Same cross-
section is used for the beam and the column. The thickness of the column web stiffener is
changing. The geometry of the examples is described in Table 5. The joint is loaded by bend-
ing moment.
The design resistance calculated by CBFEM model in IDEA StatiCa (2019) software is
compared with ROFEM. The comparison is focused on the design resistance and critical load.
The results are summarised in Table 6. The diagram in Figure 7c) shows the influence of the
thickness of the column web stiffener on the resistances and critical loads in the examined
examples. The results present very good agreement in critical load and design resistance. The
CBFEM model of the joint with web stiffener thickness 3 mm is shown in Figure 7a. The first
buckling mode of the joint is shown in Figure 7b.

Figure 6. a) CBFEM model b) First buckling mode c) Influence of width of the column web panel on
resistance and critical load.

670
Table 5. Specimens geometry for a sensitivity study the column web stiffener.
Column/beam flange Column/beam web Stiffener

bf tf hw tw ts
Sample [mm] [mm] [mm] [mm] [mm] Material

t3 400 20 600 12 3 S235


t4 400 20 600 12 4 S235
t5 400 20 600 12 5 S235
t6 400 20 600 12 6 S235

Table 6. Verification of critical load Mcr and resistance MR predicted by CBFEM to ROFEM.
Mcr αcr MR αult,k Difference

ROFEM CBFEM CBFEM ROFEM CBFEM CBFEM MCBFEM/MROFEM


Example [kNm] [kNm] [-] [kNm] [kNm] [-] [%]

t3 260 286 0,94 290 304 1,96 5


t4 511 561 1,32 419 426 1,43 2
t5 874 950 1,73 532 549 1,13 3
t6 1346 1460 2,32 580 629 1,00 8

Figure 7. a) CBFEM model b) First buckling mode c) Influence of thickness of the column web stiffener
on resistance and critical load.

SUMMARY AND ACKNOWLEDGMENT

Verification studies confirmed the accuracy of the CBFEM model for prediction of a column
web panel and a column web stiffener behaviour. Results of CBFEM are compared with the
results of the ROFEM and CM. The design procedure is verified on the ROFEM model. Pro-
cedures predict similar global behaviour of the joint. The difference in design resistance is in
all cases up to 10 %.
The procedure to design slender stiffeners in structural steel joints is proposed. It is possible
to use material nonlinear analysis (MNA) without imperfections and linear buckling analysis
(LBA) to design slender stiffeners in FEM models with a complex geometry. Based on the
distribution of strains calculated in MNA and the critical buckling factor in LBA is assessed
the buckling resistance without using second order calculation. In practical design it is recom-
mended to check the buckling resistance for critical buckling factor smaller than 3,0, see

671
(Wald et al. 2017). Otherwise is the resistance governed by reaching 5 % plastic strain for stiff-
eners with smaller slenderness.
This study was carried out with support by the Technology Agency of the Czech Republic
within the MERLION II project No. TH02020301.

REFERENCES

Abidelah, A., Bouchair, A., Kerdal, D. 2012. Experimental and analytical behavior of bolted end-plate
connections with or without stiffener, in Journal of Constructional Steel Research. 76(18): 13–27.
Dlubal RFEM 5.0, 2019. at https://www.dlubal.com/en/downloads-and-information
EN1993-1-5:2005. Eurocode 3: Design of steel structures - Part 1-5: Plated Structural Elements. Brus-
sels: CEN.
EN1993-1-8:2006, Eurocode 3: Design of steel structures - Part 1-8: Design of joints. Brussels: CEN.
Idea Statica 9.1.5, 2019. at https://www.ideastatica.com/downloads/
Kurejková, M. and Wald, F. 2017. Design of haunches in structural steel joints, in Journal of Civil
Engineering and Management. 23(6): 765–772.
Kurejková, M. and Wald F. 2014. Compressed stiffeners in structural connections, in Proceedings of the
7th European Conference on Steel and Composite Structures. Eurosteel, 317–318.
Kurejková, M., Wald, F., Kabeláč, J., Šabatka, L. 2015 Slender compressed plate in component based
finite element model, in Proceedings of 2nd International Conference on Innovative Materials., Struc-
tures and Technologies.
Laustsen B., Nielsen M., Hansen T., Gath J. 2012. Stability of brackets and stiffeners in steel structures,
in Steel Construction, 5: 138–144.
Lee, C.H., Jung, J.H., Oh, M.H., Koo, E.S. 2003. Cyclic seismic testing of steel moment connections
reinforced with welded straight haunch, in Engineering Structures. 225(14): 1743–1753.
Lee, C.H. and Uang C.M. 2001. Analytical modeling and seismic design of steel moment connections
with welded straight haunch, in Journal of Structural Engineering. 127(9): 1028–1035.
Martin, L. 1979. Methods for the limit state design of triangular steel gusset plates, in Building and
Environment. 14: 147–155.
Robinson S. 1983. Failure of steel gusset plates, The University of Aston in Birmingham, Department of
Civil Engineering and Construction.
Salmon C., Buettner D., O’Sheridan T. 1964. Laboratory investigation of unstiffened triangular bracket
plates, in Journal of the Structural Division, 90: 257–278.
Shakya S. and Vinnakota, S. 2008. Design aid for triangular bracket plates using AISC specifications, in
Engineering Journal. 45: 187–196.
Shi, Y., Shi, G., Wang, Y. 2007. Experimental and theoretical analysis of the moment–rotation behaviour
of stiffened extended end-plate connections. in Journal of Constructional Steel Research, 63: 1279–1293.
Wald, F. et al. 2017. Benchmark cases for advanced design of structural steel connections. Praha: Czech
Technical University Publishing House.
Yu, Q.S., Uang, C.M., Gross, J. 2000. Seismic rehabilitation design of steel moment connection with
welded haunch. in Journal of Structural Engineering. 126(1): 69–78.

672
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Experimental verification of lateral-torsional buckling of steel


I-beams with tapered flanges

J. Kuś
Faculty of Civil Engineering and Architecture, Opole University of Technology, Poland

ABSTRACT: A procedure for finding the critical buckling moment of a tapered beam is
proposed with the application of potential energy calculations using the Ritz method. The
respective solution allows us to obtain the critical moments initiating the lateral buckling of
the flange-tapered steel I-beam. Detailed, numerical, parametric analyses are conducted.
A typical engineering design with uniformly distributed loads is considered for three cases of
load applied to the top flange, shear centre, and bottom flange. The full-scale test was con-
ducted for a steel flange-tapered I-beam under a point load.

1 INTRODUCTION

Tapered steel columns, beams, and beam columns are increasingly used in contemporary steel
structures because of their optimal cross-section utilisation along the shape of the element.
Thus, effective engineering solutions for the problem of their capacity are needed. The design
of the beams with a tapered web or flanges is more complex than that of prismatic elements.
Thus, analytical expressions for the elastic critical loads are not readily available, and the
choice of the critical section for the application of the buckling resistance formulae is not
straightforward. The problem of determining the critical load, initiating the loss of the stabil-
ity of the beams with linearly variable web or flanges, is a current issue. However, this was
analysed by a small group of specialists.
In 2007, Andrade et al. presented a comprehensive publication on the stability of steel
beams with tapered webs. In their work, they applied the Rayleigh-Ritz method to determine
the critical load of simply supported beams and cantilevers for two types of convergence of
the web for various spans. A comparison of their results with the FEM program ABAQUS
showed good agreement. Asgarian et al. (2012) and Benyamina et al. (2013) published papers
on the stability of tapered beams, limiting their research to the problem of a tapered web. The
paper by Asgarian et al. (2012) is devoted to the issue of the stability of beams with arbitrary
open cross sections. These authors applied the power series method to determine the critical
lateral-torsional buckling of simply supported beams and cantilevers. Recently, Ozbasaran
and Yilmaz (2018) presented work on the optimisation of tapered I-beams due to lateral-
torsional buckling.
However, the disadvantage of the previous research (Kuś 2015, 2017) and other research
(Asgarian et al. 2013, Hakan 2018, Trahair 2014) is that they remain at the level of theoretical
and numerical considerations with various solutions, assumptions, and methodologies of numer-
ical solutions, allowing them to determine the critical forces initiating the loss of stability. Experi-
mental results for lateral-torsional buckling of steel members were presented by Tankova et al.
(2018). In the paper, the authors conducted two full-scale experiments on web-tapered beams.
In the paper, a procedure for calculating the critical buckling moments for flange-tapered
I-beams is demonstrated. Detailed examples are carried out. Typical engineering using uni-
formly distributed design loads is considered for loads applied to the top flange, shear centre,
and the bottom flange. In addition, the results of the experiment are presented and compared
with analytical solutions for simply supported flange-tapered I-beams.

673
2 MODEL OF A BISYMMETRIC FLANGE-TAPERED I-BEAM

2.1 Geometrical properties of tapered cross sections


Geometrical properties of beams with flange-tapered cross sections depend on parameters that
define their shape. In this paper, the βTP parameter was chosen to describe the size change of the
flanges. This parameter can be changed simultaneously in all calculations. The parameters are
described in detail in Figure 1, with the initial cross-sectional characteristics defined as follows:
 
A0 ¼ 2bf 0 tf þ hw0 tw ð1Þ

1
Iz0 ¼ b3f 0 tf ð2Þ
6
h2d0
Iω0 ¼ Iz0 ð3Þ
4

hðxÞ ¼ h0 ð4Þ
 
bðxÞ ¼ bf 0 1 þ βTP Lx ð5Þ

1 bm  bf 0
βIT0 ¼ 2bf 0 t3f þ hw0 t3w ¼ ð6Þ
3 TP bf 0

Assuming that the thickness of the web and flanges are constant along the length of the beam,
the characteristics of any geometrical cross section varying over the length can be expressed as
follows:
  x
AðxÞ ¼ A0 þ 2bf 0 tf βTP þ hw0 tw ð7Þ
L
1  x
IT ðxÞ ¼ IT0 þ 2bf 0 t3f βTP þ hw0 t3w ð8Þ
3 L
 x  x 2  x 3 
Iy ðxÞ ¼ Iy0 1 þ ðβTP Þ : þ ðβTP Þ þðβTP Þ ð9Þ
L L L
 x  x 2   x 3
Iz ðxÞ ¼ Iz0 1 þ 3βTP þ 3βTP
2
þ Iz0 β3TP ð10Þ
L L L

In these equations, A(x) is the cross-section area, IT(x) is the torsion constant of the cross sec-
tion, Iy(x) and Iz(x) are the principal moments of inertia about the y and z axes, hw0 is the
height of the web, tw is the web thickness, and tf is the flange thickness.

Figure 1. Geometrical properties of beams with tapered flanges.

674
It is assumed that the tapering of the beams analysed in this paper is limited to twice
the cross-section size at the mid-span length. When parameter βTP equals zero, the beam
is prismatic with constant values of dimensions h0 and bf0 along the beam (uniform
shape). The maximum tapering takes place when the mid-span cross section is twice as
big as the left and right supports. In this case, it is assumed that the taper parameter
takes a value equal to 1.

2.2 Total potential energy for beams with tapered cross sections
In this paper, the functional for the total potential energy of the beam, commonly available in
the literature (Kuś 2015, Benyamina et al. 2013, Andrade et al. 2010), was used.
In Figure 2, u, v, and w are the displacement components of the shear centre point in the x,
y, and z directions, and θ is the twist angle:

ðL  2  2  2 2 !
1 d2v dθ d θ
PLTB ¼ EIZ ðxÞ þ GIT ðxÞ þ EIω ðxÞ dx
2 dx2 dx dx2
0
ðL        !
1 dhðxÞ 2 dθ 2 dhðxÞ dθ d 2 θ
þ EIZ ðxÞ þ hðxÞ dx ð11Þ
2 dx dx dx dx dx2
0
ðL  2  ðL
1 d v 1  
þ My θ dx þ qz ez ðxÞθ2 dx
2 dx2 2
0 0

3 APPLICATION FOR A FLANGE-TAPERED I-BEAM UNDER A UNIFORMLY


DISTRIBUTED LOAD

The purpose of this chapter is to present the accuracy of the solutions obtained by the pro-
posed method in comparison with the finite element approach and to conduct a parametric
analysis. In accordance with the Ritz method, the system of homogeneous algebraic equations
were obtained by applying the minimisation of the functional of the total potential energy (11)
using the proper conditions. The detailed definition of equations, components, and procedures
for solving square polynomials with the unknowns +/-M can be found in the papers by Kuś
(2015, 2017). Steel beams are assumed to be made of steel with a Young’s modulus of E = 210
GPa and a modulus of Kirchoff G = 81 GPa (i.e., a Poisson’s ratio = 0.3).

Figure 2. I-section under lateral-torsional buckling.

675
Figure 3. Sketch of a simply supported beam with a uniformly distributed load.

The verifying computations were performed using the ANSYS FEM software system.
A rod ‘BEAM188’ finite element was assumed. It is a two-node element with 7 degrees
of freedom – three translational UX, UY, and UZ; three rotary ROTX, ROTY, and
ROTZ; and a seventh optional degree of freedom to restrict the freedom of cross-
sectional warping. The beam supports were modelled by blocking the translational
degrees of freedom UY = 0 and UZ = 0 and by locking the axial rotational degree of
freedom ROTX = 0. The difference between the values of the critical buckling moments
calculated using the Ritz method and those of the ANSYS program were determined
according to the following formula:
 Ritz 
Mcr  Mcr Ansys 

Δ¼   100% ð12Þ
Ansys 
Mcr

A simply supported beam with a uniformly distributed load was adopted (Figure 3).
In this case, the formula for the gradient of moments takes the following form:

 
x x2
mðxÞ ¼ 4  2 ð13Þ
L L

Using the procedure as mentioned in paragraphs two and three, the proper equations were
calculated. In the next step, the critical buckling moments for a beam with tapered flanges
were computed.
Critical buckling moments for beams with tapered flanges evaluated by the presented
method and FE model are in sufficient agreement. Relative differences do not exceed by more
than 20%, on the safety side (critical buckling moments from the FE method have a higher
value than those calculated by the presented method). As could be expected, the highest value
of a critical load is for the beams with the taper parameter βTP.

4 EXPERIMENTAL TEST ON A FLANGE-TAPERED I-BEAM

4.1 Introduction
To check the behaviour of a steel flange-tapered I-beam and the accuracy of the procedure for
calculating the critical buckling loads presented in this paper, a full-scale test was conducted.
An experiment was conducted in the design laboratory at the Faculty of Civil Engineering
and Architecture of Opole University of Technology. In the following chapter, the results
from the beam test under the point load are shown.

4.2 Specimen geometry


Due to the possibilities of the measurement system, the lateral dimensions of the beam have
been selected in such a way that the critical force initiating the loss of stability does not exceed
70% of the theoretical value. This was determined according to the procedure given in the

676
Figure 4. Plot of critical buckling moments for a simply supported beam with tapered flanges a) under
a uniformly distributed load, b) to the top flange (I), c) to the shear centre (II), and d) to the bottom
flange (III).

Figure 5. Sketch of flange-tapered I-beam with geometrical properties of the cross section.

677
Figure 6. Flange-tapered I-beam before the experiment.

article and the FEM analysis in the ANSYS for an element with a span of 5.0 m loaded with
a concentrated force in the middle of the span, applied to the upper flange. The initial analyses
have shown that, for the assumed load system, the cross section of a beam with a height of
200 mm will be the most suitable.
Because of the limited costs, it was assumed that it would be most advantageous to
make a beam from the hot-rolled HEA200 element by maintaining a transverse dimen-
sion in the middle with a nominal value of 200 m and that the initial cross section
would be halved. In this way, the taper coefficient βTP = 1 was obtained. The steel
grade of this steel beam is S235 ( fy,nom = 235 MPa), and the steel quality is JR.

4.3 Loading system


To achieve the global buckling of a laterally unrestrained flange-tapered steel I-beam,
a special testing system was used, as shown in Figure 6. The testing frame consists of two
parts, namely, the support system made from a rigid steel frame and the loading system with
a hydraulic force actuator with equipment. Vertical supports were provided at the beam ends,
as shown in Figure 7. To restrain the beam from lateral displacements but release longitudinal
displacements, a group of steel bars was placed horizontally between the supporting frame
and beam flanges. The beams could develop warping deformations freely, as longitudinal
restraints were only provided at one end. The loading system consisted of a rigid loading
frame. A concentrated point load was applied to the beams through the hydraulic jack, and its
magnitude could be measured by the load cell. The hinge allowed the beam to deflect freely in
the lateral direction. Strain gauges were placed in sections S-1 to S-5 and displacement gauges
were placed in sections S-1, S-4, and S-5, as shown in Figure 7.

Figure 7. Sketch of loading system with gauge locations.

678
4.4 Results
The point load from the hydraulic force actuator was divided between beams in two steps.
The first load step allowed for adjustments all systems (i.e. beams, supports, and steel frame)
and served for the elimination of the initial gaps. This step was performed in the force control
at the maximum value of force equal to 10 kN. The second step was to load the beam until the
critical force was reached, which was carried out in the displacement range. The loading was
stopped when an appreciable drop in the applied force was registered. This effect was achieved
when the maximum point load was 109 kN and abruptly dropped to 93 kN (568.7 seconds of
testing) and 55 kN (568.8 seconds of testing). At the end of the experiment, the tested beam
demonstrated lateral-torsional buckling, as shown in Figure 8.
The applied vertical load increases linearly with the strains of the web and flanges through
almost the entire range of the experiment. Detailed values of strains and force measured
during the experiment in the middle span of the beams are shown in Figure 10.
To compare results from the experiment, the procedure mentioned in the second and third
paragraphs of the paper was used to determine the critical buckling force for the beam under the
point load, with all the necessary assumptions. The value of the critical force obtained using the
Ritz method for the beam used in the experiment was 115 kN. In addition, the FEM analysis of
the tested beam was conducted in the ANSYS 19.2 programme. Moreover, the SHELL181

Figure 8. Displacements of the mid-span for the tested beam with lateral-torsional buckling.

Figure 9. Deformation of the tested beam with lateral-torsional buckling.

679
Figure 10. Diagram of strain values of the tested beam with lateral-torsional buckling.

elements were used to build up the prismatic web and the tapered flanges with dimensions of
elements like those in Section 4.2. The numerical model (Figure 11) was used to simulate the
global buckling behaviour of the beam in the experiment. The critical buckling load obtained
from the FEM analysis was 101 kN. The difference between the values of the critical buckling
loads obtained from the experiment and ANSYS has a satisfactory agreement of around 7%. The
difference in the buckling loads from the Ritz method and FEM analysis in this case is around
5%. The numerical model of the steel flange-tapered I-beam is shown in Figure 11a, and the
deformed beam in the numerical simulation is in Figure 11b. The detailed description and analysis
of the tests, considering the tensile test of steel samples taken from the beam after the lateral-
torsional buckling test, together with the in-depth FEM analysis, are a separate article material
and will be published in the near future.

Figure 11. FEM analysis of flange-tapered I-beam in ANSYS. (a) FEM model. (b) deformations.

680
5 CONCLUSIONS

The results of the research on the lateral-torsional buckling problem of beams with tapered
flanges have been reported. A solution to this problem using the Ritz method has been devel-
oped. It allows us to determine the critical lateral-torsional buckling loads for beams with
tapered cross sections.
Bisymmetrical I-beams with linearly variable flange widths were analysed, considering three
load cases: distributed load applied at the shear centre, at the upper flange, and at the lower
flange. The values of critical buckling loads calculated using the applied implementation of
the Ritz method and the FEM approach gave satisfactory agreement.
Furthermore, one full-scale experimental test on lateral-torsional buckling of a linearly vari-
able flange-tapered steel I-beam under a point load was performed. A load system with the
geometrical properties of a flange-tapered beam was characterised. Strains and displacement
were recorded under the loss of stability in the mid-span of beam. The critical load determined
from the experiment was compared with the critical load calculated by the Ritz method and
FEM analysis, which also showed satisfactory agreement.

REFERENCES

Andrade A., Camotim D., Borges Dinis P. 2007. Lateral-torsional buckling of singly symmetric
web-tapered thin-walled I-beams: 1D model vs. Shell FEA, Computers and Structures. vol. 85,
1343–1359.
Asgarian B., Soltani M., Mohri F. 2012. Lateral-torsional buckling of tapered thin-walled beams with
arbitrary cross-sections”, Thin-walled structures. vol. 62, 96–108.
Benyamina A.B., Meftah S.A., Mohri, F., Daya E.M. 2013. Analytical solutions attempt for lateral tor-
sional buckling of double symmetric web-tapered I-beams, Engineering Structures. vol. 56, 1207–1219.
Kuś J. 2015. Lateral-torsional buckling steel beams with simultaneously tapered flanges and web, Steel
Composite Structures: An International Journal, vol.19, No.4, 897–916.
Kuś J. 2017. Lateral-torsional buckling of steel beams with tapered flanges and web. CE/Papers.
Ozbasaran H., Yilmaz T. 2018. Shape optimization of tapered I-beams with lateral-torsional buckling,
deflection and stress constraints, Journal of Constructional Steel Research, 143, 119–130.
Tankova T., Martins J.P., Da Silva L.S., Marques L., Craveiro H.D., Santiago A. 2018. Experimental
lateral-torsional buckling behaviour of web tapered I-section steel beams. Engineering Structures, vol.
168, 355–370.
Trahair N.S. 2014. Bending and buckling of tapered steel beam structures, Engineering Structures, vol.
59, 229–237.

681
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Composite floor system with cold-formed trussed beams


and pre-fabricated concrete slab

Luiz Alberto A. de S. Leal & Eduardo de M. Batista


COPPE, Civil Engineering Program, Federal University of Rio de Janeiro, Brazil

ABSTRACT: Composite steel and concrete floor systems are usual solution for building con-
struction, conducting to reduce material consumption, improve structural strength and flexural
stiffness. Steel frame light construction systems, based on combination of thin-walled cold-formed
steel members (CFS) and panels (e.g wood, cementitious, gypsum plaster), may also benefit of
composite behavior for the case of concrete slabs. In this context, full-scale experimental tests
were conducted to evaluate the structural behavior of floor system conceived with 0.95-mm thick
CFS trussed beams and pre-fabricated concrete slabs. Innovative solution for shear connectors
was designed and tested, in order to ensure full interaction between the CFS trussed beam and
concrete slab. Thin-Walled Channel connector (TWC) is based on lipped channel CFS attached
to the top chord of the truss by self-drilling screws. Experimental results indicate efficient behavior
of TWC shear connectors with improved bending capacity of the floor system.

1 INTRODUCTION

In recent years, many studies concerned with the structural behavior of composite steel and
concrete floor systems have being reported. In this context, the cold-formed sections (CFS)
represents an excellent alternative, especially because of its automated fabrication process,
lightweight of CFS members, combined with high structural resistance. Although the referred
solution is associated to several benefits, the use of cold-formed members in composite struc-
tures requires more information in terms of steel-concrete shear connection solution, bending
and ductility capacity of the structural system, as well as the collapse mechanisms, as referred
by Hsu et al (2014), Nakamura (2002), Wehbe et al (2013).
The present research performed experimental bending tests of composite trussed beams,
designed with thin-walled lipped channel CFS and pre-fabricated concrete slabs. The obtained
results show reliable information regarding ultimate capacity, as well as the collapse mechan-
ism. The present article is addressed to the results obtained with the Thin-Walled Channel
(TWC) shear connector, composed by 0.95 mm thick lipped channel CFS member, attached
to the top chord of the truss by 4.8 mm self-drilling screws.
The floor system bending test is part of an experimental program aiming at the evaluation
of the structural response of TWC shear connectors. Additional thin-walled shear connector
solutions are under development, including the Thin-Walled Perfobond (TWP), submitted to
preliminary push-out tests, based on Eurocode 4 (2004) reported in Leal and Batista (2017).

2 FULL-SCALE EXPERIMENTAL PROGRAM

2.1 Definition of the composite floor system


The full-scale prototype is composed by two thin-walled CFS trusses, partially pre-cast concrete
slab and innovative TWC shear connectors, conceived to ensure composite behavior between
steel and concrete (Figure 1). All cold-formed members of the trusses were manufactured with

682
Figure 1. Typical details of the composite steel-concrete floor system: (a) top view and (b) cross-section.

0,95 mm steel plates and all connections, including those between the shear connectors and the
top chord, are composed by 4,8 mm diameter self-drilling screws. The chords, and diagonals of
the steel truss were designed with lipped channel C89 × 40 × 12 × 0,95 (in mm) members, com-
monly adopted in steel frame system construction.
Figures 1 and 2 show typical details of the composite floor system, with trussed beams with
375 mm height and 600 mm span between the twin-trussed beams. Pre-cast concrete ribs com-
pose the slab of the prototype, combined with Styrofoam blocks and a portion of 60 mm thick
top concrete (filled in field) above the referred blocks. The total thickness of the concrete slab
is 130 mm and concrete designed with 25 MPa nominal strength.
It is important to emphasize the slab advantageous properties, such as 420 mm spaced pre-
fabricated slab ribs combined with Styrofoam blocks, which enables reduced consumption of
concrete and lower self-weight dead load of the floor system.
Figure 1 also shows the dimensions of the composite floor system prototype, with 7800 mm
total length and 1200 mm width. The clear span of the structure during bending test was
defined with 7580 mm.
Additional details of the composite steel-to-concrete beams are shown in Figure 2, including
the connection between diagonals and chords (reinforced by gusset plates) with self-drilling
screws. The self-drilling screws connection details were conceived and designed to avoid pre-
mature collapse during the bending test.
The TWC shear connectors are composed by three components, enabling adequate shear
transfer mechanism: (a) lipped channel C89 × 40 × 12 × 0,95 mm, (b) single angle
L40 × 40 × 0,95 mm and (c) thin steel plate 92 × 150 × 0,95 mm between the shear connectors
and the top chord of the truss. As presented in Figure 2, these elements are assembled together
with 4.8 mm self-drilling screws.

Figure 2. Details of the composite beams: (a) trussed beams (b) end node of the trusses (c) TWC
connector.

683
Note, in Figure 2, that both lipped channel and single angle are attached to the top chord
of the truss with six self-drilling screws (three screws for each member). Three additional
screws interconnected lipped channel and angle components. It is important to highlight that
the shear connectors were placed over each node of the top chord, resulting in a total of
15 units of TWC.
The additional thin steel plate placed between top chord and shear connector, connected to
the chord by eight screws, provides local reinforcement, improving load capacity and stiffness
of the connection.

2.2 Preparation of the full-scale prototype


The preparation of the full-scale prototype included: (a) fabrication and assembling of the steel
members, (b) formwork activities, (c) positioning of the steel welded mesh above the Styro-
foam blocks and (d) concrete filling. Manufacturing and assembling the steel trusses was pro-
vided by steel frame system specialized company, which cooperates with the research activities.
Figure 3 shows (a) the twin trusses (laterally restrained at both ends, as well as at mid-
section of the trussed system), (b) the arrangement of the slab ribs and (c) the Styrofoam
blocks and the top steel welded mesh. It is also clear in the same figure the presence of TWC
shear connectors, attached to the top chord of the trusses.
Finally, an important aspect refers to the application of grease at the upper face of the top
chord, avoiding bond interaction between steel member and slab filled concrete. In this con-
text, the only horizontal shear transfer mechanism between slab and trusses is associated to
the TWC shear connector.

2.3 Instrumentation of the composite floor system


The instrumentation of the prototype was defined to provide reliable information regarding the
behavior of the composite floor system in terms of ultimate capacity, flexural stiffness, relative
displacements between steel truss and concrete slabs and collapse mechanism. Additionally, the
actual contribution of the steel top chord is a main goal of the experimental planning.
The adopted instrumentation included sets of strain gages at (i) bottom and top steel chords
and (ii) at the top face of the concrete slab. Strain gages were placed next to mid-span section,
inside the pure bending zone, in order to follow development of constant bending moment
distribution. Furthermore, a set of vertical displacement transducers was placed along the
structure, as well as horizontal displacement transducers to enable evaluation of possible slip-
page (relative differential displacement) between steel and concrete.
Figure 4 presents the test set-up, including (i) reaction frame, (ii) 250 kN capacity deform-
ation controlled hydraulic actuator, (iii) three loading distribution beams (main longitudinal

Figure 3. Preparation of the full-scale trussed beams.

684
Figure 4. Test set-up (a) reaction frame, hydraulic actuator and loading distribution beams, (b) vertical
displacement measuring transducers.

stiff H beam and two transversal RHS beams. Loading was transmitted to the floor system
aligned over the steel trusses, with the help of plywood plates placed in between steel distribu-
tion beams and the concrete slab.
Figure 4 shows the arrangement of the experimental test, including the position of the distri-
bution beams (total self-weight of 9,0 kN), 2350 mm extension of pure bending zone and the
simply supported condition at the prototype ends.
The arrangement of strain gauges and displacement transducers are represented in
Figures 5, 6 and 7, respectively. The top slab concrete strain gages were uniformly spaced,
as can be seen in Figure 5. This set of strain measurements are also able to evaluate the
influence of shear-lag effects (non-uniform distribution of stresses along the width of the
prototype).
The strain gauges associated to steel members were placed at the web of the top and
the bottom chord, next to the mid-sections (pure bending zone), as shown in Figure 6.
Finally, the vertical and horizontal displacements transducers distribution is observed in
Figure 7.

Figure 5. Arrangement of strain gages at the top of the concrete slab.

Figure 6. Arrangement of strain gages: (a) at top and bottom chord of the steel truss and (b) at
mid-section of the prototype.

685
Figure 7. Arrangement of the vertical and horizontal displacement transducers.

2.4 Analytical models for composite trussed beams


The bending behavior of the composite floor system was evaluated by four analytical models,
as shown in Figure 8. Models 1 and 2 take the collapse associated to elastic behavior, as con-
firmed by the adopted stress distribution. On the other hand, the Models 3 and 4 assume plas-
tic stress distribution, resulting in higher predicted bending capacity.
It is important to emphasize that (some) traditional analytical models neglect the contribu-
tion of the top chord for the equilibrium of the cross-section (ABNT NBR 8800:2008). In this
context, the predicted ultimate capacity related to Models 1 and 3 take into account internal
equilibrium forces developing only (i) at the top concrete slab thickness (above the Styrofoam
blocks) and (ii) at the bottom chord of the trusses. Models 2 and 4, on the other hand, take
into consideration the top chord contribution in elastic and plastic stress distribution,
respectively.
Table 2 presents the predicted bending moment capacity, as well as estimated hydraulic
actuator force. The self-weight bending moment, also indicated in the table, is related to the
prototype self-weight together with the set-up loading distribution beams, applied over the
structure before starting experimental data acquisition. Figure 4 illustrates the test set-up.

2.5 Experimental results

2.5.1 Bottom and the top steel chords deformation


The relation between applied force and deformations at bottom and the top chords are very
important to provide information regarding collapse mechanisms of the prototype. In add-
ition, it helps to clarify the actual contribution of the top chord in terms of bending capacity
of the composite trussed beams.

Figure 8. Analytical models for the estimation of the floor system bending capacity: (a) system
cross-section, (b) Elastic Model 1, (c) Elastic Model 2, including contribution of the top chord, (c) Plastic
Model 3, (d) Plastic Model 4, including contribution of the top chord.

686
Table 1. Mechanical and geometric properties of the cross-section of the prototype.
Steel Concrete Composite cross-section properties
Young Yield Ult Yield Concrete Secant Modular Moment Elastic Elastic
Mod (Es) Stress Stress Deform (ε) Str (fck) Mod (Ec) Ratio (η) of Inertia Mod Wt Mod Wb
[MPa] [MPa] [MPa] [μstr] [MPa] [MPa] - [cm4] [cm3] [cm3]

189000 370 436 1960 26 24271 7,8 8280 1635 182

Obs: The moment of inertia and elastic modulus were obtained using the Model 2, taking into account elastic
stress distribution and contribution of the top chord.

Table 2. Computed prediction of the bending moment capacity of the floor system and estimated
hydraulic actuator force.
Contribution Depth of Moment Self-W Additional Estimated
Analytical of the top Neutral Axis Capacity Moment Moment Actuator Force
Model chord [cm] [kN.cm] [kN.cm] [kN.cm] [kN]

1 N 4,72 6159 2898 22,2


2 Y 5,07 6750 3489 26,7
3 N 0,50 6468 3261 3207 24,5
4 Y 1,00 8261 5000 38,2

Figure 9a shows the force-deformation response of the bottom chord, with initial linear
behavior up to a load level of 10 kN, approximately. The experimental results showed increas-
ingly non-linear behavior, produced by the material non-linearity associated to concrete slab
and, especially, due to the steel yielding of the bottom chord.
Observing the results plotted in Figure 9, the recorded maximum applied force is 37.2 kN, asso-
ciated to a collapse mode, with tension rupture of the net cross-section of the bottom chord. Fur-
thermore, it is important to highlight that top chord of the trusses developed fully tension stress,
confirming neutral line inside concrete slab and possible contribution of the chord for the bending
capacity of the system. Also, tension normal stress in the chord avoid loss of effectiveness of the
thin-walled lipped channel CFS members by local or distortional buckling.
At collapse, the maximum deformation associated to the top chord is approximately 1309 μst
(see Figure 9(b)), which indicates the referred member developed stress level around 260 MPa,
only for the applied load during the experimental test. The previous deformation produced by
the self-weight of the prototype and the distribution beams is estimated (computed) as 189 μst.
Therefore, the final stress level of the top chord at collapse was 283 MPa.

Figure 9. Force vs deformation associated to the bottom and the top chord of the trusses.

687
Figure 9(a) shows bottom chord maximum deformation at collapse loading was 6609 μst
(5586 μst related to BCA06A and estimated self-weight previous deformation of 1023 μst). In
addition, the referred member reached the yield stress (370 MPa) at a load level of 17 kN, asso-
ciated to a deformation of 937 μst (measured during the experimental test). Figure 10 shows the
collapse mechanisms of the prototype, related to the net rupture of the truss “A”. Similar mode
of collapse was observed by experimental tests performed by Mujagic et al. (2010).
Based on the estimated stresses associated to the chords of the trusses and considering the
presence of 6 (six) effective shear connectors over a single truss (outside pure bending zone),
over the distance of 2735 mm illustrated in Figure 4, it is possible to estimate the forces acting
on each TWC. In this context, assuming a total shear force of 122.2 kN (50.5 kN associated to
the top chord and 71.7 related to the rupture force of the bottom chord), each connector was
subjected to 20.4 kN. The cross-sectional areas, adopted for the referred prediction, were
1.783 cm2 (gross area of the top chord) and 1.646 cm2 (net area of the bottom chord).

2.5.2 Measured deformations at the top face of the concrete slab


The relation between applied force and the measured deformation at the top face of the concrete
slab is presented in Figure 11. The obtained results show the maximum absolute deformations (at
ultimate load), associated to strain gages SL07 VA and SL07IIIB, are 664 μstr and 547 μstr,
respectively. In addition, it is possible to observe that the average absolute value, related to the
positions A and B, are 590 μstr (COV 9,3%) and 506 μstr (COV 7,5%), showing almost uniform
distribution of the compression deformation at top face of the slab and nearly null shear lag effect.
Another important aspect refers to the computed average measured deformation at concrete
slab, at ultimate load, including the contribution of the self-weight of the prototype and the distri-
bution beams. Assuming the total maximum deformation is 700 μstr (590 μstr plus estimated

Figure 10. Net rupture of the bottom chord of the truss “A”.

Figure 11. Experimental deformations at the top face of the concrete.

688
Figure 12. Strain gage deformation distribution at mid-span section of the composite floor system.

previous deformation of 110 μstr, obtained by linear proportion), the estimated stress at the top
face of the concrete is 19.4 MPa, based in the following equation from Eurocode 2, 2004:

σc ¼ fc ½1  ð1  εc =εc2 Þn
ð1Þ

where σc refers to the compressive stress of the concrete; fc is the compressive nominal strength
of the concrete; εc is the concrete strain ; εc2 is the strain associated to the compressive nominal
strength; n is a constant, which can be taken as 2,0 for fck  50MPa.

2.5.3 Cross-section measured deformations at mid-span section


The distribution of measured strain gage deformations at mid-span section, revealed import-
ant aspects, as can be seen in Figure 12. The experimental results confirmed the neutral axis is
located at the concrete, approximately 50 mm below the top face of the slab. In this context,
all the steel members are subjected to tension stresses and almost all concrete above Styrofoam
blocks is compressed. These experimental evidences are well related with Model 2 described in
Figure 8 and included in Table 2.
Additionally, the presence of a single neutral axis suggests full-interaction between the steel
truss and the partially pre-cast concrete slab, which demonstrates the lipped channel TWC
connectors were capable to provide adequate steel-concrete shear transference capacity. The
obtained results of relative displacement between steel trusses and concrete slab, measured by
horizontal displacement transducers, indicate nearly negligible value of 0.56 mm, thus con-
firming full shear interaction between materials of the composite floor system.

3 CONCLUSIONS

The full-scale experimental test of the composite trussed beams revealed important aspects in
terms of bending capacity, as well as the actual contribution of the top chord of the trusses. In
addition, it was possible to observe the innovative lipped channel TWLC shear connectors
(composed by thin-walled members and self-drilling screws) were able to provide full-
interaction, as well as adequate shear capacity to the structural systems.
In terms of ultimate bending capacity, the obtained data is well-related with analytical
Model 4, illustrated in Figure 8 and Table 2, based on the assumption of full plastification of
the composite cross section (top chord, bottom chord and concrete slab). The prediction of
depth of the neutral axis by Model 2 (elastic stress distribution, including top chord contribu-
tion) on the other hand, presented better result.
Additionally, the obtained results showed that the actual contribution of the top chord is sig-
nificant for the evaluation of the composite trusses. As mentioned before, the referred members
were associated to an estimated stress of 283 MPa, which is near from yield stress of the steel.

689
Observing the obtained results, it is important to highlight that the TWC shear connectors,
attached to the top chord of the trusses, were capable to guarantee an adequate shear transfer
mechanism between steel and concrete members. As mentioned before, the estimated force
acting on each connector was 20.4 kN and nothing was observed, during the bending test, in
terms of cracks over the concrete slab, as well as significant slippage between the trusses and
the slab.

ACKNOWLEDGEMENTS

This study was financed in part by the Coordenação de Aperfeiçoamento de Pessoal de Nível
Superior- Brasil (CAPES) - Finance Code 001. Moreover, the authors would like to acknow-
ledge GypSteel Group for the delivery of the steel CFS members.

REFERENCES

ABNT NBR 8800. 2008. Design of steel and composite structures for buildings. Associação Brasileira de
Normas Técnicas, Rio de Janeiro, Brazil (in Portuguese).
CEN, EN 1992-1-1, 2004. Eurocode 2: Design of concrete structures, Part 1-1: General Rules and Rules
for Buildings, Brussels,.
CEN, EN 1994-1-1, 2004. Eurocode 4: Design of Composite Steel and Concrete Structures, Part 1-1:
General Rules and Rules for Buildings, Brussels,.
Hsu, Cheng-Tzu Thomas; Punurai, Sun; Punurai, Wonsiri; Majdi, Yazdan. 2014. New composite beams
having cold-formed steel joists and concrete slab. Engineering Structures, vol. 71, 187–200.
Leal, L.A.A.S; Batista, E.M. 2017. Experimental investigation of new thin-walled perfobond shear
connectors. In: XXXVIII Iberian Latin-American Congress on Computational Methods in Engineering,
CILAMCE. Florianopolis, Brazil.
Mujagic, J.R.U; Easterling, W.S; Murray, T.M. 2010. Design and behavior of light composite
steel-concrete trusses with drilled standoff screw shear connections. Journal of Constructional Steel
Research, vol. 66. 1483–1491.
Nakamura, SHUN-ICHI. 2002. Bending behavior of composite girders with cold-formed steel U section.
Journal of Structural Engineering, vol. 128, 1169–1176.
Wehbe, Nadim; Bahmani, Pouria; Wehbe, Alexander. 2013. Behavior of concrete/cold-formed steel com-
posite beams: experimental development of a novel structural system. International Journal of Concrete
Structures and Materials, vol 7, 51–59.

690
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Experimental investigation of stability behavior of members


supported by sandwich panels at elevated temperature

A. Lendvai & A.L. Joó


Department of Structural Engineering, Budapest University of Technology and Economics, Budapest,
Hungary

ABSTRACT: The paper introduces experimental tests in order to examine the stability
behaviour of members supported by sandwich panels at elevated temperature.
The test specimens were 3.0 m x 3.0 m sized diaphragms, constructed of three HEA120
beams with sandwich panels installed on the outer face. After reaching desired temperature
level, the middle beam was subjected to monotonically increasing axial load until failure.
Total of 18 different test configurations were performed, in which the type of cladding (min-
eral wool or PIR core), thickness of cladding (100, 160 or 230 mm), and temperature (20, 200,
250 and 300°C) were varied.
The experimental research program was executed as part of an RFCS project named STABFI.
The principal aim of the project is to find those variables, which have primary effect on the resist-
ance of the structure, and to get data at different temperature levels for numerical modelling.

1 INTRODUCTION

Research started to investigate the stabilizing effect of sandwich panels at the beginning of
21th century (Hedman-Petursson, 2001, Höglund, 2002, ECCS, 2014). The results in previous
projects (i.e. EASIE, 2013) shown that considerable savings can be achieved for structural
members if diaphragms (i.e. sandwich panels, trapezoidal sheetings) were used for stabilizing
the whole structure, compared to the case when stability is ensured by other means such as
bracing elements. Fire limit state was out of scope of these international research projects.
The aim of this article is to identify the behaviour of steel members stabilized by sandwich
panels or trapezoidal sheeting in normal condition simulating fire. The innovative feature is to
investigate whether similar savings could be achieved in fire due to the stabilizing effect of
sandwich panels and trapezoidal sheeting.
In order to examine this phenomena, an experimental test series was executed at the Buda-
pest University of Technology and Economics, Department of Structural Engineering. Testing
in both, normal temperature and elevated temperature covered stability tests of members in 18
different configurations on four different temperature level (20, 200, 250, 300 C°).
Test results will be used for further numerical analyses, FE model validation and
a subsequent parametric study, which will compose a background for design rules and guid-
ance development, which is the principal aim of the RFCS project, STABFI.

2 TEST PROGRAM

2.1 Test set-up


The main objective of the experimental testing was to determine the stabilizing effect of cladding
panels at ambient and elevated temperatures. Testing was carried out by the Structural Laboratory
of Department of Structural Engineering, Budapest University of Technology and Economics.

691
For stability testing two types of sandwich panels and two types of trapezoidal sheetings
were chosen: 100 and 230 mm thick panels with mineral wool core, and panels with PIR core
with 100 mm and 160 mm thicknesses; sheetings with profile height of 153 mm (t = 0.9 mm)
and with 100 mm profile height (t = 0.88 mm) were applied.
The panels with mineral wool core had inner sheet thickness of 0.5 mm and outer sheet
thickness of 0.6 mm, while PIR panels had inner sheet thickness of 0.4 mm and outer sheet
thickness of 0.5 mm. Steel grade of both sheets was S280GD.
The connections of sandwich panels to the supporting beams was self-tapping stainless steel
screws with diameter 5.5 mm, and shot nails for the trapezoidal sheeting.
Figure 1 shows a typical test specimen. The overall dimensions of test specimens were
3.0 m × 3.0 m sized diaphragms, where the spacing between the three hot rolled beams were
1.5 m. The sandwich panels were installed on the outer face of the supporting members. The
simply-supported main member with the span of 3 m is located at the middle of the specimen.
The section of the beams were hot rolled HEA120 sections (steel grade S355).
On Figure 2. the general scheme of the stability tests can be seen. A loading frame was
erected in order to provide the loading force to the simply supported middle main member.
The two beams on the side of the panels were restrained in vertical direction, so they were able
to move horizontally in the direction of the applied force.
A horizontally unrestrained (movable) column was installed at the end of the loading
frame. Two, symmetrically placed hydraulic jacks were fixed to the loading frame, providing
an increasing load to the movable column, so this column was able to load the panel’s middle
member with an increasing load.

Figure 1. Typical test specimen.

692
Figure 2. General scheme of test specimen and the loading frame.

2.2 Test series


The tests were completed at four temperature levels (ambient and elevated): 20, 200, 250 and
300 °C (in case of PIR panels the maximum temperature level was 250, and in case of mineral
wool panels 300 C°). The following measurements and observations were done during testing:
• Temperatures of main member (two thermocouples per test);
• Cladding temperatures on the exposed and unexposed faces (one thermocouple and ten
temperature controller per test);
• Deflection of main member at midspan (three displacement sensors): one in plane displace-
ment, two out of plane displacements;
• Displacement at the support of main member (one displacement);
• Loading force;
• Collapse of structure.
The test matrix is shown in Table 1, there were 18 tests in total.

3 TEST ARRANGEMENT

3.1 Loading
Axial load was applied by the horizontally unrestrained, movable column to the middle
member.
Load-controlled testing machine was used for the experiment with the loading rate of cca. 5
kN/sec. The load was increased monotonically until failure, the tests were stopped at stability
failure of structural member at room temperature and at elevated temperatures as well.

693
Table 1. Test series.
Cladding thickness Temperature
Cladding type (mm) (C°)

Sandwich panel with PIR core 100 20


200
250
Sandwich panel with PIR core 160 20
200
250
Trapezoidal sheeting 100 20
(H = 100 mm, t = 0.88 mm) 200
250
Sandwich panel with mineral wool core 100 20
200
300
Sandwich panel with mineral wool core 230 20
200
300
Trapezoidal sheeting 153 20
(H = 153 mm, t = 0.9 mm) 200
300

The tests were completed as steady state meaning that temperature is kept same with
thermo-pads, and after reaching the desired temperature level the axial load was increased
until failure.

3.2 Boundary conditions


The loading column was installed to the loading frame with connecting bars. The loading
column was unrestrained in horizontal direction. Two hydraulic jacks were fixed to the load-
ing frame to provide sufficient loading force to the loading column, and pushing the loading
column toward test specimen’s middle beam.
The simply supported middle beam of the diaphragm was pinned. The side beams were
restrained in vertical direction with a hinged connection to the upper beams of the loading
frame, in order to stabilize the tested specimen in vertical direction, see Figure 2.

3.3 Heating
The specimens were tested at different temperatures of the inner steel sheet (20, 200, 250,
300 °C), as it is shown in test matrix in Table 1.
Two Weldotherm’s Standard Europa heating machines were used with a system of ceramic
heating pads to heat the inner side of sandwich panels and the middle beam.
The steel members were considered to be fire protected, therefore for a uniform temperature
distribution in a cross-section, the temperature increase Dθa;t of an insulated steel member
during a time interval was calculated using (1) given in EN 1993.1.2 and considering the mater-
ial properties of intumescent coating protection for 1 hour and the steel profile properties.
0 1
λp=  f
d A B C 
Dθa;t ¼ ca ρ p Vp @ 1 A θg;t  θa;t Dt  e10  1 ð1Þ
a f

3

The calculated temperatures of the protected steel HEA120 beam and the temperatures of the
heated steel sheet of the sandwich panel are given in Table 2.

694
Table 2. Heating temperature of supporting member.
Temperature of steel sheet Tempereature of fire protected HEA120
(C°) (C°)

20 20
200 65
250 105
300 153

Figure 3. Measured displacements and temepratures on the tested specimen.

The heating pads were placed on the sandwich panels covering 60-80 cm on both sides of
the loaded middle member, and in vertical position on alternal sides of the web of the middle
beam, as well as along the middle beam’s upper flange. The ceramic pads were placed to the
surface of the inner steel sheet, then the top of heating pads was covered by insulation. The
specimens were heated to the designed temperatures by the heating pads at a rate of 16.67
2103 per minute, then the temperature was kept constant until failure.

3.4 Measurements
During the tests displacement of the tested member, the loading force, temperature of the
inner sheet, outer sheet and the middle beam were recorded. The overall behaviour of the
sandwich panel was also monitored visually. Temperature of the specimen during heating was
recorded by coated thermocouples of type K, named as TC1, TC2, and TC3 on 3 points at the
middle member’s midspan. Another ten more temperature controllers were used to control the
temperature of the ceramic heating pads, which were attached to the steel surface by self-
drilling screws, pressing the end of thermocouples to the steel surface tightly. Displacement
was measured in 4 positions with the help of displacement transducers (see Figure 3): axial
displacement at the support (P2), in plane displacement of the middle member’s upper flange
at midspan (P1H), and out-of-plane displacement of the middle member’s upper flange at
midspan in two points (P1V1 and P1V2).

4 TEST RESULTS AND MAIN OBSERVATIONS

4.1 Typical failure modes


In this chapter the observed behaviour of sandwich panels, structural members and screws are
presented and discussed, and the main findings related to the measured force-displacement
diagrams and ultimate load of different configurations are concluded.

695
Figure 4. Typical failure mode of supporting member.

Figure 5. Typical failure modes of sheeting and screws.

The stability failure mode of the supporting member in all tests were flexural buckling
around weak axis, see Figure 4.
Beside this local buckling of sheeting around screw head at supported member was experi-
enced, as well as bending of screws in each cases, and delamination of sheeting on elevated
temperature see Figure 5.
During evaluation of test results the ultimate compressive strength of loaded members were
established, these results are summarized in Table 3. In the last column of the table decrease
of ultimate force at elevated temperature is also indicated, in comparison with the results of
configurations at ambient temperature.
In all cases load-displacement diagrams were produced in order to compare experimental
results for same configurations on different temperature levels, see Figure 6-7.

696
Table 3. Ultimate force in different configurations.
Cladding Ultimate Decrease of
Thickness Temperature force ultimate force
Cladding type (mm) (C°) (kN) (%)

Sandwich panel with PIR core 100 20 1060 100


200 673 63
250 652 62
Sandwich panel with PIR core 160 20 1001 100
200 530 53
250 655 65
Trapezoidal sheeting 100 20 856 100
(H = 100 mm, t = 0.88 mm) 200 597 70
250 902 105
Sandwich panel with mineral wool core 100 20 984 100
200 762 77
300 754 77
Sandwich panel with mineral wool core 230 20 1057 100
200 663 63
300 763 72
Trapezoidal sheeting 153 20 850 100
(H=153 mm, t=0.9 mm) 200 622 73
300 430 50

Figure 6. Load-displacement diagram of configurations with PIR sandwich panels.

5 CONCLUSIONS

The main results and observations of stability tests are introduced in this article. From the
experimental results the ultimate compressive resistance of beams restrained by sandwich
panels can be compared on ambient and elevated temperatures. Experimental results showing,
that the decrease in ultimate compressive force at temperature level 200 C° is 27-28% in case
of PIR panels, and 13-27% in case of mineral wool panels. Further decrease can be observed
at higher temperature level of 250 C°, which is between 30-49%.

697
Figure 7. Load-displacement diagram of configurations with mineral wool sandwich panels.

These results will be used as an input data for numerical simulation, FE model validation,
verification and parametric study. The aim of the research is to develop a design guidance.

ACKNOWLEDGEMENTS

This research has been financially supported by the european union, in research fund of coal
and steel (rfcs) program. the project number is 751583 and the name is steel cladding system
for stabilization of steel buildings in fire (stabfi). the authors gratefully acknowledge for the
support.

REFERENCES

EASIE, 2008–2013, Acronym Description: Ensuring Advancement in Sandwich Construction through


Innovation and Exploitation. Programme Acronym: FP7-NMP2-SE-2008. Grant agreement No
213302.
ECCS Publication No. 88, 1995, European recommendations for the application of metal sheeting acting
as a diaphragm.
EN 191993-1-2, 2005, Eurocode 3: Design of steel structures. Part 1-2: General rules - Structural fire
design.
European Recommendations on the Stabilisation of Steel Structures by Sandwich Panels, ECCS, 2014.
Hedman-Petursson, E., 2001, Column Buckling with Restraint from Sandwich Wall Elements. PhD
Thesis, Department of Civil and Mining Engineering, Division of Steel Structures, Luleå University of
Technology.
Höglund, T., 2002, Stabilisation by stressed skin diaphragm action, The Swedish Institute of Steel Con-
struction, Publication 174.

698
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Ultimate shear resistance of cylindrically curved steel panels

F. Ljubinković, J.P. Martins, H. Gervásio & L.S. Silva


ISISE, Civil Engineering Department, University of Coimbra, Portugal

ABSTRACT: In this paper, a comprehensive numerical study is performed with the aim to
investigate the buckling and the post-buckling behavior of simply supported cylindrically
curved steel panels subjected to pure shear. The main objective is to understand the influence
of geometrical parameters, such as curvature and aspect ratio, on the elastic critical load and
the ultimate shear resistance. Moreover, a new set of formulae are proposed, which enable
simple and accurate prediction of the shear buckling coefficient and the ultimate shear reduc-
tion factor for curved panels with curvature parameter and aspect ratio within the ranges of
practical interest. The proposed formulae have a form similar to the one available in EN
1993-1-5 (2006) for the prediction of the ultimate shear load of a flat panel.

1 INTRODUCTION

Thin cylindrically curved steel panels have long found their use in various branches of engineer-
ing such as Civil, Aerospace, Aeronautical, and Naval Engineering. In recent years, the use of
curved steel panels in bridge applications has become a very attractive trend, both for aesthetic
and structural reasons. However, one of the main challenges for the widespread use of curved
panels in engineering practice is the lack of adequate design guidance. Namely, the current
European design standards for bridges, EN 1993-1-5 (2006) and EN 1993-1-6 (2007), are
restricted to flat plates and cylinders and many studies (Gerard & Becker 1957, Tran et al. 2012
and Martins et al. 2013) have shown that the pre-buckling and the post-buckling behaviors of
curved panels is quite unique and different from these two extreme cases.
In the case of webs of horizontally curved bridges or transversally curved flanges of box-girders
subjected to horizontal external loads, shear stresses may reach significant values, leading to shear
buckling instability modes. Therefore, curved panels need to be adequately designed to sustain
shear stresses. A generic curved panel subject to shear is presented in Figure 1.
Commonly, the panel’s geometry is described by two non-dimensional parameters, curva-
ture Z and aspect ratio α, defined by Equation 1 and Equation 2, respectively as:

b2
Z¼ ð1Þ
Rt
a
α¼ ð2Þ
b

where a and b are the length and width of the panel, t is the thickness, R is the radius of the
panel, and finally, β is the sectorial angle of the panel (Figure 1).
The knowledge on the post-buckling behavior of the cylindrically curved panels subjected
to shear is still scarce. Only recently, there has been an increased number of numerical studies
(Featherston & Ruiz 1998, Featherston 2003, Domb & Leigh 2002 and Amani et al. 2011), in
which the non-linear post-buckling behavior of curved panels was assessed. Although all the
mentioned studies contributed to knowledge on the post-critical behavior of curved panels, no
clear guidance for the prediction of the ultimate shear load was developed, at least not for any

699
Figure 1. The geometry of a cylindrically curved panel subject to shear.

desired aspect ratio and curvature parameter. Namely, Domb & Leigh (2002) proposed
a modified expression for calculation of the shear buckling coefficient, kτ, accounting for the
initial imperfections, however, the expression is applicable only for a small range of param-
eters, i.e. for curved panels with an aspect ratio higher than α ≥ 4. Finally, Machaly et al.
(2010) also proposed expression for the calculation of the ultimate shear resistance, based on
a study in which horizontally curved I-girder web panels were investigated. However, both the
curvature parameter and the aspect ratio were kept within narrow ranges of parameters (i.e.
Z = 2 – 10 and α = 0.5 – 1.0). Finally, a new design guidance DNVGL-CG-0128 (2015) for
shells and curved panels was published for the design of the steel ship industry structures. In
a recent study by Martins et al. (2018), the semi-empirical formulae provided in these two
standards were evaluated and compared with the FEM results and the authors showed that
for the lower and intermediate values of curvatures, the results deviate significantly from the
numerical ones.
Hence, this study aims at developing a simple and reliable methodology for the computa-
tion of the ultimate shear strength of unstiffened curved panels. Moreover, the goal is to
extend the existing rules for flat plates from EN 1993-1-5 (2006) to a wide range of the curva-
ture parameter (Z ≤ 50) and aspect ratios (0.5 < α ≤ 5), addressing most practical cases of the
bridge, offshore and aeronautical applications.

2 NUMERICAL PARAMETRIC STUDY

For the numerical simulations, the FEM software package ABAQUS (2014) is used. Firstly,
the elastic buckling analysis (LBA) is performed for each model, in order to derive the eigen-
modes, which are then used as the shapes of the initial geometrical imperfections in
a geometrically and materially non-linear analysis with imperfections (GMNIA). For that
purpose, the arc-length Riks’ method was used (Riks 1979).
A linear four-node shell element with reduced integration (S4R) was used, which is com-
monly suggested for thin plated structures (Amani et al. 2011, Martins et al. 2013, Manco
et al. 2018, etc.). Based on a mesh convergence study, it was concluded that 80 FE along with
the curved edge (edge b) were sufficient to achieve numerical convergence.

2.1 Definition of geometry and material properties


As several authors showed (Tran et al. 2012, Martins et al. 2013 and Manco et al. 2018) for
the LBA analysis, the curvature parameter Z (see Equation 1) is a suitable parameter for the
definition of the panel’s geometry. This means that the width b and the thickness t may be
kept constant, where the desired Z is obtained by variation of the radius R. However when
assessing the ultimate load of a panel, additionally, the non-dimensional slenderness

700
parameter λw needs to be specified since it depends on the b/t ratio of the panel. According to
EN 1993-1-5 (2006), the slenderness parameter is given by Equation 3 as:
sffiffiffiffiffi
λw ¼ 0:76 fy b=t
¼ pffiffiffiffiffi ð3Þ
τcr 37:4ε kτ

where τcr is the elastic buckling load and ε non-dimensional parameter defined in EN 1993-1-5
(2006) as (235/fy)1/2.
The ranges of geometrical parameters varied in this study are shown in Table 1,
whereas the thickness is kept constant and equal to t = 10 mm. The material used in the
models is steel grade S355 JR, with E = 210 GPa and ν = 0.3, whereas the nominal mech-
anical properties (i.e. yield strength fy and ultimate strength fu) are adopted in accordance
with EN10025 (2004).
Finally, regarding the initial imperfections, the equivalent geometric imperfections were
adopted, with a shape affine to the first buckling mode and an amplitude equal to b/200, as
recommended in EN 1993-1-5 (2006).

2.2 Load and boundary conditions


To define the loads and boundary conditions, a cylindrical coordinate system with the origin
at the point A was adopted, with a longitudinal, tangential and radial axis oriented in x, y,
and z-direction. The displacements u, v and w correspond to x, y, and z-axis, respectively, as
presented in Figure 2.
The load is applied along the middle surface of the panels as a shell edge uniform load,
along the edges y = ± b/2, x = 0 and x = a. As for the boundary conditions, only simply sup-
ported panels are studied in this paper since the principal goal was to propose a formula that
returns lower bound of the shear resistance, as it is commonly done in the design practice,
namely in EN 1993-1-5 (2006).
To ensure the simply supported boundary condition, the following restraints were introduced
in the shell model (see Figure 2): all edges restrained in the z-direction (w = 0); points A, B,
C and D restrained in the x-direction (u = 0) and point A restrained in the y-direction (v = 0).

Table 1. Ranges of geometrical parameters.


Slenderness parameter ( λw ) Curvature (Z) Aspect ratio (α)

0.3 to 3.0 0 to 50 0.5 to 5


(step 0.1) (step 10) (step 0.5)

Figure 2. Definition of boundary conditions.

701
Figure 3. Comparison of numerical results: a) Ljubinkovic et al. (2019b) and b) Amani et al. (2011).

2.3 Verification of the model


Due to the lack of the experimental evidence and a general analytical expression that allows
the calculation of the ultimate shear resistance of the curved panels, it is decided to validate
the numerical model against the numerical results from Amani et al. (2011). Therefore, in
Figure 3, the force-displacement curves are plotted for a curved panel (Z = 15 and α = 1.0),
pffiffiffi (w0). In Figure 3, the applied shear load τ is
with various levels of the imperfection amplitudes
normalized with the yield shear load τu ¼ fy = 3, whereas the radial displacement (w) meas-
ured in the middle point of the panel is normalized with the thickness of the panel (t).
It may be observed that the behavior of panels corresponds to the one obtained by Amani
et al., although the boundary conditions were defined in a different way in these two studies.
The ultimate resistance obtained in this paper is slightly lower than the one obtained by Amani
et al., however, this might be explained by the differences in the density of the mesh. Therefore,
it may be concluded that the numerical model used in this study is validated to some extent.

3 DISCUSSION OF NUMERICAL RESULTS

3.1 Influence of geometry


Several studies (Ljubinkovic et al. 2019a, b, Featherston 2003 and Amani et al. 2011) showed
that the pre-buckling and the post-buckling behavior of cylindrically curved panels is a highly
complex problem, being a function of both the aspect ratio and the curvature parameter. The
influence of the aspect ratio on the shear buckling load is illustrated in Figure 4a, where the
shear buckling coefficient kτ is plotted against the aspect ratio for various values of the

Figure 4. Influence of the aspect ratio α on: a) kτ and b) χw (Ljubinkovic et al. 2019a, b).

702
curvature parameter. The influence of the aspect ratio on the ultimate shear resistance is pre-
sented in Figure 4b, where the shear reduction factor χw is plotted against the b/t ratio for
various values of aspect ratio, with the constant curvature (Z = 20). This was done in order to
isolate only the influence of the aspect ratio, where the area (A = bt) of the cross-section and
the curvature parameter were kept constant.
It may be observed that the higher the aspect ratio the lower are the buckling coefficient
and the reduction factor, regardless of the b/t ratio and the curvature parameter. Furthermore,
it may be noticed that the influence of the aspect ratio is significant for values up to α ≤ 3,
after which it becomes negligible, as it was shown also in NASA (1947) and proven later by
Ljubinkovic et al. (2019a).
Regarding the influence of the curvature parameter on the critical shear load and ultimate
resistance, several previously mentioned authors showed that the higher curvature yields
higher critical load, whereas the ultimate resistance decreases. An example is presented in
Figure 5, where the aspect ratio (α = 1) and the b/t = 300 were kept constant in order to show
how the curvature affects the pre-bucking and the post-buckling behavior of the curved
panels. In this graph, the applied shear load τ is normalized with the yield shear load, whereas
the in-plane displacement v measured in the point C (Figure 2) is normalized with the thick-
ness of the panel t.
As it may be noticed, the curvature increases the second moment of inertia, which leads to
a higher critical load; however, the curvature also leads to a very unstable “shell-like” behav-
ior, resulting in a reduced post-critical strength reserve.

3.2 Influence of the initial imperfections


Curved panels, having a “shell-like behavior” may show a strong sensitivity to the initial
imperfections, as proven by numerous authors (Martins et al. 2018, Ljubinkovic et al. 2019a,
Amani et al. 2011). Therefore, an imperfection sensitivity analysis was performed, where in
particular, the influence of the amplitude and the sign of the geometrical imperfections were
investigated. As for the shape of the imperfection, based on the previous studies by Feather-
ston (2003) and Amani et al. (2011), it was concluded that in the most of the cases the shape
affine to the first buckling mode leads to the lowest ultimate resistance. Therefore, in this
study, the shape of the first buckling mode was adopted.
In the parametric study, the amplitude of the imperfections was adopted to be equal to
a minimum between a/200 and b/200, as suggested in EN 1993-1-5 (2006). However, in case of
both flat and curved panels, Rusch & Lindner (2001), Tran et al. (2012) and Martins et al. (2014)
showed that these amplitudes may lead to conservative results. Rusch & Lindner (2001) showed
that the amplitude equal to b/420 is suitable for the recalculation of the Winter curve for plates
under pure compression. Hence, a comparison for these two amplitudes (b/200 and b/420) is pre-
sented in Figure 6a, for aspect ratio α = 2 and two curvature parameters, in order to understand

Figure 5. Influence of curvature Z on the behavior of a curved panel (Ljubinkovic et al. 2019b).

703
Figure 6. Influence of the initial imperfections: a) amplitude and b) sign (Ljubinkovic et al. 2019b).

how two various values of the amplitudes affect the ultimate shear resistance. Moreover, the
buckling (χw − λw ) curve provided in the EN 1993-1-5 (2006) is plotted for the sake of
comparison.
It may be observed from the figure that the difference between two amplitudes is not signifi-
cant, especially in the case of the flat panels, due to the presence of the tensile component that
reduces sensitivity to imperfections. A larger difference occurs only in a narrow range of the
slenderness parameter (0.5 < λw < 1.2), due to imperfection sensitivity of these panels since the
ultimate resistance is governed by the yield strength of the material. For higher slenderness,
the difference is negligible regardless of the curvature parameter. Moreover, a quite good
agreement is achieved between the curve proposed by EN 1993-1-5 (2006) and the numerical
results when the amplitude is equal to b/420, but also for the value equal to b/200. Finally, due
to the lack of a more precise database for the imperfection of the curved panels subjected to
a shear load, the recommended value (i.e. b/200) was chosen in this study and the formulae
are derived based on the results with this amplitude.
Apart from the amplitude, the sign of the imperfections (i.e. its direction) is also investigated,
where the “positive” (w0+) and the “negative” (w0−) imperfections are distinguished as shown in
Figure 7. Namely, the imperfection shape in case of the panels under shear may be antisymmetric
or symmetric. In the case of the antisymmetric shape, the ultimate resistance is insensitive to the
direction of the imperfections. However, in case of the curved panels with the symmetric shape of
the imperfection, the direction in which the buckling waves propagate may affect the behavior of
the panels and, thus, the ultimate resistance.
In Figure 6b, using as an example a panel with an aspect ratio equal to α = 2, a comparison
is made between these two signs of the imperfections, where the amplitude is kept constant
and equal to the previously adopted value of b/200. Based on this imperfection sensitivity ana-
lysis, it was concluded that the “positive” sign of imperfection (i.e. radially inwards), always
results in lower ultimate resistance, regardless of the curvature parameter. Therefore, the for-
mulae for the ultimate resistance are proposed considering only the “positive” sign of the
imperfections, being a less favorable case.

Figure 7. Signs of the imperfections: a) positive and b) negative (Ljubinkovic et al. 2019b).

704
4 PROPOSAL OF EXPRESSIONS

4.1 Expressions for kτ


Based on the parametric study, the authors in Ljubinkovic et al. (2019a) proposed new formu-
lae for the calculation of the shear buckling coefficient, extending the scope of the European
standard to cylindrically curved panels. The formula, like the one for the flat panel, in general
form may be written by Equation 4
 2
1
k τ ¼ C1 þ C 2 ð4Þ
α

where the parameters C1 and C2, which are functions of the Z parameter, may be calculated
using Table 2, for various aspect ratios.
The formula is applicable for the curvature parameters in the range 1 < Z ≤ 100, whereas for
Z ≤ 1, the coefficients provided in EN 1993-1-5 (2006) may be used (i.e. C1 = 4.0 and C2 = 5.34
for α ≤ 1, whereas C1 = 5.34 and C2 = 4.0 for α ≤ 1).

4.2 Expressions for χw


According to the methodology provided in the European standard for flat panels, the
ultimate shear resistance VRd of a panel with thickness t and width b is obtained using
Equation 5
pffiffiffi
VRd ¼ χw btfy = 3 ð5Þ

where the reduction factor χw is calculated as a function of the slenderness parameter λw
(see Equation 3). For flat plates (Z = 0), only one buckling curve is sufficient to define
λw − χw relation. However, in the case of curved panels, as shown by Ljubinkovic et al.
(2019b), it is necessary to define the buckling curve as a function of the curvature par-
ameter. Therefore, the authors proposed a formula, which in general form is written by
Equation 6 as:
8
> 1:0 ; λw  λw;0 9>
>
<  2 >
=
 ; λw;0 5 λw 51:1
χw ¼ Aλw þ Bλw þ C ð6Þ
> A >
>
: ; 1:1  λw 53:0 > ;
B þ λw

where λw;0 is the length of the initial plateau, i.e. for which the reduction factor is equal
to χw = 1.0, given by Equation 7 and A, B and C should be taken from Table 3.
 
λw;0 ¼ 0:4 Z  30
ð7Þ
0:3 Z 430

Table 2. C1 and C2 parameters.


α≤1 α>1

C1 = 0.214Z + 2.88 C1 = 0.096Z + 5.15


C2 = 5.343 – Z/175.6 C2 = 0.135Z + 3.18

705
Table 3. A, B and C parameters.
λw;0 5 λw 51:1 1:1  λw 5 3:0

A = -Z/57.7 – 0.48 A = (335.5 - Z)/380


B = Z/86 + 0.25 B = (3.7 + Z)/47.5
C = -Z/367 + 1.0

5 CONCLUSIONS

The paper gives a description of the ultimate shear resistance of simply supported cylindrically
curved panels, where the influence of several parameters was numerically investigated, such as
curvature parameter (from Z = 0 to Z = 50), aspect ratio (from α = 0.5 to α = 5) and initial
imperfections. In particular, it was shown that: (1) the post-critical behavior depends on both
geometrical parameters (α and Z) and not only one of them. Nevertheless, the higher the
aspect ratio and the curvature parameter, the lower is the ultimate resistance; (2) the curved
panels in the intermediate range of slenderness parameter (0.5 < λw < 1.2) show particular
sensitivity to the imperfections. Outside of this range, there is no significant difference in the
ultimate resistance between two considered amplitudes of the imperfections (b/200 and b/420);
(3) the “positive” sign of imperfection (i.e. radially inwards), always results in a lower ultimate
resistance, regardless of the curvature parameter; (4) using the same χw − λw approach for flat
plates from EN 1993-1-5 (2006), the expressions were proposed for prediction of the shear
reduction factor χw for curved panels; (5) Additionally, the expressions for the shear buckling
coefficient kτ are proposed, which may easily replace the impractical and outdated NASA
charts.

ACKNOWLEDGMENTS

Financial support from the Portuguese Ministry of Science, Technology and Higher Educa-
tion (Ministério da Ciência, Tecnologia e Ensino Superior) under the project contract Grant
PTDC/ECM-EST/1494/2014 is gratefully acknowledged. This work was also financed by
FEDER funds through the Competitivity Factors Operational Programme - COMPETE and
by national funds through FCT – Foundation for Science and Technology.

REFERENCES

ABAQUS FEA, D. S. Simulia. 2014. Dassault Systems, Version 6.14.


Amani, M. Edlund, B. L. O. and Alinia, M. M. 2011. Buckling and post-buckling behavior of unstiffened
curved plates under uniform shear. Thin-walled Structures 49 (8), 1017–1031.
Batdorf, S. B. 1947. A simplified method of elastic-stability analysis for thin cylindrical shells I – Don-
nell’s Equation. NACA Technical Report No. 1341.
CEN (2006). EN1993-1-5:2006 - Eurocode 3: Design of steel structures, Part 1–5: Plated Structural Elem-
ents, European Committee for Standardization, Brussels.
CEN (2007). EN1993-1-6:2007 - Eurocode 3: Design of steel structures, Part 1–6: Strength and Stability
Shells Structures, European Committee for Standardization, Brussels.
DNVGL (2015). DNVGL-CG-0128. Buckling Class Guideline. Norway, DNVGL.
Domb, M. M. & Leigh, R.L. 2002. Refined design curves for shear buckling of curved panels using non-
linear finite element analysis. In: 43rd AIAA/ASME/AHS/ASC Structures, Structural Dynamics and
Materials Conference, Denver, U.S.A, Paper #2002–1257.
Featherston, C. A. & Ruiz, C. 1998. Buckling of curved panels under combined shear and compression,
Proc. Intern. Mech. Eng., Part C: Journal of Mechanical Engineering Science, 212 (3), 183–196.
Featherston, C. A. 2003. Imperfection sensitivity of curved panels under combined compression and
shear. International Journal of Non-Linear Mechanics, 38, 225–238.

706
Gerard, G. and Becker, H. 1957. Handbook of Structural Stability: Part III – Buckling of Curved Plates
and Shells. NACA Technical Report No. 3783.
Ljubinkovic, F., Martins, J.P., Simões da Silva, L. 2019a. Eigenvalue analysis of cylindrically curved
steel panels under pure shear. Thin-walled Structures. (in print 2019).
Ljubinkovic, F., Martins, J.P., Simões da Silva, L. 2019b. Ultimate load of cylindrically curved steel
panels under pure shear stress. Thin-walled Structures. (in print 2019).
Machaly, E.B., Safar, S.S., and Abdel-Aal, E.A. 2010. Shear strength of horizontally curved plate girder web
panels with transverse stiffeners. Journal of engineering and applied science, Vol. 57, No. 4, pp. 257–276.
Manco, T., Martins, J.P., Rigueiro C. and Simões da Silva, L. 2018. Semi-analytical model for the predic-
tion of the post-buckling behavior of unstiffened cylindrically curved steel panels under uniaxial
compression. Marine Structures 59 387–400.
Martins, J. P., Simões da Silva, L. and Reis, A. 2013. Eigenvalue analysis of cylindrically curved panels under
compressive stresses – Extension of rules from EN1993-1-5. Thin-Walled Structures, 68, pp. 183–194.
Martins J, Simões da Silva L, Reis A. 2014. Ultimate load of cylindrically curved panels under in-plane
compression and bending – extension of rules from EN 1993-1-5.Thin-Walled Struct;77:36–47.
Martins, J.P., Ljubinkovic, F., Simões da Silva, L., Gervásio, H. 2018. The behavior of thin-walled cylin-
drically curved steel plates under generalized in-plane stresses. A review. Journal of Constructional
Steel Research 140:191–207.
Riks E. 1979. An incremental approach to the solution to the solution of buckling and snapping
problems. Int. J. Solids Struct, 15:524–551.
Rusch A. and Lindner J. 2001. Tragfähigkeit von beulgefährdeten Querschnittselementen unter Berück-
sichtigung von Imperfektionen (in German), Stahlbau, Vol. 70, No. 10, pp. 765.
Tran, K., Davaine, L., Douthe, C. and Sab, K. 2012. Stability of curved panels under uniform axial
compression. Journal of Constructional Steel Research, Vol.69, No 1, pp. 30–38.

707
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Nonlinear finite element analysis of delta hollow flange girders


subjected to patch loading

Nelson Loaiza & Carlos Graciano


Departamento de Ingeniería Civil, Universidad Nacional de Colombia, Medellín, Colombia

Euro Casanova
Departamento Ingeniería Civil y Ambiental, Universidad del Bío-Bío, Concepción, Chile

ABSTRACT: Recent investigations related to plate girders design have demonstrated that is
possible to increase the flexural and shear resistance by using delta hollow flange girders. This type
of reinforcement also increases the torsional rigidity of the plate girder reducing lateral deform-
ation. However, the impact of delta hollow flange girders subjected to patch loading has received
little attention. Therefore, this paper aims at studying the postbuckling behaviour and comparing
the patch loading resistance of longitudinally stiffened and delta hollow flange girders, evaluating
the influence of different geometrical parameters such as the panel aspect ratio, the patch loading
length and the position of the longitudinal stiffener. To this purpose, a nonlinear finite element
model was developed considering geometrical and material imperfections. The computed ultimate
resistances are validated against experimental results found in the literature. Based on the results,
design recommendations are presented for hollow flange girders under patch loading.

1 INTRODUCTION

Delta hollow flange girders (DHFG) are used in bridge construction to increase shear resistance
and prevent lateral torsional buckling of I-section girders. This type of arrangement also
enhances the eccentric load carrying capacity by reinforcing the flange of the girder. Numerous
investigations have been devoted to study the buckling behavior and the structural response of
DHFG subjected to bending moment and shear forces. Avery & Mahendra (1996) carried out
a numerical research to study the effect of transverse stiffeners to control the lateral buckling of
simply supported DHFG under uniform bending moment.
Later on, using nonlinear finite element modeling, numerical analyses were conducted to
evaluate the lateral buckling behavior and ultimate flexural resistance of DHFG (Pi &Trahair
1997, Avery et al. 2000, Dong & Sause 2009). Recently, El Masri & Lui (2018) performed
a numerical and analytical investigation to evaluate the effect of the cross section properties
on the lateral torsional buckling of plate girder with only one delta hollow flange. From their
results some closed-form solutions for cross section properties of DHFG are derived.
Arabzadeh and Varmazyari (2009) conducting nonlinear finite element analysis presented
a parametric evaluation to compare the resistance of DHFG and longitudinally stiffened gir-
ders subjected to eccentric patch loading. The authors concluded that for eccentric patch load-
ing the resistance of the plate girder is considerable increased when a delta hollow flange is
used. Moreover, a similar type of arrangement was used by Navarro-Manso and collaborators
(2014, 2015), in this case a triangular cell along the loaded flange of the girder is employed to
increase the patch loading resistance in slender and high depth steel panels, the results showed
that web stresses are reduced a 30% when this type of reinforcement is used. However, there is
little information about the behavior of DHFG subjected to centric patch loading and the
influence on other geometrical parameters such as the panel aspect ratio, the patch loading
length and the position and size of the delta hollow flange.

708
Therefore, this paper aims at studying the effect of DHFG subjected to patch loading. To
this purpose, a nonlinear finite element model is developed. Next, a parametric study is carried
out to evaluate the effect of the panel aspect ratio, and position and cross section configur-
ations of the DHFG on the patch loading resistance. Subsequently, the results are compared
against those obtained using a longitudinal flat stiffener. Finally, some design recommenda-
tions for DHFG are presented.

2 FINITE ELEMENT MODEL

2.1 Geometric and material properties


To investigate the influence of DHFG subjected to patch loading, a nonlinear finite element
model was developed using the software ANSYS (2018). It must be mentioned that in this
model only a delta hollow flange is used in the loaded flange. Figure 1 shows the nomencla-
ture employed to describe the geometrical dimensions of the studied plate girders, the type
of delta hollow flange, and the longitudinal flat stiffener. The components of the girder were
modelled using shell elements S181, a four-node element with six degrees of freedom at
each node.
Moreover, an elastoplastic material with a Young’s modulus E = 200 GPa and Poisson’s
ratio ν = 0.3 is considered to model the material nonlinearities. The Riks method was chosen
to capture the structural response related to the postbuckling behavior of the girder (Riks
1979). It is worth mentioning, that initial geometrical imperfections were taken into account
during this analysis. In this case, all nodes of web panel were modified to obtain an initial
curvature similar to cosine functions. This type of initial imperfection resembles the first buck-
ling mode of a girder subjected to patch loading.

Figure 1. Notation for plate girders subjected to patch loading (notation): (a) Plate girder; (b) symmet-
ric DHFG; (c) asymmetric DHFG; (d) flat stiffener.

709
2.2 Boundary conditions
As a result of the symmetry of the geometry and boundaries restrictions, only a half of the
girder was modelled. Transverse stiffeners at the end of the girder were replaced using kine-
matic constraints, allowing only in-plane rotation. The patch load was applied over an equiva-
lent length of ss/2 on the loaded flange, allowing only vertical displacement and restricting all
rotations and out-of-plane displacements.

2.3 Validation of the model


To demonstrate the validity of the finite element model, the computed ultimate resistances are
compared against the experimental results of a longitudinal stiffened plate girder subjected to
patch loading obtained by Dubas & Tschamper (1990). Table 1 presents the dimensions and
material properties of the girder, it must be mentioned that for all tested girders a flat longitudinal
stiffener of 90 × 6 mm was used, and with maximum initial web imperfection of w0 = 7mm.
After conducting a convergence study, a mesh of 5650 elements was selected. As observed
in Table 1, a good correspondence between the numerical and experimental ultimate resistance
to patch loading is achieved, with ratio of FuFEM/FuEXP that varies from 0.94 to 1.10.

3 PARAMETRIC STUDY

In this section, two studies are conducted to investigate the influence of different geometrical
parameters on the postbuckling behavior and the ultimate resistance of DHFG subjected to
patch loading. First, a comparative analysis is done to contrast various configurations of DHFG
and the effect of the panel aspect ratio, size and position of the delta hollow flange. Next,
a second study is carried out with the purpose of comparing the DHFG against single longitu-
dinal stiffened girders. Table 2 presents the fixed geometric and material properties used during
this analysis, Additionally, an initial imperfection with a amplitude of wo = tw was employed, this
dimension meets the tolerance found in the EN1993-1-5 (2006).

3.1 Delta hollow flange girders


As mentioned above, a parametric evaluation is performed in order to investigate the influ-
ence of the following parameters (Figure 1):
• panel aspect ratio a/hw = [1.0, 2.0]
• relative depth of the DHFG d/hw = [0.1, 0.2, 0.3]
• relative width of the DHFG b´/bf = [0.25, 0.50]
• thickness ratio tst/tw = [1.00, 1.25, 1.50, 1.75, 2.00]

Table 1. Geometric and material properties of the plate girder used by Dubas & Tschamper (1990).
a [mm] hw [mm] tw [mm] fyw [MPa] bf [mm] tf [mm] fyf [MPa] ss [mm] b1 [mm] FuFEM/FuEXP

1760 1000 3.80 375 150 8.35 281 40 200 0.94


1760 1000 3.80 358 150 8.35 328 240 200 1.10
1760 1000 3.80 371 150 12.00 283 40 150 0.99
1750 1000 3.80 380 150 12.00 275 240 150 1.10

Table 2. Geometric and material properties of used in the parametric study.


hw [mm] tw [mm] fyw [MPa] bf [mm] tf [mm] fyf [MPa] ss [mm]

1000 4.0 345 250 8.0 345 200

710
It is worth noticing that in this evaluation symmetric and asymmetric DHFG were also
investigated. To this purpose, a set of 120 numerical simulations were performed, 60 for sym-
metric DHFG (Figure 1b), and other 60 for asymmetric DHFG (Figure 1c). Figures 2 and 3
show the patch loading resistance Fu of DHFG in terms of the thickness ratio tst/tw for sym-
metric and asymmetric configurations.
For both cases, of symmetric and asymmetric DHFG, the ultimate resistance Fu proportion-
ally increases with tst/tw, this is clearly observed for small panel aspect ratios a/hw = 1 and all
relative depths d/hw (see Figures 2a, 2b, 3a, 3b). However, for a/hw = 2, the influence of tst/tw on
the ultimate resistance is only significant for d/hw ≥ 0.2. It must be noted, that when the panel
ratio increases from a/hw = 1 to 2, Fu decreases an average of 20% and 14% for symmetric and
asymmetric DHFG, respectively.
Moreover, Fu is clearly affected by d/hw, this is particularly observed for symmetric DHFG
with a/hw = 2 (see Figure 2c, 2d). Nevertheless, for asymmetric DHFG, Fu increases consider-
ably when the relative depth rises from d/hw = 0.1 to 0.2. On the other hand, there is no signifi-
cant influence of b´/bf on the ultimate resistance, this is appreciated in Figures 2, 3, where Fu
increases only a 7% when b´/bf is increased from 0.25 to 0.50.
Concerning the DHFG configuration, the symmetric arrangement represents the best
option to increase Fu. This is shown in Figures 2c, 3c, where the ultimate resistance of
a symmetric DHFG for tst/tw = 2 and d/hw = 0.3 (Fu = 1425.5 kN) is 1.80 times the resistance
of an asymmetric DHFG (817.5 kN).
Figures 4, 5 depict the von Mises stress distribution in terms of the relative depth d/hw and
relative width b´/bf for symmetric and asymmetric DHFG, respectively.
For symmetric DHFG, the most stressed area is widely distributed on the lower panel of
the web, specifically in the region where the delta hollow flange is attached to the web.

Figure 2. Ultimate resistance Fu vs. relative thickness tst/tw (symmetric DHFG).

711
Figure 3. Ultimate resistance Fu vs. relative thickness tst/tw (asymmetric DHFG).

Analyzing Figure 4c, 4f, it can be seen that when d/hw ≥ 0.2 two regions of plastic collapse
are observed, one in the web panel and another in the loaded flange. Consequently, the
ultimate resistance is significantly increased. On the other hand, for asymmetric DHFG the
results show that the most stressed areas are located in the web panel. For short relative
depths d/hw = 0.10, the results are similar to those obtained for symmetric DHFG, where the
maximum stress is achieved in the area of the web panel beneath the delta hollow flange
(see Figure 5a, 5b). However, for d/hw = 0.20 and 0.30 high stress level are concentrated in
the directly loaded panel (see Figure 5c, 5f).

3.2 Comparison with longitudinal flat stiffeners


A set of additional 10 numerical simulations were carried out to contrast the postbuckling
behavior and ultimate resistance of DHFG with longitudinally stiffened girders. Since the
optimum position for longitudinally stiffened girders subjected to patch loading is attained
when the stiffener is placed near b1/hw= 0.1 (Walbridge & Lebet 2001, Graciano & Edlund
2002), for practical purpose the same dimension is used for the relative depth d/hw of DHFG.
It is worth mentioning, that the width-to-thickness ratio of the stiffener was set to bst/tst = 15.
Figure 6 presents a comparison between the ultimate resistance of symmetric and asymmetric
DHFG, and a longitudinally stiffened girder in terms of the thickness ratio tst/tw.
For the type of stiffening investigated herein, Fu increase with the ratio tst/tw. Moreover, for
d/hw = 0.1, the best configuration is a symmetric DHFG. This could be appreciated in Figure 6,
where the ultimate resistance of a symmetric DHFG for tst/tw = 2 is around 1.35 times the resist-
ance of a longitudinally stiffened girder. An increment is also found for asymmetric DHFG, for
tst/tw = 2 the resistance is around 1.21 times the ultimate resistance of a stiffened girder. Figure 7

712
Figure 4. von Mises stress distribution of symmetric DHFG at ultimate load level (a/hw = 2.00, tst/tw = 2).

shows a comparison of the stress distribution of unstiffened plate girder and the three types of
reinforcement studied herein. As shown in Figure 7a, for an unstiffened girder with panel ratio
of a/hw = 2.00 the most stressed areas are concentrated in two short regions in the web panel.
One in the region near the loaded flange, and a second one at one-third of the web panel depth.
For DHFG and the longitudinally stiffened girder, it is observed that the areas of maximum

713
Figure 5. von Mises stress distribution of asymmetric DHFG at ultimate load level (a/hw = 2.00, tst/tw = 2).

stresses are widely distributed on the web panel. For symmetric DHFG (Figure 7b), the max-
imum stress area is in the region beneath the lower panel and the reinforcement, while for asym-
metric DHFG and the longitudinally stiffened girder (see Figure 7c and 7d), two areas of
maximum stress values are defined, one in the lower panel and another in the directly loaded
panel.

714
Figure 6. Ultimate resistance Fu in terms of tst/tw for DHFG and longitudinally stiffened girder
(a/hw = 2).

Figure 7. von Mises stress distribution for DHFG and longitudinally stiffened girder at ultimate load
level (a/hw = 2.00, tst/tw = 2).

715
4 CONCLUSIONS

The resistance and postbuckling behavior of DHFG subjected to patch loading was investi-
gated herein. The results showed that compared to longitudinally stiffened girders, both sym-
metric and asymmetric DHFG significantly increased their ultimate resistance. It was also
demonstrated that the best configuration to increase the resistance occurs when a symmetric
DHFG of a relative depth of d/hw = 0.3 is used. Moreover, there is a significant control of the
stress levels when a symmetric DHFG is employed. Finally, additional numerical simulation
with different longitudinal stiffener positions should be carried out in order to extend the com-
parison with DHFG.

REFERENCES

Ansys. 2018. Ansys Release 19 Elements Reference.


Arabzadeh, A. & Varmazyari, M. 2009. Strength of I-girders with delta stiffeners subjected to eccentric
patch loading. Journal of Constructional Steel Research 65: 1385–1391.
Avery, P. & Mahendran, M. 1996. Finite element analysis of hollow flange beams with web stiffeners,
13th International Specialty Conference on Cold-Formed Steel Structures, St. Louis, 17 October 1996.
Missouri U.S.A.
Avery, P., Mahendran, M. & Nasir, A. 2000. Flexural capacity of hollow flange beams. Journal of Con-
structional Steel Research 53: 201–223.
Dong, J. & Sause, R. 2009. Flexural strength of tubular flange girders. Journal of Constructional Steel
Research 65: 622–630.
Dubas, P & Tschamper, H. 1990. Stabilité des âmes soumises à une charge concentrée et à une flexion
globale. Construction Métallique 27(2): 25–39.
El Masri, O.Y. & Lui, E.M. 2018. Cross section properties and elastic lateral-torsional buckling capacity
of steel delta girders. International Journal of Steel Structures 1–18.
ENV 1993- 1-5. 2006. Eurocode 3: design of steel structures, part 1–5: general rules, supplementary rules
for planar plated structures without transverse loading.
Graciano, C & Edlund, B. 2002. Nonlinear FE analysis of longitudinally stiffened girder webs under
patch loading. Journal of Constructional Steel Research 58(9): 1231–1246.
Navarro-Manso, A., del Coz Díaz, J.J., Alonso-Martínez, M., Blanco-Fernández, E. & Castro-Fresno,
D. 2014. New launching method for steel bridges based on a self-supporting deck system: FEM and
DOE analyses. Automation in Construction 44: 183–196.
Navarro-Manso, A., del Coz Díaz, J.J., Alonso-Martínez, M., Castro-Fresno, D. & Alvarez Rabanal, F.
P. 2015. Patch loading in slender and high depth steel panels: FEM–DOE analyses and bridge launch-
ing application. Engineering Structures 83: 74–85.
Pi, Y.-L. & Trahair, N.S. 1997. Lateral-distortional buckling of hollow flange beams. Journal of Struc-
tural Engineering 123(6): 695–702.
Riks, E.1979. An incremental approach to the solution of snapping and buckling problems. International
Journal of Solids and Structures 15: 529–551.
Walbridge, S & Lebet, JP. 2001. Patch loading tests of bridge girders with longitudinal web stiffeners.
Rapport d’essais École Polytechnique Fédérale de Laussane. ICOM 447. [In French].

716
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Critical loads of semi-rigid columns subjected to non-linear


temperature distributions

T. Ma & L. Xu
University of Waterloo, Waterloo, Ontario, Canada

ABSTRACT: A numerical method is presented for calculating the critical load of a semi-
rigidly connected column subjected to non-linear temperature distributions and applied grav-
ity loads in fire. The modulus of elasticity in a column can vary longitudinally due to many
factors, such as the presence of vertical gas temperature gradients in fires, localized heating
and damage to insulation in fire. An analytical equation for calculating the critical load of
a semi-rigidly connected column with up to three segments of varying temperatures in fire is
presented. The assumption of semi-rigid connections in the proposed method is realistic and
practical, but can be simplified to include the ideal cases of pinned and fixed connections. The
effect of initial out-of-plumbness, which increases deflections under applied loads, on the
inelastic behavior of columns is also considered. The proposed method is demonstrated via
a brief numerical example and the results are verified via finite element analysis.

1 INTRODUCTION

Longitudinal temperature gradients may exist in columns in many cases, such as the vertical gas tem-
perature gradients in fires are considered (Xu & Zhuang, 2014), or in the case of a post-earthquake
fire where segments of insulation have become damaged during the earthquake (Arablouei &
Kodur, 2016). As the Young’s Modulus and yield stress of steel varies with temperature (Lie, 1992),
the buckling loads of such columns cannot be computed with assuming uniform material properties.
Moreover out-of-plumbness column imperfections can occur in the columns during construction or
as a result of deformations during seismic events, which increase deflections and reduce inelastic
buckling loads. A numerical method is presented for calculating the critical loads of columns con-
taining up to three segments of differing temperatures. The critical load is defined in this paper as
either the load at which bifurcation is imminent (for concentrically loaded columns), or the load cor-
responding to a critical deflection defined by the onset of yielding in any segment of the column
(for eccentrically loaded columns). A brief numerical example is presented to demonstrate the
efficiency of the propose method, and the results are verified via finite element analysis.

2 BACKGROUND

For reasons of complexity, the use of segmented members in fire-structural analysis is generally
avoided by assuming uniform temperature distributions in members. However, significant differ-
ences in the results can be realized with use of the proposed method, especially when the elastic
moduli of segments in the columns differ significantly. Temperatures can vary longitudinally
through members when they are subjected to localized heating or when vertical temperature gradi-
ents are considered in room fires, as considered by Xu & Zhuang (2014), who proposed a two-
segmented column model in their storey-based stability approach analysis method. Delamination
of insulation can occur along segments of members due to explosions or seismic damage
(Arablouei & Kodur, 2016; Braxtan & Pessiki, 2011; Wang et al., 2013), especially at locations of
maximum flexure. The model of insulation delamination often results in members that can be

717
divided into up to three segments of varying temperatures during fire-structural analysis. The heat-
ing of steel reduces the modulus of elasticity and yield stress (Lie, 1992). The Eurocode 3 (BSI,
2005) is currently among the most widely accepted numerical models for estimating the properties
of steel materials in fire, and will be adopted in the calculations throughout this paper. The pro-
posed method applies the Hoblit (1951) method of determining the buckling loads of stepped col-
umns towards semi-rigidly connected columns in fire. Although a theoretical basis for the
determination of the buckling loads of such columns is provided in Hoblit (1951), the computa-
tional procedures provided are insufficient to solve for the buckling loads of semi-rigidly connected
columns. Moreover, the effect of out-of-plumbness imperfections towards increasing the deflec-
tions of segmented semi-rigid columns and reducing the inelastic buckling load has not yet been
considered. As such, the proposed method is complete for determining the critical load of semi-
rigidly connected segmented column, which may contain column imperfections.

3 CONCENTRICALLY LOADED COLUMN

The case of a concentrically loaded stepped column in Figure 1 is first considered. The column
is assumed to be sufficiently braced such that no lateral sway will occur at its ends. The rota-
tion, z'(x), between segments is also assumed to be continuous. Since the column is concentric-
ally loaded, the critical load will be either the rotational buckling or yielding loads. Where P is
the compressive axial load, V is the transverse end reaction, and Mu and Ml are the end
moments. Let L = Ll + Lm + Lu and I be the moment of inertia. The rotational springs at the
ends of the columns are assumed to be linear in Equation 1.

Ru z0 ðLÞ ¼ Mu ð1aÞ

Rl z0 ð0Þ ¼ Ml ð1bÞ

Due to the presence of multiple segments along the length of the column with differing elastic
moduli, the buckling load must be solved using the Hoblit (1951) method. With neglecting the

Figure 1. Three-segmented column subjected to concentric loading.

718
effects of shear and axial deformations, the internal moment in each of the three sections of
the column are given in Equation 2.

d2z
El I ¼ PzðxÞ þ Ml þ Vx; 0  x  Ll ð2aÞ
dx2
d2z
El I ¼ PðzðxÞ  zðLl ÞÞ þ MðLl Þ þ Vx; Ll  x  Ll þ Lm ð2bÞ
dx2
d2z
Eu I ¼ PðzðxÞ  zðLl þ Lm ÞÞ þ MðLl þ Lm Þ þ Vx; Ll þ Lm  x  L ð2cÞ
dx2

By solving the system of differential equations in Equation 2, the bending moment, rotation
and deflection of the column can be obtained at the upper and lower ends of the column, as
well as at the intermediate points between adjacent segments. The displacement quantities,
z(x), are given in Equation 3.

zð0Þ ¼ 0 ð3aÞ

Mð0Þ Mð0Þ z0 ð0ÞV=P VLl


zðLl Þ ¼ zð0Þ þ  Cl þ Sl þ ð3bÞ
P P l =L P

MðLl Þ MðLl Þ z0 ðLl Þ  V=P VLm


zðLl þ Lm Þ ¼ zðLl Þ þ  Cm þ Sm þ ð3cÞ
P P m =L P

MðLl þ Lm Þ MðLl þ Lm Þ z0 ðLl þ Lm Þ  V=P VLu


zðLÞ ¼ zðLl þ Lm Þ þ  Cu þ Su þ ð3dÞ
P P u =L P

zðLÞ ¼ 0 ð3eÞ

where the coefficients ϕl, ϕm, ϕu, Sl, Sm, Su, Cl, Cm and Cu are all shown in Equation 4. As no
lateral sway is assumed, the boundary conditions in Equations 3a and 3e must be satisfied.
rffiffiffiffiffiffiffiffi
   P
l ¼ pffiffiffiffi ; m ¼ pffiffiffiffiffiffi ; u ¼ pffiffiffiffiffi ;  ¼ L ð4aÞ
μl μm μu E0 I
     
l Ll m Lm u Lu
Sl ¼ sin ; Sm ¼ sin ; Su ¼ sin ð4bÞ
L L L
     
l Ll m Lm u L u
Cl ¼ cos ; Cm ¼ cos ; Cu ¼ cos ð4cÞ
L L L

where the elastic moduli of the segments are related by E = μE0, and E0 is the elastic modulus
at ambient temperature. The rotation quantities, z’(x), are given in Equation 5.

z0 ð0Þ ¼ θ ð5aÞ
 
Mð0Þ l V
z ðLl Þ ¼
0
Sl þ ðz0 ð0Þ  V=PÞCl þ ð5bÞ
P L P
 
MðLl Þ m V
z0 ðLl þ Lm Þ ¼ Sm þ ðz0 ðLl Þ  V=PÞCm þ ð5cÞ
P L P

719
 
MðLl þ Lm Þ u V
z0 ðLÞ ¼ Su þ ðz0 ðLl þ Lm Þ  V=PÞCu þ ð5dÞ
P L P

where θ is arbitrarily chosen, since any value of which can satisfy the equilibrium equation in the
buckled configuration, unless the bottom end is fixed, in which case θ = 0. Finally, the flexural
quantities, M(x), are given in Equation 6, with substitution of Equation 1 for the end moments.
0
Mð0Þ ¼ Ml ¼ Ru z ðLÞ ð6aÞ

MðLl Þ ¼ Mð0ÞCl  PLSl ðz ð0Þ  V=PÞ=l


0
ð6bÞ

MðLl þ Lm Þ ¼ MðLl ÞCm  PLSm ðz0 ðLl Þ  V=PÞ=m ð6cÞ

MðLÞ ¼ MðLl þ Lm ÞCu  PLSu ðz0 ðLl þ Lm Þ  V=PÞ=u ð6dÞ

MðLÞ ¼ Mu ¼ Ru z0 ðLÞ ð6eÞ

Solving for V expressed as a function of only the given quantities, produces Equation 7.
V ¼ θðRl β1 þ LPβ2 Þ=ðLβ3 Þ ð7aÞ

β1 ¼ l m u ð1  Cl Cm Cu Þ þ 2l u Sl Sm Cu þ 2l m Sl Cm Su þ l 2m Cl Sm Su ð7bÞ

β2 ¼ l m Cl Cm Su þ l u Cl Sm Cu þ m u Sl Cm Cu  2m Sl Sm Su ð7cÞ

β3 ¼ l m u þ 2m Sl Sm Su  l m Cl Cm Su  l u Cl Sm Cu  m u Sl Cm Cu ð7dÞ

Any value of P satisfying all fourteen equations in the system comprising of Equations 3, 5
and 6 is a buckling of the column, of which the minimum positive and non-zero value corres-
ponds to the fundamental mode and is denoted as Pb. However, there is no explicit solution
due to the non-linearity of the system with respect to P. As such, a simple root-finding
method such as the Newton-Raphson method (Ypma, 1995) may be used. Note that as mul-
tiple buckling loads corresponding to the different buckling modes may be found, the use of
multi-starting (György & Kocsis, 2011) is recommended. A brief computational procedure for
determining the buckling load, Pb, is presented as follows:
1. Take an initial trial value of P.
2. Evaluate the resulting values of M(L) in (Eq. 6e) and z'(L) in (Eq. 5d) via
Equations 3, 5 and 6, and then the residual of (Eq. 6e), ϵ = M(L) + Ruz'(L).
3. Vary P until ϵ sufficiently close to zero using root finding methods, and report the result-
ing buckling load.
4. Repeat Steps 1 through 3 for a sufficient number of trial values of P in varying orders of
magnitude to ensure that the minimum buckling load corresponding to the fundamental
buckling load is fond, rather than the buckling loads corresponding to higher buckling
modes.
In addition to rotational buckling, the yielding of the column occurs when the stress in the
column exceeds the yield stress of its weakest segment. The critical load, Pu, of the column is
shown in Equation 8.
Pu ¼ Pb  Py ð8aÞ

Py ¼ minffy;l ; fy;m ; fy;u gA ð8bÞ

where fy,l, fy,m, and fy,u correspond to the yield stress in each of the lower, middle and upper
segments, respectively, and A is the cross-sectional area. Py is an upper bound for Pb since the
elastic modulus diminishes to zero at P = Py.

720
4 IMPERFECT COLUMN

Consider now the case of a column subjected to out-of-plumbness, Δ0, in Figure 2. The out-of-
plumbness function, z0(x), is linearly varying from zero to Δ0 as shown, and the resulting
storey drift angle is θ0. Also, V = –KbΔ, where Kb is the stiffness of the lateral bracing.
Since an eccentricity function, z0(x) = Δ0x/L, now exists in the column, the deformations of
the column are determinate and no longer arbitrary. The solution to the Euler-Bernoulli dif-
ferential equations in each segment are shown in Equation 9.

d 2 zl
El I ¼ Pðzl ðxÞ þ z0 ðxÞÞ þ Ml  Vx; 0  x  Ll ð9aÞ
dx2
d 2 zm
El I ¼ Pðzm ðxÞ þ z0 ðxÞÞ þ MðLl Þ  Vx; Ll  x  Ll þ Lm ð9bÞ
dx2
d 2 zu
Eu I ¼ Pðzu ðxÞ þ z0 ðxÞÞ þ MðLl þ Lm Þ  Vx; Ll þ Lm  x  L ð9cÞ
dx2

By solving the differential equations in Equations 9, the displacements z in each segment are
presented in Equation 10.
   
l l θl Rl D0 V
zl ðxÞ ¼ D k1 cos x þ k2 sin x þ  x  x; 0  x  Ll ð10aÞ
L L P L P
   
m m θl Rl D0 V
zm ðxÞ ¼ D k3 cos x þ k4 sin x þ  x  x; Ll  x  Ll þ Lm ð10bÞ
L L P L P

Figure 2. Three-segmented out-of-plumb column subjected to axial loading.

721
   
u u θl Rl D0 V
zu ðxÞ ¼ D k5 cos x þ k6 sin x þ  x  x; Ll þ Lm  x  L ð10cÞ
L L P L P

where the integration constants k1through k6 can be solved by applying the boundary and
compatibility conditions in (Eq. 11a) and (Eq. 11b-c), respectively.
0
zl ð0Þ ¼ 0; zu ðLÞ ¼ D; zl ð0Þ ¼ θl ; z0u ðLÞ ¼ θu ð11aÞ

zl ðLl Þ ¼ zm ðLl Þ; zm ðLl þ Lm Þ ¼ zu ðLl þ Lm Þ ð11bÞ


0
zl ðLl Þ ¼ z0m ðLl Þ; z0m ðLl þ Lm Þ ¼ z0u ðLl þ Lm Þ ð11cÞ

Solving for the constants yields the solution in Equation 12.


2 3 2 3
k1 α11 α12 α13
6 k2 7 6 α21 α22 α23 7 2 3
6 7 6 7 Ru Rl
6 k3 7 6 α33 7
6 7 ¼ 1 6 α31 α32 74 LPRl 5 ð12Þ
6 k4 7 Dk 6 α41 α42 α43 7
6 7 6 7 LPRu
4 k5 5 4 α51 α52 α53 5
k6 α61 α62 α63

where the denominator coefficient Dk and the elements of the α matrix are given in the appen-
dix. The bending moment is obtained by substituting (Eq. 11) into (Eq. 10) in Equation 13.
    
l l
MðxÞ ¼ PD k1 cos x þ k2 sin x ; 0  x  Ll ð13aÞ
L L
    
m m
MðxÞ ¼ PD k3 cos x þ k4 sin x ; Ll  x  Ll þ Lm ð13bÞ
L L
    
u u
MðxÞ ¼ PD k5 cos x þ k6 sin x ; Ll þ Lm  x  L ð13cÞ
L L

where the deflection, Δ, is expressed in Equation 14.

D0 ðψ  1Þ
D¼ ð14aÞ
P ðψ  1Þ
1 þ KL

9ru rl a5  32 ½rl ð1  ru Þτu þ ru ð1  rl Þτl


a1
ψ¼ ð14bÞ
4 τ l τ u ð1  rl Þð1  ru Þa1 þ 9ru rl a2  3 ½rl ð1  ru Þτ u a3 þ ru ð1  rl Þτ l a4

Equations for the coefficients a1 through a5, end fixity factors ru and rl, and adjustment factors τu
and τl are provided in the appendix. In the presence of initial column imperfections, the onset of
yielding will occur with excessive deflections before buckling shortly thereafter. The critical deflec-
tion of the column is hereby defined as the deflection corresponding to the initiation of yielding in
the section within any segment. The critical load, Pu, of the imperfect column is thus the load that
results in the critical deflection corresponding to the initiation of yielding. This failure condition is
shown in Equation 15, in which the inequalities shown will always be satisfied.

Pu ¼ minfPu;l ; Pu;m ; Pu;u g  Psw ðKb Þ  Pb ð15Þ

where Pu,i, Pu,m and Pu,u are determined by solving (Eq. 16a-c) individually. Psw is the upper
bound sway load of the column, obtained by solving for the lowest positive value of P which

722
results in a denominator of zero in (Eq. 14a). It can be shown that Pu asymptotically approaches
Psw as Δ0 approaches zero. Pb is the rotational buckling load of the column obtained via Section 3
for a fully braced column (Kb = ∞). Psw asymptotically approaches Pb as Kb approaches infinity.

Pu;l =A þ Mmax;l ðPu;l ; Tl Þ=Sx  fy ðTl Þ ¼ 0 ð16aÞ

Pu;m =A þ Mmax;m ðPu;m ; Tm Þ=Sx  fy ðTm Þ ¼ 0 ð16bÞ

Pu;u =A þ Mmax;u ðPu;u ; Tu Þ=Sx  fy ðTu Þ ¼ 0 ð16cÞ

where Sx is the section modulus and Mmax is the absolute maximum bending moment in each
segment which can readily be obtained via discretization (no closed form solution is possible).
Since the bending moment is a function of P, either root finding or the iteration of the axial
load is required to determine Pu in each segment. Alternatively, for any given axial load the
column has not yielded if the left-hand sides of (Eq. 16a-c) are all negative and the inequalities
of (Eq. 15) are satisfied.

5 NUMERICAL EXAMPLE

A brief numerical example is hereby provided to demonstrate the use of the proposed method
for a typical column. Consider a steel (E0 = 200 GPa) column of length L = 4.877 m with a
W310 × 60 section (I = 129 × 106 mm4, A = 7,610 mm2). Damage to the insulation has occurred
at the upper and lower ends of the column, such that Ll = 1.877 m, Lu = 1.0 m and Lm = 2.0 m.
Following the damage, a room fire occurs, resulting in temperatures Tl = 336°C,
Tu = 224°C and Tm = 56°C. The rotational rigidities of the connections on either end are Rl = 0
and Ru = 6.852 × 107 Nm. The rotational buckling load of the column was determined via
Section 3, and resulted in Pb = 2,120 kN. A simple finite element analysis was conducted in
ABAQUS to verify the result using wireframe B23 Euler-Bernoulli elements to simulate the seg-
ments of the column. By utilizing the buckling linear perturbation analysis feature, the resulting
critical load was verified to be 2,120 kN.
Suppose now an out-of-plumbness imperfection Δ0 with varying magnitude is introduced to
the column, as well as a lateral brace Kb varying from zero (unbraced) to infinity (fully-
braced). The loads at which the critical deflection corresponding to the onset of yielding is
reached, Pu, as well as the lateral sway load, Psw, were determined using the proposed method
in Section 4. The sway load is plotted as a function of the lateral bracing in Figure 3, and Pu is
plotted versus Δ0 for varying values of Kb in Figure 4.
Based on the figures, the effect of out-of-plumbness imperfections on Pu becomes more sig-
nificant when less lateral bracing is provided. As long as Δ0 > 0, Pu is less than the sway load,
Psw. If full lateral bracing is provided, then Pu and Psw approach the rotational buckling load,
Pb. Note that in Figure 4 some inflection points exist in the plotted curves, and correspond to
the transition points where the segment governing the onset of yielding in the column
changes. For example, with Kb = 750 kN/m, the yielding of the middle segment governs for
Δ0 < 0.029L whereas the yielding of the upper segment governs for Δ0 ≥ 0.029L.

Figure 3. Effect of lateral bracing on sway load.

723
Figure 4. Effect of out-of-plumbness on the onset of yielding.

The results were once again validated via finite element analysis using a 2D element column
model with B23 elements. The resulting beam stresses (based on the assigned cross-sectional prop-
erties) and section moments of the column in the numerical example up to before the onset of yield-
ing were computed using the finite element method and were found to match the reported values.

6 CONCLUSION

A method was derived for determining the theoretical buckling load of a semi-rigidly connected
column, which can easily be accomplished via the computational procedure provided. The com-
putational procedure involves the use of simple root-finding algorithms such as the Newton-
Raphson method (Ypma, 1995). Where out-of-plumbness column imperfections exist, the critical
load corresponding to the deflection that results in the onset of yielding can also be computed
with ease in the proposed method. A numerical example is provided to demonstrate the efficiency
and applicability of the proposed method. It is shown that the load corresponding to the onset
of yielding in the column will always be less than the buckling load as long as out-of-plumbness
imperfections exist in the column. Moreover, the introduction of lateral bracing helps to improve
the performance of the column by reducing the effect of column imperfections on the critical
load, as well as increasing the sway buckling load. The results of the numerical example were
validated by finite element analysis. Given that programming of the recommended computation
procedures can easily be done using the proposed method, the buckling analysis of such columns
can be completed with ease in less time than with utilizing a finite element software package.

REFERENCES

Arablouei, A. & Kodur, V.K.R. 2016. Effect of fire insulation delamination on structural performance of
steel structures during fire following an earthquake or an explosion. Fire Safety Journal 84(2016): 40-49.
Braxtan, N.L. & Pessiki, S. P. 2011. Postearthquake fire performance of sprayed fire-resistive material on
steel moment frames. Journal of Structural Engineering 137(9): 946–953.
BSI. 2005. BS EN 1993-1.2: 2005 Eurocode 3, Design of steel structures, Part 1.2: General rules – Struc-
tural fire design. London: British Standards Institution.
György, A. & Kocsis, L. 2011. Efficient multi-start strategies for local search algorithms. Journal of Arti-
ficial Intelligence Research 41(2011): 407–444.
Hoblit, F.M. 1951. Buckling load of a stepped column. Journal of Aeronautical Sciences 18(2): 124–126.
Lie, T. 1992. Structural Fire Protection. New York: American Society of Civil Engineers.
Wang, W-Y., Li, G-Q. & Kodur, V. 2013. Approach for modeling fire insulation damage in steel
columns. Journal of Structural Engineering 139(4): 491–503.
Xu, L. & Zhuang, Y. 2014. Storey stability of unbraced steel frames subjected to non-uniform elevated
temperature distribution. Engineering Structures 62-63(2014): 164–173.
Ypma, T.J. 1995. Historical development of the Newton-Raphson method. SIAM Review 37(4):
531–551.

724
APPENDIX
The variables for computing the bending moment function of a three-segmented column are
given as follows.

Dk ¼ Rl Ru ðα01 þ α02 þ α03 Þ þ L2 P2 α04 þ LPðRl þ Ru Þα04 þ LPRl α05 þ LPRu α06 ð17aÞ

α01 ¼ 2l m u ð1  Cl Cm Cu Þ ð17bÞ

α02 ¼ l 2m Cl Sm Su þ 2l m Sl Cm Su þ l 2u Cl Sm Su þ …


ð17cÞ
2l u Sl Sm Cu þ m 2u Sl Cm Su þ 2m u Sl Sm Cu

α03 ¼ 2l 2u Sl Sm Su  2l m u Sl Cm Cu  l 2m u Cl Sm Cu  l m 2u Cl Cm Su ð17dÞ

α04 ¼ l m Cl Cm Su þ l u Cl Sm Cu þ m u Sl Cm Cu  2m Sl Sm Su ð17eÞ

α05 ¼ l 2m Cl Sm Su þ 2l m Sl Cm Su þ 2l u Sl Sm Cu  l m u Cl Cm Cu ð17fÞ

α06 ¼ l 2u Cl Sm Su þ 2m u Sl Sm Cu þ m 2u Sl Cm Su  l m u Cl Cm Cu ð17gÞ

α11 ¼ l m u ð1  Cl Cm Cu Þ  l 2u Cl Sm Su  m 2u Sl Cm Su  2m u Sl Sm Cu ð17hÞ

α12 ¼ l m Cl Cm Su  l u Cl Sm Cu  m u Sl Cm Cu þ 2m Sl Sm Su ð17iÞ

α13 ¼ 0 ð17jÞ

α21 ¼ l m u Sl Cm Cu  l 2u Sl Sm Su þ m 2u Cl Cm Su þ 2m u Cl Sm Cu ð17kÞ

α22 ¼ l m Sl Cm Su  l u Sl Sm Cu þ m u Cl Cm Cu  2m Cl Sm Su ð17lÞ

α23 ¼ 2 3 ð17mÞ

α31 ¼ l m u ðCl Cml  Cu Cmm Þ  l 2u Su Smm  2l u Sl Sml ð17nÞ

α32 ¼ l m Su Cmm  l u Cu Smm ð17oÞ

α33 ¼ l u Cl Sml þ m u Sl Cml ð17pÞ

α41 ¼ l m u ðCu Smm  Cl Sml Þ þ l 2u Su Cmm þ 2l u Sl Cml ð17qÞ

α42 ¼ l m Su Smm þ l u Cu Cmm ð17rÞ

α43 ¼ l u Cl Cml  m u Sl Sml ð17sÞ

α51 ¼ l m u ðCl Cm Cum  Cuu Þ þ 2l u Sl Sm Cum  2l m Sl Cm Sum  l 2m Cl Sm Sum ð17tÞ

α52 ¼ l m Suu ð17uÞ

α53 ¼ l m Cl Cm Sum þ l u Cl Sm Cum þ m u Sl Cm Cum þ m 2 Sl Sm Sum ð17vÞ

α61 ¼ l m u ðSuu  Cl Cm Cum Þ  l 2 u Sl Sm Sum þ 2l m Sl Cm Cum þ l 2m Cl Sm Cum ð17wÞ

α62 ¼ l m Cuu ð17xÞ

725
α63 ¼ l m Cl Cm Cum  l u Cl Sm Sum  m u Sl Cm Sum  m 2 Sl Sm Cum ð17yÞ

Some additional coefficients are introduced in Equation 17 and shown in Equation 18.
     
 m Ll m ðLl þ Lm Þ u ðLl þ Lm Þ
Sml ¼ sin ; Smm ¼ sin ; Sum ¼ sin ; Suu ¼ sin u
L L L
ð18aÞ
     
m L l m ðLl þ Lm Þ u ðLl þ Lm Þ
Cml ¼ cos ; Cmm ¼ cos ; Cum ¼ cos ; Cuu ¼ cos u
L L L
ð18bÞ

The variables for computing the deflection of a three-segmented column with out-of-
plumbness imperfections are presented as follows.

a1 ¼ 2m Sl Sm Su  l m Cl Cm Su  m u Sl Cm Cu  l u Cl Sm Cu ð19aÞ

a2 ¼ 2l m u Sl Cm Cu þ l 2m u Cl Sm Cu þ l m 2u Cl Cm Su  2l 2u Sl Sm Su ð19bÞ

a3 ¼ 2l m Sl Cm Su þ l 2m Cl Sm Su þ 2l u Sl Sm Cu  l m u Cl Cm Cu ð19cÞ

a4 ¼ l 2u Cl Sm Su þ m 2u Sl Cm Su þ 2m u Sl Sm Cu  l m u Cl Cm Cu ð19dÞ

a5 ¼ a3 þ a4 þ 2l m u ð19eÞ

1
rl ¼ ð19fÞ
3E0 I
1 þ LR
l τl

1
ru ¼ ð19gÞ
3E0 I
1 þ LR
u τu

     3 !  !
1 Ll 3 1 Ll þ Lm 3 Lu 1 Ll þ Lm 3
τl ¼ þ þ þ 1 ð19hÞ
μl L μm L L μu L

     3 !  3 !
1 Lu 3 1 Lu þ Lm 3 Ll 1 Lu þ Lm
τu ¼ þ þ þ 1 ð19iÞ
μu L μm L L μl L

726
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

The stability of semi-braced steel frames containing members with


stepped segments

T. Ma & L. Xu
University of Waterloo, Waterloo, Ontario, Canada

ABSTRACT: A new method is presented for evaluating the stability of a semi-braced steel
frame containing stepped members. Stepped members consist of multiple segments with dif-
fering sectional or material properties, such as the cross-sectional area, moment of inertia,
or elastic modulus. For instance, steel members subjected to non-linear elevated temperature
distributions contain longitudinally varying modulus of elasticity. The proposed method is
in the form of a storey-based lateral stiffness equation for semi-rigidly connected members
with up to three segments. A structural frame becomes laterally unstable when its lateral
stiffness diminishes to zero. The efficiency of the proposed method is demonstrated via an
example of a semi-braced steel frame containing stepped members, whereby the critical grav-
ity load causing instability of the frame is computed. The results obtained in the example
were also validated via finite element analysis.

1 INTRODUCTION

The structural analysis of stepped members has been studied for decades (Thomson, 1949;
Hoblit, 1951; Castiglioni, 1984; Xu & Zhuang, 2014), owing to its applicability in aeronautical
sciences (Thomson, 1949; Hoblit, 1951) and most recently in structural-fire analysis (Xu &
Zhuang, 2014). A stepped member contains multiples longitudinal segments of differing sec-
tional or material properties. That is, the elastic modulus, E, moment of inertia, I, and/or the
cross-sectional area, A, may differ between segments, but are assumed to be constant within
each segment. In fire-structural engineering, stepped column analyses can be applied for when
longitudinal temperature gradients exist in columns (Xu & Zhuang, 2014), or when the delamin-
ation of insulation occurs due to seismic or blast loading prior to fire (Arablouei & Kodur,
2016; Braxtan & Pessiki, 2011; Wang et al., 2013). A new computational method is presented
for calculating the lateral stiffness of a semi-braced storey frame containing semi-rigidly con-
nected members with up to three segments. To be clear, the terms “semi-braced” and “semi-
rigid” in this paper refer to the presence of lateral bracing and semi-rigid or rotational spring
connections in the frame, respectively. The use of semi-rigidly connected members is useful as
most of beam-to-column connections can be realistically classified as semi-rigid connections. It
is also a generalization that accommodates the cases of idealized connections. Lateral instability
occurs when the lateral stiffness of an unbraced or semi-braced frame diminishes to zero (Xu,
2001). As such, the elastic critical sway buckling loads or temperatures can be calculated using
the proposed method. A numerical example is provided to demonstrate the efficiency of the pro-
posed method towards the stability analysis of a two-bay frame containing segmented members.
The proposed method is an exact solution and is verified via finite element analysis.

2 DERVIATION OF EQUATIONS

The proposed methodology requires three major components: (1) evaluation of the lateral
stiffness of a three-segmented column, and subsequently the lateral stiffness of a semi-braced

727
frame; (2) computing of the rotational restraint provided by connecting beams, which are also
stepped members; and (3) calculating of the rotational buckling load, which is the upper
applicability limit for the proposed lateral stiffness equation and is already derived in Ma &
Xu (2019).

2.1 Column lateral stiffness equation


Consider the semi-rigidly connected column shown in its deformed state in Figure 1, with P-Δ
effects being considered. The column contains three segments, denoted by the subscripts
l, m and u for the lower, middle, and upper segments, respectively. The cross section at any
location x is assumed to be uniform. The rotation between segments is also assumed to be
continuous.
P is the compressive axial load, Y is the transverse force, and Mu and Ml are the end
moments. Let L = Ll + Lm + Lu. The moment-rotation relationship at the ends of the columns
are assumed to be linear in Equation 1.

Rl l ¼ Ml ð1aÞ

Ru u ¼ Mu ð1bÞ

where Rl and Ru are the rotational stiffnesses of the lower and upper connections, respectively.
In the consideration of semi-rigid connections, the end fixity factor, r, is popular as it maps
values of R to a finite domain, whereby r = 1 corresponds to a fixed connection (R = ∞), r = 0
corresponds to a pinned connection (R = 0). Values of r between zero and unity correspond to

Figure 1. Three-segmented column subjected to second order effects.

728
semi-rigid connections. The end fixity factor was first defined by Monforton & Wu (1963) and is
modified in Equation 2 to consider the segmentation of the column via a τ factor. The expression
for τ was obtained via the principle of virtual work in following the procedure of Zhuang (2013).

     3 !   !
1 1 Ll 3 1 Ll þ Lm 3 Lu 1 Ll þ Lm 3
rl ¼ ; l ¼ þ þ þ 1
3ðEIÞ0 μl L μm L L μu L

LRl l
ð2aÞ
   3  3 !  3 !
1 1 Lu 3 1 Lu þ Lm Ll 1 Lu þ Lm
ru ¼ ; u ¼ þ þ þ 1
3ðEIÞ0 μu L μm L L μl L

LRu u
ð2bÞ

where (EI)0 is a reference value of EI whereby μ = EI/(EI)0 is the relative stiffness of each
segment. The external equilibrium equation in Equation 3 is first obtained by taking the sum
of moments in Figure 1.

Mu þ Ml ¼ YL þ PΔ ð3Þ

Internal equilibrium is then established using the Euler-Bernoulli assumption via Equation 4.

d 2 yl
ðEIÞl ¼ Ml  Pðyl ðxÞÞ  Yx; 0  x  Ll ð4aÞ
dx2
d 2 ym
ðEIÞm ¼ Ml  Pðym ðxÞÞ  Yx; Ll  x  Ll þ Lm ð4bÞ
dx2
d 2 yu
ðEIÞu ¼ Ml  Pðyu ðxÞÞ  Yx; Ll þ Lm  x  L ð4cÞ
dx2

The boundary and compatibility conditions for the column are listed in Equation 5.

yl ð0Þ ¼ 0 ð5aÞ

yu ðLc Þ ¼ Δ ð5bÞ

y0l ð0Þ ¼ l ð5cÞ

y0u ðLc Þ ¼ u ð5dÞ

yl ðLl Þ ¼ ym ðLl Þ ð5eÞ

ym ðLl þ Lm Þ ¼ yu ðLl þ Lm Þ ð5fÞ

y0l ðLl Þ ¼ y0 m ðLl Þ ð5gÞ

y0m ðLl þ Lm Þ ¼ y0u ðLl þ Lm Þ ð5hÞ

Solving the system of differential equations in Equation 4 with applying the boundary condi-
tions in Equation 5 yields the lateral stiffness of the column, S, in Equation 6.

729
 
Y 0 2 ðEIÞ0 1
S¼ ¼ ð6aÞ
Δ L3 1

9ru rl a5  30 2 ½rl ð1  ru Þu þ ru ð1  rl Þl


a1
¼ ð6bÞ
0 4 l u ð1  rl Þð1  ru Þa1 þ 9ru rl a2  30 2 ½rl ð1  ru Þu a3 þ ru ð1  rl Þl a4

a1 ¼ 2m Sl Sm Su  l m Cl Cm Su  m u Sl Cm Cu  l u Cl Sm Cu ð6cÞ

a2 ¼ 2l m u Sl Cm Cu þ l 2m u Cl Sm Cu þ l m 2u Cl Cm Su  2l 2u Sl Sm Su ð6dÞ

a3 ¼ 2l m Sl Cm Su þ l 2m Cl Sm Su þ 2l u Sl Sm Cu  l m u Cl Cm Cu ð6eÞ

a4 ¼ l 2u Cl Sm Su þ m 2u Sl Cm Su þ 2m u Sl Sm Cu  l m u Cl Cm Cu ð6dÞ

a5 ¼ a3 þ a4 þ 2l m u ð6eÞ

where the modified loading coefficients, ϕl, ϕm and ϕu are related to the reference axial load
factor of the column, ϕ0 via modification by μ of the corresponding segment in (Eq. 7a). The
trigonometric functions Sl, Sm, Su, Cl, Cm and Cu are defined in (Eq. 7b-c).
sffiffiffiffiffiffiffiffiffiffiffiffi
P 0 0 0
0 ¼ L; l ¼ pffiffiffiffi ; m ¼ pffiffiffiffiffiffi ; u ¼ pffiffiffiffiffi ð7aÞ
ðEIÞ0 μl μm μu
     
l Ll m Lm u Lu
Sl ¼ sin ; Sm ¼ sin ; Su ¼ sin ð7bÞ
L L L
     
l Ll m L m u L u
Cl ¼ cos ; Cm ¼ cos ; Cu ¼ cos ð7cÞ
L L L

Assuming that all the columns in a storey frame experience the same deflection (valid when
rigid floor diaphragms are present), it can be shown that the lateral stiffness of a semi-braced
storey frame, ΣS, is the sum of contributions of the individual columns, Si, plus the total bra-
cing stiffness, Sb, given in Equation 8.
! " 2  #
X
n X
n 0;i ðEIÞ0;i 1
S ¼ Si þ Sb ¼ þ Sb ð8Þ
i¼1 i¼1 Li 3 i 1

Note that the values of μ can be set to be equal between adjacent segments to ensure uniform-
ity of properties between segments. In this way, columns with less than three segments can be
modeled as well. The storey becomes unstable when ΣS diminishes to zero (Xu, 2001).

2.2 Beam-to-column rotational restraining stiffness


The rotational restraints provided to the ends of columns are usually provided by either ground
supports or beams. The values of Ru and Rl are derived in this section for cases where segmented
beams contribute to the rotational restraints. First, the value of R at an end of a column is the
sum of rotational rigidities provided by all of the in-plane connecting beams (in most cases, up
to two) via Equation 9 (Xu, 2001). Conversely, if the lower ends of the columns are connected
to column bases, then the corresponding value of R may be directly specified (Xu, 2001).
X
ml X
mu
Rl ¼ Rjl ; Ru ¼ Rju ð9Þ
jl ¼1 ju ¼1

730
where Rj is the rotational stiffness contribution to the column provided by beam j, and m is
the number of contributing beams. The value of Rj is given in Equation 10, derived using
a similar procedure to Monforton & Wu (1963).

6ðEIÞj;0 rN 2F μL μM μR ð1  rF Þ þ 2lNN rF þ lNF rF v
Rj ¼ ð10Þ
Lj 4lA þ rN lB þ rF lC  rN rF lD

where the subscripts N and F correspond to the near and far ends of the beam with respect to
the column in consideration. (EI)j and Lj refer to the properties of beam j. The coefficients λA,
λB, λC, and λD are given in (Eq. 11a-d) and are defined such at λA = λD = 1 and λB = λC = 0 in
the case of a single segment, uniform beam (μN = μM = μF = 1). The coefficients λNN, λNF, κN
and κF depend on the properties and lengths of each segment of the member, given in (Eq.
11e-h). v is the ratio between the rotation of the far end to the rotation of the near end in the
buckled shape, and may assumed to take the value of unity (Xu & Liu, 2002), which corres-
ponds to the asymmetrical buckling case.

lA ¼ μL μM μR N F ð11aÞ

lB ¼ 4F ðN  N μN μM μF Þ ð11bÞ

lC ¼ 4N ðF  F μN μM μF Þ ð11cÞ


lD ¼ 4½F N þ N F  N F μN μM μF
 …

3L4b;j L4N μM μF =μN þ L4M μN μF =μM þ L4F μN μM =μF þ …
4LN L3M μF þ 4LN L3F μM þ 4LM L3N μF þ …
ð11dÞ
4LM L3F μN þ 4LF L3N μM þ 4LF L3M μN þ …
6L2N L2M μF þ 6L2N L2F μM þ 6L2M L2F μN þ …

12L2N LM LF μM þ 12LN L2M LF μM þ 12LN LM L2F μM
1  3
lNN ¼ LN μM μF þ L3M μN μF þ L3F μN μM þ …
Lb;j 3
ð11eÞ
3LN L2M μN μF þ 3LN L2F μN μM þ 3LM L2N μN μF þ …
3LF L2M μN μF þ 3LF L2N μF μM þ 3LM L2F μM μN þ 6LN LM LF μN μM Þ
1 
lNF ¼ L3N μM μF þ L3M μN μF þ L3F μN μM þ
Lb;j 3
ð11fÞ
3LN L2M μN μF þ 3LN L2F μN μM þ 3LM L2N μM μF þ …
3LF L2M μN μF þ 3LF L2N μF μM þ 3LM L2F μN μM þ 6LN LM LF μN μF Þ
1 h 4 
N ¼ 4 LN μM μF þ L4M μN μF þ L4F μN μM þ …:
Lb;j
 
3 LN L3M μM μF þ LN L3F μM μF þ LM L3F μN μF þ …
 
4 LM L3N μM μF þ LF L3N μM μF þ LF L3M μN μF þ …
 
LN L3M μN μF þ LN L3F μN μM þ LM L3F μN μM þ … ð11gÞ
 
6 L2N L2M μM μF þ L2N L2F μM μF þ L2M L2F μN μF þ …
 
9 LN L2M LF μM μF þ LN LM L2F μM μF þ …
 
12 L2N LM LF μM μF þ …
 i
3 LN L2M LF μN μF þ LN LM L2F μN μF

731
1
 4 
F ¼ 4
LN μM μF þ L4M μN μF þ L4F μN μM þ …
Lb;j
 
3 LM L3N μN μF þ LF L3N μN μM þ LF L3M μN μM þ …
 
4 LN L3M μN μF þ LN L3F μN μM þ LM L3F μN μM þ …
 
LM L3N μM μF þ LF L3N μM μF þ LF L3M μN μF þ … ð11hÞ
 
6 L2N L2M μN μF þ L2N L2F μN μF þ L2M L2F μN μM þ …
 
9 LN L2M LF μN μM þ LF LM L2N μM μN þ …
 2 
12 LF LM LN μM μN þ …
 
3 LN L2M LF μN μF þ LF LM L2N μN μF

The left and right end fixity factors of the beam, rL and rR, follow the same form as Equation 2,
except that they are functions of the connections themselves in Equation 12.

     3 !  !
rL;0 L 1 LL 3 1 LL þ LM 3 LR 1 LL þ LM 3
rL ¼ ; L ¼ þ þ þ 1
1  rL;0 ð1  L Þ μL Lj μM Lj Lj μR Lj
ð12aÞ
     !  !
rR;0 R 1 LR 3 1 LR þ LM 3 LL 3
1 LR þ LM 3
rR ¼ ; R ¼ þ þ þ 1
1  rR;0 ð1  R Þ μR Lj μM Lj Lj μL Lj
ð12bÞ

where rL,0 and rR,0 are the end fixity factors of the beam if it was uniform with EI = (EI)0,
and the subscripts L, M and R correspond to the left, middle and right segments of the beam.

2.3 Column rotational buckling load


The applicability of the lateral stiffness equation in Equation 6 is for 0 ≤ P ≤ Pu, where Pu is the
rotational buckling load of the column, which can be obtained by root finding procedures as
demonstrated in Ma & Xu (2019) or derived from the method of Hoblit (1951). Note that except
for lean-on columns (that is, columns with rl = ru = 0), the value of S in Equation 6 asymptotic-
ally approaches negative infinity as P approaches Pu. For lean-on columns, S becomes discon-
tinuous at P = Pu and the column buckles individually, regardless of the conditions of the
other columns in the frame. For P < Pu, the value of S decreases monotonically with increasing
axial loads. Pu is also the elastic buckling load of the column if it is braced at ends.

3 NUMERICAL EXAMPLE

Consider now the case of the two-bay frame shown in Figure 2. In a particular design fire
scenario, a localized fire in Bay 1 causes the peak member segment temperatures shown in the
figure. All segments extend a third of the length of each member. The frame is diagonally
braced in both sway directions, with the total bracing stiffness Sb taking various values, and is
assumed to be unaffected by the fire.
In this example, (EI)0 = E0I where E0 = 200 GPa since the moment of inertia remains uniform
in each member. The following sections are used: W310 × 60 (I = 129 × 106 mm4, A =
7,610 mm2), W410 × 67 (I = 245 × 106 mm4, A = 8,600 mm2) and W250 × 33 (I = 48.9 × 106
mm4, A = 4,570 mm2). Column 1 is rigidly connected to the ground (rl = 1) while Columns 2 and
3 are pinned to the ground (rl = 0), and the reference beam end fixity factors are assumed to be rL,
6
0 = rR,0 = 0.8 (corresponding to constant end connection stiffnesses of R = 80.38 × 10 Nm/rad).
In calculating the modulus of elasticity at elevated temperatures, the tangent modulus method

732
Figure 2. Three-segmented out-of-plumb column subjected to axial loading.

stipulated in Eurocode 3 (BSI, 2005) was adopted with an ambient yield stress of 350 MPa. The
frame is assumed to be thermally unrestrained, so no adjustments are made to the applied loads to
account for thermal restraining forces. The axial loads are distributed in proportion to tributary
area as shown. For values of Sb ranging between zero (no diagonal bracing) and infinity (full lat-
eral bracing), the elastic buckling load, P, of the frame was determined in each case by increasing
P until Equation 8 diminished to zero. A finite element analysis was conducted in ABAQUS to
verify the results using B23 Euler-Bernoulli element for each stepped segment, with calibrating the
value of v to represent the actual buckling shape, and the results were found to be exact. The
obtained axial load ratios of the columns imminent during their rotational buckling loads are plot-
ted in Figure 3 below. As the finite element results are exact to the proposed method, only one
curve is plotted in Figure 3.
It was found that without lateral bracing (Sb = 0), the critical load of the frame was
P = 695 kN, and that with full lateral bracing (Sb → ∞) the critical load of the frame
approached P → Pu = 1,000 kN, corresponding to an improvement of up to 44% in the critical
load if bracing were provided. In all cases, Column 1 was imminent to its rotational buckling
load during instability since it was subjected to the higher temperature. When the lateral bracing
stiffness reached a certain value, which in this example was about Sb = 106 N/m, the critical

Figure 3. Sway load ratio versus diagonal bracing stiffness of frame.

733
load of the frame would no longer be significantly improved with increase of the lateral bracing
stiffness. That is, the value of P at the point of failure is 986 kN with Sb = 106 N/m, which is
within 1.4% of the critical load of Column 1 being fully braced. On the other hand, as shown in
Figure 3, if less than Sb = 104 N/m were provided, the critical load does not improve signifi-
cantly compared to the unbraced case. The critical load in the case of Sb = 104 N/m is only
700 kN, corresponding to an improvement of only 0.7% from the unbraced case which has a
critical load of 695 kN.
Based on the results of the example, it is concluded that the capacity of a semi-braced frame
with segmented members will be within a certain threshold of the fully braced capacity once
a critical bracing stiffness is provided. In contrast, if the bracing stiffness is less that of the critical
one, the buckling capacity of the frame will be significantly influenced by the bracing stiffness. It
was also verified via finite element analysis that the proposed method yields an exact theoretical
solution for lateral bucking capacity of a semi-rigidly connected, semi-braced segmented frame.
Given that programming of the recommended computation procedures can easily be done, the
instability analysis of such frames can be completed with ease in less time than with utilizing
a finite element software package.

4 CONCLUSION

A method was proposed to evaluate the lateral stiffness of a semi-braced frame containing semi-
rigidly members containing up to three segments of varying material and/or cross-sectional prop-
erties. Derivations of the column lateral stiffness and rotational rigidity provided by connecting
beams were detailed. A numerical example was presented to demonstrate the use of the proposed
method towards the critical lateral load analysis of a two-bay frame containing segmented mem-
bers as a result of elevated segment temperatures in fire. The results show that the stiffness of the
diagonal bracing can significantly improve the buckling capacity of the frame, but only if
a sufficient bracing stiffness is provided. The capacity will also asymptotically approach that of
a fully braced frame as the bracing stiffness is increased. The results of the example were also val-
idated via finite element analysis and shown to be theoretically exact. The proposed method can
conveniently implemented in daily design tools such as Microsoft Excel for engineering practice.

REFERENCES

Arablouei, A. & Kodur, V.K.R. 2016. Effect of fire insulation delamination on structural performance of
steel structures during fire following an earthquake or an explosion. Fire Safety Journal 84(2016): 40–49.
Braxtan, N.L. & Pessiki, S.P. 2011. Postearthquake fire performance of sprayed fire-resistive material on
steel moment frames. Journal of Structural Engineering 137(9): 946–953.
BSI. 2005. BS EN 1993-1.2: 2005 Eurocode 3, Design of steel structures, Part 1.2: General rules – Struc-
tural fire design. London: British Standards Institution.
Castiglioni, C.A. 1984. Stepped columns: a simplified design method, Fritz Laboratory Reports, Paper 1481.
Hoblit, F.M. 1951. Buckling load of a stepped column. Journal of Aeronautical Sciences 18(2): 124–126.
Ma, T. and Xu, L. In Press, 2019. Buckling loads of semi-rigid columns subjected to non-linear temperature
distributions, Proceedings of the International Colloquium of Stability and Ductility of Steel Structures 2019,
Prague.
Thomson, W.T. 1949. Matrix solution of the n-section column, Journal of Aeronautical Sciences 16(10):
623–624.
Wang, W-Y., Li, G-Q. & Kodur, V. 2013. Approach for modeling fire insulation damage in steel
columns. Journal of Structural Engineering 139(4): 491–503.
Xu, L. 2001. The buckling loads of unbraced PR frames under non-proportional loading. Journal of Con-
structional Steel Research 58(2002): 443–465.
Xu, L. & Liu, Y. 2002. Storey-based effective length factors for unbraced PR frames. Engineering Journal
39(1): 13–29.
Xu, L. & Zhuang, Y. 2014. Storey stability of unbraced steel frames subjected to non-uniform elevated
temperature distribution. Engineering Structures 62-63(2014): 164–173.
Zhuang, Y. 2013. Storey-based stability analysis of unbraced steel frames at ambient and elevated tem-
peratures (PhD Thesis). Waterloo: University of Waterloo.

734
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

A quasi-static nonlinear analysis for assessing the fire resistance of


steel 3D frames exploiting time-dependent yield surfaces

D. Magisano, F. Liguori, L. Leonetti & G. Garcea


Dipartimento di Modellistica per l’Ingegneria, Università della Calabria, Rende, Italy

ABSTRACT: In this work, we evaluate the axial force-biaxial bending yield surface of steel
sections as a Minkowski sum of ellipsoids by a section analysis carried out once and for all.
Its time-dependent expression is derived accounting for the strength reduction of the mater-
ials, which is a function of the fire duration. The equilibrium state of 3D frames with such
yield conditions, once discretized using beam finite elements, is formulated as a nonlinear vec-
torial equation defining a curve in the hyperspace of the discrete variables and the fire dur-
ation. A generalized path-following strategy is proposed for tracing this curve. It can be seen
as an optimisation method, which furnishes a sequence of safe states at increasing fire dur-
ations according to the lower bound theorem of the limit analysis up to the limit fire duration,
that is the time of exposure which leads to the structural collapse. The methodology represents
a global nonlinear analysis for assessing the fire resistance of 3D frames, able to take account
of the stress redistribution. Numerical examples are given to illustrate and validate the
proposal.

1 INTRODUCTION

The evaluation of the carrying capacity of a structure im'plies not only situations of normal
service conditions but also exceptional loadings. An important aspect is to ensure the overall
structural integrity during fire events Buchanan and Abu (2017). Usually, frame structures
exhibit a relevant overstrength, that is their ultimate capacity can be significantly higher than
the elastic limit. For this reason, the material nonlinear analysis is a necessary tool for the
structural engineer. A widely employed approach formulates the cross-section yield criterion in
terms of generalized stresses, usually axial force and bending moments. For steel sections, the
strain limit is sufficiently large to allow the use the classical plasticity theory Sessa et al. (2018);
Leonetti et al. (2015). A point cloud of generalized yield stresses can be obtained by assigning
the corresponding collapse mechanisms, that is the position and orientation of the neutral axis
at the collapse states. The Minkowski sum Fogel and Halperin (2007) of ellipsoids represents
an interesting alternative for the approximation of particular convex shapes known as zonoids,
such as the cross-section yield surface Bleyer and De Buhan (2013); Leonetti et al. (2015);
Sessa et al. (2018); Magisano et al. (2018). A few ellipsoids can be used to describe the yield
surface of homogeneous and composite cross-sections with great accuracy, also in the case of
non-smooth shapes. The yield surface at an assigned fire duration can be obtained by simply
contracting the ambient temperature yield surface accounting for its strength reduction. The
time-dependent yield criteria can be easily used to check the building safety by means of local
strength checks of the sections. However, the significant ductility and overstrength of 3D build-
ings, allow a stress redistribution and can make the sectional check extremely conservative.
Although this fact is well known, a global fire analysis accounting for the stress redistribution
is not a standard approach of structural check and design, but is usually limited to specific
case studies as in MacNeill and McAllister (2008), which are often tackled with complex simu-
lations, unsuitable for the ordinary work of the structural engineer. To deal with this lack, in
this work we propose an efficient quasi-static nonlinear analysis for assessing the global safety

735
of 3D steel frames in conditions of fire based on beam finite elements. It is a strain-driven
incremental strategy which evaluates a sequence of safe states for an increasing fire duration.
The time-dependent yield surface together with a finite element beam model allows us to for-
mulate the equilibrium condition of the structure as a nonlinear system of equations defining
a curve in the hyperspace of the discrete variables and the fire duration. A generalized path-
following strategy is proposed for solving step-by-step the global nonlinear equilibrium equa-
tions. The methodology can be seen as an optimisation method Garcea and Leonetti (2011),
which furnishes a sequence of safe states at increasing fire durations according to the lower
bound theorem of the limit analysis up to the limit fire duration, that is the time of exposure
which leads to the structural collapse. At each step of the analysis the nonlinear internal forces
are obtained by an elastic predictor-return mapping process based on the closest point projec-
tion (CPP) scheme on the yield surfaces at the current fire duration Magisano et al. (2019).

2 MECHANICS OF STEEL SECTIONS IN FIRE

In this section, we describe the mechanical model for reinforced concrete sections in fire. In
particular, we define the strength reduction as a function of the fire duration and the section
yield surface in terms of axial force and bending moments.

2.1 Strength reduction for the steel


The fire temperature and, then, the temperature of the steel member depends on the fire dur-
ation. The heating leads to a reduction of the steel strength which becomes a function of the
fire duration t, that is

fyT ½t
¼ k½t
fy ð1Þ

where fy is the steel strength at ambient temperature and k½t


2 ½0; 1
is the strength reduction
factor. For simplicity, it is possible to assume a uniform temperature distribution over the
cross-section. In this case k represents the average strength reduction of the section. The
reduction factor vs fire duration law can be obtained following the Eurocode method summar-
ized in Buchanan and Abu (2017). A typical behavior for protected HE300A columns exposed
on all four sides and protected IPEA300 beams carrying a concrete floor slab and then
exposed on three sides is reported in Figure 1.

2.2 Section kinematics and statics


Let us consider a cylinder occupying a reference configuration B of length ‘ confined by the
lateral boundary denoted by ∂B and two terminal bases O0 and O‘ . The cylinder is referred to

Figure 1. Strength reduction vs fire duration.

736
a Cartesian frame ðO O; x1 ≡s; x2 ; x3 Þ with unit vectors fe1 ; e2 ; e3 g and e1 aligned with the cylin-
der axis. In this system, we denote with X ¼ se1 þ x the position of a point P, where s is an
abscissa which identifies the generic cross-section Os of the beam, while x ¼ x2 e2 þ x3 e3 is the
position of P inside Os .
The displacement field u½X
of the model is expressed, as usual, as a rigid motion of the
section

u ½X
¼ u0 ½s
þ j½s
^ x ð2Þ

where u0 ½s
and ’½s
are the mean translation and rotation of the section and the operator ^
denotes the cross product. The kinematics assumed in Eq. (2) allows us to evaluate, using
a linear Cauchy continuum, the stress-strain work W in terms of the generalized strains and
stresses on the section as
Z
W :¼ N ½s
T ε½s
þ M½s
T χ½s
Þds
ðN ð3Þ

where the generalized strains ε and χ are defined as

ε½s
¼ u0 ;s ½s
þ e1 ^ j½s
; χ ¼ j½s
;s ; ð4Þ

a comma stands for derivative and N ½s


¼ fN1 ; N2 ; N3 g and M½s
¼ fM1 ; M2 ; M3 g are the
resultant force and moment. Finally, the elastic constitutive law is expressed as

ε N FNN FNM
¼F ;F¼ ð5Þ
χ M F TNM FMM

where the coefficients of the cross-section compliance matrix F can be obtained as in Leonetti
et al. (2015); Garcea et al. (2016).

2.3 The cross-section yield surface


Following we denote with O the beam section domain with ðx2i ; x3i Þ its coordinates. The
material is assumed to be elastic-perfectly plastic with the plastic admissibility condition
expressed in terms of normal stress only as fyT  σ11  fyT , where the normal stress is assumed
positive in tension. fyT depends on the fire duration t according to (2.1). Omitting the depend-
ence on s for a clearer exposition, we introduce the plastic mechanism of the cross-section as

n ¼ f2_ 1 ; χ_ 2 ; χ_ 3 g ð6Þ

which defines the position and orientation of the neutral axis for the collapse state from the
condition

ε_ 11 ≡2_ 1 þ x3 χ_ 2  x2 χ_ 3 ¼ 0: ð7Þ

The yield stress vector τy collecting the generalized section resultants associated with n by the
Drucker condition, at a given fire duration t, is

2 3 8 R
Ny1 >
< Ny1 ¼ k½t
fy Ωc dOc
R
τy ½n; t
¼ 4 My2 5with My2 ¼ k½t
fy Ωc x3 dOc
R ð8Þ
>
: My3
My3 ¼ k½t
fy Ωc x2 dOc

737
Figure 2. HE300A: Minkowski sum and yield points in the normalized space.

Equation (8) allows the evaluation, for an assigned fire duration t, of the set of generalized
yield stress τyk ½t
≡ τy ½nk ; t
associated to the mechanism nk , simply by assuming uniaxial stress
fields reaching their maximum strength capacity in each region, either in tension or in
compression.

2.4 Construction of yield surface at ambient temperature as a Minkowski sum of ellipsoids


Omitting the dependence on the fire duration, the cross-section yield points at ambient tem-
perature can be fitted using a Minkowski sum of ellipsoids as described in Magisano et al.
(2018), that is

X CI n
τ ½ n
¼ τI ½n
τI ½ n
¼ cI þ pffiffiffiffiffiffiffiffiffiffiffiffiffiffi : ð9Þ
I nT CI n

with τI ½n
representing the contribution of a portion of the cross-section to the overall surface
τ½n
for a generic normal vector n, whose components are the variables of the parametrization.
An example of fitting is reported in Figure 2. The results of the numerical tests section are
referred to approximations based on nine ellipsoids.

2.5 Account of the time-dependent strength reduction


The time-dependent yield surface can be then obtained by assigning a scaling law for the sur-
face at ambient temperature as a function of the fire duration.
X
τ½n; t
¼ k½t
τI ½n
: ð10Þ
I

Note that the approach is possible also for nonuniform temperature distributions over the sec-
tion by considering different kI ½t
Magisano et al. (2019).

3 THE FINITE ELEMENT QUASI-STATIC ANALYSIS FOR 3D FRAMES


SUBJECTED TO FIRE

In the following the finite element beam model for the incremental fire analysis is described.

3.1 The 3D beam finite element


The beam finite element adopted (see Leonetti et al. (2015)) uses an interpolation of the gener-
N ; M
T ¼ Dt ½s
βe , where the interpolation matrix Dt ½s
is obtained satisfying
alized stresses ½N

738
the equilibrium equations on the element for zero body forces exactly. Body load effects are
then included exactly as a “particular solution”. This means that N and the torsional moment
component M1 are constant, while the two flexural components M2 ½s
and M3 ½s
of M½s
are
linear with s and linked to the shear resultants so that N2 ‘ ¼ ðM3 ½‘
 M3 ½0
Þ and
N3 ‘ ¼ ðM2 ½‘
 M2 ½0
Þ. The internal work becomes

W ≡ N T ðu0 ½‘
 u0 ½0
Þ þ M½‘
T j½‘
 M½0
T j½0
¼ dTe QTe βe ð11Þ

allowing us to directly obtain the discrete form of W without any FEM interpolation for the
kinematic variables. The vectors collecting the kinematics de and static βe finite element gener-
alized parameters and the compatibility operator Qe are defined as
2 3 2 3
N 2 3 ‘ e T1 0 ‘ e T1 0
6 M2 ½0
7 u0 ½0
6 eT ‘ eT2 e T3 0 7
6 7 6 3 7
6 M3 ½0
7 6 j ½0
7 1 6  e T2 ‘ e 3 e T2 0 7
βe ¼ 6 7
6 M2 ½‘
7; de ¼ 6 7
4 u0 ½‘
5; Qe ¼ 6 7: ð12Þ
6 7 ‘66 e3
T
0 e T3 ‘ e T2 7
7
4 M3 ½‘
5 j ½‘
4 eT 0 e T2 ‘ e T3 5
2
M1 0 ‘ eT1 0 ‘ e T1

3.2 The linear elastic problem


The linear elastic problem can be formulated as the stationarity of the Hellinger-Reissner
functional HR that at the element level can be written as

1
HR ¼ dTe QTe βe  βTe Fe βe  dTe pe
2

where pe is the element contribution of the external loads and the elastic compliance matrix of
the element Fe is obtained from the equivalence

Z T ! Z
N FNN FNM N
ds ¼ βTe Fe βe ; Fe ¼ Dt ½s
T F Dt ½s
ds: ð13Þ
‘ M FTNM FMM M ‘

The stationarity of HR with respect to the stress variables furnishes the discrete elastic consti-
tutive law

βe ½d e
¼ E e Qe de with Ee ¼ F1
e ð14Þ

which allows us to express the elastic problem in terms of displacement variables only.
The stationarity condition with respect to d e furnishes the equilibrium equations on the
element as

QTe βe ½de
 pe ¼ 0 ð15Þ

which, in the elastic case, become

Ke de  pe ¼ 0 with Ke ¼ QTe Ee Qe :

739
3.3 Stress update for time-dependent yield conditions
An elastic behavior with respect to shear effects as well as torsion is assumed. The section
yield function f ½s; t; τ½s

is then defined in a 3D space involving axial force N1 and bending


moments M2 and M3 collected in vector τ½s
¼ fN1 ; M2 ; M3 g. The plastic admissibility condi-
tion on the cross-section s then becomes

f ½s; t; τ½s

 0: ð16Þ

The update of the stress is obtained, in a strain-driven way, by means of a closest point projec-
tion (CPP) which corresponds to a backward Euler scheme for integrating the constitutive
law. Starting from a known state de ½t0
; βe ½t0
the stress parameters βe ½ βe ½t0
; t; Δde
for an
assigned displacement increment Δde and a given fire duration t are obtained by solving, for
each element, the optimization problem
 T  
minimize 1
2 βe  βe Fe βe  βe
ð17Þ
subject to f ½0; t; τ½0

 0
f ½‘; t; τ½‘

 0

where βe ¼ βe ½t0


þ Ee Qe Δde is the elastic predictor and, when the admissibility condition
is checked on the beam end nodes only, the generalized normal stress vectors of the two
end sections are extracted directly from βe as τ½0
¼ P0 βe and τ½‘
¼ P‘ βe by the extrac-
tion operators P0 and P‘ . Note that since the mixed finite element always satisfies the
local equilibrium, the stresses at both end sections are coupled with each other and,
then, the CPP has to be performed at the element level. In this case the element equilib-
rium equations (15) modify as

QTe βe ½ βe ½t0
; t; Δde
 pe ¼ 0: ð18Þ

A specialized stress update strategy for the Minkowski description of the yield condition
can be adopted Magisano et al. (2018). The main idea is to reformulate the CPP prob-
lem using the normal vector as primal unknown. This is convenient since the dimension
of the problem is fixed and independent of the number of ellipsoids used in the approxi-
mation. An elastic predictor-return mapping scheme is used. Both the admissibility check
and the finite element return mapping are formulated as small nonlinear systems of
equations of fixed size.

3.4 Incremental nonlinear fire analysis


Once the finite element assemblage has been carried out, the equilibrium condition of a RC
building subjected to fire can be written as

r½d; t
¼ s½d; t
 p ¼ 0 ð19Þ

This system of nonlinear equations represents a curve in the hyperspace d-t, which can be
traced in a path-following manner.
The curve can exhibit a limit fire duration, that is the time of exposure which leads to
structural collapse. For this reason it is not convenient to use a time controlled scheme
since Eq. (19) could not have a solution, that is no equilibrium state, for a given fire dur-
ation. We propose instead the use of a generalized arc-length method. The equilibrium
equations are completed with the additional constraint g½d; t
 ξ ¼ 0, which defines
a surface in ℝNþ1 . Assigning successive values to the control parameter ξ ¼ ξ ðkÞ the solu-
tion of the nonlinear system

740
2 3
r ½ d; t

R½ξ
≡4 g½ d; t
 ξ 5 ¼ 0 ð20Þ

defines a sequence of points (steps) zðkÞ ≡ fdðkÞ ; tðkÞ g belonging to the equilibrium path. Start-
ing from a known equilibrium point z0 ≡ zðkÞ , the new one zðkþ1Þ is evaluated correcting a first
extrapolation z1 ¼ fd1 ; t1 g by a sequence of estimates z j by a Newton–Raphson iteration

J Δz ¼  Rj
ð21Þ
z jþ1 ¼ z j þ Δz

where Rj ≡R½zj
and J is the Jacobian of the non-linear system (20) at z j or its suitable esti-
mate. The simplest choice for g½d; t
is the linear constraint corresponding to the orthogonal
hyperplane

nd ≡ M ð d j  dðkÞ Þ
nTd ðd  d j Þ þ nt ðt  t j Þ ¼ Δξ where ð22Þ
nt ≡ μ ðtj  tðkÞ Þ

M and μ being some suitable metric factors Magisano et al. (2017a,b), Δξ an assigned incre-
ment of ξ and

∂R½z
Kt st
J≡ ¼ : ð23Þ
∂z z j nTd nt

The time-controlled scheme can be recovered assuming g½d; t


¼ t, but it is not convenient as
previously discussed. The solution of Eq.(21) is conveniently performed in a partitioned way
as follows
(
n T K rj
Δt ¼ nt dnT tKt st
d ð24Þ
Kt Δ d ¼ Δts t  rj

in order to exploit the symmetry and the band structure of the tangent stiffness matrix Kt . The
points zðnÞ evaluated by the scheme are, by definition, equilibrated and plastically admissible
at time tðnÞ . In other words, they satisfy the hypotheses of the lower bound theorem of the
limit analysis. Furthermore, in Magisano et al. (2019), it was proven that when
tðnþ1Þ  tðnÞ ¼ 0 the structure is just at the point of failure because the hypotheses of the upper
and lower bound theorems of the limit analysis are satisfied simultaneously.

4 NUMERICAL TESTS ON 3D BUILDINGS IN FIRE

The time-dependent yield surfaces and the incremental strategy are employed for assessing the
safety of steel structures exposed to fire.

4.1 A simple frame


This example regards the simple 3D frame reported in Figure 1. The load over the floor is
uniformly distributed over the four beams. The fire scenario considered is the one described in
Figure 1.
In Figure 4 the fire duration-displacement curve for the assigned distributed load is
reported. The curve is characterized by a significant inital portion with zero displacements.
This means that the load is largely inside the initial domain at ambient temperature. 1:75

741
Figure 3. Simple 3D frame: geometry, loads and cross-sections.

Figure 4. Simple 3D frame: equilibrium paths.

hours are required to observe the first plastic deformations while the limit fire duration is
equal to 2:23. Different cross-section discretizations are employed. The results provided by
our fire analysis in terms of limit fire duration are assessed by means of a comparison with
a standard elasto-plastic analysis where the yield surfaces are kept constant and the load is
amplified by a factor λ. In particular, we show that the limit load provided by the elasto-
plastic analysis coincides with the applied one (λ ¼ 1) when the yield surface of the fire
exposed sections is evaluated at the limit fire duration.

5 CONCLUSIONS

In this work, we showed a numerical procedure for constructing the axial force-biaxial bend-
ing yield surface of steel sections in fire in a simple, accurate and efficient way. The strategy is
based on a particular Minkowski sum of ellipsoids. The yield conditions depends on the fire
durations. The equilibrium condition of the structure, once discretized using beam finite elem-
ents, has been formulated as a nonlinear vectorial equation defining a curve in the hyperspace
of the discrete variables and the fire duration. We proposed a strain-driven incremental strat-
egy for tracing this curve in a path-following quasi-static manner. This kind of analysis takes
account of the stress redistribution and provides the limit duration, that is the time of expos-
ure which leads to the structural collapse. The step-by-step procedure furnishes a sequence of

742
safe states at increasing fire durations according to the lower bound theorem of the limit ana-
lysis. When a limit fire duration exists, the approach can be framed as an optimisation prob-
lem very similar to the static limit analysis one. The main difference is that the loads are kept
constant and the fire duration, which leads to a contraction of the yield surfaces, replaces the
load factor as objective function. Future developments will focus on accounting for other sig-
nificant structural behaviors as the effects of large deformations Garcea et al. (2017).

REFERENCES

J. Bleyer and P. De Buhan. Yield surface approximation for lower and upper bound yield design of 3D
composite frame structures. Computers and Structures, 129:86–98, 2013.
Andrew H. Buchanan and Anthony K. Abu. Structural design for fire safety. Wiley, United Kingdom,
second edition, 2017.
E. Fogel and D. Halperin. Exact and efficient construction of Minkowski sums of convex polyhedra with
applications. CAD Computer Aided Design, 39(11):929–940, 2007.
G. Garcea and L. Leonetti. A unified mathematical programming formulation of strain driven and inter-
ior point algorithms for shakedown and limit analysis. International Journal for Numerical Methods in
Engineering, 88(11):1085–1111, 2011. doi: 10.1002/nme.3188.
G. Garcea, R. Gonçalves, A. Bilotta, D. Manta, R. Bebiano, L. Leonetti, D. Magisano, and
D. Camotim. Deformation modes of thin-walled members: A comparison between the method of gen-
eralized eigenvectors and generalized beam theory. Thin-Walled Structures, 100:192–212, 2016. doi:
10.1016/j.tws.2015.11.013.
G. Garcea, L. Leonetti, D. Magisano, R. Gonçalves, and D. Camotim. Deformation modes for the
post-critical analysis of thin-walled compressed members by a Koiter semi-analytic approach. Inter-
national Journal of Solids and Structures, 110-111: 367–384, 2017. doi: 10.1016/j.ijsolstr.2016.09.010.
Leonardo Leonetti, Raffaele Casciaro, and Giovanni Garcea. Effective treatment of complex statical and
dynamical load combinations within shakedown analysis of 3D frames. Computers & Structures,
158:124–139, 2015. ISSN 0045-7949. doi: 10.1016/j.compstruc.2015.06.002.
Robert S. MacNeill and Therese P. McAllister. Global Structural Analysis of the Response of World
Trade Center Building 7 to Fires and Debris Impact Damage, Federal Building and Fire Safety Investiga-
tion of the World Trade Center Disaster (NIST NCSTAR 1-9A). 2008.
D. Magisano, L. Leonetti, and G. Garcea. How to improve efficiency and robustness of the Newton
method in geometrically non-linear structural problem discretized via displacementbased finite
elements. Computer Methods in Applied Mechanics and Engineering, 313: 986–1005, 2017a. doi:
10.1016/j.cma.2016.10.023.
D. Magisano, L. Leonetti, and G. Garcea. Advantages of the mixed format in geometrically nonlinear
analysis of beams and shells using solid finite elements. International Journal for Numerical Methods in
Engineering, 109(9):1237–1262, 2017b. doi: 10.1002/nme.5322.
D. Magisano, F.S. Liguori, L. Leonetti, and G. Garcea. Minkowski plasticity in 3D frames: Decoupled
construction of the cross-section yield surface and efficient stress update strategy. International Journal
for Numerical Methods in Engineering, 116(7):435–464, 2018. doi: 10.1002/nme.5931.
D. Magisano, F.S. Liguori, L. Leonetti, D. de Gregorio, G. Zuccaro, and G. Garcea. A quasi-static non-
linear analysis for assessing the fire resistance of reinforced concrete 3d frames exploiting
time-dependent yield surfaces. Computers & Structures, 212:327–342, 2019. ISSN 0045-7949. doi:
10.1016/j.compstruc.2018.11.005.
S. Sessa, F. Marmo, L. Rosati, L. Leonetti, G. Garcea, and R. Casciaro. Evaluation of the capacity sur-
faces of reinforced concrete sections: Eurocode versus a plasticity-based approach. Meccanica, 53
(6):1493–1512, Apr 2018. ISSN 1572-9648. doi: 10.1007/s11012-017-0791-1.

743
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Fire design of class 4 tapered steel beams with the general


method—a proposal

É. Maia, C. Couto, P. Vila Real & N. Lopes


RISCO – Department of Civil Engineering, University of Aveiro, Aveiro, Portugal

ABSTRACT: The EN1993-1-2 provides several approaches to the stability verification of


steel members but no fire design rules are provided for the case of tapered steel members.
With that in mind, current proposals/design methodologies for tapered beams at normal tem-
perature have been adapted to fire in a previous work by the authors. Out of all approaches
analysed, the modified General Method was the one recommended for the safety-check of
a tapered beam in fire, although the methodology was deemed overly conservative.
The conservativeness of the adaptation lied in its inability to correctly take into account the
loading conditions and the resistance provided by the taper. As such, the present work aims at
expanding upon these conclusions by proposing improvements to the precursor adaptation in
the abovementioned areas. These minor adaptations to the EN 1993-1-2 buckling curve have
proven to be effective at improving the accuracy of the methodology.

1 INTRODUCTION

Non-uniform (tapered) steel members are widely used in steel construction due to their struc-
tural efficiency as material savings can be achieved. The existing fire design methodologies of
the EN 1993-1-2 (EC3) (CEN, 2005a) are limited to uniform members and no rules are pro-
vided for non-uniform members. At room temperature, although some guidelines are given in
Annex B of the EN 1993-1-5 (CEN, 2006), the out-of-plane stability check of non-uniform
members should be performed using the General Method (GM) given in clause 6.3.4 of
EN1993-1-1 (CEN, 2005b).
However, this methodology has been shown to not be reliable in some cases (Marques
et al., 2014) and scientific consensus on its use for the design of non-uniform beams has not
yet been reached. Regardless, the General Method is the only option for the verification of
non-prismatic members at normal temperature according to the EN 1993-1-1 (CEN, 2005a).
For the design in fire situation, the verification is unclear since no definition of
a methodology is provided in the fire part of the norm (CEN, 2005a). Preliminary investiga-
tions on the adaptation of the General Method for the fire design of tapered beams (Couto
et al., 2014a) and beam-columns (Couto et al., 2015) have demonstrated safe results but, in
those references, only one bending diagram was considered. In the scope of the project Taper-
Steel (FCT, 2012), some of the limitations/inconsistencies of tapered member design have
been addressed by Marques et al. (2013), leading to the proposal of a new thorough design
methodology for the stability of tapered beams at normal temperature. The latter, along with
the General Method, has recently been adapted to fire by the authors (Couto et al., 2019) for
the fire design of tapered beams but was deemed unsafe at elevated temperatures. On the
other hand, the GM produced safe predictions of resistance under these conditions.
Although the abovementioned adaptation of the GM was on the safe side, undesirable levels
of conservativeness have been noted by the authors. With that in mind, this investigation aims
at expanding upon the work of Couto et al. (2019) by proposing and validating improvements
to the original adaptation of the GM, in order for it to more accurately perform the safety
check of tapered beams at elevated temperatures. This is done, first, by including, in the original

744
formulation, the reduction of steel material properties with the temperature. Then, a parametric
study is defined for the two reference cross-sections investigated by Couto et al. (2019), for dif-
ferent temperatures, taper ratios and loading conditions. The ultimate capacity of the tapered
beams in fire has been obtained by means of geometrically and materially non-linear analyses
with imperfections (GMNIA) in SAFIR (Franssen and Gernay, 2017) using shell finite elements
and considering local imperfections in the numerical model. Finally, as the conservativeness of
the adaptation lied in its inability to correctly take into account the loading conditions and the
added resistance provided by the taper, improvements to the out-of-plane buckling curve of the
EN 1993-1-2 (CEN, 2005a) are proposed and validated.
With the improved buckling curve, the General Method is, on average, 9% more accurate at
predicting the stability of tapered beams in fire with a lower scatter of results. Furthermore,
the proposal fulfils the three reliability criteria of Kruppa (1999), which attests its robustness.

2 GENERAL METHOD

2.1 Normal temperature


The stability of uniform beams in EC3-1-1 (CEN, 2005b) is checked by the application of
clause 6.3.2. Regarding the stability of a non-uniform member, this clause is not applicable as
the evaluation of the buckling resistance of such members lies outside the range of application
of the interaction formulae. For those cases, verification should be performed using clause
6.3.4 (denoted as General Method or GM).
The out-of-plane stability of members is verified by ensuring that condition (1) is verified.
χop  αult;k
1 ð1Þ
γM1

where χop is the reduction factor taken from the EN 1993-1-1 (CEN, 2005b) for the non-
dimensional slenderness λop , obtained in (3), and analytically describes the reduction in resist-
ance due to out-of-plane (“op”) instability phenomena. αult;k is the minimum load amplifier of
the design loads in order to reach the characteristic resistance of the most critical cross-section
of the structural component considering only the in-plane behaviour of the member. For
a beam it can be obtained with Eq. (2).
My;Rk
αult;k ¼ ð2Þ
My;Ed
rffiffiffiffiffiffiffiffiffiffi
λop ¼ αult;k ð3Þ
αcr;op

where αcr;op is the minimum amplifier for the in-plane design loads to reach the elastic critical
resistance of the structural component in regards to the lateral/lateral-torsional buckling.

2.2 Adaptation to fire


The adaptation of this methodology to fire is done by accounting for the reduction in steel
material properties with the temperature, namely the yield strength fy;θ and Young’s Modulus
Ea;θ with the reduction factors ky;θ and kE;θ , respectively. Then, given the direct proportional-
ity of the load amplifier to the yield strength and of the critical amplifier to the Young’s
Modulus, the following expressions can be deduced:
rffiffiffiffiffiffiffiffiffiffiffiffiffi sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi sffiffiffiffiffiffiffiffi
λop;θ αult;θ;k ky;θ  αult;k  ky;θ
¼ ¼ ¼ λop  ð4Þ
αcr;θ;op kE;θ  αcr;op kE;θ

745
The load amplifier of the in-plane design loads – for the fire situation – to reach the character-
istic resistance in fire situation may be determined with Eq. (5).

My;θ;Rk
αult;θ;k ¼ ð5Þ
My;fi;Ed

The final condition to be fulfilled in fire situation for the General Method is thus,

χop;fi  αult;θ;k
1 ð6Þ
γM;fi

The reduction factor χop;fi , for the particular case of lateral-torsional buckling (LTB) in fire situ-
ation can be calculated with Eq. (7) following the principles of the EN 1993-1-2 (CEN, 2005a),
and taking λLT;θ ¼ λop;θ , as

1
χop;fi ¼ χLT;fi ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð7Þ
LT;θ þ 2LT;θ  λ2LT;θ

with

1  qffiffiffiffiffiffiffiffiffiffiffiffiffiffi
LT;θ ¼  1 þ αλLT;θ þ λ2LT;θ with α ¼ 0:65  235=fy ð8Þ
2

3 NUMERICAL MODEL

The Geometrically and Material Non-Linear Analyses with Imperfections (GMNIA) were
performed using the finite element software SAFIR (Franssen and Gernay, 2017). After
a sensitivity analysis, the chosen numerical model had a mesh with 12 elements in the flange,
18 in the web and 100 along the length. The loads were applied to the model using nodal
forces and “Fork-support” conditions were modelled by restraining the vertical displacements
of the bottom flange and the out-of-the plane displacements of the web. The present numerical
model has previously been validated experimentally (Prachar et al., 2016) and used by the
authors (Couto et al., 2016; Maia et al., 2016).
The geometric imperfections have been introduced in the model by changing the nodal
coordinates to represent the worst scenario for the assessment of member resistance.
This has been considered as the shape given by the modes of a linear buckling analysis
(LBA) calculated with the software Cast3M (CEA, 2015). In addition, the recommenda-
tions given in the Annex C of EN1993-1-5 (CEN, 2006) were followed: i) a combination
of local (Figure 1) and global (Figure 2) modes were considered by taking the lowest
mode as the leading imperfection and the other one reduced to 70%, ii) the amplitude of
the imperfections was chosen as 80% of the fabrication tolerances given in the EN1090-2
(CEN, 2011), accordingly, an amplitude of 80% of L/750 for the global mode and 80%
of b/100 for the local mode was used, where L is the member span and b is the flange
width or web height, depending on which the maximum displacement occurs in the cor-
responding buckling mode. For the material imperfections, the pattern of residual stres-
ses of welded cross-sections used by Maia et al. (2016) was included in the numerical
simulations, following the recommendations of the ECCS (1984, 1976). An example col-
lapse shape for a tapered member subjected to a triangular bending moment (ψ ¼ 0) is
presented in Figure 3.

746
Figure 1. Local eigenmode for a γh ¼ 4 IPE200.

Figure 2. Global eigenmode for a γh ¼ 4 IPE200.

Figure 3. Collapse shape for a γh ¼ 4 HEB300 member subjected to a ψ ¼ 0 end moment at 550ºC.

747
4 NUMERICAL STUDY

4.1 Parametric analysis


Table 1 specifies the parameters considered in the numerical study performed in the scope of
the original work of Couto et al. (2019), which have been chosen based on the investigation
done in Marques et al. (2013). As in the former, steel grade S235 has been selected. Also, four
loading conditions have been considered in respect to the bending diagrams, namely ψ ¼ 1,
ψ ¼ 0 and ψ ¼ 1 for the relation between end moments, as well as uniformly distributed load
(designated as “UDL”).
Since the treatment of shear failure is out of the scope of the present work, the numerical
simulations presenting shear-driven collapse have not been considered.

4.2 Accuracy of the GM for the design of tapered beams in fire (Couto et al., 2019)
The accuracy of the adapted General Method has previously been assessed by the authors
(Couto et al., 2019) by comparing the reduction factor for lateral-torsional buckling χMethod;θ ,
obtained in Eq. (7), against its numerically obtained counterpart χFEM;θ given by expression (9).
MUlt
χFEM ¼  ð9Þ
MRd xIc;M

where MRd xIc;M is the cross-sectional bending moment resistance at the first-order critical
cross-section, which corresponds to the one with the maximum utilization ratio. Also, given
the expected high tendency to local buckling in the case of fire (Maraveas, 2018), when appro-
priate, the cross-sectional resistances have been obtained according to the procedure proposed
in Couto et al. (2014b) for Class 3 and 4 cross-sections.
The comparison of the adapted GM against the numerical results of Couto et al. (2019) is
shown in Figure 4. Values below the solid line represent an underestimation of out-of-plane resist-
ance by the GM, compared to the numerical results, and thus are considered to be on the safe
side. In that work, the authors have demonstrated undesirable levels of conservativeness mainly
in terms of the methodology predicting the influence of different loadings, as well as taper ratios.

4.3 Proposal
In this sub-section, the proposed improvements to the General Method (specifically in terms
of the EN 1993-1-2 buckling curve) are presented. This has been done by adding a slenderness
plateau and adjusting it according to both the taper ratio, as well as the loading typology.
Expression (8) is adjusted as follows:

1    qffiffiffiffiffiffiffiffiffiffiffiffiffiffi
LT;θ ¼  1 þ α  λLT;θ  λLT;0;θ þ λ2LT;θ with α ¼ 0:65  235=fy ð10Þ
2

Table 1. Parameters considered in the analyses.


Height-taper Bending  
Reference cross-section ratio diagrams Temperature Member length λLT xI
c

- - - (ºC) (m) -

IPE 200 γh ¼1 ψ ¼ 1 350 0.5 3.5 6.5 10.0 [0;4]


HEB 300 γh ¼2 ψ¼0 450 1.0 4.5 7.0 12.0
γh ¼3 ψ¼1 550 2.5 5.0 8.0 13.0
γh ¼4 UDL 700 3.0 6.0 9.0 14.0

748
where the new slenderness plateau λLT;0;θ is obtained as follows:
8
>
> 0:1γh þ 0:2 for ψ ¼ 1
<
λLT;0;θ 0:14 for ψ ¼ 0
¼ ð11Þ
> 0:2γh  0:2
> for ψ ¼ 1
:
0:05γh þ 0:08 forUDL

A similar comparison to that of Figure 4 is presented in Figure 5 for the General Method with
the improved EN 1993-1-2 out-of-plane buckling curve in fire, according to Eq. (10). The
diagonal comparison shows that most data points are approximately less than 10% on the
unsafe side and, only in some instances (of high values of slenderness) above 25% on the safe
side, therefore displaying a satisfactory agreement against numerical results.
Figures 6 and 7 illustrate the improvements to the current EN 1993-1-2 out-of-plane buck-
ling curve using χ  λ plots for tapered beams subjected to bi-triangular and uniform bending
moments, respectively. For these two specific sub-sets, the proposal follows the numerical
results more closely than the EN 1993-1-2 which is overly conservative.

Figure 4. Accuracy of the GM compared to the numerical results of Couto et al. (2019).

Figure 5. Comparison of the GM with the proposed buckling curve against numerical results.

749
Figure 6. Comparison of the GM and the proposal against numerical results for γh ¼ 3 beams subjected
to a bi-triangular bending moment distribution.

Figure 7. Comparison of the GM and the proposal against numerical results for γh ¼ 4 beams subjected
to a uniform bending moment distribution.

4.4 Statistical analysis


In order to better understand the significance and the variability of the numerical data,
a general comparison between the adapted General Method from Couto et al. (2019) and the
present proposal are presented and organized in Table 2. In this table, the ratio
χMethod;θ =χFEM;θ is analysed, meaning that a ratio lower than 1.00 represents a safe result for
the respective approach, in comparison to FEM results. A similar analysis providing a more
in-depth evaluation of the proposal is presented in Table 3.
The mean (“Mean” of Table 2) is on the safe side and 9% closer to the reference resistance
and a lower standard deviation is also obtained for the proposal, in comparison to the General

750
Table 2. Statistical evaluation of the ratio χMethod;θ =χFEM;θ for the GM and the proposal.
Sub-set/case N. of cases Mean St. Dev. Coef. Var. Min. Max. %Unsafe %(>1.15)

General Method 1608 0.81 0.082 10.1% 0.60 1.08 0.6% 0.0%
Proposal 1608 0.90 0.076 8.4% 0.63 1.08 6.4% 0.0%

Table 3. Statistical evaluation of the ratio χProposal;θ =χFEM;θ – analysis by sub-sets.


Sub-set/case N. of cases Mean St. Dev. Coef. Var. Min. Max. %Unsafe %(>1.15)

All 356 0.93 0.049 5.3% 0.76 1.05 4.5% 0.0%


γh = 1 106 0.93 0.048 5.1% 0.76 1.05 4.7% 0.0%
ψ = −1
γh = 2 96 0.91 0.041 4.5% 0.79 1.00 1.0% 0.0%
γh = 3 82 0.93 0.045 4.8% 0.78 0.98 0.0% 0.0%
γh = 4 72 0.94 0.060 6.4% 0.78 1.03 13.9% 0.0%
All 404 0.88 0.081 9.3% 0.63 1.08 7.2% 0.0%
γh = 1 113 0.83 0.052 6.3% 0.66 0.93 0.0% 0.0%
ψ=0
γh = 2 110 0.91 0.069 7.6% 0.63 1.05 8.2% 0.0%
γh = 3 94 0.92 0.084 9.2% 0.76 1.08 17.0% 0.0%
γh = 4 87 0.84 0.072 8.5% 0.73 1.06 4.6% 0.0%
All 444 0.88 0.080 9.0% 0.64 1.08 4.1% 0.0%
γh = 1 125 0.91 0.077 8.5% 0.64 1.08 8.0% 0.0%
ψ=1
γh = 2 112 0.85 0.051 5.9% 0.71 0.96 0.0% 0.0%
γh = 3 111 0.87 0.074 8.5% 0.64 0.97 0.0% 0.0%
γh = 4 96 0.90 0.100 11.1% 0.64 1.05 8.3% 0.0%
All 404 0.92 0.073 8.0% 0.75 1.06 9.9% 0.0%
γh = 1 112 0.91 0.067 7.3% 0.76 1.04 6.3% 0.0%
UDL γh = 2 100 0.93 0.069 7.4% 0.79 1.06 13.0% 0.0%
γh = 3 97 0.92 0.073 7.9% 0.78 1.04 11.3% 0.0%
γh = 4 95 0.91 0.081 8.9% 0.75 1.03 9.5% 0.0%

Method. Furthermore, the following three reliability criteria laid out by Kruppa (1999) are
fulfilled by the proposal:
1. The mean is on the safe side;
2. No result of the proposal is more than 15% on the unsafe side (“>1.15%”);
3. Less than 20% of the points are unsafe (“>1.00%”) for every sub-set.

5 CONCLUSIONS

This paper describes a numerical/parametric investigation on the proposal of the General


Method for the stability check of unrestrained tapered steel beams in fire. Expanding upon
the findings of a previous work by the authors (Couto et al., 2019), in regards to the adapta-
tion of the General Method to fire, an improvement to the methodology has been proposed in
the scope of this work. The following conclusions can be taken from this study:
• The adaptation of the General Method to fire has yielded overly conservative results
(Couto et al., 2019) for the safety check of tapered beams, specifically regarding the inabil-
ity of the methodology to correctly account for the added resistance provided by the taper,
as well as for different loadings;
• In the scope of this work, a parametric study similar to that of Couto et al. (2019) was
defined and the ultimate capacity of tapered beams was obtained using GMNIA simula-
tions in SAFIR (Franssen and Gernay, 2017);

751
• In order to improve the EN 1993-1-2 out-of-plane buckling curve, a slenderness plateau
has been added, based on the taper ratio and the bending diagram;
• Although minor, these modifications have proven to be effective at improving the accuracy
of the methodology, as attested by the statistical analysis. Additionally, the proposal fulfils
the three reliability criteria of Kruppa (1999).

ACKNOWLEDGEMENTS

This research work was partially funded by the Portuguese Government through the FCT
(Foundation for Science and Technology) under the PhD grant SFRH/BD/114838/2016
awarded to the first author and the Post-doc grant SFRH/BPD/114816/2016, awarded to
the second author.

REFERENCES

CEA, 2015. CAST 3M is a research FEM environment; its development is sponsored by the French
Atomic Energy Commission <http://www-cast3m.cea.fr/>.
CEN, 2011. EN 1090-2:2008+A1, Execution of steel structures and aluminium structures - Part 2: Tech-
nical requirements for steel structures.
CEN, 2006. EN 1993-1-5, Eurocode 3: Design of steel structures - Part 1-5: General rules - Plated struc-
tural elements. Belgium.
CEN, 2005a. EN 1993-1-2, Eurocode 3: Design of steel structures - Part 1-2: General rules - Structural
fire design. Belgium.
CEN, 2005b. EN 1993-1-1, Eurocode 3: Design of steel structures - Part 1-1: General rules and rules for
buildings. Belgium.
Couto, C., Duarte, P., Vila Real, P., Lopes, N., 2015. Verification of web tapered beam-columns in case
of fire using the general method of Eurocode 3, in: IFireSS 2015. Coimbra, pp. 79–86.
Couto, C., Maia, É., Vila Real, P., Lopes, N., 2019. Stability check of tapered steel beams in fire.
J. Struct. Fire Eng. https://doi.org/10.1108/JSFE-01-2019-0002.
Couto, C., Vila Real, P., Ferreira, J., Lopes, N., 2014a. Numerical validation of the General Method for
structural fire design of web tapered beams, in: Eurosteel 2014.
Couto, C., Vila Real, P., Lopes, N., Zhao, B., 2016. Numerical investigation of the lateral torsional buck-
ling of beams with slender cross-sections for the case of fire. Eng. Struct 106, 410–421.
Couto, C., Vila Real, P., Lopes, N., Zhao, B., 2014b. Effective width method to account for the local
buckling of steel thin plates at elevated temperatures. Thin-Walled Struct. 84, 134–149.
ECCS, 1984. Ultimate limite state calculation of sway frames with rigid joints. Publication No. 33. Euro-
pean Convention for Constructional Steelwork Technical Committee No. 8.
ECCS, 1976. Manual on Stability of Steel Structures. Publication No. 22. European Convention for Con-
structional Steelwork Technical Committee No. 8.
FCT, 2012. TaperSteel, Stability design of non-uniform steel members - PTDC/ECM-EST/1970/2012
(2012-2014).
Franssen, J.-M., Gernay, T., 2017. Modeling structures in fire with SAFIR ® : Theoretical background
and capabilities. J. Struct. Fire Eng. 8, 300–323.
Kruppa, J., 1999. Eurocodes - Fire parts: Proposal for a methodology to check the accuracy of assess-
ment methods. CEN TC 250, Horizontal Group Fire, Document no. 99/130.
Maia, É., Couto, C., Vila Real, P., Lopes, N., 2016. Critical temperatures of class 4 cross-sections.
J. Constr. Steel Res. 121, 370–382. https://doi.org/10.1016/j.jcsr.2016.02.017
Maraveas, C., 2018. Local buckling of steel members under fire conditions: a review. Fire Technol. 1–30.
https://doi.org/https://doi.org/10.1007/s10694-018-0768-1.
Marques, L., Simões da Silva, L., Greiner, R., Rebelo, C., Taras, A., 2013. Development of a consistent
design procedure for lateral-torsional buckling of tapered beams. J. Constr. Steel Res. 89, 213–235.
https://doi.org/10.1016/j.jcsr.2013.07.009
Marques, L., Simões da Silva, L., Rebelo, C., 2014. Review of the general method in EC3-1-1 as a global
stability verification procedure, in: Eurosteel 2014.
Prachar, M., Hricak, J., Jandera, M., Wald, F., Zhao, B., 2016. Experiments of Class 4 open section
beams at elevated temperature. Thin-Walled Struct. 98, 2–18. https://doi.org/10.1016/j.tws.2015.04.025

752
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Progressive collapse assessment of storage racks due to localized


failures. Explicit consideration of dynamic effects

I. Marginean
Department of Steel Structures and Structural Mechanics, Politehnica University Timisoara, Romania

F. Dinu & D. Dubina


Department of Steel Structures and Structural Mechanics, Politehnica University Timisoara, Romania
Laboratory of Steel Structures, Romanian Academy, Timisoara Branch, Romania

ABSTRACT: Storage rack structures are light metal structures, with elements made of thin-
walled sections, mostly cold formed, and with pinned or partial-strength/semi-rigid connections
(upright base, beam-upright). Due to low redistribution capacity and special gravity load condi-
tions (storage structure), any local damage in elements or connections can spread to neighboring
elements, generating extensive damage or even generalized collapse (progressive collapse). The
analysis and design of the rack structures is laborious, especially due to the characteristics of the
sections and connections (thin-walled sections, perforations on uprights, nonlinear response of
connections). Because of these peculiarities, reference cannot be made to usual structural design
recommendations and standards and additional rules are required for modeling, analysis and
design of racks. The accidental impact of transport and lifting equipment (forklift trucks) is one of
the main sources of risk for the stability and integrity of rack structures. In the study, robustness
of Selective Pallet Rack structures (SPR) in case of accidental impact was evaluated. The response
was compared with the more conventional column (upright) loss approach employed by means of
both static and dynamic analyses. Structural configurations vary by considering different proper-
ties of connections (upright base, beam-upright). The results have shown that upright removal can
generate global failure, if the stiffness of connections is low (pinned or semi-rigid). Also, upright
loss approach can provide satisfactory results with less computational effort, but there are add-
itional effects that require further corrections.

1 INTRODUCTION

Rack structures are lightweight metal systems, pre-fabricated, used to store various goods. Stand-
ardized details and prefabricated components and connections (including upright base anchoring)
allow a fast assembly and greater flexibility in use. Among different systems and configurations
found on the market, Selective Pallet Rack (SPR) is the most common system (Figure 1a).
Another growing segment of warehouses are known as rack-supported buildings, in which racking
structure is part of the construction together with side and roof cladding (Figure 1b). Storage
racks are usually made from thin walled profiles, cold-formed or hot rolled and connected with
mechanical fasteners. In the transverse direction (cross-aisle), the frames are usually braced, while
in the longitudinal direction (down-aisle), braces can be placed in one or more spans. SPR-type
structures are widely used in warehouses and distribution centers. They are also used in commer-
cial spaces, presenting a risk to personnel and clients in the event of accidents. Major accidental
loading situations that may give rise to local damage of SPR structures are: i) inadequate distribu-
tion or overloading with live (storage) loads; ii) accidental impact; iii) earthquakes; iv) localized
fire. SPR structures, but also other types of storage structures, are low redundancy systems,
making them especially vulnerable in case of local damage. Considering the special gravity load
conditions (storage structure), local damage can propagate to neighboring elements, and may even
lead to a generalized collapse (or progressive-collapse) (Figure 1c).

753
Figure 1. Steel pallet racks: a) selective pallet rack (SPR); b) example of a rack supported building,
Ploiesti, Romania; c) partial collapse due to local failure of an upright due to forklift impact.

Considering the importance of the connections (upright base, beam-upright), numerous studies
focused on the influence of connection properties (strength, stiffness, and ductility) on rack per-
formances under gravity and seismic loads (K. M. Bajoria & Talikoti, 2006; Kamal M. Bajoria,
Sangle, & Talicotti, 2010; Baldassino, Bernuzzi, & Zandonini, 2000; Dai, Zhao, & Rasmussen,
2018; Shariati et al., 2018). Other studies focused on the improvement of the overall response by
vertical and horizontal bracings (Abdel-Jaber, Beale, & Godley, 2006; Dubina, Sigauan, & Dinu,
2015; Jacobsen & Tremblay, 2017; Ng, Beale, & Godley, 2009; Rosin et al., 2009; Yin, Tang, Li, &
Zhang, 2018). Rather few studies investigated the forklift impact, most of them focusing on sys-
tems other than SPR (ex. drive-in) (Freitas, Souza, & Freitas, 2010; Gilbert & Rasmussen, 2011a,
2011b; Zhang, Gilbert, & Rasmussen, 2011). Also, there are no specific standards dealing with
assuring SPR structural robustness to mitigate progressive collapse due to impact, as European
structural design recommendations and standards for SPR structures, i.e. EN 15512 provisions
(CEN, 2009), do not provide a methodology for structural design and verification to impact.
The study presented in this paper evaluated the robustness of SPR structures in the event of
an accidental impact from a forklift. The response was evaluated by explicit impact modeling
(Dubina, Marginean, & Dinu, 2019) and the results were compared with notional column
(upright) removal approach using static and dynamic analyses (Department of Defense DoD,
2016). Structural configurations vary with the type of connections (upright base, beam-upright).

2 ROBUSTNESS OF SPR SYSTEMS AT ACCIDENTAL IMPACT

2.1 Description of the geometry, sections and loading procedure


The configurations considered in the study are based on a classic SPR system, see Figure 1a. In
such a system, the horizontal forces that develop in the down-aisle frames (front, rear) are trans-
ferred to the vertical longitudinal bracings arranged in the rear frame. The stabilizing effect of the
longitudinal vertical bracings is also transferred to the front uprights by means of horizontal bra-
cings, while the stability of the transversal frames is ensured by the transverse bracings. The refer-
ence system is shown in Figure 2. The structure has four spans of 2.0 m in the longitudinal
direction and two transverse spans of 1.1 m arranged in back to back double rows. The first beam
is positioned at a height of 1.375 m from the ground, while the distance between the current beams
is 1.2 m. The steel used for beams, uprights and diagonals is S355 (fy = 355 N/mm2). The upright
is made of an omega profile with perforations, see Figure 3a. The beam has a rectangular tubular
section and dimensions 130 x 45 x 1.5 mm. Cross-aisle braces are made of 40 x 40 x 3 mm square
hollow sections and are fixed with pinned connections. X diagonal spine braces, located at end
spans of rear (interior) down-aisle frames (2 and 3 in Figure 2), and top plan braces (in all spans)
are made of rods with an area of 283.84 172 mm2. Run space links, used to connect the two mod-
ules at the level of each beam, are also made of 40x40x3 mm square hollow sections.
The reference structure was designed for a characteristic gravity load of 12.5 kN/m, uni-
formly distributed on the beams. For this value, a load factor λ = 1 will be further considered

754
Figure 2. Geometry and layout of the reference SPR: a) general view; b) cross-aisle frame; b) outer
down-aisle frames (1, 4); c) inner down-aisle frames (2, 3).

Figure 3. Cross-section details for uprights (a) and beams (b).

as reference. Frame out-of-plumb imperfections have been modelled as equivalent imperfec-


tions according to EN 15512 provisions (CEN, 2009).

2.2 Evaluation of progressive collapse resistance


To evaluate the ability of the structure to resist progressive collapse after a local damage,
a comparative robustness study of SPRs was performed. The main parameters of the study
(see Table 1) are described below.
The upright connection at the base was modeled either as fully rigid (Ru) or as nominally
pinned (Pu). For the beam-to-upright connections, also two types of connections were used. They
were classified in relation with the rigidity and flexural capacity of the beam, i.e. a low rigidity
connection (Lr) with 2.5EIb/Lb and a high rigidity (Hr) connection with 12 EIb/Lb, where Ib and
Lb are the moment of inertia and length of the beam. The low rigidity connection has a strength
ratio of 0.2 Mj,Rd/Mb,Rb (connection moment capacity/beam moment capacity). The high rigidity
connection is also partial strength, with Mj,Rd/Mb,Rb=0.5. According to EN 1993-1-8 [37], both
configurations are classified as semi-rigid. The two rigidity levels are relevant for the current prac-
tice in the field as indicated in [9].
The spine bracing in the down-aisle frame was made in X configuration, with the braces located
in the two exterior front bays (see Figure 2). To consider the effect of horizontal plan bracings,
racks were configured with X plan braces (Xh). Damage scenario includes the loss or impact of an
upright from the middle front frame (M), located at first-story level, see Figure 4.
Numerical analyses were performed using Extreme Loading for Structures ELS computer code
(Applied Science International, 2018). For simulating the accidental impact, three methods of ana-
lysis were applied. First two are based on notional column (upright) removal, and employ

755
Table 1. Structural parameters for robustness analysis.
Configurations Parameters

Upright base connection Pinned upright support (Pu)


Rigid upright support (Ru)
Beam-to-upright connection High rigidity (Hr)
Low rigidity (Lr)
Spine Braces (Longitudinal) X vertical braces on the 2 external bays (Xs)
Top braces (Horizontal) X braces at top (Xh)
Upright loss/impact scenario Middle perimeter upright (M)
Type of analysis Upright removal with Nonlinear Static Procedure (NSP)
Upright removal with Nonlinear Dynamic Procedure (NDP) *
Explicit simulation of forklift impact on upright (IMP) *

* for Ru-Hr and Pu-Lr configurations

Figure 4. Plan view with the position of upright for loss/impact scenarios.

nonlinear static and dynamic analyses, respectively. In the static notional upright removal method
(NSP – nonlinear static procedure), the capacity of the structure is evaluated by a non-linear static
analysis, where the upright is deleted from the model and the structure is subjected to increasing
levels of gravity loading (push-down analysis). This analysis does not take into consideration any
dynamic effects (inertia effects). Therefore, a dynamic non-linear analysis was also performed
(NDP – nonlinear dynamic procedure). First, the gravity load is applied on the structure using
a static analysis, then, in the second stage, the upright is removed almost instantaneously (duration
of 0.001 seconds). The analysis is performed for increasing levels of gravity loads until the failure
of the structure.
To directly capture the structural response and the interaction with the impacting body (fork-
lift), a third type of analysis was also performed, i.e. dynamic explicit forklift impact (IMP). The
damage resulting from the impact depends on the direction or angle of the collision. A diagram
showing the correct forklift truck manoeuvres as well as common mistakes during pallet removal
and placement is shown in Figure 5a (Gilbert & Rasmussen, 2011b). However, due to narrow aisle
spacing, in many accidents the direction of the impact is oblique, and therefore the impact force
will have components on both cross-aisle and down-aisle directions (Figure 5b). In the study, the
angle of impact was considered 45º with respect to the cross-aisle direction.
In the analysis, the impacting body (i.e. forklift) is allowed to slide only on the horizontal plane
and has mass assigned to account for the weight of the forklift and of the lifted load. The planar
dimensions of the plate simulating the forklift are 1600x1150 mm, while the plate modelling the
forks is 1200 mm long and 700 mm wide. A plate thickness of 80 mm is considered in the impact
zone. The forklift mass is 3.5 t, while the mass carried by the forks is 1.5 t respectively. The elem-
ents (solid) representing the forklift are removed after the impact, such that they would not inter-
fere with the upright, considering that the vehicle would be stopped and removed after the impact.
The impact object is sliding at 300 mm from the ground level. The analysis is performed for
increasing levels of gravity loads until the failure of the structure.
As the numerical analyses are demanding in terms of computational effort, only two con-
figurations have been considered in the dynamic analyses (NDP, IMP), i.e. Pu was linked to
Lr (Pu-Lr), while Ru was linked to Hr (Ru-Hr).

756
Figure 5. Forklift truck impact with the rack: a) correct forklift truck manoeuvres and common mistakes
during pallet removal and placement [30]; b) deformations produces by oblique forklift impact in a real
accident.

2.3 Results
The application of the NSP for upright removal shows that beam-to-upright connection stiff-
ness (Lr vs Hr) is less influential than the upright base fixing condition (Pu vs. Ru), and only
Ru configurations resist progressive collapse (Figure 6). As expected, the stiffer configuration
(Ru-Hr) has the highest capacity to resist the loss of the upright.
Figure 7 shows the gravity load multiplier (λ) vs vertical displacement for configurations Ru-Hr
and Pu-Lr, all three types of analysis (NSP, NDP, IMP). The dynamic amplification following the
upright removal, calculated as a ratio of static force over the dynamic one (NSP vs NDP) for
a certain vertical displacement, is less than 20 % for both Ru-Hr and Pu-Lr configurations. Com-
pared to NDP, the capacity is very much reduced if the forklift impact (with the parameters speci-
fied in this study) is explicitly modelled (IMP). The main reason is the axial force development in
the upright during forklift impact at large deformation stage. This axial force induces vertical
deflections in the adjoining structure even before the complete separation of the upright (see
points 1 and 2 in Figure 7). Thus, as seen from Figure 8, the upright is initially in compression,
then it is bent due to impact and axial compression force starts to diminishes, then changes to
tension due to development of catenary force and increases until upright fails. At final stage, the
axial force in the upright is completely released (upright is removed due to impact).
Figure 9 shows in detail the failure initiation for the Pu-Lr structure under forklift impact.
After the loss of the upright, the adjacent upright from the same cross-aisle frame (rear frame)
buckles due to load redistribution, then the unsupported cross-aisle frame drops vertically and

Figure 6. Gravity load multiplier vs vertical displacement, nonlinear static procedure NSP.

757
Figure 7. Gravity load multiplier vs vertical displacement: a) configuration Ru-Hr; b) configuration Pu-Lr.

Figure 8. Upright response during impact: a) normalized forces at the upright base during impact; b)
deformations of the upright during impact and distribution of axial force in uprights (note: blue is com-
pression, and red is tension).

pull the adjacent frames inward because of the connected beams and top plan braces. Ultim-
ately, the collapse propagates to the entire structure, see Figure 10.
As seen from Figure 11, the explicit modeling of impact for fully restrained upright base vertical
displacement (IMP-fixed) leads to higher vertical displacements when compared with NDP (and
NSP, obviously), due to the tensile force that is developed in the impacted upright and “pulls” the

758
Figure 9. Collapse initiation for Pu-Lr structure in case of (M) scenario (colored scale indicate deform-
ations on cross-aisle direction), explicit impact analysis a) t=0.49 s b) t=0.57 s c) t=0.62 s d) t=0.67 s.

Figure 10. Collapse propagation for Pu-Lr structure in case of explicit impact analysis.

Figure 11. Comparison between NSP, NDP and explicit impact IMP for fixed and free upright base
vertical displacement restraint, Ru-Hr-M configuration.

structure downward. However, if the upright base vertical displacement is completely released
(upright base is free to move vertically), the tensile force does not develop and the vertical displace-
ments significantly reduce and become close to the static ones. For intermediate levels of restraint,
it is expected that the response will bound between the two margins (i.e. free and fixed).

759
3 CONCLUSIONS

The paper presents the results of a numerical study that aimed to assess the robustness of selective
pallet rack structures (SPR) in case of an accidental forklift impact. Analyzes were performed with
the Extreme Loading for Structures program, using both notional upright removal approach and
explicit forklift impact. For the scenario involving the middle front frame upright (M scenario),
the dynamic analysis (NDP) shown that all structures are vulnerable, and progressive collapse can
be prevented only for the stiffer structural configurations (Ru-Hr).
When forklift impact is explicitly modeled considering the characteristic gravity load, progres-
sive collapse is initiated for all configurations, including the Ru-Hr structure, for which the
notional upright removal with dynamic analysis did not result in collapse. The main reason for
this reduced capacity is the development of tensile axial forces in the upright during impact, which
pulls down the structure and amplifies the effect of the gravity loads. This additional effect
depends on the deformation capacity of the upright and the vertical load resistance of the upright
base connection. Given the lack of provision for the design against impact of rack structures, fur-
ther developments are necessary.

REFERENCES

Abdel-Jaber, M., Beale, R.G., & Godley, M.H.R. 2006. A theoretical and experimental investigation of
pallet rack structures under sway. Journal of Constructional Steel Research, 62(1–2), 68–80.
Applied Science International. 2018. Extreme Loading for Structures Theoretical Manual, Version 6.
Bajoria, K.M., & Talikoti, R.S. 2006. Determination of flexibility of beam-to-column connectors used in
thin walled cold-formed steel pallet racking systems. Thin-Walled Structures, 44(3), 372–380.
Bajoria, Kamal M., Sangle, K.K., & Talicotti, R.S. 2010. Modal analysis of cold-formed pallet rack
structures with semi-rigid connections. Journal of Constructional Steel Research, 66(3), 428–441.
Baldassino, N., Bernuzzi, C., & Zandonini, R. 2000. Performance of joints in steel storage pallet racks.
Department of Mechanical and Structural Engineering, University of Trento.
CEN. 2009. EN 15512 Steel static storage systems—Adjustable pallet racking systems—Principles for
structural design.
Dai, L., Zhao, X., & Rasmussen, K.J.R. 2018. Flexural behaviour of steel storage rack beam-to-upright
bolted connections. Thin-Walled Structures, 124, 202–217. https://doi.org/10.1016/j.tws.2017.12.010
Department of Defense DoD. (2016). UFC 4-023-03. Design of Buildings to Resist Progressive Collapse.
Dubina, D., Marginean, I., & Dinu, F. 2019. Impact modelling for progressive collapse assessment of
selective rack systems. Thin-Walled Structures, (Special Issue ICTWS 2018-in print).
Dubina, D., Sigauan, A., & Dinu, F. 2015. Stability and Robustness of Steel Storage Rack Systems.
CONSTRUIEŞTE CU „STEEL”, 9–31. Cluj Napoca: QUALMEDIA.
Freitas, A.M., Souza, F.T., & Freitas, M.S. 2010. Analysis and behavior of steel storage drive-in racks.
Thin-Walled Structures, 48(2), 110–117.
Gilbert, B.P., & Rasmussen, K.J. 2011a. Determination of the base plate stiffness and strength of steel
storage racks. Journal of Constructional Steel Research, 67(6), 1031–1041.
Gilbert, B.P., & Rasmussen, K.J.R. 2011b. Impact tests and parametric impact studies on drive-in steel
storage racks. Engineering Structures, 33(5), 1410–1422. https://doi.org/10.1016/j.engstruct.2011.01.017
Jacobsen, E., & Tremblay, R. 2017. Shake-table testing and numerical modelling of inelastic seismic
response of semi-rigid cold-formed rack moment frames. Thin-Walled Structures, 119, 190–210.
https://doi.org/10.1016/j.tws.2017.05.024
Ng, A.L.Y., Beale, R.G., & Godley, M.H.R. 2009. Methods of restraining progressive collapse in rack
structures. Engineering Structures, 31(7), 1460–1468.
Rosin, I., Calado, L., Proenca, J., Carydis, P., Mouzakis, H., Castiglioni, C., . . . Negro, P. 2009. Storage
racks in seismic areas. Report EUR, 23744.
Shariati, M., Tahir, M.M., Wee, T.C., Shah, S.N.R., Jalali, A., Abdullahi, M.M., & Khorami, M. 2018.
Experimental investigations on monotonic and cyclic behavior of steel pallet rack connections. Engin-
eering Failure Analysis, 85, 149–166. https://doi.org/10.1016/j.engfailanal.2017.08.014
Yin, L., Tang, G., Li, Z., & Zhang, M. 2018. Responses of cold-formed steel storage racks with spine
bracings using speed-lock connections with bolts II: Nonlinear dynamic response history analysis.
THIN WALLED STRUCTURES, 125(1), 89–99.
Zhang, H., Gilbert, B.P., & Rasmussen, K.J. 2011. Drive-in steel storage racks. II: Reliability-based
design for forklift truck impact. Journal of Structural Engineering, 138(2), 148–156.

760
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Elastic buckling strength of lipped channel section beam restrained


on upper flange subjected to bending

H. Masuda & K. Ikarashi


Architecture and Building Engineering, Tokyo Institute of Technology, Japan

ABSTRACT: The existing Japanese code for thin-walled structures underestimates the
allowable stress for distortional buckling. The objective of this paper is to propose a close-
form expression of the elastic buckling stress corresponding to the distortional buckling mode
for thin-walled lipped channel section members subjected to bending based on a semi-
analytical study. First, we performed finite element analyses to understand the buckling
behavior. Then, we built the model for the energy method, which is one of the theoretical and
effective methods to find what parameter determines the elastic buckling eigenvalues.

1 INTRODUCTION

The elastic buckling stress and the nominal strength of cold-formed steel lipped channel mem-
bers per specific conditions have been provided in AISI Specification (AISI 2016). Lipped
channel sections acting as a single member exhibit one of three elastic bucking modes when
subjected to uniform bending, local, distortional or global buckling modes, the half wave-
length of which is short, medium, and long, respectively. Appendix 2 of the AISI Specification
provides the method to calculate the elastic buckling stress for each mode. In particular,
a formula for the distortional mode is proposed based on a semi-analytical approach where
the compression lipped flange is considered to be a beam attached to a rotational spring
(Schafer 2006). This formula includes benefit of rotational stiffness provided by sheathing and
so on if and only if it attaches to the compression side.
Members used as a roof or wall system have also been studied by many researchers (Peköz
and Soroushian 1982, Laboube 1986, Wibbenmeyer 2009). Their achievements are summar-
ized in AISI-S100 Section I6.2. This section applies to a beam with the tension flange attached
to the deck and the compression flange unbraced, for example, a roof purlin or wall girt sub-
jected to wind suction. The nominal strength in this case has been empirically derived accord-
ing to a procedure wherein static and uniform suction pressure is applied on the sheathing.
The formula for the nominal strength in this case is provided as a reduction from that of
a fully braced member on the compression flange, i.e., the nominal strength for the local buck-
ling multiplied by the reduction factor.
In Japan, earthquake load usually determines the required capacity of structures and let
a inflection point appear on a beam, i.e., make the compression flange unbraced. As the load con-
figuration in this case differs from that specified in AISI-S100, the specification may not be fully
applied to structures in Japan. In addition, the existing Japanese code does not consider restraining
effects provided by sheathing, despite the fact that panel members are expected to prevent the ten-
sion flange from rotating and significantly benefit the elastic buckling eigenvalue, given that previ-
ous studies about H-shaped beam restrained on the top flange by concrete slabs come to a the
similar conclusion (Ikarashi 2018). The objective of this study is to provide a closed-form expres-
sion for elastic buckling stress regarding mainly a degree of constraint on the tension flange and
briefly the moment gradient and boundary conditions at ends of the cold-formed steel lipped chan-
nel sections. To begin with, finite element analyses are performed to understand fundamental

761
buckling behavior. Then, a model for the energy method is established regarding characteristics
specific to lipped channel sections, which are asymmetrical around the weak axis.

2 FINITE ELEMENT ANALYSIS

2.1 Overview of finite element analysis


This section describes finite element models for lipped channel members subjected to a bending
restrained on the top flange. We briefly take into account effects of the support condition at the
ends and the moment gradient. ABAQUS 2017 is employed as an analysis tool.
The models are shown in Figure 1. We gave these models a boundary condition which restricts
the rotational and horizontal displacement at the tension web-flange juncture instead of an inter-
mediate of flange to exclude a reduction in rotational stiffness due to out-of-plane bending of the
top flange. This is because that it is mentioned that rotational stiffness depends on the position of
connections (Gao and Moen, 2012). Finally two boundary conditions at ends are considered, pin
ends (free to warp) or fixed ends (prevented from warping and rotating around y-axis).

2.2 Results of finite element analysis


This section addresses the results of finite element analyses. Figure 2 shows the relationship
between the length-to-depth ratio λ and the critical stress σcr . The members without restraint
on the flange exhibit either local (referred to as LB), distortional or global (referred to as GB)
modes. On the contrary, the members with restraint exhibit the local or one of two distor-
tional modes (referred to as DB1 and DB2 respectively). The deformation diagrams are
shown on the bottom left of Figure 2. Restricting rotation of the tension flange is confirmed
to change the dominant mode and increase the critical stress. The next section addresses the
BD2 mode interacting with the GB mode.

3 THE ENERGY METHOD

3.1 Overview
This section presents an overview of a semi-analytical approach with the energy method (Ritz
method) for lipped-channel sections restrained on the tension flange. A similar configuration
has been studied on H-shaped sections where it is assumed that only the web undergoes out-of
-plane deformation and beam theory is applied to the flanges. Establishing a model for the
lipped-channel section here employs the same assumption.
The energy method includes two procedures: assuming virtual displacement field and equat-
ing the strain energy as the external work calculated based on the assumed configuration. This
means that the configuration should be able to deal with any possible configurations. As

Figure 1. Analysis model for ABAQUS.

762
Figure 2. Results of finite element analysis.

shown in Figure 3, it is observed from finite element analyses that vertical displacement
occurs during buckling in the DB2 mode for lipped channel sections, as opposed to H-shaped
sections. Therefore, this configuration must be taken into account for more accurate results.
In addition to the vertical degrees of freedom, distribution of the longitudinal strain should be
accurately expressed. Figure 4 shows a schematic distribution of the longitudinal strain of two
types of section shape when they exhibit DB2 mode. For an H-shaped section, the neutral axis
of in-plane bending of the compression flange coincides with the web-flange junction and thus
there is no longitudinal strain induced in web, whereas the distribution is complicated for
lipped-channel sections. These two features should be taken into account when constructing
models, as shown in Figure 5. We consider three restraining springs attached to the tension
web-flange junction.

3.2 Derivation of total potential energy


In this section, the strain energy and external work are derived. The strain energy is expressed
as shown in Equation 1. It is a summation of U0 ,U1 ,Uw and Us , which denote strain energy in
the tension lipped flange, compression lipped flange, web, and restraining springs, respect-
ively. Us includes terms due to rotational, torsional, and horizontal springs. The strain energy
is finally normalized and expressed as a functional with w0 ,w1 , 0 , 1 and v as shown in Equa-
tion 2, where vector u and matrices C, D and B are defined in Table 1. The first term of the
strain energy is obtained by integrating E x 2 =2 over the area, where E and x denotes the
Young’s modulus and the longitudinal strain, respectively. The second term comes from
Saint-Venant torsion and the torsional spring attached to the tension flange. The third term is
due to the out-of-plane deformation of the web and the rotational or horizontal springs. The

Figure 3. Deformation of section during DB2.

763
Figure 4. Schematic diagram of longitudinal strain.

Figure 5. Model for energy method.

out-of-plane configuration of theweb w is assumed to be a cubic function in terms of direction


in y-axis, as shown in Equation 3.

U ¼ U0 þ U1 þ Uw þ Us ð1Þ
21 3
ð ð1 T ð1
Et 4 d 2 uT d 2 u du du
U¼ 3 C dξ þ λ2 D dξ þ λ4 uT Bu dξ 5 ð2Þ
2λ dξ 2 dξ 2 dξ dξ
0 0 0

 y 2 y 3    y 3 
y 2
wðx; yÞ ¼ w0 ðxÞ 1  3 þ2 þ w1 ðxÞ 3 2
 h h  h h 
y y 2 y 3  y 2 y 3
ð3Þ
þh0 ðxÞ  2 þ þ h1 ðxÞ  þ
h h h h h

The external work is expressed as shown in Equation 4. It is summation of T0 ,T1 and Tw ,


which denote the external work done to the tension lipped flange, compression lipped flange,
and web, respectively. The work done by shear stress is assumed to be negligible since the
DB2 mode occurs with relatively long beams and then the shear stress is negligible compared to
the longitudinal stress. The external work is finally normalized and expressed as a functional
with w0 ,w1 , 0 ,1 and v as shown in Equation 5, where matrix X is defined at Table 1.

T ¼ T0 þ T1 þ Tw ð4Þ
2 3
ð1
σt 4 2 duT du 5
T ¼ 3 λ ð1  ð1 þ M1 =M0 Þξ Þ X dξ ð5Þ
2λ dξ dξ
0

The total potential energy P is defined as shown in Equation 6. Substituting Equation 2 and 5
into Equation 6, the normalized expression for the total potential energy is obtained as shown
in Equation 7. The expression can be more simplified as shown in Equation 8 when w0 ,w1 ,0 ,1
andv are proportional to each other.

764
Table 1. Variable used for Normalization
2 3 2 3 2 3
C 11 C12 C13 C 14 C15 D11 D12 D13 D14 D15 B11 B12 B13 B14 B15
6 C 21 C22 C23 C 24 C25 7 7 6 D21 D22 D23 D24 7
D25 7 6 B21 B22 B23 B24 B25 7
6 6 6 7
C ¼6
6 C 31 C32 C33 C 34 C35 7 6
7; D ¼6 D31 D32 D33 D34 D35 7 6
7; B ¼6 B31 B32 B33 B34 B35 7
7
4 C 41 C42 C43 C 44 C45 5 4 D41 D42 D43 D44 D45 5 4 B41 B42 B43 B44 B45 5
C 51 C52 C53 C 54 C55 D51 D52 D53 D54 D55 B51 B52 B53 B54 B55
2 3 2 3 2 3
X 11 X 12 X 13 X 14 X 15 w0 ðξ Þ w0
6 X 21 7 6 7 6 7 u¼
6 X 22 X 23 X 24 X 25 7 6 w1 ðξ Þ 7 6 w1 7 uf ðξ Þ;
X ¼6 X 35 7; u ðξ Þ¼ 6 h0 ðξ Þ 7; u ¼ 6 h0 7 T C
u u ¼ UC;
6 X 31 X 32 X 33 X 34 7 6 7 6 7
4 X 41 X 42 X 43 X 44 X 45 5 4 h1 ðξ Þ 5 4 h1 5  u ¼ UD; 
uT D u ¼ UB; 
uT B uT X
u ¼ TX
X 51 X 52 X 53 X 54 X 55 vðξ Þ v

β3 β2 ðβ þ 2γÞ2 β2 βγ βγ2 1 þ 6β þ 6γ  12γ2 þ8γ3


C 22 ¼ þ β2 γ  ; C 25 ¼ C 52 ¼ þ  ; C 55 ¼
3 4ð1 þ 2β þ 2γÞ 4 2 2 12

G  t 2 Gðβ þ γÞ  t 2 Gðβ þ γÞ  t 2 1
D22 ¼ ; D33 ¼ þ kt ; D44 ¼ ; X 22 ¼ þβ þ γγ2
3E h 3E h 3E h 6

x y l b c  1  t 2 kr kt kh h E
ξ ¼ ;η ¼ ; λ ¼ ; β ¼ ; γ ¼ ;K ¼ ; kr ¼ ; kt ¼ ; kh ¼ ;G ¼
l h h h h 12ð1 Þ h
2 Eht Eh3 t Et 2ð1 þ Þ

P¼U T ð6Þ
21 3
ð 2 T ð1 T ð1
Et d u d u
2
du du
P¼ 34 C dξ þ λ2 D dξ þ λ4 uT Bu dξ 5
2λ dξ 2 dξ 2 dξ dξ
2
0 0
30 ð7Þ
ð1
σt duT
du
 3 4λ2 ð1  ð1 þ M1 =M0 Þξ Þ X dξ 5
2λ dξ dξ
0
2 3
ð1  2 2 ð1  2 ð1
Et 4 d f df
P ¼ 3 UC dξ þ UD λ2 dξ þ UB λ4 f 2 dξ 5
2λ dξ 2 dξ
2 0 1 0
3 0
ð8Þ
ð  2
σt df
 3 4TX λ2 ð1  ð1 þ M1 =M0 Þξ Þ dξ 5
2λ dξ
0

3.3 Comparing the results


In this section, the eigenvalue analysis results obtained by finite element analysis and the
energy method are compared to validate the latter method. Let normalized stiffness of springs
kr , kt , kh all be zero for the corresponding model for unrestrained sections and zero,
1:0  1012 , 1:0  1012 , respectively, for restrained sections. Figure 6 shows the results.
The eigenvalues obtained by the two methods do not coincide with each other in the range
where members exhibit LB mode in the finite element analysis. This is because the model built
for this study does not consider the out-of-plane deformation of lipped flanges. In range of
GB, DB1 or DB2 mode, on the contrary, the model for the energy method has approximate
validity except for the case where M1 =M0 equals to 1.0.

765
Figure 6. Comparison of two methods (the legend is the same as that in Figure 2).

3.4 Assessment of eigenvalue considering stiffness of torsional spring


In this section, we find principle elements of matrices in Equation 8 and simplify the expres-
sion of total potential energy. The scalar value UC ¼  uT C
u in Equation 8 is the summation of
T
all elements of matrix Cuu . For example, by substituting the eigenvectors  u obtained by the
energy method into these expression, elements are obtained for the section of
h ¼ 140; b ¼ 40; c ¼ 12; t ¼ 1:6; λ ¼ 12 as shown in Table 2. The number of half waves in the
longitudinal direction is one in these cases. For the unrestrained section, taking (2,2), (2,5),
(5,2), (5,5) elements of matrix CuuT and summing up results in 0.96 times as the summation
of all elements, and we can thus conclude that these four are dominating elements. Likely
(2,2) element of matrix X  u
uT turns out to be a principle element. These approximations are
also true for the restrained section and other slender sections, for example,
h ¼ 300; b ¼ 30; c ¼ 12; t ¼ 1:6; λ ¼ 8. However, the same procedure makes a non-negligible
error for stocky sections such as h ¼ 80; b ¼ 50; c ¼ 20; t ¼ 1:6; λ ¼ 20.
Next we think of approximating the out-of-plane deformation of the web. We first deal
with a slender section mentioned above. When no restraining effect is provided and thus the
member exhibits the GB mode, it is approximated that w  0 ¼ 0, w
 1 ¼ h0 ¼ h1 and thus
 ðηÞ ¼ η. In contrast, when the tension flange is fully restrained and the member exhibits
w
the DB2 mode, it is approximated that w  0 ¼ 0, h0 ¼ 0 3 w1 ¼ 2h1 and thus
 ðηÞ ¼ 3=2η2  1=2η3 . This configuration is similar to that of a cantilevered beam loaded lat-
w
erally on the free end. It is assumed that the configuration under incomplete restraint can be
expressed as a superposition of these two extreme configurations as shown in Equation 9.
However, this assumption is not true for stocky sections. In this case h1 overly exceeds 3=2w1
and thus the superposition mentioned above cannot be applied.
  
 ðηÞ ¼ ð1  aÞη þ a 3=2η2  1=2η3 w
w 1 ð9Þ

It is clear that sections with small h=b exhibit more complicated behavior since the simplifica-
tion and the superposition mentioned above cannot be applied. However, members subjected to
flexure usually have large h=b, and hence the two assumptions about the principle elements and
the out-of-plane deformation of the web are supposed to hold. Thus, we can apply these
assumptions to the later sections. Based on these assumptions, UC ,UD ,UB and TX in Equation 8
are approximated as in Equation 10 through 13, whose symbols are defined in Table 1.

  2
 21 þ 2C25 w
UC ¼ C22 w  1v þ C55v2 ¼ C22  C25
2
=C55 w 1 ð10Þ
n o
UD ¼ D22 þ D33 ð1  aÞ2 þ D44 ð1 þ 0:5αÞ2 w  21 ð11Þ

 2w
UB ¼ 3Ka  21 ð12Þ

766
Table 2. Elements of matrices obtained by eigenvalue analyses.
for unrestrained section of h ¼ 140; b ¼ 40; c ¼ 12; t ¼ 1:6; λ ¼ 12;
uT ¼ ½   0:031 1 0:966 1:068   0:128

2 3 2 3
0:001 0:009 0:001 0:000 0:14 0:001 0:000 0:000 0:000 0:000
6 0:009 0:387 0:004 0:031 0:457 7 6 0:007 0:000 7
CuuT 6 7 X uuT 6 0:000 1:005 0:000 7
¼ 6 0:001 0:004 0:002 0:000 0:006 7 6
7; T u ¼ 6 0:000 0:000 0:025 0:000 0:015 7
uT Cu 6 4 0:000 0:031 0:000 0:002 0:007 5 u X 4 0:000 0:007 0:000 0:031
7
0:000 5
0:014 0:457 0:006 0:007 0:486 0:000 0:000 0:015 0:016 0:000

for fully restrained section of h ¼ 140; b ¼ 40; c ¼ 12; t ¼ 1:6; λ ¼ 12;


uT ¼ ½ 0:000 1:000 0:000 1:545 0:134

2 3 2 3
0:000 0:000 0:000 0:000 0:000 0:000 0:000 0:000 0:000 0:000
6 0:000 1:337 0:000 0:044 0:461 7 6 0:000 0:009 0:000 7
CuuT 6 7 X uuT 6 0:000 0:913 7
¼ 6 0:000 0:000 0:000 0:000 0:000 77; u ¼6 0:000 0:000 0:025 0:000 0:000 7
uT Cu 64 0:000 0:044 0:000 0:004 0:01 5
 T X
u 6
4 0:000 0:009 0:000
7
0:059 0:023 5
0:000 0:461 0:000 0:01 0:514 0:000 0:000 0:015 0:023 0:000

 21
TX ¼ X22 w ð13Þ

3.4.1 Member with pinned boundary at ends and subjected to uniform bending
To begin with, members with pinned boundary at ends and subjected to uniform bending are
analyzed. The deformation function along the x - or ξ-axis is assumed to be f ðξÞ ¼ sinðmπξÞ
in this case, where m is a natural number. Substituting this function and M1 =M0 ¼ 1 into
Equation 8, Equation 14 is obtained. Furthermore, equating the total potential energy as
zero, temporal critical stress is obtained,
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ffi as shown in Equation 15. This value has a local min-
imum with m at mπ=λ ¼ 4 UB =UC . This minimum value is now a function of α, which gives
the smallest stress at the value shown in Equation 16.
When the normalized torsional stiffness kt approaches infinity and hence α approaches to 1.0,
a member can seem to be fully restrained and the critical stress can be expressed as shown in
Equation 16, where the first term can be neglected as p it ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
is smallffi compared to the second. It is
observed that the critical depth-to-length ratio λcr ¼ π 4 UC =UB tends to infinity and only one
half-wave appears when α equals to 0.0. In this case, λ is always smaller than λcr and returning the
local minimum in terms of m is not appropriate; thus, calculating the critical stress by substituting
m ¼ 1 into Equation 14 and differentiating it with α is necessary. This is even true for the case
where α does not equal 0.0. When α is not equal to 0.0 and moreover λ is larger than λcr , the
critical stress is given by Equation 15 and 16. This observation achieved by the above simplified
method coincides with that of the non-simplified energy method, as shown in Figure 7. Moreover,
the critical stress obtained by Equations 15 and 16 is approximately the same value as that of the
non-simplified method when α does not equal 0.0 and λ is larger than λcr , as shown in Table 3.
 mπ 2 σtðmπÞ2
EtðmπÞ2 mπ 2
P¼ UC þ UD þ UB  TX ð14Þ
4λ λ λ 4λ
mπ 2  2
λ
UC þ UD þ UB
λ mπ
σ¼E
TX n o qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
pffiffiffiffiffiffiffiffiffiffiffiffiffiffi 
D22 þD33 ð1  αÞ2 þD44 ð1 þ 0:5αÞ2 þ2 C 22 C 225 =C 55 3Kα  2
U D þ2 U C U B
E ¼E
TX X 22
ð15Þ

767
Table 3. Comparison between the simplified method and the non-simplified method
(a) h ¼ 140; b ¼ 40; c ¼ 12; t ¼ 1:6 (b) h ¼ 300; b ¼ 30; c ¼ 12; t ¼ 1:6

kt α λcr σcr * σcr ** kt α λcr σcr * σcr **


1:0  100 0.999 12.6 419.1 417.3 1:0  100 0.999 9.3 86.6 93.6
1:0  102 0.945 13.0 408.3 407.8 1:0  102 0.994 9.3 86.3 93.1
1:0  103 0.472 18.4 310.7 315.1 1:0  103 0.938 9.6 84.0 89.6
1:0  104 0.000 ∞ 49.9 63.7 1:0  104 0.380 15.1 60.6 62.7
1:0  105 0.000 ∞ 15.1 29.1 1:0  105 0.000 ∞ 9.9 11.3
0:0 0.000 ∞ 11.3 25.2 0:0 0.000 ∞ 3.1 4.6

* as of Equation 15 and 16
** as of the non-simplified method (λ = 50 )

0 qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
  ffi1
D33 D44 =2 C 22 C 225 =C 55 3K 
α ¼ max@0; A ð16Þ
D33 þD44 =4
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
  qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 
fD22 þ2:25D44 gþ2E C 22 C 225 C 55 3K  2E C 22 C 225 C 55 3K 
σcr ¼ ’ ð17Þ
X 22 X 22

3.4.2 Considering of moment gradient and boundary condition at ends


When a member is fully restrained and no rotation occurs at the tension junction, the tor-
sional term UD can be neglected in Equation 8. Therefore, the total potential is expressed as
shown in Equation 18. This is similar to that of a beam on an elastic foundation subjected to
distributed axial force. The solution for the buckling eigenvalue in this simplified case is
shown as lines in Figure 8 for various load or boundary conditions. The deformation function
f ðξÞ along the x- or ξ-axis is assumed as shown in Equation 19, which depends on boundary
condition at the ends. It is observed that the results of this simplified energy method approxi-
mately coincide with that of finite element analyses for various load and boundary conditions
for the range wherein the member exhibits the DB2 mode. It can be concluded from this
pffiffiffiffiffiffiffiffiffiffiffiffiffi
observation that the normalizedcritical stress σcr =ð2 UC UB =TX Þ for the DB2 mode depend
on only normalized length-to-depth ratio λ=λcr , load and boundary conditions.
2 3
ð1 2 !2 ð1 " ð1  2 #
Et 4 d f  4 f dξ 5 σt df
P ¼ 3 ðC 22 C 225 =C 55 Þ dξþ3Kλ 2
X 22 λ2 ð1ð1þM 1 =M 0 Þξ Þ dξ
2λ dξ 2 0 2λ 3
0 dξ
0
ð18Þ

Figure 7. Critical buckling stress with torsional spring of various stiffness.

768
Figure 8. Critical buckling stress on various load or boundary condition.

8 n
> P
>
< sinðiπξÞ for pin ends
f ðξ Þ ¼ i¼1
P ð19Þ
>
>
n
: sinðπξÞ sinðiπξÞ for fixed ends
i¼1

4 CONCLUSION

In this study, the model for the energy method was established based on characteristics observed
in preceding finite element analyses. By taking advantage of this theoretical method, a closed-
form expression of the critical buckling stress involving torsional stiffness was derived even
though its application is limited to the most basic condition. Furthermore, it was confirmed that
critical stress under various load or boundary conditions for a fully restrained section can be
expressed as a function of normalized length-to-depth ratio λ=λcr and basic critical stress. Since
this study dealt with completely straight member, no initial imperfection is included.
In this study, rotational stiffness (not torsional) remained zero for ease despite rotational
spring being expected to contribute more to the increase in the critical stress than torsional
spring. Moreover, applications are limited to sections with large h=b so equations needs to be
adaptable for general dimension of lipped-channel section.

REFERENCES

American Iron and Steel Institute. 2016. North American Specification for the Design of Cold-Formed
Steel Structural Members 2016 Edition.
Gao, T. and C.D. Moen. 2012. Predicting Rotational Restraint Provided to Wall Girtsand Roof Purlins
by Through-Fastened Metal Panels, Thin-Walled Structures Vol. 61, pp. 145–153, 2012.
Ikarashi, K. and Sano, T. 2018. Influence of upper flange restraint condition on lateral buckling behavior
of H-shaped beam, Journal of Structural and Construction Engineering (Transactions of AIJ), Vol.83,
No. 745, pp. 491–501, March 2018.
LaBoube, R. a. 1986. Roof Panel to Purlin Connection: Rotational Restraint Factor, Proceedings of
theIABSE Colloquium on Thin-Walled Metal Structures in Buildings, Stockholm, Sweden, 1986.
Peköz, T.B. and P. Soroushian. 1982. Behavior of C- and Z-Purlins Under Wind Uplift, Proceedings of
the Sixth International Specialty Conference on Cold-Formed Steel Structures, University of Missouri-
Rolla, Rolla, MO, November 1982.
Schafer, B.W., A. Sarawit and T. Peköz 2006. Complex Edge Stiffeners for Thin Walled Members, Jour-
nal of Structural Engineering, ASCE, Vol. 132, No. 2, February 2006.
Wibbenmeyer, K. 2009. Determining the R Values for 12 Inch Deep Z-Purlins and Girts With Through-
Fastened Panels Under Suction Load, Thesis presented to the Missouri University of Science and Tech-
nology in partial fulfillment of the requirements for the degree of Master of Science in Civil Engineering,
Rolla, MO, 2010.

769
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Experimental study on SCFs of empty SHS-SHS T-joints subjected


to static out-of-plane bending

F.N. Matti & F.R. Mashiri


Western Sydney University, Penrith, Australia

ABSTRACT: This paper presents an experimental study on stress concentration factors (SCFs)
of empty (unfilled) square hollow section (SHS) T-joints specimens. Square hollow sections are
extensively used for the columns and truss of numerous structures such buildings, bridges, cranes,
glass houses and towers. However, there is limited investigation on SCFs of empty SHS-SHS
T-joints under static out-of-plane bending loads which needs to be explored. As a result, there are
no design guidelines for empty SHS-SHS T-joints under out-of-plane bending. To address this
research gap in the literature and develop the fatigue design rules of thin-walled SHS T-joints,
four empty SHS-SHS T-joints were tested under static out-of-plane bending loads on the brace.
The four empty T-joints are made up of empty SHS brace and empty SHS chord members. The
non-dimensional parameters of the SHS-SHS T-joints are β, 2γ and τ. The range of each
non-dimensional parameter of the SHS-SHS T-joint specimens used in this investigation are
0.25 ≤ β ≤ 1, 25 ≤ 2γ ≤ 33.33 and 0.75 ≤ τ ≤ 1. The distribution of the SCFs and the peak SCF at
the hot spot locations (line A, B, C, D and E) of empty T-joints were determined by attaching
strain gauges on the T-joint specimens. In summary, the maximum SCFs of the four SHS-SHS
T-joints under out-of-plane bending commonly occurred along line A, B or C.

Keywords: Stress Concentration Factors, Square Hollow Sections, Empty T-Joints,


Experimental Study

1 INTRODUCTION

Thin-walled welded T-joints made up of steel tubular SHS brace and chord are widely used in
engineering structures such as truss bridges and high rise buildings. This connection could be sub-
jected to cyclic out-plane bending loads. Limited numerical and experimental investigations on
unfilled SHS-SHS tubular T-joints under out-of-plane bending are available. CIDECT Design
Guide 8 (Zhao et al. 2001) does not provide SCF formulas and graphs for predicting the SCFs for
empty SHS-SHS T-joints subjected to out-of-plane bending. Therefore, the experimental studies
carried out will assist with the development of fatigue design rules of steel thin-walled welded tubu-
lar SHS T-joints.
Cheng et al. (2015) carried out a numerical investigation on SCFs of square bird-beak SHS
T-joints under axial loading. Tong et al. (2014) carried out SCFs investigation on seven diamond
bird-beak SHS T-joints under axial loading and in-plane bending. Chen & Wang (2015) carried
out a numerical study and presented a FE model of diamond bird-beak SHS T-joints under axial
compression. Feng & Young (2010) conducted a numerical study and design on SHS and rect-
angular hollow section (RHS) stainless steel tubular T and X-joints with concrete-filled chords
under static compression loading. Tong et al. (2012) tested eight empty circular hollow sections
(CHS) to SHS T-connections under axial loading and in-plane bending and developed 3D FE
models to determine the SCFs. In addition, Mashiri et al. (2004) carried out fatigue tests on CHS
brace to SHS chord T-connections and measured stress distributions at hot spot locations, where
cracks initiated and propagated resulting fatigue failure.

770
Mashiri & Zhao (2010) tested empty and concrete-filled SHS-SHS steel tubular T-joints
under in-plane bending. Matti & Mashiri (2018) used ABAQUS software to model identical
T-joints used in Mashiri & Zhao (2010) experimental investigation to obtain SCFs and com-
pare the FEA results with the experimental results. It was found that the maximum SCFs in
concrete-filled SHS T-joints are lower than the maximum SCFs in empty SHS T-joints.
The aim of this investigation is to study the effect of the non-dimensional parameters on the
SCFs and identify the parameters that influence the fatigue strength of weld connection in
SHS-SHS T-joints. The objective of this paper is to determine the strain concentration factors
(SNCFs), SCFs and the hot spot locations in the empty SHS-SHS T-joint connections under
out-of-plane bending.

2 TEST SPECIMENS

Four unfilled SHS-SHS T-joint specimens were tested experimentally. Table 1 lists the four
T-joints connections with their dimensions, non-dimensional parameters, brace and chord lengths.
The non-dimensional parameters are defined as follows: β is the ratio of the brace to chord width
(b1/b0), 2γ is ratio of the chord width to chord wall thickness (b0/t0) and τ, is the brace to chord
wall thickness ratio (t1/t0). The SHS T-joints have non-dimensional parameters, β ranging from
0.25 to 1, 2γ ranging from 25 to 33.33 and τ ranging from 0.75 to 1.0. As shown in Table 1, the
steel SHSs used in this investigation are cold-formed and have steel grade of C350LO which
comply with AS1163-1991 (Standards Australia Online 2009).
The external corner radii for each T-joint specimen were calculated according to OneSteel
Market Mills (2004) specifications. The design yield stress (fy) of the SHS braces and chords
used in this investigation is 350 MPa. The design tensile strength of the SHS braces and
chords is 430 MPa in accordance to AS 4100-1998 (Standards Australia Online 1998). The
Young’s Modulus and the Poisson’s ratio of the steel SHSs is 200 GPa and 0.3, respectively.
Figure 1 shows the stress-strain distribution of the steel used in this investigation.
The chord of each T-joint specimen is welded to a brace. Each chord is also welded to two
265  265  20 end plates. The brace of each T-joint connection is welded to a 200  200  20

Table 1. Connection series used in the experimental investigation.

Chord Brace size Length


Non-dimensional
do  bo  to d1  b1  t1 parameters Brace Chord

Series mm  mm  mm mm  mm  mm β 2γ τ mm mm

S6S1 100  100  4 SHS 25  25  3 SHS 0.25 25 0.75 500 600


S6S2 100  100  4 SHS 40  40  3 SHS 0.40 25 0.75 500 600
S6S4 100  100  4 SHS 75  75  3 SHS 0.75 25 0.75 500 600
S5S5 100  100  3 SHS 100  100  3 SHS 1.00 33.33 1.00 500 600

Figure 1. Stress strain relationship of steel.

771
top plate. The SHS chord is connected to the SHS brace at right angles. The sizes of the welds
were determined in accordance to AS 4100-1998 (Standards Australia Online 1998). The leg
length (tw) and throat thickness (tt) of each SHS T-joint specimen is 6 mm and 4.24 mm, respect-
ively. The nominal tensile strength of the weld metal (fuw) is 480 MPa, AS 4100-1998 (Standards
Australia Online 1998).

3 EXPERIMENTAL STUDY

3.1 Instrumentation
Single and strip strain gauges were attached onto four empty SHS T-joint specimens listed in
Table 1 for the measurement of strains. The hot spot locations (line A-E) and single strain
locations are shown in Figure 2. Three (3) single strain gauges and five (5) strip strain gauges
were attached onto each SHS-SHS T-joint specimen. Each strip strain gauge consists of
5-element single strain gauges. The gauge resistance of the strain gauges is 120 ± 0.5 Ω.
Prior to attaching the strain gauges, the paint and oxidized layer at the location of each
strain gauge was removed using sanding paper (Metal Cloth). M-Prep Conditioner A MCA-2
(16-OZ bottle) was applied on the locations of the strain gauges. The M-Prep Conditioner
A MCA-2 (16-OZ bottle) was used in order to remove the oxidized layer. Then, M-Prep Neu-
tralizer 5A was applied on the same locations to remove the acid of the M-Prep Conditioner
A. The strain gauges were attached using Cyanoacrylate Strain Gauge Glue.
As shown in Figures 2 and 3, strip strain gauges were attached along the chord and the brace
intersection at lines A, B, C, D and E. As recommended by Zhao et al. (2001), the minimum
distance of the first strain gauge closest to the weld toe is 0.4t mm where t is the tube wall thick-
ness. If 0.4t is less than 4 mm, the minimum distance of the first strain gauge closest to the weld
toe will then equals to 4 mm as stated by Zhao et al. (2001). Therefore, the first strain gauge of
the strip strain gauge was installed at 4 mm from the weld toe since 0.4t is less than 4 mm.

Figure 2. Locations and labels of strip and single strain gauges.

772
Figure 3. Attaching the strip strain gauges.

3.2 Test setup and loading procedure


The T-joint specimens are connected to the testing rig via M12 bolts of grade 8.8 snug tight.
The SHS-SHS T-joint specimens are pin connected at both end plates. Each end plate com-
prised of 4 × 13 mm holes and bolted to the end brackets to support the T-joint specimens.
The testing of the SHS-SHS T-joint specimens was setup under out-of-plane bending loads on
the brace. The test setup for out-of-plane bending displayed in Figures 4 and 5 was used for
the measurement of the strain distribution. The strain gauges were connected to the data
logger to automatically record the strain gauge readings. Out-of-plane bending loads were
applied to each T-joint specimen. The loads applied to each specimen fall within the elastic
response region of the load-deformation curve of the SHS-SHS T-joint connections. The loca-
tion of this loading condition is shown in Figures 4 and 5.
Sixteen cycles of loads were applied on the brace of each specimen. These sixteen cycles of
loads comprise of four increased levels of the quasi-static loads followed by cycling at the 4th
load eight times and then again the first four cycles of loading levels. As shown in Figure 6,
for specimen S6S1, the levels of the quasi-static loads increased from 0.2 kN to 0.26 kN at
0.02 kN intervals and cycled at 0.26 kN nine times and finally quasi-static loads were applied
from 0.2 kN to 0.26 kN at 0.02 kN intervals. This process is used in order to stabilize
the loading process and obtain accurate results. The loads that were applied to each empty
SHS-SHS T-joint specimen are shown in Table 2. These loads are within the elastic response
range of the chord and brace member load-deformation curve under static loading. This will
allow the researcher to carry out the fatigue testing correctly as exceeding the plastic limit
load will result in permanent deformation or specimen failure.

3.3 Stress concentration factor (Scf)


The SCFSHS at a hot spot location was calculated using Eq. (1) as recommended by Zhao
et al. (2001).
SCFSHS; A-E ¼ 1:1  SNCFSHS ð1Þ

The SNCF for each load case is defined as the ratio of the hot spot strain (HSSN) and nom-
inal strain. Since strains were measured at four (4) load levels, four (4) SNCFs were calculated
for each strip strain gauge and for each location from line A to E. The strain concentration
factor for the connection (SNCFSHS) for each strip strain gauge was determined by averaging
the four SNCFs.

773
Figure 4. Test setup.

Figure 5. Test rig: out-of-plane bending.

The hot spot strain for each load case was calculated using quadratic extrapolation. Figure 7
compares the linear and quadratic extrapolation methods in determining the hot spot strains for
a typical T-joint specimen along line A. The quadratic hot spot strains are higher than the linear
hot spot strains. In rectangular hollow section (or SHS), Zhao et al. (2001) recommended the
use of quadratic extrapolation method as the distribution of the stresses is non-linear. Mashiri
& Zhao (2010) used quadratic extrapolation for the determination of the hot spot stresses.
Therefore, in this investigation, quadratic extrapolation method was used for determining the
SCFs at the hot spot locations since the distribution of the strains at the intersection is non-
linear. The nominal strains for each specimen under out-of-plane bending were calculated
through the use of linear extrapolation method. The extrapolation points are located on the ten-
sion side of the brace; at the load location, top quarter points of the brace, at the middle of the
brace, bottom quarter points of the brace and at the brace end.

774
Figure 6. Out-of-plane bending load applied on the brace of empty SHS S6S1 T-joint.

Table 2. Out-of-plane loads applied to each SHS T-joint specimen.


Out-of-plane bending load

kN

Loading Stage S6S1 S6S2 S6S4 S5S5

Stage 1 0.20 0.20 0.60 0.50


Stage 2 0.22 0.25 0.72 1.50
Stage 3 0.24 0.35 0.92 3.00
Stage 4 0.26 0.42 1.32 4.60

Figure 7. Determination of hot spot strain (HSSN) using linear and quadratic extrapolation methods.

775
4 RESULTS AND DISCUSSION

The distribution of the SCF in the four empty SHS T-joint connections is shown in Table 3 and
Figure 8. Each Figure displays a plot of SCF against the hot spot locations for each T-joint
specimen under out-of-plane bending load on the brace. The maximum experimental SCFs in

Table 3. Experimental SCFs in unfilled T-joints under out-of-plane bending.


SCF (quadratic) in Empty T-joint connections under
out-of-plane bending

Series name Line B Line C Line D Line A Line E Maximum SCF

S6S1 2.29 1.67 0.46 1.45 0.71 Line B – 2.29


S6S2 4.60 2.73 1.13 3.55 1.88 Line B – 4.60
S6S4 4.84 4.14 1.69 5.76 0.99 Line A – 5.76
S5S5 4.18 6.23 2.09 4.34 1.59 Line C – 6.23

Figure 8. SCFs in empty SHS-SHS T-joints under out-of-plane bending: (a) Empty S6S1; (b) Empty
S6S2; (c) Empty S6S4; and (d) Empty S5S5.

776
unfilled S6S1, S6S2, S6S4 and S5S5 under out-of-plane bending occurred along line B (2.29),
B (4.60), A (5.76) and C (6.23), respectively. The non-dimensional parameter, β influences the
maximum SCFs under out-of-plane bending. The maximum SCFs occurred at different loca-
tions of interest due to the change of non-dimensional parameters β. The maximum SCFs in
S6S1 and S6S2 occurred along line B, where line B in these two specimens is located on the
chord’s top face. However, the maximum SCFs in S6S4 and S5S5 did not occur along line B,
where line B in S6S4 and S5S5 is located on the external corner of the chord and chord’s front
face, respectively. The values of the maximum SCFs in this experimental study increase with the
increased value of β. Since CIDECT Design Guide 8 (Zhao et al. 2001) does not provide SCF
formulas and trends for SHS T-joints under out-of-plane bending, numerical SCFs will be deter-
mined in the future using ABAQUS software to verify the experimental results.

5 CONCLUSIONS

An experimental study on four unfilled SHS-SHS T-joint specimens was carried out. The
T-joint connections were tested under static out-of-plane bending. Strain gauges were installed
onto each specimen in order to determine the SCFs at the hot spot locations (Line A-E). The
maximum SCF of specimen S6S1 and S6S2 occurred on the chord along line B. In addition,
the peak SCF of specimen S6S4 and S5S5 under out-of-plane bending occurred along line
A and C, respectively. In the future, numerical investigation on unfilled SHS T-joints identical
to the T-joints used in this experimental investigation will be carried out under axial loads, in-
plane bending and out-of-plane bending load on the braces. In addition, experimental, numer-
ical and parametric investigations will be conducted on SHS-SHS T-joint connections with
concrete-filled chord to study the benefits of concrete filling in reducing the maximum SCFs
and improving the overall performance of the T-joint connections.

REFERENCES

Cheng, B. Qian, Q. & Zhao, X.L. 2015. Numerical investigation on stress concentration factors of square
bird-beak SHS T-joints subject to axial forces, Thin-Walled Structures, vol. 94, pp. 435 – 45.
Chen, Y. & Wang, J. 2015. Numerical study and design equations of square and diamond bird-beak SHS
T-joints under axial compression, Thin-Walled Structures, vol. 97, pp. 215 – 24.
Feng, R. & Young, B. 2010. Design of concrete-filled stainless steel tubular connections, Advances in
Structural Engineering, vol. 13, no. 3, pp. 471 – 92.
Mashiri, F.R. & Zhao, X.L. 2010. Square hollow section (SHS) T-joints with concrete-filled chords sub-
jected to in-plane fatigue loading in the brace, Thin-Walled Structures, vol. 48, no. 2, pp. 150 – 58.
Mashiri, F.R. Zhao, X.L. & Grundy, P. 2004. Stress concentration factors and fatigue behaviour of welded
thin-walled CHS–SHS T-joints under in-plane bending, Engineering Structures, vol. 26, no. 13, pp.
1861–75.
Matti, F.N. & Mashiri, F.R. 2018. Numerical study on SCFs of empty and concrete-filled SHS-SHS
T-joints under in-plane bending, Ninth International Conference on Advances in Steel Structures, vol. 2,
pp. 875 – 86.
OneSteel Market Mills. 2004. Cold formed structural hollow sections & profiles, 4th edn.
Standards Australia Online. 1998. Steel structures, AS 4100-1998, viewed 27 January 2018, SAI Global
database.
Standards Australia Online. 2009. Cold-formed structural steel hollow sections, AS 1163-2009, viewed
24 March 2018, SAI Global database.
Tong, L. Fu, Y. Liu, Y. & Zhao, X.L. 2014. Stress concentration factors of diamond bird-beak SHS
T-joints under brace loading, Thin-Walled Structures, vol. 74, pp. 201 – 12.
Tong, L.W. Zheng, H.Z. Mashiri, F.R. & Zhao, X.L. 2012. Stress-concentration factors in circular
hollow section and square hollow section T-connections: experiments, finite-element analysis, and
formulas, Journal of Structural Engineering, vol. 139, no. 11, pp. 1866 – 81.
Zhao, X.L. Herion, S. Packer, J.A. Puthli, R.S. Sedlacek, G. Wardenier, J. Weynand, K. van
Wingerde, A.M. & Yeomans, N.F. 2001. Design guide for circular and rectangular hollow section
welded joints under fatigue loading, TüV-Verlag.

777
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Experimental study of cold-formed high strength steel circular


hollow sections

X. Meng & L. Gardner


Department of Civil and Environmental Engineering, Imperial College London, London, UK

ABSTRACT: An experimental study into the cross-sectional behaviour of cold-formed high


strength steel circular hollow sections (CHS) is described in this paper. A total of six CHS
profiles with steel grades of S700 was examined, spanning from Class 1 to 4 (in compression)
according to Eurocode 3. The investigation consisted of twelve tensile coupon tests, six stub
column tests, fifteen short beam-column tests, six four-point bending tests and six three-point
bending tests. The obtained experimental results revealed that the studied CHS were all cap-
able of reaching the plastic cross-section capacities under the considered loading scenarios,
and their resistances under moment gradients were shown to be on average 8% higher than
those under constant moments.

1 INTRODUCTION

High strength steels with grades above S460 are being increasingly used in structural engineering
due to the combined benefits of high strength-to-weight ratio, lighter structural components as
well as reduced overall cost and carbon emissions. Two types of heat treatments – quenching
and tempering and thermomechanical rolling, are commonly used for producing high strength
steels. With the advances in production techniques, structural high strength steels with yield
strengths over 1100 MPa are now commercially available. Recent years have witnessed signifi-
cant developments in the area of high strength steel tubular sections, which are available mainly
as hot-finished and cold-formed products. High strength steel circular hollow sections (CHS)
are the focus of the present study. A number of experimental studies have been carried out on
high strength steel CHS, such as Pournara et al. (2017) on hot-rolled CHS and Zhao (2000) and
Ma et al. (2016) on cold-formed CHS. Overall though, the experimental data on high strength
steel CHS are still rather scarce, and further tests are considered necessary.
This paper presents a comprehensive experimental investigation into the cross-sectional
behaviour of cold-formed high strength steel CHS. The programme comprised twelve tensile
coupon tests, six stub column tests, fifteen short beam-column tests, six four-point bending
tests and six three-point bending tests, covering the load cases of axial compression, combined
loading, uniform bending moments and moment gradients.

2 EXPERIMENTAL INVESTIGATION

2.1 General
This section describes a comprehensive experimental investigation into the cross-sectional
behaviour of cold-formed high strength steel CHS. Six CHS profiles, all cold-formed from ther-
momechanically rolled S700 steel, were studied with the following nominal cross-section sizes
(outer diameter × thickness in mm): 168.3 × 4, 139.7 × 4, 139.7 × 5, 139.7 × 6, 139.7 × 8 and
139.7 × 10. The examined cross-sections covered Class 1 to Class 4 in compression and Class 1
to Class 3 in bending according to the slenderness limits specified in prEN 1993-1-1:2018.

778
2.2 Tensile coupon tests
The engineering stress-strain curves and key mechanical properties of the tested materials
were obtained through tensile coupon tests. In total, twelve tensile coupons, with two
extracted from each cross-section at 90° to the welding seam, were prepared and tested. The
tensile coupon tests were performed using an Instron 250 kN hydraulic testing machine. Two
linear electrical resistance strain gauges, along with a video extensometer, were used to meas-
ure axial strains. Note that the coupon specimens exhibited significant initial longitudinal
curvatures upon extraction from the CHS due to the release of bending residual stresses; their
influence was inherently incorporated into the obtained stress-strain curves as the specimens
straightened during testing. The hydraulic testing machine was driven under displacement
control, with the displacement rate chosen in accordance with EN ISO 6892-1:2016.
The obtained engineering stress-strain curves are plotted in Figure 1. The key mechanical
properties, including the Young’s modulus E, the yield stress fy (taken as the 0.2% proof
stress), the ultimate tensile stress fu, the ultimate strain εu and the percentage elongation at
fracture based on the standard gauge length εf, are reported in Table 1.

2.3 Local imperfection measurements


The distributions of local geometric imperfections were measured for all the stub column and
short beam-column specimens. The specimens were seated on a pair of V-shaped blocks,
which were aligned and clamped onto the bed of a milling machine. The specimens were then
moved along the length using the milling machine, with its bed serving as a reference flat
plane. A linear variable displacement transducer (LVDT) was fixed on top to record the devi-
ations at 2 mm intervals. For each specimen, the measurement was carried out along twelve
equally spaced longitudinal lines at 30° intervals around the circumference.

Figure 1. Average engineering stress-strain curves from tensile coupon tests.

Table 1. Average measured material properties from tensile coupon tests.


E fy fu εu εf fu/fy εu/εy
2 2 2
Cross-section N/mm N/mm N/mm % % - -

CHS 168.3 × 4 211700 720.0 823.4 5.9 15.3 1.144 17.4


CHS 139.7 × 5 213300 742.4 842.3 4.7 15.9 1.135 13.5
CHS 139.7 × 4 212500 729.7 843.3 4.7 14.6 1.156 13.7
CHS 139.7 × 5 207900 779.0 866.7 2.8 13.7 1.113 7.5
CHS 139.7 × 8 205700 784.8 866.8 2.9 13.9 1.105 7.5
CHS 139.7 × 10 205600 787.6 877.5 2.2 13.8 1.114 5.7

779
Figure 2. Typical measured local imperfection patterns along one longitudinal line.

Typical measured deviations are illustrated in Figure 2. End flaring can be clearly observed
from the results, which is primarily attributed to the release of bending residual stresses. There-
fore, only the data within the central 80% of the length were used to record representative local
imperfections. Linear regression analysis was performed for each set of the measurement data,
and the deviations with respect to the fitted line were considered as the local geometric imper-
fections. The local imperfection amplitude ωl of a specimen was taken as the maximum devi-
ation of all measurements around its perimeter, as summarised in Tables 2 and 3 for the stub
columns and short beam-columns respectively.

2.4 Stub column tests


A total of six stub column tests, one on each of the CHS profiles, was carried out to investi-
gate their local buckling responses and cross-sectional capacities under axial compression. The
nominal length of the stub columns was chosen as three times the outer diameter, as recom-
mended by Ziemian (2010). The ends of the specimens were machined flat and square prior to
testing. The measured geometric properties of the stub column specimens are summarised in
Table 2, where D is the outer diameter, t is the wall thickness and L is the specimen length.
The axial compression tests were performed using an Instron 3500 kN hydraulic testing
machine, as shown in Figure 3. The stub columns were compressed between two parallel end
platens of the testing machine whereby fixed boundary conditions were achieved. Hardened
end plates were placed between the end platens and specimens to avoid damage to the platens.
Steel ring stiffeners were mounted at each end of the specimens to prevent any distortion or
local bending at the end cross-sections. The instrumentation consisted of four strain gauges at
the mid-height at 90° intervals plus four LVDTs around the specimens. The stub column tests
were conducted under displacement control at a constant rate of 0.5 mm/min.

Table 2. Measured geometric properties and key results from stub column tests.
D t L ωl D/(tε2) Npl Nu δu Nu/Npl

Cross-section ID mm mm mm mm - kN kN mm -

CHS 168.3 × 4 CSC1 168.42 3.95 505.31 0.12 130.5 1471.1 1634.6 5.1 1.111
CHS 139.7 × 4 CSC2 140.10 3.98 420.28 0.10 111.2 1263.6 1419.1 5.9 1.123
CHS 139.7 × 5 CSC3 140.42 4.91 420.08 0.08 88.9 1524.1 1798.5 7.1 1.180
CHS 139.7 × 6 CSC4 139.82 6.04 420.02 0.15 76.7 1978.4 2261.5 8.2 1.143
CHS 139.7 × 8 CSC5 140.06 7.85 420.25 0.12 59.6 2557.4 2919.6 10.6 1.142
CHS 139.7 × 10 CSC6 140.29 9.93 419.99 0.10 47.4 3202.4 3760.0 13.9 1.174

780
Table 3. Measured geometric properties and key results from short beam-column tests.
D t L ωl en em Nu Δu φu Mu,tot

Cross-section ID mm mm mm mm mm mm kN mm - kNm

CHS 168.3 × 4 CSBC1-1 168.35 3.93 505.04 0.19 10 10.37 1400.6 4.92 1.35 21.4
CHS 168.3 × 4 CSBC1-2 168.37 3.93 505.09 0.25 25 25.02 1194.7 4.86 1.33 35.7
CHS 168.3 × 4 CSBC1-3 168.43 3.93 505.14 0.16 40 40.22 985.2 4.98 1.54 44.5
CHS 168.3 × 4 CSBC1-4 168.07 3.93 505.35 0.31 75 76.02 726.8 5.88 1.78 59.5
CHS 168.3 × 4 CSBC1-5 168.59 3.91 504.60 0.29 150 151.01 455.9 6.77 1.94 71.9

CHS 139.7 × 4 CSBC2-1 140.06 3.95 419.15 0.14 10 9.94 1129.2 5.42 1.87 17.3
CHS 139.7 × 4 CSBC2-2 140.11 3.94 419.42 0.16 25 25.02 924.4 5.92 2.05 28.6
CHS 139.7 × 4 CSBC2-3 140.02 3.94 420.01 0.33 50 49.43 715.4 7.43 2.45 40.7
CHS 139.7 × 4 CSBC2-4 140.13 3.94 420.16 0.18 85 83.64 505.6 6.70 2.41 45.7
CHS 139.7 × 4 CSBC2-5 140.08 3.95 419.80 0.17 170 171.24 302.7 7.81 2.49 54.2

CHS 139.7 × 5 CSBC3-1 140.38 4.88 420.24 0.11 10 9.72 1417.2 5.99 1.89 22.3
CHS 139.7 × 5 CSBC3-2 140.35 4.86 420.83 0.11 25 24.51 1136.6 6.61 2.31 35.4
CHS 139.7 × 5 CSBC3-3 140.34 4.89 419.66 0.13 50 49.69 887.8 7.23 2.45 50.5
CHS 139.7 × 5 CSBC3-4 140.37 4.88 420.16 0.15 85 83.84 648.4 8.16 2.54 59.7
CHS 139.7 × 5 CSBC3-5 140.37 4.86 420.00 0.17 170 168.93 381.6 8.78 3.08 67.8

Figure 3. Stub column test setup (left) and typical failure modes (right, scale in mm).

The normalised load-end shortening curves are displayed in Figure 4, where Npl is the plas-
tic load and δpl is the elastic end-shortening at Npl (equal to fyL/E). The ultimate axial load
Nu, the end shortening at the peak load δu, the local slenderness D=ðtε2 Þ, where ε2 ¼ 235=fy ,
the plastic load Npl and the ratio of Nu to Npl are reported in Table 2. All the stub columns
failed by the formation of a bulge near the end, as shown in Figure 3. All tested sections,
including those falling into the Class 4 range, were capable of reaching Npl. Furthermore, with
the increase of the local slenderness, an increasing susceptibility to local buckling can be
observed in Figure 4.

2.5 Short beam-column tests


The local buckling behaviour and load-carrying capacities of the tested cross-sections under
combined compression and bending were investigated through short beam-column tests. In
total, fifteen short beam-column tests were carried out, five on each of the CHS 168.3 × 4,
CHS 139.7 × 4 and CHS 139.7 × 5 profiles. The nominal length of the specimens was also
taken as three times the outer diameter, as for the stub columns. The measured dimensions
and local imperfection amplitudes of the short beam-column specimens are reported in
Table 3.

781
Figure 4. Normalised load-end shortening curves from stub column tests.

The short beam-column tests were conducted in an Instron 2000 kN hydraulic testing
machine, as shown in Figure 5. A pair of knife edges were employed at the top and bottom of
the rig, by which pin-ended boundary conditions were provided. The test specimens, with end
plates welded onto both ends, were bolted eccentrically onto the wedge plates. The initial load-
ing eccentricities were carefully measured prior to each test, as reported in Table 3, where en
and em are the nominal and measured loading eccentricities respectively. Strain gauges were
attached to the mid-height of the specimens at the most compressive and tensile (or least com-
pressive) fibres. The mid-height lateral deflections were measured using an LVDT on the ten-
sile (or least compressive) side, and two inclinometers were attached onto the top and bottom
wedge plates to monitor the end rotations. The tests were performed under displacement con-
trol at a constant rate of 0.5 mm/min.
The load-deformation curves from the short beam-column tests are plotted in Figure 6, and
the key experimental results are summarized in Table 3, where Nu is the peak load, Δu and φu
are the mid-height lateral deflection and end rotation at the peak load respectively and Mu,tot is
the total bending moment at the peak load. The second-order effects were incorporated into the
total bending moment Mtot, as given by Equation (1). All the short beam-columns failed due to
local buckling on the most compressive side, with a typical failure mode displayed in Figure 5.

Mtot ¼ N ðem þ ΔÞ ð1Þ

Figure 5. Short beam-column test setup (left) and typical failure mode of CSBC2-4 (right, scale in mm).

782
Figure 6. Load-deformation curves from short beam-column tests.

2.6 Four-point bending tests


Four-point bending tests were conducted on the examined CHS profiles to investigate their
local buckling responses, moment resistances and rotation capacities under constant bending
moments. In total, six four-point bending tests, one on each cross-section, were performed.
The measured cross-section dimensions are summarised in Table 4.
The bending tests were performed using an Instron 2000 kN hydraulic testing machine; the
test setup is shown in Figure 7. The beam specimens were seated symmetrically on a pair of
roller supports at a distance of 1850 mm, which provided simply-supported boundary condi-
tions, with an overhang of 75 mm at each end. The vertical load was applied through
a spreader beam to two steel rollers as two equal vertical loads onto the specimens. The dis-
tance between the two rollers was equal to 720 mm for the CHS 168.3 × 4 profile and equal to
600 mm for the remainder. The loading points of the specimens were stiffened through wood
blocks inside the tubes and saddle-shaped steel seats between the rollers and tubes, with
rubber sheets placed between the seats and the specimens to ensure a full contact. Three
LVDTs were mounted at the mid-span and two loading points to measure the vertical dis-
placements, and four inclinometers were attached onto the steel saddles to measure the rota-
tions at the loading points. In addition, a Digital Image Correlation (DIC) technique, using
a two-camera StrainMaster DIC system from LaVision, was employed to monitor the deform-
ation within the flexural span.
The normalised moment-average curvature curves are plotted in Figure 8, where Mpl is the
plastic moment resistance and κpl is the elastic curvature at Mpl (equal to Mpl/EI, where I is
the second moment of area). Table 4 summarises the key experimental results, including the
local slenderness D=ðtε2 Þ, the ultimate bending moment Mu, the average curvature at the ultim-
ate moment κu, the rotation capacity R and the ratio of the ultimate moment to the plastic

Table 4. Measured geometric properties and key results from four-point bending tests.
D t D/(tε2) Mpl Mu κu R Mu/Mpl

Cross-section ID mm mm - kNm kNm × 10-4/mm - -

CHS 168.3 × 4 C4B1 168.33 3.92 131.6 76.3 80.0 1.79 2.90 1.048
CHS 139.7 × 4 C4B2 140.24 3.91 113.3 54.0 57.1 2.61 3.58 1.059
CHS 139.7 × 5 C4B3 140.47 4.88 89.4 65.5 71.3 3.15 5.39 1.089
CHS 139.7 × 6 C4B4 139.70 6.03 76.8 84.0 88.7 3.47 6.02 1.056
CHS 139.7 × 8 C4B5 140.16 7.83 59.8 107.7 113.3 3.57 7.08 1.052
CHS 139.7 × 10 C4B6 140.20 9.92 47.4 132.9 142.8 4.19 >7.99* 1.075

* Test terminated before moment dropped below Mpl on descending branch.

783
Figure 7. Four-point bending test setup.

Figure 8. Normalised moment-curvature curves from four-point bending tests (left) and typical failure
mode of C4B4 (right, scale in mm).

moment capacity Mu =Mpl . The rotation capacity R was determined from the four-point bending
tests according to Equation (2), where κpl,u is the curvature at which the applied moment
dropped to Mpl on the descending branch. All the four-point bending specimens failed within
the central flexural region; a typical failure mode is shown in Figure 8. The tested sections range
from Class 1 to 3 in bending; however, all specimens failed at an ultimate moment above Mpl.

κpl;u
R¼ 1 ð2Þ
κpl

2.7 Three-point bending tests


Having examined the behaviour of the CHS under constant bending moment, three-point
bending tests were carried out to study their structural responses under moment gradients.
A total of six three-point bending tests, one test on each of the CHS profiles, was conducted.
The measured cross-section dimensions are summarised in Table 5.
The three-point bending tests were performed using a similar test setup to that for the four-
point bending tests, but with a modified loading configuration, as shown in Figure 9. A fixed

784
Table 5. Measured geometric properties and key results from three-point bending tests.
D t D/(tε2) Mpl Mu θu R Mu/Mpl

Cross-section ID mm mm - kNm kNm ° - -

CHS 168.3 × 4 C3B1 168.35 3.90 132.3 76.0 84.9 2.61 1.93 1.118
CHS 139.7 × 4 C3B2 140.12 3.92 112.9 54.0 61.4 3.65 2.61 1.137
CHS 139.7 × 5 C3B3 140.47 4.89 89.2 65.6 77.0 4.42 >3.68* 1.173
CHS 139.7 × 6 C3B3 139.56 6.02 76.8 83.7 96.5 5.15 4.62 1.153
CHS 139.7 × 8 C3B5 140.16 7.84 59.7 107.8 124.9 6.25 >6.09* 1.158
CHS 139.7 × 10 C3B6 140.22 9.90 47.5 132.7 154.4 6.85 >3.18* 1.164

* Tests terminated before moment dropped below Mpl on descending branch.

Figure 9. Three-point bending test setup.

roller was mounted directly beneath the hydraulic jack to apply a point load to the mid-span
of the specimens. Wood blocks and saddle-shaped steel seats were utilised at the loading
points to prevent local failure under the concentrated loading. An LVDT was mounted at
mid-span to measure the vertical deflection, while three inclinometers were attached to the
steel seats to monitor the rotations at the mid-span and both ends. DIC was also employed to
monitor the deformation near the higher moment regions (i.e. near the mid-span).
The normalised maximum moment-end rotation curves are plotted in Figure 10, where M is
the maximum (i.e. mid-span) bending moment and θpl is the elastic end rotation when M is
equal to Mpl. The test results are summarised in Table 5, where θu is the end rotation at Mu and
R is the rotation capacity calculated using Equation (3). All specimens failed by development of
local buckling in the highest moment region close to the mid-span, as shown in Figure 10. The

Figure 10. Normalised maximum (i.e. mid-span) moment-end rotation curves from three-point bending
tests (left) and typical failure mode of C3B4 (right, scale in mm).

785
ultimate bending moments obtained from the three-point bending tests are on average 8%
higher than those from the four-point bending tests, due primarily to the favourable effects
from the moment gradients compared with the constant bending moments.

θpl;u
R¼ 1 ð3Þ
θpl

3 CONCLUSIONS

An experimental investigation into the cross-sectional behaviour of cold-formed S700 circular


hollow sections (CHS) has been conducted. Six CHS profiles have been studied, ranging from
Class 1 to 4 in accordance with prEN 1993-1-1:2018. In total, twelve tensile coupon tests, six
stub column tests, fifteen short beam-column tests, six four-point bending tests and three-
point bending tests have been performed and are reported. The key test results reveal that all
the examined CHS profiles were capable of reaching the plastic cross-section capacities under
various loading scenarios including axial compression, bending and combined loading. For
the case of bending, their capacities under moment gradients were shown to be on average 8%
higher than those under constant moments. The obtained experimental data will be used for
the validation of numerical models and the development of the design guidance for high
strength steel CHS in future work.

ACKNOWLEDGEMENTS

This research was funded by the Research Fund for Coal and Steel (RFCS) under Grant
Agreement No. RFCS-2015-709892. The test materials were provided by SSAB, whose sup-
port is gratefully acknowledged.

REFERENCES

EN ISO 6892-1:2016. Metallic materials - Tensile testing - Part 1: Method of test at room temperature.
The British Standards Institution.
Ma, J. L., Chan, T. M. & Young, B. 2016. Experimental investigation on stub-column behavior of
cold-formed high-strength steel tubular sections. Journal of Structural Engineering 142(5): 04015174.
prEN 1993-1-1:2018. Eurocode 3 — Design of steel structures — Part 1-1: General rules and rules for build-
ings. European Committee for Standardisation.
Pournara, A. E., Karamanos, S. A., Mecozzi, E. & Lucci, A. 2017. Structural resistance of high-strength
steel CHS members. Journal of Constructional Steel Research 128: 152–165.
Zhao, X. L. 2000. Section capacity of very high strength (VHS) circular tubes under compression. Thin-
Walled Structures 37(3): 223–240.
Ziemian, R. D. 2010. Guide to Stability Design Criteria for Metal Structures. John Wiley & Sons.

786
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Towards automated identification of structural steel components


from 3D-point clouds to subsequent GMNA-stability-analysis

C. Merkl & A. Taras


Institute of Structural Engineering, University of the Bundeswehr Munich, Germany

ABSTRACT: Due to the growing computational performance and advancing image-


processing capabilities, comprehensive geodetic data acquisition of structures has become pos-
sible. Such a digital inventory of buildings is performed either through photogrammetry or
through laser scanning, and it results in 3D point clouds. Until recently, however, their goal-
oriented further processing posed problems for civil engineers in terms of actual information
acquisition, data compatibility and data transfer to general software used in day-to-day prac-
tice. Thereby, the main issue persists in a reasonably quick and intuitive deduction of the
required specific information for the engineering task, e.g. performing the structural reevalua-
tion of an existing bridge structure. To carry out such an analysis, structural engineers require
very specific information of individual structural components as well as of their assembly to
generate the structural model and subsequently solve it. In steel construction, for example,
this modelling is primarily based on beam and shell elements, which first require the reduction
of captured three-dimensional information to surfaces and further to one-dimensional or two-
dimensional component descriptions. As an initial step towards addressing the described
problem, this work presents an algorithmic approach to extracting characteristic geometric
information of a selected steel component from its point cloud, which is required for
a subsequent finite element model. For this purpose, the point cloud data of previously laser
scanned channel sections is used. In this data, basic geometric elements, in particular planes
and cylinders are automatically recognised and measured via computer vision tools. After-
wards, the acquired information is transferred to input data for the FE-software ABAQUS
and solved. Finally, a numerical study shows the robustness of this procedure.

1 INTRODUCTION

Current modelling in steel construction is largely based on column and beam elements. A fact
that won`t be changed by the new trend of BIM modelling (Nöldgen 2016). Equivalent to most
calculation software, the main member information is entered in form of one-dimensional elem-
ents with a defined end and beginning. However the visualization itself progress a more real
three-dimensional representation. This coincides both with the framework-based mechanical
modelling of Eurocode 3 (DIN EN 1993-1-1 2010), and with most structural analysis software.
However, the calculation and evaluation of already existing elements and structures requires
anew, partly intricate model construction from planning documents.
In order to simplify the analysis of existing components, the following paper makes an ini-
tial contribution towards automated recognition and evaluation for the special case of labora-
tory profiles modelled as shell elements. Thus, it follows the recent tendencies to incorporate
a more comprehensive, general two-dimensional modelling approach based on shell elements
for connections (Wald et al. 2016) and the development plans for a more comprehensive FE-
based design (EN 1993-1-14). On the path towards this advanced FEM concept in steel con-
struction, researches currently apply the GMNIA modelling technique as defined in part 1-6
of Eurocode 3 (DIN EN 1993-1-6 2010) to calibrate the FEM models to the conducted experi-
ments for new design approaches of stability driven steel components. Some exemplary

787
publications presenting this approach are (Winterstetter et al. 2001), (Zhao et al. 2017) and
(Ladinek et al. 2018).
However, as this technique requires the quantitative knowledge of the geometric imperfec-
tions, its implementation is heavily dependent on manual analysis and, therefore, very time
consuming. This paper attends to an automated GMNA analysis as a prerequisite for the
future goal of an automated GMNIA analysis just based on the preceding measurement of the
real steel component`s geometry in form of a three-dimensional point cloud.
Hereby, the paper starts with an overview over previously developed concepts and
approaches to automatically transfer scan data to a FEM analysis based on computer vision
methods to assist in the identification of geometric features of point clouds. Then it depicts
the applied scanning technology to detect the steel profiles outer surface geometry. After that
a procedural approach implemented in the proprietary numerical computing software
MATLAB (MATLAB 2018c) is described to obtain the main geometric properties of the steel
component in question, which will further be inserted into afore prepared calculation scripts
in the proprietary FEM software ABAQUS (Abaqus 2016). The paper concludes with an out-
look on the next implementation steps to automate the GMNIA analysis persistently for steel
components in research applications.

2 CURRENT DEVELOPMENTS AND LITERATURE STUDY

Computer Vision is a currently rapidly developing field of research that deals with the auto-
mated recognition of objects in digitally measured data, such as point clouds. In civil engineer-
ing, this leads to the interaction of pattern recognition, computational engineering and classical
geodesy, and thus to a possible, direct conjunction to current 3D-BIM modelling and hence-
forth FEM analysis. As a result of this, the present literature spreads vastly with a bit diverse
focus based on the respective research focuses of the authors, yet a lack of more specified litera-
ture stemming from the applied structural engineering disciplines is apparent. Two particular
areas, in which a relatively large number of applications have already been conducted, are vari-
ous cultural heritage constructions, i.e. (Barazzetti et al. 2015) or (Ioannides et al. 2018), and
the observation of changes occurring on sites or in facilities via automated point cloud analysis
for present geometric components, e.g. (Tang et al. 2014) or (Braun et al. 2018).
In contrast, this work focuses primarily on industrial and laboratory applications. Here, for
new components and structures, the basic three-dimensional objects, including their material
and shape properties, are generated directly on the computer. This is defined as-designed
modeling.
In the same direct sense, however, it is not possible to analyze existing components or structures.
In their case, the development of precise LiDAR (Light Detection and Ranging) systems sup-
ported by capable computational software rendered the collection of point cloud based representa-
tions of an object`s visible surface possible under reasonable effort, e.g. (Vosselman et al. 2010).
However, the meaningful assignment of these large amounts of data to the underlying real
objects posed and continues to pose a challenge, to which the three previously mentioned dis-
ciplines are dedicated. The latter is denoted by as-is or as-built modeling (Fathi et al. 2015).
An elaborate overview in applied point cloud analysis to structures` as-built state can be
found in (Riveiro et al. 2016).
Purposeful developments applying this to steel structures have been published in (Cabaleiro
et al. 2014), (Laefer et al. 2017), and (Riveiro et al. 2018). In contrast to this paper, these pub-
lications neither include two dimensional shell modelling by focusing either on one dimen-
sional wire structures or on three dimensional volumetric structures nor are they dealing with
stability analysis. For the specific case of bridge structures, automated analysis concepts
toward modelling via volumetric FEM-elements are presented in (Zhang et al. 2014), (Conde-
Carnero et al. 2016), and (Yan et al. 2017). In all the cases mentioned above, computer vision
tools regarding pattern recognition and machine learning, e.g. (Torr et al. 2000), are applied
to identify structural components without requiring a 3D-CAD- or BIM-model to compare it
to, in contrast to (Bosché 2010).

788
A conceptually similar approach based on an automated transfer of the manually prepared
point cloud into a NURBS surface for further analysis was conducted in (Lemeŝ 2010). How-
ever, this thesis does not yet include an entire point cloud object detection process and is more
focused on FEM-Scan comparison based on an automatically generated NURBS
representation.

3 OBTAINING THE POINT CLOUDS

3.1 Investigated steel profiles


This paper investigates point clouds depicting ideal I and H profiles and of real, cold-formed
edge U profiles. Both profile types are shown in Figure 1. The former were generated for test-
ing purposes from CAD models of the standardized profiles in the propriatary computer-
aided design software Autodesk Inventor (Autodesk 2016), and the latter were recorded for
the calibration of FEM models to tested specimens regarding their stability behaviour under
eccentric loading.
The actual project regarding the cold-formed channel sections dealt with the determination of
the maximum load-bearing capacity of eccentrically pressure-loaded U-profiles without lips. It
studied the channel sections` behavior under eccentric loading by caused by both: compression
and simultaneously occurring bending. The findings have been published in (Merkl et al. 2017).

3.2 Scanning devices


The sample surfaces were measured using a 3D hand-held scanner system. The actual scanning
was cooperatively performed by the neighboring institute of engineering geodesy with a Leica
T-Scan TS50-B profile scanner (Hexagon 2013). The point clouds were then manually cleared of
evident error data, such as the laboratory floor. This could be conducted via the GPL-licensed
3D point cloud processing software CloudCompare (CloudCompare 2018). So, the remaining
valid points constituting each object`s point cloud were exported in form of ASCI-files containing
simply each point’s x-y-z-coordinates. Additionally, CloudCompare also permits the comparison
either between two point clouds or between point cloud and STL-(stereolithography)-mesh with
the latter representing a common and possible CAD-export format.
In (Merkl 2018) this has been performed to obtain quantitative data on the existent imper-
fections. This would potentially constitute a possible foundation for importing imperfections
into a GMNIA-analysis, but it still requires pre-alignment steps to be entered manually.
Figure 2 shows thereby obtained exemplary imperfection data.
An eventual solution to this may result from extending the CloudCompare build by an even
more automated alignment in combination with a supplementary export of the quantified point-
to-point mapping differences. When this were to be setup in batch process callable, e.g. via

Figure 1. Profile types analyzed within this paper: a) CAD-model of an I-section, b) ideal point cloud
generated from the CAD-model c) photo of channel section specimens actually scanned by a 3D
hand-held scanner.

789
Figure 2. Measured imperfections on the channel section: a) STL-model of a scanned member,
b) CAD-model of the channel section under investigation, c) color chart representing the imperfections.

MATLAB, then CloudCompare could be used as a sophisticated intermediate solution for obtain-
ing quantitative imperfection analyses for further GMNIA studies. Unfortunately, with the pre-
sent lack of this option, a fully automated GMNIA analysis cannot yet be presented.

3.3 Current manual transformation of point cloud data to FEM


Thus, the present transformation of point clouds into computable FEM meshes originates
from a reverse engineering process performed manually in the proprietary software Geomagic
Design X (Geomagic Design X 2018). Figure 3 depicts an exemplary process chain of this,
which results in the surface that can be cross-linked in a FE program. The feasible connection
of that prevailing progression between STL-mesh and FEM-mesh resides with a NURBS-
mesh (Piegl et al. 1997) of a smoothness of at least class-C0 in between individual NURBS-
patches. At present, that needs to be manually derived from the point cloud via Geomagic
Design X. Figure 3d) displays such an adequate mesh and the consecutive one to the right the
FEM-mesh generated after importing it. Together with an elastic-plastic material behavior
created from accompanying tensile tests, GMNIA calculations can thus be carried out, which
contained the actual imperfections of the tested specimens. The last picture in Figure 3 finally
shows the result of the calculation.
The acquisition and processing of the real geometry thus contains potential, especially for
the test-based investigation of the load-bearing capacities of thin-walled profiles. Further, it
enables parameter studies (LBA and GMNIA) of ideal component geometries to be calibrated
not only on the test results themselves, but also on the imperfections measured and, hence,
included in the subsequent FEM simulations.

Figure 3. Sequence of transformation from point clouds into a FEM model: a) actual test specimen,
b) scanned point cloud, displayed in STL triangle format, c) colors highlighting recognizable imperfections,
d) waterproof NURBS surface, e) FEM mesh in ABAQUS, f) ABAQUS calculation results (Hertle 2017).

790
FEM-results originating from this manual transformation sequence devised by the authors
have been presented during the DASt-Colloquium 2018, and the process has meanwhile been
used at the institute several times in industrial and research projects, e.g. (Müller 2018), and
(Toffolon 2018). It is similar to (Conde-Carnero et al. 2016). From these experiences with the
time-consuming, yet interesting and insightful results being produced, follows the need for
automation.

4 IDENTIFICATION ALGORITHM

4.1 Selection of analysis procedure


MATLAB was chosen as the implementation environment, because it already provides basic
computational objects and analysis algorithms in its Computer Vision Toolbox (MATLAB
2018a). The point cloud data itself was imported via MATLAB’s Import Module (MATLAB
2018b), followed by direct conversion to a point cloud object.
In order to identify profiles, it is necessary to extract basic geometric shapes from the point
clouds (e.g. flat areas or cylindrical segments). Here, the functions already contained in
MATLAB (pcfitplane and pcfitcylinder), which are based on the M-estimator SAmple Consen-
sus (MSAC) of a variation of the RANdom SAmple Consensus (RANSAC) (Torr et al.
2000), were used due to their direct feasibility. For this it is, however, necessary to align the
point clouds beforehand according to the base coordinate planes, which is shown subse-
quently. Another alternative used in some of the literature sources presents the Principle Com-
ponent Analysis (Huber et al. 2005), which may be investigated in the future.
The chosen methodology was already applied in (Anil et al. 2012), but, due to the LiDAR
scan, incomplete and erroneous data, it revealed results with sometimes considerable devi-
ations. However, originating from the complete and hardly faulty point clouds obtained
under laboratory conditions, the procedure is considered to behave robust against the conse-
quential effects that occurred before.

4.2 Alignment via point cloud volume minimization


In general, point clouds are imported in correspondence to the global coordinate system, in
which the scanning system initially records them. Therefore, most clouds are located some-
where in space and exhibit an arbitrary oblique orientation with respect to the base coordinate
planes.
Given the previous knowledge that the point cloud embodies a single, basically longitudinal
steel construction profile, it is possible to achieve alignment by minimizing the total volume
occupied by the point cloud.
First, the center of gravity of the point cloud is calculated and its subsequent repositioning
into the origin of a local analysis coordinate system takes place. Now the whole cloud can be
rotated around all three coordinate axes one after the other, whereby the angle of rotation is
determined which minimizes the total volume and at the same time also the associated axis
interval. In this paper the following sequence took place with the specification of the axes of
rotation: y, z, and finally x. The resulting alignment is depicted in Figure 4 for visual clarifica-
tion using a colored cuboid.

4.3 Identification
In order to carry out the profile identification and following quantification, histograms of the
point distribution corresponding to the coordinate axes are first created, since a distinction
can be made on the basis of the peak value positions in these distributions as to whether it
contains an I or a U profile. Exemplarily, this is shown in Figure 5.
Now, the MSAC via MATLAB`s incorporated functions may be applied to recognize for
instance the planes contained within the point clouds. Here, the preceding alignment also

791
Figure 4. Alignment rotations: a) Positioning of the point cloud in the origin, b) Rotation around the
y-axis, c) Rotation around the z-axis, d) Final rotation around the x-axis.

Figure 5. Histograms to demonstrate the identification of web and flange surfaces for I- and U- profiles:
a) webs of the I-section, b) flanges of the I-section, c) flanges of the U-section, d) web of the U-section.

supports this process step, because now searches parallel to the principle coordinate planes
suffice. Results for both, the investigated channel and I-sections are displayed in Figure 6.
As computational result, MATLAB generates plane objects already containing quantifiers
such as width and length. So, a few additional analytical geometric plane distance calculations
remain to obtain enough dimensions for the profile in question to be compared to database
values of design code cross-sections (e.g. all parts of DIN 1025 2009).

Figure 6. Recognized planes: a) of I-sections in y-direction, b) of I-sections in z-direction, c) of


U-sections in y-direction, and d) of U-sections in z-direction.

792
Figure 7. Calculation result in ABAQUS: a) profile loaded close to the web, b) profile loaded at the
edges of the flanges.

5 FE-ANALYSIS

The previously obtained dimensions can now be transferred as parameters to the control files of
a FEM program. In case of this paper, the calculations have been implemented in ABAQUS by
preparing parameterized calculation scripts (Input-Files) for different steel profile types. Now only
the parameter values have to be transferred to them with the help of MATLAB, which is done by
a structured mesh programming and the assembly of prepared text modules to a finished INP file.
Then, the subsequent FEM calculation is automatically started through the system command
input of the latter. For the case of the U-profiles, finally achieved results are shown in Figure 7.

6 OUTLOOK

This paper showed an automated execution of GMNA calculations exclusively based on the
point cloud data of the profiles, suitable for laboratory based stability investigations of steel pro-
files. The next step will be to integrate imperfection data and thus achieve GMNIA calculations.
In the future, the robustness of the developed implementation should be improved and
made more flexible in order to be able to automatically process further profile types and their
different point clouds in an identical manner. Furthermore, the reviewed literature offers pos-
sibilities to improve the identification algorithm by testing other existing analysis strategies.

REFERENCES

Abaqus. 2016. Reference manual, version 6.16. Simulia, Dassault Systéms. France.
Anil E., Raghuram S., Akinci B. 2012. Challenges of Identifying Steel Sections for the Generation of
As-Is BIMs form Laser Scan Data. Gerontechnology 11 (2), 317.
Autodesk. Autodesk Inventor. Version 20.0. Autodesk. San Rafel, CA, USA.
Barazzetti, L. Banfi, F. Brumana, R. Gusmeroli, G. Previtali, M. Schiantarelli, G. 2015.
Cloud-to-BIM-to-FEM: Structural simulation with accurate historic BIM from laser scans. Simulation
Modelling Practice and Theory, Vol. 57, 71– 87.
Bosché, F. 2010. Automated recognition of 3D CAD model objects in laser scans and calculation of as-built
dimensions for dimensional compliance control in construction. Adv. Eng. Inform., 24, pp. 107–118.
Braun A., Tuttas S., Stilla U., Bormann A. 2018. BIM-Based Progress Monitoring. In: Building Information
Modeling. Springer.
Cabaleiro M., Riveiro B., Arias P., Caamano JC., Vilán JA. 2014. Automatic 3D modelling of metal
frame connections from LiDAR data for structural engineering purposes. ISPRS Journal of Photo-
grammetry and Remote Sensing, 96, 47–56.
Cloud Compare. 2018. CloudCompare (version 2.9.1) [GPL software]. Retrieved from http://www.cloud
compare.org/.
Conde-Carnero B., Riveiro B., Arias P. Caamano JC. 2016. Exploitation of Geometric Data provided by
Laser Scanning to Create FEM Structural Models of Bridges. Journal of Performance of Constructed
Facilities, 30 (3).
DIN EN 1025. 2010. Warmgewalzte I-Träger. Deutsches Institut für Normung. Berlin.

793
DIN EN 1993-1-1. 2010. Eurocode 3. Design of steel structures – Part 1-1: General rules and rules for
buildings. CEN – European Committee for Standardization. Brussels.
DIN EN 1993-1-6. 2010. Eurocode 3. Design of steel structures – Part 1-6: Strength and Stability of Shell
Structures. CEN – European Committee for Standardization. Brussels.
Fathi, H. Dai, F. Lourakis, M. 2015. Automated as-built 3D reconstruction of civil infrastructure using com-
puter vision: Achievements, opportunities, and challenges. Advanced Engineering Informatics Vol. 29 (2),
149–161.
Geomagic Design X. 2018. User Guide. 3D Systems. USA.
Hertle T., Garsch M., Gebbeken N. 2018. BIM/Bauwerksüberwachung. Der Prüfingenieur 53.
Hexagon Metrology. 2013.Leica T-Scan TS 50-A. Hexagon Metrology. Germany. www.hexagon
metrology.com.
Huber M., Rousseeuw P.J., Vanden Branden K. 2005. ROBPCA: A new approach to robust principal
component analysis, Technometrics 47 (1), 64–79.
Ioannides, M. Fink, E. Brumana, R. Patias, P. Doulamis, A. Marins, J. Wallace, M. 2018. Digital
Heritage. Nicosia. 7th International Conference EuroMed 2018 Part I.
Laefer D., Truong-Hong L. 2017. Toward automatic generation of 3D steel structures for building
information modelling. Automation in construction Vol 74, 66–77.
Ladinek, M. Niederwanger, A. Lang, R. Schmid, J. Timmers, R. Lener, G. 2018. The strain-life applied to
welded joints: Considering the real weld geometry. Journal of Constructional Steel Research Vol. 148,
180–188.
Lemeŝ, S. 2010. Validation of Numerical Simulations by digital scanning of 3D sheet metal objects. Phd
Thesis. Ljubljana.
MATLAB. 2018a. Computer Vision System ToolboxTM Reference. The MathWorks, Inc.
MATLAB. 2018b. MATLAB® Data Import and Export. The MathWorks, Inc.
MATLAB. 2018c. MATLAB Programming Fundamentals. The MathWorks, Inc.
MATLAB. 2018d. MATLAB® 3-D Visualization. The MathWorks, Inc.
Merkl C., Taras A., Toffolon A., 2017. Experimental and numerical study of the behaviour of eccentrically
loaded cold-formed channel sections-determination of an optimal lead eccentricity. Copenhagen, Eurosteel
2017.
Merkl C. 2018, Analyse des Verformungsverhaltens und der Traglast von exzentrisch druckbeanspruch-
ten Kant-U-Profilen bezüglich lokaler Vorverformungen. 21. DASt-Kolloquium. Kaiserslautern.
Müller A., Hausmann B., Taras A. 2018. Study on the influence of imperfections and strain hardening on
the buckling strength of spiral-welded aluminum circular hollow sections. Eigth International Confer-
ence on Thin-Walled Structures – ICTWS 2018. Lisbon.
Nöldgen, M. 2016. BIM im Brücken- und Ingenieurbau. Springer Fachmedien. Wiesbaden.
Piegl L., Tiller, W. 1997. The NURBS Book. Springer-Verlag. Heidelberg.
Riveiro B., Solla M., et al. 2016. Non-Destructive Techniques for the Evaluation of Structures and Infra-
structure. Taylor & Francis Group. London.
Riveiro B., Cubreiro G., Conde B., Cabaleiro M., Lindenbergh R., Soilán M., Caamano JC. 2018. Auto-
mated Calibration of FEM Models using LiDAR Point Clouds. The International Archives of the
Photogrammetry, Remote Sensing and Spatial Information Sciences, Volume XLII-2. Italy.
Tang P., Huber D., Akinci B., Lipman R., Lytle A. 2010. Automatic reconstruction of as-built building
information models form laser-scanned point clouds: A review of related techniques. Automation in
Construction 19 (7), 829–843.
Toffolon A., Taras A. 2018. Numerical and Experimental Studies for the Development of Direct Strength
Design Rules for Locally and Globally Slender Hollow Sections. Proceedings of the Annual Stability
Conference Structural Stability Research Council. Baltimore.
Torr, P., Zisserman A. “MLESAC: A New Robust Estimator with Application to Estimating Image
Geometry.” Computer Vision and Image Understanding. 2000.
Vosselman G. Maas H. 2010. Airborne and Terrestrial Laser Scanning. Whittles Publishing. CRC Press.
Scotland.
Wald F. et al. 2016. Benchmark cases for advanced design of structural steel connections. Ideastatica. Praha.
Winterstetter T., Schmidt H. 2001. Shell buckling strength of tubes under combined axial, radial and
shear loading. Tubular Structures IX.
Yan Y., Guldur B., Hajjar, J. 2017. Automated Structural Modelling of Bridges form Laser Scanning.
Structures Congress 2017. Denver.
Zhang G., Vela P., Brilakis I. 2014. Automatic generation of as-built geometric civil infrastructure
models from point cloud data. Computing in Civil and Building Engineering. 406–413.
Zhao X., Tootkaboni M., Schafer B. 2017. Laser-based cross-section measurement of cold-formed steel
members: model reconstruction and application. Thin-Walled Structures 120, 70–80.

794
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Static effects of modular structures made of containers

Ondřej Miller, Vít Křivý, David Mikolášek & Přemysl Pařenica


Faculty of Civil Engineering, VSB-TU, Ostrava, Czech Republic

Richard Cuřín
IMECON Containers, a.s.

ABSTRACT: When designing and static-assessing modular buildings, it is necessary to take


into account some specificities and differences related to the static effect of the load-carrying
structure made of interconnected steel containers. The objective of container module manu-
facturers is to achieve a broad application of their products in building construction. The
load-carrying structure of containers used in building construction was created by the modifi-
cation of the original containers used mainly in rail and ship transport. The necessary condi-
tion for the application of the products in building construction is to ensure the mechanical
resistance and stability of the supporting structure in accordance with the requirements of the
applicable standards. This paper therefore points to selected static problems associated with
the use of containers when designing multi-storey modular buildings.

1 INTRODUCTION

Modular container systems were developed for fast and time-saving construction. Modular con-
struction is based on basic assumptions: the prefabrication of components in a production envir-
onment, the mobility of the components for simple transport to the site and the variability of the
components to create the necessary composition of the building. Modules can be combined to
create different assemblies according to the purpose and location of the building. An example of
a multi-storey modular building composed of containers is shown in Figure 1. For more detailed
information about the possibilities of application of modular container structures, see [1], [2].
The vertical load is transferred from the floor or roof beams to the knees of the frames through
the longitudinal girders. The transfer of load effects between adjacent containers takes place only
at the corners, the joint statically acting as an articulated joint. Compression forces in the joint are
transmitted by contact; tension forces and shear forces are transmitted by mechanical coupling.
All the structural elements are welded together. The basic dimensions and structural elements
of the container are shown in Figure 2 (on the right). The corresponding cross-section proper-
ties of the applied elements are shown in Table 1. The subject of this paper is the issue of ensur-
ing sufficient horizontal stiffness of the structure, see chapter 2. The issue of static assessment of
the knees of the rigid frame is only briefly indicated in chapter 3 at the end of the paper.

2 HORIZONTAL STIFFNESS OF THE STRUCTURE

The horizontal stiffness of the structure is primarily ensured by rigid joints between the columns
and the girders of the individual containers. The main load-carrying elements of the container are
relatively subtle. The horizontal stiffness of the open frame. i.e. the frame without a moulded
sheet cladding. is significantly affected by this fact. The stiffness of the adjoining containers is
insignificant. and therefore. an articulated joint is assumed in a static beam model (vertical con-
nection is usually modelled using a short. rigid beam with an articulated joint at the end). With

795
Figure 1. Multi-storey modular building constructed from containers [1]

Figure 2. Load-carrying elements (on the right) and mechanical coupling of containers [1] (on the left)

respect to these facts. it is clear that meeting the condition of the Serviceability limit state (SLS)
related to the maximum permissible value of horizontal object deformation may be the decisive
criterion limiting the use of containers for multi-storey constructions.
With respect to the required layout of the building. it is usually possible to have some walls
with continuous moulded sheet cladding reinforced with a wall grid (the solution employed by
IMECON Containers. a.s.). The cladding of wall is made of trapezoidal sheet metal, which is
anchored by rivets 4.0 mm at a distance of 246 mm to the wall grid. The wall grid is made of
aluminium thin-walled U-profile with a thickness of wall 1 mm. The geometry of the wall grid
is shown in Figure 3. Due to the low stiffness of the container frame in the horizontal direc-
tion. it is therefore possible to use the cladding as an additional stiffening diaphragm.
In order to determine the real horizontal stiffness of the cladding wall of the container.
a programme of experimental tests on a standard container wall system was carried out in
cooperation with the container manufacturer IMECON Containers a.s. (for the comparison.
the stiffness of the separate steel frame of the container was verified). An extensive experimen-
tal programme using specimens of real dimensions took place at the premises of the Ostrava
branch of TZÚS Praha. s.p.
The tested container wall was fastened to a pre-set gripping cage. After the deformation
and strain gauges had been installed. a horizontal load was applied in the upper corner. The
wall was loaded until it was completely damaged: the wall was completely unloaded and then

796
Table 1. Structural elements and the cross-section properties.
Figure Name and type A [m2] Wel.y [m3] Wel.z [m3] Figure Name and type A [m2] Wel.y [m3] Wel.z [m3]

CS1-Q1 1.9061e-03 5.3666e-05 5.1441e-05 CS2-Q2 1.5131e-03 4.2695e-05 4.3389e-05

CS3-Q5 2.2729e-03 8.5231e-05 2.0553e-05 CS4-Q6 2.0011e-03 1.1713e-04 5.8411e-05

CS5-Q8 2.8877e-03 7.2267e-05 1.3597e-04 CS6-U75 5.0066e-04 1.0485e-05 1.8096e-06

CS7-I80 7.5700e-04 1.9500e-05 3.0000e-06 CS8-U80 1.3800e-04 3.3292e-06 5.0168e-07


Figure 3. Geometry of the wall grid (on the left) and aluminium thin-walled U-profile of wall grid (on
the right)

Figure 4. Concepts and basic dimensions of the physical test (on the left) and the damaged wall after
completion of the measurement (on the right)

reloaded several times in the area of the elastic deformations. The concepts and basic dimen-
sions of the physical test and photographs of the selected wall after completion of the meas-
urement are shown in Figure 4.
During the step-loading of the cladded walls. local buckling of the sheet appeared first. Fail-
ure of the self-tapping screws and rivets occurred with increasing deformation. The final
damage was caused by the crack in the weld of the plastic-deformed knee of the rigid frame.
see Figure 5.
The results of the experimental measurement show that the wall formed by profiled sheet
(TR9/123-0.55 in the experiment) contributed significantly to the increase in the horizontal stiff-
ness of the structure. Comparison of the experimentally determined data for the non-cladded
and cladded wall of the container is shown in Figure 6.

Figure 5. Crack in the weld of the knee of the rigid frame (selected examples)

798
Figure 6. Comparison of the experimentally determined data for horizontal stiffness (green – cladded
wall. blue – non-cladded wall)

Experimentally found data can be used to verify numerical models. With the use of avail-
able static software. it is possible to create a shell model of a wall cladded with a profiled sheet
with a reinforcing wall grid. The shell model shown in Figure 7 (above) was created in SCIA
Engineer and the initial stiffness of the numerical and experimental model was equal. It is pos-
sible to determine the contribution of the shear stiffness to the overall horizontal stiffness of
the container wall with sufficient accuracy using numerical models.
The determined value of the additional horizontal stiffness can then be implemented into
the overall beam model of the structure. for example. using fictitious rigid diagonals ending in
axial flexible support. see Figure 7 (in the middle). The stiffness of the flexible support can be
tuned by comparing the deformations of the shell and the beam model. which must be identi-
cal for both models.
The substitution of wall stiffness by using fictitious diagonal elements with axial flexible
support is correct only if the horizontal load applied to the wall of the container does not
exceed the value corresponding to the elastic effect of the wall. As the experimental loading
first resulted in local buckling of the profiled sheet. the horizontal load limit can be derived
from the results of the non-linear stability analysis performed on the above shell model. For
the container wall shown in Figure 7 loaded at the top by a horizontal force F = 5 kN. the
non-linear stability calculation determined the critical factor acr = 3.05 for the first mode of
buckling: the deflection shape is shown in Figure 7. When verifying the suitability of the
beam model. it is necessary to verify that the horizontal force acting on the cladding wall
(the horizontal force can be determined from the reactions) does not exceed the value guar-
anteeing the elastic effect of the cladding wall Fmax = acr · F = 3.05 · 5 = 15.3 kN (in the
static calculations of the various container assemblies made by the authors of this paper.
this criterion has always been met).

3 STATIC ASSESSMENT OF THE KNEES OF THE RIGID FRAME

The results of the static analysis carried out using the spatial beam models of the container
assemblies show that the main supporting elements of the containers are most used statically
in the area of the knees. The elements are designed to almost 100% of their load-carrying

799
Figure 7. Numerical model of the cladded wall of the container (top – numerical model using shell elem-
ents. middle – beam wall model with fictitious rigid diagonals ending with axial flexible support.
bottom – loss of stability caused by horizontal force F = 5 kN for the first mode of buckling: acr = 3.05)

capacity according to the applicable standards. The structural design of the welded knees is so
complicated that it cannot be modelled with sufficient precision using beam models.
Based on the practical experience of the authors of this paper. a conservative recommenda-
tion for carrying out a detailed analysis of the structure in the knee area is provided. if one of
the attached members in the knee is statically designed for more than 80% load-carrying cap-
acity determined in accordance with the applicable standards. The authors of the paper have
good experience in these cases using numerical models combining beam and shell elements –
the whole structure is considered as a beam model except the knee of the rigid frame and the
appropriate length of the attached beams. see Figure 8. Geometrically and Materially Non-
linear Analysis with Imperfections (GMNIA) [2] was used for the calculation of the load
effects (for example using SCIA Engineer).
Imperfections introduced into the shell-beam numerical model included local and global
imperfection of the beam model and local imperfection of the shell model. The local

800
Figure 8. Numerical model combining finite beam and shell elements

Figure 9. Knee of the rigid frame – course of stress (on the left) and total strain (on the right) on the
outer surface of the shell model determined by GMNIA

imperfection of the beam model was obtained from the standard EN 1993-1-1, Tab 5.1. The
global imperfection of the beam model was obtained from the stability analysis. The local
imperfection of the shell (buckling) was obtained from the stability analysis and size of imper-
fection was determined in accordance with EN 1993-1-5, Annex C.
For the Ultimate limit state (ULS) assessment of parts of the structure modelled with shell
elements. the criteria for allowable total and plastic strain given in Standard EN 1993-1-5 can
be used. The first criterion determines the recommended value for the total principal strain on
the wall surface (i.e.. including the momentum strain of the walls) by ε  0:05 ð5%Þ. Another
criterion recommends the limits of the permanent (plastic) component of the strain on the
median plane (membrane deformation) by εm:pl  0:002 ð0:2%Þ.
The material properties of steel were implemented into the structural analysis in accordance
with Annex C of EN 1993-1-5. The accuracy of the numerical models was verified by compar-
ing the outputs of the numerical analysis with the results of the tests. To model material
behaviour in numerical models. it is advisable to use the real stress-strain diagram correspond-
ing to the tensile test results of steel. Therefore. specimens for the tensile test according to EN
ISO 6892-1 were taken from the structure designed for the physical testing. Tensile testing was
carried out in an accredited laboratory of Arcelor Mittal Ostrava. Ten specimens were tested
using the Instron 4210 laboratory press. The results of the tensile test are shown in Figure 10.
The average value of the yield strength of the steel was 270 MPa. The structures of the con-
tainers are made of S235JR steel grade. The difference in the yield strength in relation to the
nominal value of fy = 235 MPa given in EN 1993-1-1 is common for the S235 steel grade.
The comparison between the real behaviour of the non-cladded container wall resulting
from the physical testing and the results of the numerical analysis is shown in Figure 11. The
non-cladded container wall described by curves referred to as physical test 01 and 02. The
yellow curve describes the behaviour of a numerical model of a non-cladded wall with the

801
Figure 10. Stress-strain diagrams for ten steel samples

Figure 11. Comparison of results of physical tests with results of numerical analysis

implementation of the real stress-strain diagram, which showing very good agreement with
the results of experimental tests. Thus, it is possible to state that nonlinear numerical analysis
using finite shell elements can be used to accurately model the real behaviour of the load-
carrying structure of modular buildings composed of containers.
Figure 11 also shows the differences resulting from the use of various stress-strain diagrams
(green curve describes elastic-plastic diagram with strain hardening and blue curve describes
elastic-plastic diagram without strain hardening). As expected, the most conservative results are
achieved using the stress-strain diagram for ideal plastic behaviour (blue curve). The red arrow
shows the state of the structure corresponding to the principal strain on the surface ε = 0.05

802
4 CONCLUSIONS

The main aim of the article was to share practical experience with the static design of
modular buildings from containers. In static assessment of the various multi-storey
modular systems developed by IMECON Containers. a.s., particular attention has to be
paid to the issue of ensuring and verifying sufficient horizontal stiffness of the objects
and to the assessment of the ultimate load condition in the most loaded frame corners
of the container assembly. Standard static programs can be used for design modelling.
A beam model of the whole structure can be used to design the structure. and for more
complex numerical analysis a shell model can be used. Due to the material properties of
steel obtained from experiments conducted on the numerical model. numerical models
approximate their static behaviour to the real behaviour of the structure. Container
modular assemblies developed by IMECON Containers. a.s., verified using detailed
numerical and experimental modelling. meet the relevant requirements for the reliable
operation of the load-bearing steel structure.

ACKNOWLEDGEMENTS

The paper has been produced in cooperation with representatives of the Institute of Steel Struc-
tures. s.r.o.. researchers from the Faculty of Civil Engineering of VŠB-TU Ostrava and the manu-
facturer of IMECON Containers. a.s. The authors would like to thank the management of
IMECON Containers. a.s. for allowing the necessary experimental tests and the Ministry of Edu-
cation. Youth and Sports for financial support in the framework of the Student Research Grant
Competition of the Technical University of Ostrava under identification number SP2019/140.

REFERENCES

[1] https://www.imecon.cz/
[2] Křivý. V., Kubzová. M. Static analysis and numerical modeling of steel cylindrical tanks weakened by
corrosion. Modelování v mechanice. 2017. p. 31–32. ISBN 978-80-248-4010-9.
[3] KALKIN A. Quik Build: Adam Kalkin´s ABC of container architecture. London. 2008.
[4] ČSN EN 1993-1-5 Eurocode 3: Design of steel structures – Part 1-5: Plated structural elements.
Prague. ČNI 2006.
[5] ČSN EN 1993-1-1Eurokód 3: Design of steel structures – Part 1-1: General rules and rules for build-
ings. Prague. ČNI 2007.
[6] ČSN EN 1993-1-8Eurokód 3: Design of steel structures: Design of joints. Praha. ČNI. 2006.
[7] Virdi K.S. et al. Numerical simulation of semi-rigid connections by the finite element method. Report
of Working Group 6. Numerical Simulation COST C1. Brussels Luxembourg. 1999.
[8] Šabatka L. Wald F. Kabeláč J. Kolaja D. Vnitřní síly působící na styčník a globální analýza ocelo-
vých konstrukcí. Konstrukce. 3/2016. 54–57.
[9] Wald. F. et al. Benchmark cases for advanced design of structural steel connections. Praha: Česká tech-
nika – nakladatelství ČVUT. 2016. ISBN 978-80-01-05826-8.
[10] Da Silva Simoes L. Towards a consistent design approach for steel joints under generalized loading.
Journal of Constructional Steel Research. 64. 1059–1075. 2008.

803
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Calibration of European web-crippling equations for cross-sections


with one web

T. Misiek
Breinlinger Ingenieure, Tuttlingen, Germany

A. Belica
Astron Buildings S.A., Diekirch, Luxembourg

ABSTRACT: Current design equations given in EN 1993-1-3:2006 for calculating the web-
crippling resistance of cross-sections with one web were copied from AISI’s Specification for
the design of cold-formed steel structural members, taking into account the different safety
concepts: The load and resistance factor design (LRFD), with resistance factors ϕw applied to
nominal values, and the limit state design with partial factors γM applied on characteristic
values as defined in EN 1990. Furthermore, the web-crippling equations of subsequent edi-
tions of the AISI Specifications were completely revised based on the web-crippling data col-
lected and evaluated by Beshara & Schuster (2000/2006).
This paper presents the results of a calibration of a generalized web-crippling equation to
be used for cross-sections with one web (i.e. C- and Z-sections), for built-up I-sections made
of two back-to-back connected C-sections, for nested Z-sections and for C- and Z-sections fas-
tened to cleats with no gap between the lower flange and the supporting structure.
The constant factors of the generalized web-crippling equation were calibrated to comply
with the safety concept described in EN 1990:2002. That way the paper gives an introduction
as well as background information on proposed changes and amendments in EN 1993-1-3.

1 INTRODUCTION

1.1 Initial situation


As part of the ongoing revision of EN 1993-1-3, the regulations on the load-bearing capacity
of profiles (purlins, wall rails, roof and ceiling battens, etc.) under local transverse forces were
also subjected to critical review. The current rules are based on the 1996 edition of AISI’s Spe-
cification for the design of cold-formed steel structural members. In the following editions, how-
ever, these were fundamentally revised. The question therefore arose as to the extent to which
the regulations in EN 1993-1-3 should also be adapted accordingly.
Due to the different safety concept, a direct adoption is not possible. For this reason,
a separate approach was developed using a database of about 1350 single test results.

1.2 Sections
Subject of present investigations were
– C-sections, both with stiffened and unstiffened flanges (Figure 1)
– Built-up I-sections, made of two C-sections (Figure 2)
– Z-sections with stiffened flanges (Figure 3)
– nested Z-sections at intermediate supports (Figure 4)

804
Figure 1. Stiffened and unstiffened C-sections.

Figure 2. Built-up I-sections.

Figure 3. Stiffened Z-sections.

805
Figure 4. Nested -sections.

Figure 5. End one-flange loading (EOF) with c ≤ 1.5⋅h.

Figure 6. Internal one-flange loading (IOF) with c > 1.5⋅h.

Figure 7. End two-flange loading (ETF) with c ≤ 1.5⋅h and e ≤ 1.5⋅h.

Figure 8. Internal two flange loading (ITF) with c > 1.5⋅h and e ≤ 1.5⋅h.

806
Nested Z-sections as shown in Figure 4, may have equal or unequal widths of flanges. This
application can predominantly be found at intermediate supports of purlins.

1.3 Loading conditions


Resistance to local forces depends on the distance to a beam edge, but also on whether an
opposing load acts on the opposite side of the section or not. This leads to four configur-
ations, which differ with respect to the load position.
The American codes use the designations
– end one-flange loading (EOF),
– interior one-flange loading (IOF),
– end two-flange loading (ETF), and
– interior two-flange loading (ITF)
which are not common in Europe. Nevertheless, they are used in the present text because they
allow an easy capture of the loading conditions.

1.4 Fastening conditions


In most applications, sections are fastened to the supporting structure, see Figure 9. However,
at points of load introduction, resistance for the unfastened case is relevant. Therefore both
resistances are required.
Linear sections may also be connected to the supporting structure by cleats; the section is
connected to the cleat by bolts through the web (Figure 10). Usually, there is a gap between the
lower flange and the supporting structure, and the support reaction is transferred via the bolts.
If the gap is sufficiently large, web-crippling resistance does not have to be checked. In the
case of large or oversized bolt holes, providing considerable clearance for the bolts, the lower

Figure 9. Fastening to supports.

Figure 10. Fastening by cleats.

807
flange may settle and the support reaction is then transferred via contact. However, with
a cleat of sufficient stiffness, web rotation is prevented and the deformation behavior corres-
ponds to that observed for the built-up I-sections. Although a built-up I-section consists of
two C-sections, this applies also for single and nested Z-sections because with cleats the effects
of section geometry and the resulting differences in deformation behavior become negligible.

2 TEST DATA BASE AND ITS EVALUATION

For the revision, no continuous theoretical-mechanical model was used to describe the resist-
ance to transverse forces. Instead of that, a suitable trial function was chosen, whose coeffi-
cients were calibrated using test results. The procedure is therefore similar to the procedure
chosen by Beshara &Schuster (2000/2006) for the development of the regulations in the North
American specification for the design of cold-formed steel structural members (2001).

Table 1. Stiffened C-sections – parameter range.


t fy r/t ss/t h/t
C-sections, stiffened [mm] [MPa] – – –

1a EOF fastened 0.7–2.8 220–512 1.4–9.3 19–129 73–236


1b unfastened 0.8–2.8 234–641 1.2–5.0 10–141 40–261
2a IOF fastened 0.6–3.1 160–516 1.6–7.5 32–167 82–184
2b unfastened 0.6–3.1 160–581 1.5–7.5 10–167 66–257
3a ETF fastened 0.7–3.1 220–534 1.6–13 20–129 81–207
3b unfastened 1.0–3.0 302–581 1.6–4.0 10–104 65–263
4a ITF fastened 0.6–3.1 260–534 65–263 20–167 81–208
4b unfastened 0.6–3.1 260–581 1.6–5.0 10–167 66–261

Table 2. Unstiffened C-sections – parameter range.


t fy r/t ss/t h/t
C-sections, stiffened [mm] [MPa] – – –

1a EOF fastened 1.5–6.0 457–534 1.0–3.0 17–66 38–132


1b unfastened 0.9–6.0 250–450 0.8–4.1 5.2–140 19–197
2a IOF fastened 0.6–6.0 449–534 0.9–3.0 8.4–66 25–130
2b unfastened 0.9–6.0 250–450 0.8–4.3 5.1–62 19–196
3a ETF fastened* – – – – –
3b unfastened 1.0–6.1 250–550 0.5–2.1 5.2–62 19–197
4a ITF fastened* – – – – –
4b unfastened 1.1–6.1 250–550 0.5–1.9 5.2–62 19–198

*no data available

Table 3. Built-up I-sections – parameter range.


t fy r/t ss/t h/t
Built-up I-sections [mm] [MPa] – – –

1 EOF fastened and unfastened 1.2–3.8 208–432 1.0–5.0 6.8–123 29–257


2 IOF fastened and unfastened 1.2–3.1 208–436 1.0–5.0 8.1–83 65–254
3 ETF fastened and unfastened 1.2–3.8 208–432 1.0–2.7 6.8–65 29–268
4 ITF fastened and unfastened 1.2–3.8 208–325 1.0–2.7 6.8–65 29–262

808
Table 4. Stiffened Z-sections – parameter range.
t fy r/t ss/t h/t
Z-sections, stiffened [mm] [MPa] – – –

1a EOF fastened 1.4–2.6 392–508 3.0–4.8 26–45 77–162


1b unfastened 1.4–2.6 392–508 3.0–4.8 26–45 77–162
2a IOF fastened* – – – – –
2b unfastened* – – – – –
3a ETF fastened 1.1–1.5 323–446 4.8–13 20–55 82–208
3b unfastened** – – – – –
4a ITF fastened 0.9–1.5 323–446 1.6–13 20–107 82–209
4b unfastened** – – – – –

* no data available
** no data available, may not be used unfastened (rotates)

Table 5. Nested Z-sections – parameter range.


t fy r/t ss/t h/t
Nested Z-sections [mm] [MPa] – – –

1 IOF fastened and unfastened 1.4–3.0 367–482 1.7–6.3 42–118 58–150

The collection of test results compiled by Beshara & Schuster (2000/2006) was used for cali-
bration. However, this collection was supplemented by many other test results which have
been published since then, allowing closing gaps in important applications. The following
tables show the parameter range covered by the test results.
The calibration is based on procedures given in EN 1990, Annex D, for the statistical determin-
ation of resistance models combined with a least square approach. The reliability level according
to EN 1990 and a desired partial factor γM1 were prescribed in advance: Therefore the partial
safety factor has not been determined for a given theoretical resistance rt, but the coefficients of
the resistance model were adapted in such a way that in combination with the desired partial
factor γM1 = 1.1 the required reliability level (reliability class RC2 for 50 years and β = 3.8) is
reached. For local transverse forces, γM1 applies because there is no post-critical resistance (see
also provisions in EN 1993-1-5 for patch loading). The numerical value 1.1 was chosen because it
will most likely be the recommended value in the new editions of both EN 1993-1-1 and EN 1993-
1-3. A detailed description of the evaluation procedure can be found in Misiek & Belica (2018).

3 DESIGN APPROACH

3.1 Trial function


The empirical approach for calculating the web-crippling resistance of linear sections with
a single web is based on a generalized function with √(E⋅fyb) and h to cover slenderness effects.
pffiffiffiffiffiffiffiffiffiffiffiffiffi  rffiffi  rffiffiffiffi rffiffiffi!
t2  E  fyb r ss h
Rw;Rd ¼ K   1  Kr   1 þ Ks   1  Kh  ð1Þ
γM1 t t t

where t = design core sheet thickness; E = Young’s modulus; fyb = basic yield strength;
h = total depth; r = internal radius of corners; and ss = nominal length of stiff bearing.
Coefficients were designated with letter Ki to ease differentiation from letter Ci used in AISI
provisions.

809
Figure 11. Nominal length of the stiff bearing ss for fastening via cleats.

With a cleat of sufficient stiffness the coefficients Ki given for built-up I-sections may be used
for both C-sections and Z-sections with (additionally) fastened or unfastened flanges, provided
that the nominal length of the stiff bearing ss is taken as the minimum value of the length of the
stiff bearing and the width of the cleat:
 
ss ¼ min bf ; bcleat  200 mm ð2Þ

where bf = length of the stiff bearing; and bcleat = width of the cleat, see Figure 11.

3.2 Coefficients Ki
Coefficients are determined for
– C-sections, both with stiffened and unstiffened flanges

Table 6. Stiffened C-sections – coefficients Ki.


C-sections, stiffened K Kr Ks Kh

1a EOF fastened 0.266 0.165 0.155 0.032


1b unfastened 0.251 0.211 0.148 0.039
2a IOF fastened 0.627 0.151 0.098 0.036
2b unfastened 0.594 0.143 0.049 0.033
3a ETF fastened 0.200 0.109 0.142 0.046
3b unfastened 0.291 0.383 0.095 0.041
4a ITF fastened 0.558 0.102 0.053 0.028
4b unfastened 1.202 0.232 0.000 0.051

Table 7. Unstiffened C-sections – coefficients Ki.


C-sections, stiffened K Kr Ks Kh

1a EOF fastened 0.157 0.074 0.231 0.024


1b unfastened 0.085 0.188 0.640 0.044
2a IOF fastened 0.193 0.045 0.219 0.000
2b unfastened 0.222 0.002 0.120 0.000
3a ETF fastened see unfastened
3b unfastened 0.075 0.092 0.278 0.026
4a ITF fastened see unfastened
4b unfastened 0.263 0.076 0.126 0.037

810
Table 8. Built-up I-sections – coefficients Ki.
Built-up I-sections K Kr Ks Kh

1 IOF fastened or unfastened 0.580 0.163 0.0660 0


2 EOF fastened or unfastened 0.179 0 0.225 0
3 ITF fastened or unfastened 0.768 0.179 0.0699 0.0335
4 ETF fastened or unfastened 0.439 0.292 0.0528 0.0344

Table 9. Stiffened Z-sections – coefficients Ki.


Z-sections, stiffened K Kr Ks Kh

1a EOF fastened 0.162 0.094 0.239 0.029


1b unfastened 0.120 0.000 0.024 0.008
2a IOF fastened 0.324* 0.094 0.239 0.029
2b unfastened 0.240* 0.000 0.024 0.008
3a ETF fastened 0.308 0.000 0.075 0.049
3b unfastened** – – – –
4a ITF fastened 0.606 0.061 0.082 0.035
4b unfastened** – – – –

* 2-times EOF
** may not be used unfastened (rotates)

Table 10. Nested Z-sections – coefficients Ki.


Nested Z-sections K Kr Ks Kh

1 IOF fastened or unfastened 0.235 0.200 0.187 0

Table 11. Stiffened C-sections – application range.


t fyb
C-sections, stiffened [mm] [MPa] r/t ss/t h/t

fastened 0.6–3.1 160–600 ≤ 10 ≤ 170 ≤ 270


unfastened 0.6–3.1 160–600 ≤5 ≤ 170 ≤ 270

Table 12. Unstiffened C-sections – application range.


t fyb
C-sections, unstiffened [mm] [MPa] r/t ss/t h/t

fastened 0,6–6,0 250–600 ≤4 ≤ 100 ≤ 200


unfastened 1.0–6.0 250–600 ≤4 ≤ 100 ≤ 200

Table 13. Built-up I-sections – application range.


t fyb
Built-up I-sections [mm] [MPa] r/t ss/t h/t

fastened or unfastened 1.0–4.0 200–450 ≤ 5,0 ≤ 100 ≤ 270

811
Table 14. Stiffened Z-sections –application range.
t fyb
Z-sections, stifened [mm] [MPa] r/t ss/t h/t

fastened 1.0–3.0 320–500 ≤ 10 ≤ 100 ≤ 200


unfastened 1.5–3.0 320–500 ≤5 ≤ 100 ≤ 200

Table 15. Nested Z-sections – application range.


t fyb
Nested Z-sections [mm] [MPa] r/t ss/t h/t

fastened or unfastened 1.0–3.0 320–500 ≤ 10,0 ≤ 100 ≤ 200

– Built-up I-sections, made of two C-sections


– Z-sections with stiffened flanges
– Nested Z-sections at intermediate supports
for different loading conditions and fastening conditions as described above.

3.3 Application range


Since the coefficients Ki of the trial function were derived from tests and not from a closed
mechanical model, the application range of the design rules is limited to the parameter range
covered by the tests. The application range differs for the individual configurations, but
importance has been attached to ensure consistency in each case: The application range of the
design rules for simple Z-sections corresponds to that for nested Z-sections.
A general limitation ss ≤ 200 mm for the nominal length of stiff bearing applies. For larger
lengths of the stiff bearing, the approach is applicable, but ss should be reduced to 200 mm.

4 SUMMARY

Design provisions for web-crippling will be subject to changes in the next revision of EN
1993-1-3 as some deficits in target safety level had become apparent. An approach is sought
which is both simpler as well as more realistic with regards to failure loads. The design provi-
sions presented here are a proposal developed by two members of the Working Group CEN
TC250/SC3/WG3 responsible for the revision work.

REFERENCES

AISI/COS/NASPEC 2001: North American specification for the design of cold-formed steel structural
members. Washington: AISI
Beshara, B. & Schuster, R. M. 2000/2006. Web crippling data and calibrations of cold formed steel mem-
bers. Research Report RP00-2. Washington: AISI.
EN 1993-1-3:2006 + AC:2009: Eurocode 3: Design of steel structures – Part 1-3: General rules – Supple-
mentary rules for cold-formed members and sheeting. Brussels: CEN.
EN 1990:2002 + A1:2005 + A1:2005/AC:2010: Eurocode: Basis of structural design. Brussels: CEN.
Misiek, Th. & Belica, A. 2018. European web-crippling equations – Fundamentals of reliability analysis
and equations for built-up I-sections and nested Z-sections. In J. Lange, F. Rädel (eds.), Festschrift
Jörg Lange. Darmstadt: Institut für Stahlbau & Werkstoffmechanik.
Misiek, Th. & Belica, A. 2019. Calibration of European web-crippling equations for cold-formed C- and
Z-sections. Steel Construction – Design and Research 13(1): 31–43.
Specification for the design of cold-formed steel structural members. 1996. Washington: AISI.

812
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Explanatory notes to buckling design of longitudinally welded


aluminium compression members

T. Misiek
Breinlinger Ingenieure, Tuttlingen, Germany

B. Norlin & T. Höglund


Royal Institute of Technology, Stockholm, Sweden

ABSTRACT: As part of the ongoing revision of the Eurocodes, the regulations in EN 1999-
1-1 on the buckling of aluminium compression members were also subjected to a critical
review. This resulted in the need to revise the regulations for non-welded compression members
and for longitudinally welded compression members. For the design provisions of longitudin-
ally welded aluminium compression members, a more extensive revision was necessary. The
design concept was completely revised. In the course of the revision, some quite complex
design models were discussed with which the cross-section types, residual stress distributions as
well as size and position of the heat-affected zone (HAZ) were to be taken into account. In
order not to complicate the application of the standard too much (ease of use), a clearly simpli-
fied procedure was finally chosen. The following contribution will justify these simplifications

1 INTRODUCTION

As part of the revision of EN 1999-1-1, a number of investigations were carried out on longi-
tudinally welded aluminium columns, which served as the basis for supplementing and revis-
ing the normative regulations.
The buckling resistance of aluminium columns depends merely on initial bow (geometrical
imperfection), residual stresses (structural imperfection) and stress-strain relationship. Fur-
thermore, size and position of the HAZ due to heat introduction during welding affect buck-
ling resistance of longitudinally welded columns. The realistic determination of the buckling
resistance is therefore complex. In order to maintain the normative regulations in an appropri-
ate manner, simplifications had to be made. These are explained and justified below. The
focus here is on the discussion of the influence of the cross-section type and thus its specific
residual stress distribution and on local softening in the HAZ.
The following explanations are based on numerical investigations in which geometrically
and materially nonlinear analyses with imperfections were carried out on longitudinally
welded aluminium columns. In these investigations, the geometry of the cross-sections, the
axis around which the column buckles, and the stress-strain relationship were varied. The
latter are based on fictitious materials that can, however, be assigned to real materials, see
Misiek et al. (2019).

2 INFLUENCE OF CROSS-SECTION TYPE AND RESIDUAL STRESSES

2.1 Introduction
For steel members the residual stresses are the most important factor and, as the magnitude
and distribution of the residual stresses depends on the cross-section type, there are different

813
Figure 1. Section types: Residuals stresses in a section of type a (left) and superposition of residual
stresses and stresses from loading in a section of type b (right).

buckling curves for different cross-section types in EN 1993-1-1. The residual stresses for
extruded aluminium profiles are very small. However, the stress-strain relationship is strongly
non-linear from the origin, which is why in EN 1999-1-1 the choice of buckling curve for
extruded profiles depends on the alloy and temper expressed by the buckling class (BC). For
longitudinal welded aluminium members, however, there are residual stresses where the distri-
bution is similar as in welded steel members although compared to the yield strength the mag-
nitude in aluminium is smaller than in steel. Nevertheless the influence on the buckling
resistance is similar as for steel so the resistance should depend on the so-called cross section
types a and b.
In a section of type b, the longitudinal weld leads to residual compressive stresses in the
most stressed (usually the outermost) cross-sectional areas, so that the buckling resistance is
reduced. Figure 1 shows this using the example of an I-section made of three plates. When the
stresses in the outer tips of the flanges reach the yield strength, they lose their stiffness. The
cross-section’s moment of inertia around the y-axis is linearly dependent on the length bel of
the remaining area, but for the moment of inertia around the z-axis the length bel enters in the
third power.

Iel;y ¼ f ðbel Þ ð1Þ

and
 
Iel;z ¼ f b3el ð2Þ

The influence is larger for z-axis buckling than for y-axis buckling.
In a section of type a welds are smaller and are located in the web so that the compressive
force resulting from the residual stresses balancing the tension force in the area around the
weld extends over the entire flanges and large parts of the web. Figure 1 shows this for
a cross-section composed of two T-sections and a web plate. The compressive residual stresses,
especially at the flange tips, are smaller than those of section type b.

2.2 Numerical investigations


Numerical investigations were carried out on four different cross-sections, see Figure 2, with
materials I to XV, see Misiek et al. (2019). Residual stresses and their distribution were mod-
elled based on the measurement results published by Benson (1992) and Mazzolani (1994) and
normalized to 0.2% proof strength, see Figure 3.

814
Figure 2. Cross-sections under investigation.

Figure 3. Residual stress distribution.

815
Each configuration was also examined without residual stresses, so that its effects can be
evaluated. In total, 720 longitudinally welded columns were investigated.

2.3 Evaluation and conclusion


Figures 4–6 show the effect of the residual stresses on the buckling resistance, expressed by the
ratio of the reduction factors χRS (with residual stresses) and χ (without residual stresses) both
obtained from the numerical investigations.
Overall, there is no general increase or reduction in buckling resistance in buckling classes
A and B due to residual stresses. Further differentiation shows that some cross-sections bene-
fit from the residual stresses, while some suffer from them. The effect is most pronounced at λ
= 1.0. However, the deviation of ±10% is negligible. But it also turns out that the correlations
explained in the introduction are of a more theoretical nature. The effect is reduced for real
aluminium cross-sections, as can already be seen from the cross-sections used for the numer-
ical investigations where w stands for weak axis buckling and s for strong axis buckling:
– Pw: compressive residual stresses in the flange reduce the buckling resistance.
– Tw: compressive residual stresses at the outer flange edges reduce the buckling resistance.
– Cw: Tensile and compressive residual stresses in the flange balance each other out, i.e.
residual stresses have comparatively low effects.
– Ps: Compressive residual stresses in the flange are more than compensated for by the tensile
residual stresses in the area near the flange of the web, thus slightly increasing the buckling
resistance due to the residual stresses.
– Ts: Tensile residual stresses dominate the flange; residual stresses thus increase the buckling
resistance.
Contrary to what was expected, a negative effect of the residual stresses for z-axis buckling is
also apparent in the cross-section P.
In the case of buckling class C, on the other hand, a general reduction in the buckling resist-
ance as a result of the residual stresses is apparent. The deviation from the reference value is
up to -20% (the width of the scatter band thus remains the same) which covers the entire
range of medium to large slenderness.

Figure 4. Effect of residual stresses for materials of buckling class A.

816
Figure 5. Effect of residual stresses for materials of buckling class B.

Figure 6. Effect of residual stresses for materials of buckling class C.

At the end, however, the transition between the cross-section types is fluent. For this
reason, prEN 1999-1-1:2018 did not introduce any cross section types. This is also because it
is to be regarded as critical if an increasing buckling resistance due to the residual stresses is
included in the design method.
Figure 7 shows the approach of a uniform buckling curve without using any cross-section
types for buckling class C, which is most strongly affected by residual stresses, but where the
size and position of the HAZ also have the least influence. On the basis of the results shown in

817
Figure 7. Buckling curve cw (α = 1.005 and λ0 = 0.234) and FE-results for cross-section types a (filled
triangle) and cross-section type b (unfilled triangle).

Figure 6, where the cross-section Cw is hardly negatively influenced by the residual stresses,
cross-section Cw had been filed under cross-section type a.

3 INFLUENCE OF SIZE AND POSITION OF HAZ

3.1 Introduction
The softening in the heat-affected zone results in a composite cross-section in which not only dif-
ferent 0.2% proof strength but also different stress-strain relationships (different exponents n in
the Ramberg-Osgood-law) apply to the individual cross-section parts. The question arises as to
how this can be taken into account when determining the buckling resistance using the equivalent
member method, i.e. without having to resort to a much more complex second order analysis.

3.2 Numerical investigations


The influence of the location and position of the HAZ was first investigated numerically.
These investigations were carried out on I-cross-sections, see Figure 8, and limited to materials
I and III. No residual stresses have been modelled, allowing focusing on the effect under inves-
tigation. The investigations are limited to λ ≈ 0.4 and λ ≈ 1.0, but cover buckling around both
axes. In total, 96 longitudinally welded columns were investigated.

3.3 Evaluation and design approach


The evaluation showed that it is possible to determine the bearing capacity using the equiva-
lent member method. The reduction factor for buckling resistance is calculated using

1
χ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi  1; 0 ð3Þ
2
 þ 2  λ

818
Figure 8. Cross-section with varying size and position of HAZ.

with
 
A1 W   2
 ¼ 0:5  1 þ   α  λ  λ0 þ λ ð4Þ
W1 A

where
 
A1 ¼ A  AHAZ  1  ρo;HAZ ð5Þ
 
W1 ¼ W  WHAZ  1  ρo;HAZ ð6Þ

and AHAZ = area of the HAZ cross-section, WHAZ = section modulus of the HAZ cross-section,
and ρo,HAZ = reduction factor (ratio between 0.2% proof strength in HAZ and in parent material).
Imperfection factor and plateau length are calculated as follows:

WHAZ  
α ¼ αi þ  αi;HAZ  αi ð7Þ
W

and

AHAZ  
λ0 ¼ λ0;i þ  λ0;i;HAZ  λ0;i ð8Þ
A

with i = a, b, c for the different buckling classes and buckling curves and associated imperfec-
tion factor and plateau lengths αi, αi,HAZ, λ0,i and λ0,i,HAZ. Figure 9 shows a comparison of
buckling resistance Nb,R,calc calculated using this approach and the results Nb,R,FE from the
numerical investigations.

819
Figure 9. Comparison of buckling resistance Nb,R,calc and Nb,R,FE.

Further investigations suggest that for buckling class C, αc ≈ αc,HAZ and λ0,c ≈ λ0,c,HAZ may be
applied because material properties for both base material and HAZ are similar and the buckling
curve c determined in Misiek et al. (2019) seems to be capable (i.e. is so conservative) to also
cover the effect of residual stresses. As a further simplification, which allows a reduction of the
number of parameters, such that αa,HAZ = αb,HAZ = αc,HAZ and λ0,a,HAZ = λ0,b,HAZ = λ0,c,HAZ can
be used. The investigations in this regard are presented elsewhere.
A simplified design approach using buckling curves can be used if IHAZ/I ≈ AHAZ/A applies to
the section values and large concentrations of welds at the outermost fibres of the cross-section
are avoided. This approach is followed in both EN 1999-1-1:2007 and prEN 1999-1-1:2018
because of ease of use. Figure 7 shown in the previous Section 2.3 may serve as an example.

4 SUMMARY

In the case of alloys and tempers of buckling classes A and B, the effect of the reduction in
strength in the HAZ dominates over the residual stresses. In the case of alloys and tempers of
buckling class C, an effect of the cross-section type and the residual stresses is apparent. Due to
the smooth transition between the theoretical cross-section types a and b, this can be taken into
account by a correspondingly conservative definition of the buckling curve.

REFERENCES

EN 1999-1-1:2007: Eurocode 9: Design of aluminium structures – part 1-1: General structural rules. Brus-
sels: CEN.
prEN 1999-1-1:2018, Final draft: Eurocode 9: Design of aluminium structures – part 1-1: General struc-
tural rules. Brussels: CEN.
Misiek, Th., Norlin, B. & Höglund T. 2019. A look at European buckling curves for aluminium
members. Steel construction - design and research 12 (2): (in press).
Benson, P.G. 1992. Shear buckling and overall web buckling of welded aluminium girders. Stockholm:
Royal Institute of Technology, Division of Steel structures.
Mazzolani, F. 1994. Aluminium Alloy Structures. Boca Raton: CRC Press.

820
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Buckling of circular hollow section stainless steel columns in fire

A. Mohammed
Department of Civil and Environmental Engineering, Brunel University London, London, UK

S. Afshan
University of Southampton, UK

ABSTRACT: Stainless steel tubular sections are being used in engineering applications due
to their unique combination of strength, ductility, durability and aesthetic appeal. In addition,
stainless steel offers high retention of strength and stiffness at elevated temperatures compared
to its counterpart carbon steel. From the limited existing test data on stainless steel circular
hollow sections (CHS) columns, it has been observed that the current Eurocode 3 provisions
can be unconservative in their capacity predictions, which use the design formulations for
carbon steel members, despite stainless steel having different material behaviour. A numerical
modelling study has been undertaken using the finite element (FE) package ABAQUS to
study the stability of cold-formed stainless steel CHS columns at elevated temperatures. The
FE models were first validated against experimental data from the literature and subsequently
used to perform parametric studies at member level. The results have been compared with
existing design provisions provided in EN 1993-1-2 (2005) and the Design Manual for Struc-
tural Stainless Steel (2017). A new flexural buckling curve for cold-formed CHS stainless steel
columns in fire is proposed on the basis of the FE results. Statistical analyses set out by
Kruppa (1999) were carried out to check the accuracy of the new proposal.

1 INTRODUCTION

The use of stainless steel in structural engineering applications is growing due to the material
corrosion resistance, attractive appearance, low lifecycle cost and fire resistance, along with
enhanced product availability and improved design guidance. EN 1993-1-4 (2006) provides
supplementary design rules for stainless steel structures and refers to EN 1993-1-2 (2005) for
the fire design, where the same guidelines as those for carbon steel are also adopted for stain-
less steel. This paper focuses on the behaviour and design of stainless steel circular hollow sec-
tion (CHS) columns in fire based on numerical simulation study. Firstly, numerical models
were developed to replicate existing test data. Secondly, upon validation of the numerical
models, parametric studies were performed to expand the available results over a wide range
of member slenderness and elevated temperatures. The numerical results were compared to
the existing codified fire design rules. Amendments to the current codified design procedures
are proposed, utilising the numerical results, and reliability analysis is carried out to check the
suitability.

2 DEVELOPMENT AND VALIDATION OF FINITE ELEMENT MODELS

2.1 Introduction
Finite element models of cold-formed stainless steel CHS columns were developed and valid-
ated against the results of flexural buckling tests at room temperature by Zhao et al. (2016)
and Buchanan et al. (2018). The validated models were employed to carry out a numerical

821
parametric study of cold-formed CHS columns at elevated temperatures under isothermal
loading conditions, where the columns were subjected to a constant temperature, while the
applied load was increased until failure. This was achieved by employing the reduced stress-
strain responses of the material corresponding to the temperature of the modelled columns.
The details of the modelled tests is reported in Table 1, which includes the nominal cross-
section size, member length (L), stainless steel grades and failure load of the columns (Nu). All
tests were pinned at both ends.

2.2 General modelling assumptions


The measured geometric dimensions of the tested columns were utilised in the FE models to
replicate the corresponding test specimen. Shell elements with reduced integration and finite
membrane strain, S4R (ABAQUS, 2016) were used. The mesh size along the longitudinal and
circumferential directions of the CHS columns was equal to the cross-section thickness; which
was found to give the best combination of accuracy and computational efficiency. The meas-
ured engineering stress-strain curves were represented by the two stage Ramberg-Osgood
material model (Gardner & Ashraf, 2006), which were converted to true stress σtrue and loga-
rithmic plastic strain εpl
ln by means of Equations (1) and (2), where σnom and εnom are the engin-
eering stress and strain, respectively and E is the Young’s modulus. Residual stresses were not
incorporated into the numerical models, due to the negligible influence of membrane residual
stresses and the inherent presence of bending residual stresses in the measured stress-strain
responses employed (Cruise & Gardner, 2008).

σtrue ¼ σnom ð1 þ εnom Þ ð1Þ


σtrue
εpl
ln ¼ lnð1 þ εnom Þ  ð2Þ
E

Initial global and local geometric imperfections were incorporated into the numerical
models in the form of the lowest global and local buckling mode shapes, which were deter-
mined by conducting an elastic eigenvalue buckling analysis. Two global imperfection amp-
litudes, the measured value (ωg) and a fraction of the column length L/1000, where L is the
length of the column, and two local imperfection amplitudes, t/10 and t/100, where t is the
section thickness, were considered (Zhao et al., 2016 and Buchanan et al., 2018). Owing to
the symmetry in the geometry and the boundary conditions of the models, only half of the
section, and half of the member length was modelled. A static Riks analysis procedure was
used to trace the load-deformation response of each of the modelled columns and to deter-
mine their failure load.

Table 1. Summary of literature column tests used for validation of FE models.


Cross-section Reference L (mm) Grade Nu (kN)

CHS 60.5 × 2.8 Zhao et al. (2016) 1450 EN 1.4301 90.5


CHS 76.3 × 3 1450 EN 1.4301 146.0
CHS 106 × 3 550 EN 1.4432 267.0
CHS 106 × 3 1150 EN 1.4432 248.8
CHS 106 × 3 3080 EN 1.4432 150.8
CHS 88.9 × 2.6 400 EN 1.4462 425.2
CHS 88.9 × 2.6 Buchanan et al. (2018) 1650 EN 1.4462 243.4
CHS 88.9 × 2.6 3080 EN 1.4462 100.5
CHS 80 × 1.5 700 EN 1.4512 111.1
CHS 80 × 1.5 900 EN 1.4512 105.8
CHS 80 × 1.5 1600 EN 1.4512 77.9

822
Table 2. Comparison of tests and FE results for varying imperfection levels.
ωg + t/100 L/1000 + t/100 ωg + t/10 L/1000 + t/10

Nu,FE= δu,FE= Nu,FE= δu,FE= Nu,FE= δu,FE= Nu,FE= δu,FE=


Cross-section Nu,test δu,test Nu,test δu,test Nu,test δu,test Nu,test δu,test

CHS 60.5 × 2.8 0.99 1.31 0.93 1.65 0.99 1.57 0.93 1.68
CHS 76.3 × 3 0.99 1.28 0.94 1.66 0.99 1.31 0.94 1.70
CHS 106 × 3 1.07 0.36 1.07 0.36 1.07 0.37 1.07 0.37
CHS 106 × 3 0.88 0.60 0.92 1.39 0.91 1.49 0.92 1.47
CHS 106 × 3 0.96 0.65 0.96 0.64 0.96 0.64 0.96 0.61
CHS 88.9 × 2.6 0.98 0.74 0.99 0.69 0.98 0.74 0.99 0.69
CHS 88.9 × 2.6 1.03 0.73 1.02 0.46 1.03 0.74 1.01 0.72
CHS 88.9 × 2.6 1.06 1.33 1.06 1.31 1.06 1.33 1.06 1.31
CHS 80 × 1.5 1.05 0.65 1.03 0.64 1.05 0.53 1.03 0.66
CHS 80 × 1.5 1.00 0.60 1.02 0.35 1.03 0.68 1.03 0.79
CHS 80 × 1.5 1.10 1.09 1.12 0.65 1.12 0.65 1.12 0.69
Mean 1.01 0.85 1.01 0.89 1.02 0.91 1.00 0.97
COV 0.06 0.40 0.06 0.57 0.06 0.46 0.06 0.49

2.3 Validation of numerical models


The modelling procedures described were applied to replicate the test responses of stainless
steel columns reported in Table 1. Table 2 presents the ratios of the numerical to test ultimate
loads (Nu,FE=Nu,test) and mid-height deflection at ultimate load (δu,FE=δu,test) for the varying
combinations of global and local imperfection amplitudes employed. The result indicate that
the failure loads are generally well predicted for all four considered combinations of global
and local imperfection amplitudes. Figure 1(a) and (b) depict the test and numerical load-mid
height lateral deflection responses using the measured global imperfection amplitude and t/10
for the local imperfection amplitude; as these achieved the best overall agreement. Overall, it
may be concluded that numerical models are capable of simulating the behaviour of the tests,
and are suitable for performing parametric studies.

3 PARAMETRIC STUDY

Upon validation of the numerical models for stainless steel CHS columns, a series of paramet-
ric studies were performed to generate further structural performance data over a wide range

Figure 1. Experimental and numerical load-mid height deflection curves.

823
Table 3. Summary of the parametric study variables.
Grade Section (D × t) λc λθ

Austenitic CHS 100 × 8 0.18 0.1–2.0


Duplex CHS 100 × 8 0.19 0.1–2.0
Ferritic CHS 100 × 8 0.16 0.1–2.0

of member slenderness λθ and elevated temperatures θ. Table 3 provides the examined param-
eters. The room temperature cross-section slenderness of the modelled CHS, λc, determined
from Equation (3), is also included in Table 3, which were within the fully effective section
range according to EN 1993-1-4 cross-section classification limits. In Equation (3), fcr,c is the
elastic critical buckling stress for a circular hollow section, f2 is the stress at 2% total strain,
D and t are the outer diameter and thickness, respectively, E is the Young’s modulus and ν is
the Poisson ratio. For modelling convenience, the parametric study numerical models were
performed isothermally, where the material properties for a given temperature θ were incorp-
orated into the FE models, akin to applying a uniform temperature θ, and the applied load
was increased until failure at the following temperatures: 200, 400, 600 and 800°C. Since time
dependent effects, e.g. creep were not incorporated in the developed numerical models, this
method was deemed sufficient.

qffiffiffiffiffiffi  0:5 qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi0:5


f 2D
λc ¼ f2
f cr;c ¼ 3ð1ν2 Þ ð3Þ
2tE

The material properties recommended by Afshan et al. (2019) at room temperature for numer-
ical parametric studies of cold-formed stainless steel CHS were employed together with the
reduction factors provided in Table 8.1 of the Design Manual for Structural Stainless Steel
(DMSS) (2017) for Austenitic I, Duplex II and Ferritic II grades. The two-stage Ramberg-
Osgood material model was then used to develop full-range stress-strain curves at discrete tem-
peratures of 200, 400, 600 and 800°C using the 2% total strain method provided in Clause 8.5 of
the Design Manual for Structural Stainless Steel (2017). The values for nθ were taken as the
room temperature values for n provided in the Design Manual for Structural Stainless Steel
(2017) and values for mθ,2 were determined using Equation (4) as recommended in Clause 8.5 of
the Design Manual for Structural Stainless Steel (2017). The global imperfection amplitude was
taken as L/1000, where L is the column length, in accordance with the permitted out of straight-
ness tolerance in EN 1090-2 (2008). The local imperfection amplitude was set to 0.008D, where
D is the diameter, in accordance with EN 1993-1-5 Annex C (2006) recommendation.

f 0:2;θ
mθ;2 ¼ 1 þ 2:8 ð4Þ
f u;θ

where f0.2,θ is the 0.2% proof stress at temperature θ and fu,θ is the ultimate tensile stress at
temperature θ.

4 ANALYSIS OF RESULTS AND DESIGN RECCOMENDATIONS

4.1 Eurocode 3: Part 1.2 design method


From EN 1993-1-2 (2005), the buckling design resistance Nb,fi,t,Rd at time t of acompres-
sion member with a uniform temperature θ is determined from Equation (5) for Class 1-3
cross-sections.

824
χfi Ak2;θ f y
Nb;fi;t;Rd ¼ γM;fi for Class 1; 2 and 3 cross-sections ð5Þ

where, χfi is the flexural buckling reduction factor in fire which is given by Equation (6), A is
the area of the cross-section, k2,θ is the reduction factor for strength at 2% total strain at tem-
perature θ, fy is the room temperature 0.2% proof strength and γM,fi is the partial resistance
factor.

1 h i qffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2
χfi ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi with jθ ¼ 0:5 1 þ αλθ þ λθ with α ¼ 0:65 235=fy ð6Þ
2
jθ þ j2 θ  λθ

The non-dimensional elevated temperature slenderness λθ is defined by Equation (7), where, λ
is sthe non-dimensional member slenderness at room temperature, and kE,θ is the reduction
factor for the Young’s modulus at temperature θ.
 0:5
k2;θ
λθ ¼ λ for Class 1; 2 and 3 cross-sections ð7Þ
kE;θ

4.2 Design manual for structural stainless steel design method


The Design Manual for Structural Stainless Steel (2017) design buckling resistance Nb,fi,t,Rd at
time t of a compression member with a uniform temperature θ, is given by Equation (8) for
Class 1, 2 and 3 cross-sections.
χfi Ak0:2;θ f y
Nb;fi;t;Rd ¼ γM;fi for Class 1; 2 and 3 cross-sections ð8Þ

where k0.2,θ is reduction factor for the 0.2% proof strength at temperature θ. In addition,
the design method uses the same flexural buckling curves as for room temperature as
given by Equation (9), with a fixed imperfection factor (α) and plateau length (λ0) for all
temperatures, as provided in Table 6.1 of the Design Manual for structural stainless
steel. The non-dimensional elevated temperature member slenderness λθ is defined by
Equation (10).

1 h   2
i
χfi ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi  1:0 with jθ ¼ 0:5 1 þ α λθ  λ0 þ λθ ð9Þ
2
jθ þ j2 θ  λθ

 0:5
k0:2;θ
λθ ¼ λ for all Classes of cross-section ð10Þ
kE;θ

4.3 Lopes et al. method


A numerical study on axially loaded stainless steel welded I-section columns in fire was per-
formed by Lopes et al. (2010) where modifications to the flexural buckling design formula-
tions of EN 1993-1-2 (2005) were proposed. The amendments included, (1) the imperfection
factor (α) being defined as a function of temperature as given by Equation (11) and (2) the
introduction of a β parameter in the non-dimensional buckling reduction factor χfi and φθ for-
mulations as defined in Equation (12), which provides different buckling curves for different

825
elevated temperatures. For welded I-section column a variety of β and α values were proposed
for different stainless steel grades (Lopes et al., 2010) and buckling axis.
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffisffiffiffiffiffiffiffiffi
235 E kE;θ
α¼η ð11Þ
f y 210000 k2;θ

1 h 2
i
χfi ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi  1:0 with jθ ¼ 0:5 1 þ αλθ þ βλθ ð12Þ
2
jθ þ j2 θ  βλθ

4.4 Analysis of results and design proposal


This section presents and compares the results obtained from Section 3 with the existing structural
fire design guidelines utlising the reduction factors from the Design Manual for Structural Stain-
less Steel (2017) provided in Sections 4.1 and 4.2. Figure 2 compares the EN 1993-1-2 (2005)
buckling curve for Class 1 CHS columns in fire with the FE results, where the ultimate FE load is
normalised by the cross-sectional yield load (Ak2,θfy) and plotted against the non-dimensional
member slenderness (λθ), defined by Equation (7). The EN 1993-1-2 (2005) buckling curve gener-
ally over-predicts the buckling resistance of the CHS austenitic and duplex columns, while it pro-
vides closer predictions to the FE results for the ferritic columns. Figure 3 present the DMSS

Figure 2. Comparison of FE results with EN 1993-1-2 (2005) buckling curve.

Figure 3. Comparison of FE results with design manual for structural stainless steel (2017).

826
Table 4. Comparison between the FE and predicated resistances.
Material Nu,FE=Nu,pred EN 1993-1-2 (2005) DMSS (2017) Proposed

No. 50 50 50
Mean 0.910 1.047 1.051
Austenitic COV 0.108 0.094 0.054
Max 1.060 1.392 1.158
Min 0.749 0.895 0.942
No. 50 50 50
Mean 0.930 1.055 1.095
Duplex COV 0.119 0.083 0.055
Max 1.085 1.250 1.191
Min 0.696 0.865 0.959
No. 56 56 56
Mean 1.085 1.049 1.081
Ferritic COV 0.063 0.082 0.066
Max 1.216 1.200 1.197
Min 0.966 0.901 0.960

buckling curve and the FE results, where the ultimate load is normalised by the cross-sectional
yield load (Ak0.2,θfy) and plotted against the non-dimensional member slenderness (λθ), given by
Equation (10). The Design Manual for Structural Stainless Steel (2017) under-predicts the resist-
ance of the stockier columns with low elevated temperature member slenderness values as it limits
the cross-section resistance to the squash load based on the 0.2% proof stress, though it gives
improved predictions than the EN 1993-1-2 (2005) method for high slenderness ranges. Table 4
presents a summary of the numerical comparison ratios between the ultimate loads obtained
from FE (Nu,FE) and predicted resistance from design methods (Nu,pred) in terms of mean, coeffi-
cient of variation (COV), maximum and minimum values, and No. is the number of columns.
Buckling curves of the same form as the Lopes et al.’s formulation, presented in Section 4.3,
were fitted to the normalised FE data to extend its application to cold formed stainless steel
CHS columns, and are shown in Figure 4. The Lopes et al. (2010) method uses the 2%
strength (k2,θfy) in its buckling curve formulation and in that respect allows harmonization
between carbon steel and stainless steel design rules. The use of the α parameter, as given in
Equation (11), which is a function of elevated temperature strength and stiffness reduction
factors enables different buckling curves for different temperatures, which is required for
stainless steel columns. In addition, the introduction of the β parameter in the χfi equation
allows the shape of the buckling curve to better represent the FE data. The β and α parameters
were calibrated against the austenitic, duplex and ferritic CHS stainless steel columns gener-
ated herein and are reported in Table 5.

4.5 Reliability analysis


Reliability analysis according to the method proposed by Kruppa (1999) was performed to
assess the accuracy of the existing and the proposed design method to predict the buckling
resistance of cold-formed austenitic, duplex and ferritic stainless steel CHS columns in fire.
A summary of the safety assessment results is provided in Table 6. The reliability method pro-
posed by Kruppa (1999) sets out three criteria as described hereafter.
1. The predicted theoretical resistances shall not be on the unsafe side beyond 15% of the
test or FE results.
2. A maximum of 20% of the predicted theoretical resistance result shall be on the unsafe side.
3. The mean value of all percentage differences between predicted theoretical results and
test or FE results shall be on the safe side.

827
Figure 4. Comparison of FE results with the proposed buckling curves for (a) austenitic, (b) duplex and
(c) ferritic columns.

Table 5. Proposed β and η parameters for cold-formed CHS


columns.
Material Section β η

Austenitic CHS 0.7 1.3


Duplex CHS 0.8 1.2
Ferritic CHS 1.0 0.5

Table 6. Summary of the parametric study variables.


Material Criterion EN 1993-1-2 (2005) DMSS (2017) Proposed

Criterion 1 40.0% Fail 0.0% Pass 0.0% Pass


Austenitic Criterion 2 76.0% Fail 34.0% Fail 20.0% Pass
Criterion 3 0.112 Fail -0.037 Pass -0.046 Pass
Criterion 1 26.0% Fail 2.0% Fail 0.0% Pass
Duplex Criterion 2 70.0% Fail 26.0% Fail 8.0% Pass
Criterion 3 0.092 Fail -0.045 Pass -0.084 Pass
Criterion 1 0.0% Pass 0.0% Pass 0.0% Pass
Ferritic Criterion 2 8.9% Pass 28.6% Fail 16.1% Pass
Criterion 3 -0.075 Pass -0.040 Pass -0.071 Pass

828
5 CONCLUSION

In this paper, a numerical study investigation was conducted on stainless steel CHS
column in fire which included a detailed description of the numerical models and valid-
ation results. Validated models were utilised to perform parametric studies which
included columns of three main stainless steel grades, at different elevated temperatures
and member slenderness between 0.1-2.0. The FE results were analysed and it was shown
that the EN 1993-1-2 (2005) and the Design Manual for Structural Stainless Steel (2017)
methods give inaccurate and unsafe predictions for stainless steel CHS flexural buckling
resistance in fire. New buckling curves in line with the Lopes et al.’s method, for cold-
formed CHS columns in fire were proposed on the basis of numerical results, which were
shown to give an accurate prediction. The accuracy and suitability of the proposed buck-
ling curves was verified by means of reliability criteria.

REFERENCES

ABAQUS 6.16. 2016. Dassault Systmes Simulia Corp. USA.


Afshan, S., Zhao, O. & Gardner, L. 2019. Standardised material properties for numerical parametric
studies of stainless steel structures and buckling curves for tubular columns. Journal of Constructional
Steel Research 152: 2 –11.
Buchanan, C., Real, E. & Gardner, L. 2018. Testing, simulation and design of cold-formed stainless steel
CHS columns. Thin-Walled Structures 130: 297 –312.
CEN E 10219-2. 2006. Cold formed welded structural hollow sections of non-alloy and fine grains
steels – part 2: Tolerances. Dimensions and Sectional properties, Brussel, Belgium.
Cruise, RB. & Gardner, L. 2008. Residual stress analysis of structural stainless steel sections. Journal of
Constructional Steel Research 64: 352 –366.
EN 1090-2. 2008. Execution of steel structures and aluminium structures - Part 2: Technical requirements
for steel structures. Brussels: European Committee for Standardization.
EN 1993- 1-2. 2005. Eurocode 3: Design of steel structures - Part 1-2: General rules - Structural fire
design. Brussels: European Committee for Standardization.
EN 1993- 1-4. 2006. Eurocode 3: Design of steel structures - Part 1-4: General rules for stainless steels.
Brussels: European Committee for Standardization.
EN 1993- 1-5. 2006 Eurocode 3: Design of steel structures - Part 1-5: Plated structural elements. CEN,
Brussels.
Gardner, L. & Ashraf, M. 2006. Structural design for non-linear metallic materials. Engineering Struc-
tures 28: 926 –934.
Kruppa, J. 1999. “Eurocodes-Fire parts: Proposal for a methodology to check the accuracy of assessment
methods,” CEN TC 250, Horizontal Group Fire, Document no: 99/130.
Lopes, N., Vila Real, P., Silva L.S. & Franssen, J-M. 2010. Axially Loaded Stainless Steel Columns in
Case of Fire. Journal of Structural Fire Engineering 1: 43 –60.
SCI. 2017. Design Manual for Structural Stainless Steel. Fourth Edition.
Zhao, O., Gardner, L. & Young, B. 2016. Testing and numerical modelling of austenitic stainless steel
CHS beam-columns. Engineering Structures 111: 263 –274.

829
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

The capacity of bolted cold-formed steel connections in bending

S.M. Mojtabaei, J. Becque & I. Hajirasouliha


Department of Civil and Structural Engineering, The University of Sheffield, Sheffield, UK

ABSTRACT: Cold-formed steel (CFS) sections find extensive use in moment-resisting portal
frames, which are commonly encountered in industrial buildings. The structural performance of
a moment-resisting portal frame, however, depends significantly on the behaviour of the
column-to-rafter connections. This paper investigates the capacity of these connections, which
are commonly implemented with a gusset plate bolted to the webs of back-to-back channel sec-
tions, when subjected to a pure bending moment. Detailed Finite Element (FE) models,
accounting for material non-linearity, geometric imperfections and realistic bolt behaviour were
developed using the ABAQUS software and were validated against experiments on bolted CFS
apex connections. Using the validated model parametric studies were carried out to investigate
the effects of the cross-sectional dimensions and thickness of the connected members, the bolt
group length and the geometric arrangement of the bolts in the connection. On the basis of this
data a design equation was proposed which accounts for local buckling of the CFS beam web
adjacent to the first bolt row as the dominant failure mode.

Keywords: cold-formed steel sections, bolted moment connection, FE analysis, design equation

1 INTRODUCTION

Cold-formed steel (CFS) structural elements possess the advantages of high strength-to-weight
and high stiffness–to-weight ratios, a relatively straightforward and adaptable manufacturing
process, easy handling, transportation and installation, and an increased construction speed.
They are extensively used as purlins, girts (side rails), stud walls, mezzanine floors and storage
racks, and increasingly also in moment-resisting portal frames in halls for industrial or commu-
nity use. The structural performance of CFS moment-resisting frames depends to a large extent
on the flexural behaviour of the connection zone. The connections in CFS portal frames typically
consist of an edge-stiffened plate (sometimes referred to as a ‘bracket’) to which the webs of the
frame members are bolted. This configuration has the advantage of an easy assembly on site, but
necessitates the transfer of the full moment through the web. This makes the beam (rafter) web
susceptible to local buckling failure immediately outside the connection. However, other failure
modes are possible, such as bolt shear or bearing failure, or failure of the gusset plate due to lat-
eral-torsional buckling (Lim and Nethercot, 2004). Lim and Nethercot (2004) demonstrated
experimentally as well as numerically that the bolt group length lb (i.e. the distance between the
outer bolt rows) has a dominant effect on the capacity of bolted CFS connections. As part of
their research work they also proposed an empirical equation to predict the flexural capacity of
CFS moment connections in conventional back-to-back channel beams with a given bolt group
length and bolt group configuration, provided local buckling is the governing failure mode.
Noteworthy experimental research on CFS bolted connections has also been carried out by
Chung & Lau (1999), Wong & Chung (2002) and Sabbagh et al. (2012).
The present study aimed to further investigate local web buckling in CFS connections through
numerical means over a wide range of variables, including the cross-sectional dimensions of the
beam, its thickness, the bolt group configuration and the bolt group length. The here presented
results are an extension of previous work reported by the authors (Mojtabaei et al., 2019) in that

830
a more comprehensive range of section geometries is considered, additionally varying the distance
between the web, where the load is introduced by the connection, and the centroid of the individ-
ual channels. New design equations are also proposed.

2 FINITE ELEMENT MODEL

The ABAQUS (2014) finite element software was used to create detailed models of CFS
bolted moment connections, accounting for material nonlinearity, geometric imperfections
and realistic bearing behaviour of the bolts. The FE model was first validated against the tests
conducted by Lim & Nethercot (2003).

2.1 Bolt modelling


“Discrete fastener” elements, which are available from the standard ABAQUS library, were used
to model the bolt behaviour. For the purpose of validating the FE model the bearing behaviour
of the bolt against the channel web was obtained from lap shear test conducted by Lim and Neth-
ercot (2003). Figure 1 shows the measured load-elongation behaviour. In further parametric stud-
ies, which included plate thicknesses different from those reported by Lim & Nethercot, resulting
in a different bearing behaviour, the equations proposed by Fisher (1964) were used:
h iλ
RB ¼ Rult 1  eμðδbr =25:4Þ with Rult ¼ 2:1  d  t  Fu ð1Þ

In the above equation δbr is the bearing deformation (in mm), Rult is the ultimate bearing strength,
t is the web thickness, d is the bolt diameter and RB is the bearing load. Fu is the tensile strength
of the web material and μ ¼ 5 and λ ¼ 0:55 are regression coefficients borrowed from Uang et al.
(2014). It should be noted that bolt failure was excluded as a failure mode in the parametric stud-
ies by proper design, i.e. by selecting a large enough bolt diameter. In addition, the bearing behav-
iour of the bolts, as encoded in Eq. (1), greatly affected the rotational stiffness of the connection,
but did not have any significant influence on the local buckling capacity of the beam web, as
expected. Nevertheless, modelling the bolt behaviour accurately in order to obtain matching stiff-
ness predictions with the experiment was deemed an important part of the validation process.

2.2 Element type, material properties and boundary conditions


An 8-noded quadrilateral shell element with reduced integration (S8R) was selected from the
ABAQUS library and was used to model both the CFS beam and the gusset plate. A mesh sen-
sitivity analysis indicated that a mesh size of 20 × 20 mm2 provided a good balance between
model accuracy and computational time and this mesh was used throughout the study.

Figure 1. Bearing behaviour of a single bolt (Lim and Nethercot 2004).

831
A bi-linear stress-strain model was used with an elastic modulus E = 210 GPa, followed by
a linear hardening range with a slope of E/100. During the validation process values of the yield
stress and the ultimate stress were used as reported from coupon tests by Lim & Nethercot (2003):
fy = 358 MPa and fu = 425 MPa for the channel sections, and fy = 341 MPa and fu = 511 MPa for
the gusset plate.
Figure 2 illustrates the FE model and indicates the loading and boundary conditions. The
gusset plate was fixed (clamped) at its end and connected to back-to-back channels by means of
the fastener elements. At the loaded end of the beam, the rotational degrees of freedom about
the x and y-axis of all points in the cross-section were coupled to those of the centroid and an
increasing rotation was applied. The beam was supported in the out-of-plane direction to pre-
vent lateral-torsional buckling.

2.3 Imperfections
Due to the absence of lateral-torsional buckling, only a local or a distortional imperfection was
incorporated into the model, depending on which mode had the lower critical buckling stress.
The shape of the geometric imperfections was determined by carrying out an elastic buckling
analysis in ABAQUS and retaining the lowest eigenmode. For cross-sections which a thickness
less than 3 mm (which corresponds to the range of validity of the results reported Schafer and
Pekӧz 1998), the amplitude of the imperfection shape was taken as 0.34t or 0.94t, depending on
whether the imperfection was local or distortional, respectively. These imperfection magnitudes
correspond to the 50% value of the Cumulative Distribution Function and represent the ‘most
likely’ imperfections. When the thickness of the cross-section was larger than 3mm, the imper-
fection magnitude was determined based on the equation proposed by Walker (1975).

2.4 Sensitivity study


A sensitivity study was conducted to investigate the effect of the modelled beam length,
expressed as a multiple of the cross-section depth: le = ωh on the connection capacity. This inves-
tigation was carried out for two different bolt group lengths: lb/h = 0.5 and lb/h = 3. The results
are shown in Figure 3 and it is seen that the modelled beam length may have quite a significant
effect on the predicted capacity. However, its influence becomes negligible for ω ≥ 6. Therefore,
a length (le) equal to 6h was used in the parametric studies.

2.5 Validation
Table 1 provides a comparison between the FE predicted capacities MFE and the experimen-
tally capacities MTest obtained by Lim & Nethercot (2003). It is seen that good agreement was
achieved, with an average ratio of MFE/MTest of 1.02 and a standard deviation of 0.018.

Figure 2. FE model of the bolted CFS moment connection, including loading and boundary conditions.

832
Figure 3. Sensitivity study of the cantilever beam length (le) for two different bolt group lengths
(lb/h = 0.5 and lb/h = 3).

Table 1. Comparison between the connection capacities obtained from the experiment
(Lim and Nethercot, 2003) and FE.
lb MTest MFE
connection (mm) lb =h (kN.m) (kN.m) MFE/MTest

1 315 0.94 75 76.01 1.01


2 390 1.16 77.5 80.72 1.04
3 465 1.38 82.5 82.5 1.00
4 615 1.83 87.5 88.02 1.01
Average 1.02
St. dev 0.018

3 PARAMETRIC STUDIES

Table 2 provides an overview of the parametric studies. Three different back-to-back lipped-
channel geometries were selected. Table 2 lists their dimensions: h is the depth of the web, b is
the flange width and c is the width of the lip. All dimensions where measured along the heart
line of the cross-section. The rationale behind selecting these geometries is that they achieved

Table 2. Selected variables for parametric studies of bolted CFS moment connections.
Variables
Dimensions

h b c

Beam section Number


Bolt group length Bolt configuration mm mm mm (t = 1, 2, 4 and 6 mm) of bolts

300 50 25 Channel 1
hb lb
¼ 0:8; ¼ ½0:5; …; 3
Rectangular (2 × 2) 250 75 25 Channel 2 4
h h 200 100 25 Channel 3
300 50 25 Channel 1
hb lb
¼ 0:8; ¼ ½0:5; …; 3
Rectangular (3 × 3) 250 75 25 Channel 2 9
h h 200 100 25 Channel 3
300 50 25 Channel 1
hb lb
¼ 0:8; ¼ ½0:5; …; 3
Rectangular (4 × 4) 250 75 25 Channel 2 12
h h 200 100 25 Channel 3

833
a systematic shift of the location of the centroid away from the web. The distance between the
heart line of the web and the centroid will be denoted by the eccentricity X and the aim was to
study the influence of this parameter on the connection capacity. The three cross-sectional
geometries were combined with four different thicknesses: 1, 2, 4, and 6 mm. As also shown in
Table 2, three different bolt arrays (2 × 2, 3 × 3 and 4 × 4), each with relative bolt group lengths
ðlb =hÞ varying from 0.5 to 3, were considered. In total, 330 nonlinear FE models were developed
to investigate the effects of the abovementioned variables on the flexural capacity of the connec-
tions. To facilitate pre- and post-processing of the results ABQUS scripts, based on the pro-
gramming language Python, were used (Abaqus/CAE User’s Manual, 2014).
In all parametric studies the yield stress fy, the elastic modulus E and the Poisson’s ratio ν
of the steel material were taken as 313 MPa, 210 GPa and 0.3, respectively.
The moment capacity of a CFS back-to-back channel beam adjacent to a bolted connection
(Mc) can be expressed as:

Mc ¼ RMu ð2Þ

where R is a reduction factor accounting for the fact that local web buckling adjacent to the
connection may prevent the full cross-sectional bending strength Mu from being reached. Mu
was obtained from FE analysis for the back-to-back channels under consideration and the
resulting values are listed in Table 3. A model similar to the one shown in Figure 2 was used,
but with the bolted connection removed. Instead the beams were loaded in uniform bending by
applying equal and opposite rotations at both ends. The length of the FE model used to
determine Mu was taken as three times the distortional buckle half-wave length, as suggested by
Shifferaw and Schafer (2012). The distortional buckle half-wave length was calculated using the
CUFSM software (Li and Schafer, 2010), which implements the elastic finite strip method.
The aim of the parametric studies was to express the reduction factor R as a function of the
relevant parameters, namely the bolt group configuration, the bolt group length, the cross-
sectional shape of the beam and its thickness. The influence of the cross-sectional shape and thick-
ness were represented through the eccentricity X, as well as through the slenderness parameter λ,
calculated as follows:
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
λ¼ My =Mcr ð3Þ

In the above equation My = Zf fy and Mcr = Zf fcr are the yield moment and the elastic local
buckling moment of the cross-section, respectively. Zf is the section modulus about the axis of

Table 3. Cross-sectional properties and capacity of channel beams.

t X λ Mu
Cross-section (mm) (mm) (-) (kN.m)

1 11.11 2.047 15.26


2 11.11 1.034 44.05
Channel 1
4 11.11 0.572 107.15
6 11.11 0.436 168.00
1 20.83 1.752 13.08
2 20.83 0.877 41.81
Channel 2
4 20.83 0.575 96.13
6 20.83 0.446 147.72
1 33.33 1.844 10.78
2 33.33 0.970 32.89
Channel 3 4 33.33 0.658 80.86
6 33.33 0.514 123.96

834
Table 4. Proposed reduction factors for bolted CFS moment connections.
Eq. No. Bolt arrays Reduction factor (R)
1 1 1 1 0:64 1 0:81
R ¼ 1:05 þ ð0:03λÞ: logðLb =hÞ þ ð1:14  1:19ð Þ0:05 Þ:ð Þ1:7 : þ ð1:41 þ 1:83ð Þ Þ:ð Þ :λ
(4) 2×2 λ λ ðLb =hÞ X=Lb X=Lb
1 1 1 1 0:82 1 0:61
R ¼ 1:09 þ ð0:05λÞ: logðLb =hÞ þ ð0:18  0:24ð Þ0:46 Þ:ð Þ1:17 : þ ð0:77 þ 0:69ð Þ Þ:ð Þ :λ
(5) 3×3 λ λ ðLb =hÞ X=Lb X=Lb
1 1 1 1 0:05 1 0:54
R ¼ 1:04 þ ð0:02λÞ: logðLb =hÞ þ ð0:75 þ 0:70ð Þ0:07 Þ:ð Þ1:34 : þ ð0:12  0:59ð Þ Þ:ð Þ :λ
(6) 4×4 λ λ ðLb =hÞ X=Lb X=Lb
bending and fcr is the critical elastic buckling stress of the section, which was calculated using
the CUFSM software. Table 3 lists the cross-sectional slenderness λ and the eccentricity X for
the different cross-sections considered in this study.

4 PROPOSED DESIGN EQUATIONS

Surface fitting techniques in MatLab (Mathworks 2011) were used to develop design equations
based on the results of the parametric studies. Separate design equations were proposed for each
bolt group configuration and they are listed in Table 4. Figures 3–5 graphically compare the
predicted reduction factors R with the FE analysis results for the 2 × 2, 3 × 3, and 4 × 4 bolt
arrays, respectively. It is seen that good agreement is obtained.
The FE results indicate that increasing the bolt group length is most beneficial for the connec-
tion strength. This can be explained by the fact that the horizontal components of the bolt forces
(which locally increase the longitudinal compressive stresses in the web and promote local buck-
ling) are much larger in shorter bolt groups. On the other hand, the connection capacity is
inversely related to the eccentricity X. This is expected, as a more significant shear lag effect ori-
ginates for larger X values. This shear lag effect is caused by the fact that the bending moment is
introduced into the web and subsequently has to spread out into the flanges. A larger reduction
in capacity is observed for smaller thicknesses (or larger slenderness values), as the cross-section
becomes more prone to local buckling.

5 SUMMARY AND CONCLUSIONS

This paper describes a study of the capacity of bolted CFS moment-resisting connections,
focusing on local buckling of the beam web adjacent to the connection as the governing failure
mode. This type of failure arises because of the thin-walled nature of the connected members,
but also because of the common practice in CFS portal frame construction of only connecting
the web to the gusset plate, while leaving the flanges unconnected. A numerical study was con-
ducted using detailed finite element models which accounted for geometric and material non-

Figure 4. Flexural capacity of CFS bolted moment connections with rectangular 2 × 2 bolt configur-
ation and various bolt group lengths and cross-sectional eccentricities.

836
Figure 5. Flexural capacity of CFS bolted moment connections with rectangular 3 × 3 bolt configur-
ation and various bolt group lengths and cross-sectional eccentricities.

Figure 6. Flexural capacity of CFS bolted moment connections with rectangular 4 × 4 bolt configur-
ation and various bolt group lengths and cross-sectional eccentricities.

linear behaviour, geometric imperfections and realistic bolt behaviour. The model was first
verified against previously reported tests on bolted CFS apex connections. Parametric studies
were carried out to investigate the influence of the cross-sectional geometry and thickness
(reflected in the slenderness parameter λ and the eccentricity X), the bolt array configuration
and the bolt group length on the ultimate capacity. Based on the results of these studies design
equations were proposed to predict the flexural capacity of CFS bolted connections

837
REFERENCES

ABAQUS/CAE. 2014. User's manual version 6.14–2, USA.


Chung, K.F. & Lau, L. 1999. Experimental investigation on bolted moment connections among cold
formed steel members. Engineering Structures, 21, 898–911.
Fisher, J.W. 1964. On the Behavior of Fasteners and Plates with Holes. Fritz Engineering Laboratory,
Department of Civil Engineering, Lehigh University.
Li, Z. & Schafer, B.W. 2010. Buckling analysis of cold-formed steel members with general boundary con-
ditions using CUFSM conventional and constrained finite strip methods. 20th International Specialty
Conference on Cold-Formed Steel Structures, St. Louis, USA.
Lim, J.B.P. & Nethercot, D.A. 2003. Ultimate strength of bolted moment-connections between
cold-formed steel members. Thin-Walled Structures, 41, 1019–1039.
Lim, J.B.P. & Nethercot, D.A. 2004. Finite element idealization of a cold-formed steel portal frame.
Journal of Structural Engineering, 130, 78–94.
Mathworks. 2011. Matlab R2011a, Natick, MA, USA.
Mojtabaei, S.M., Becque, J. & Hajirasouliha, I. 2019. Behaviour of cold-formed steel moment connec-
tions. 9th International Conference on Steel and Aluminium Structures (ICSAS19). Bradford, UK
Sabbagh, A.B., Petkovski, M., Pilakoutas, K. & Mirghaderi, R. 2012. Experimental work on
cold-formed steel elements for earthquake resilient moment frame buildings. Engineering Structures,
42, 371–386.
Schafer, B.W. & Peköz, T. 1998. Computational modeling of cold-formed steel: characterizing geometric
imperfections and residual stresses. Journal of Constructional Steel Research, 47, 193–210.
Shifferaw, Y. & Schafer, B.W. 2012. Inelastic Bending Capacity of Cold-Formed Steel Members. Journal
of Structural Engineering, 138, 468–480.
Uang, C.M., Sato, A., Hong, J.K. & Wood, K. 2010. Cyclic Testing and Modeling of Cold-Formed Steel
Special Bolted Moment Frame Connections, Journal of Structural Engineering, 136, 953–960.
Walker, A. C. 1975. Design and analysis of cold-formed sections, Halsted Press.
Wong, M.F. & Chung, K.F. 2002. Structural behaviour of bolted moment connections in cold-formed
steel beam-column sub-frames. Journal of Constructional Steel Research, 58, 253–274.

838
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Study on the deformation and rotation capacity of HSS hollow


sections

A. Mueller & A. Taras


Institute of Structural Engineering, Bundeswehr University Munich, Germany

ABSTRACT: With increasing use of high-strength steel grades the need for more accurate
design specifications arises, with the aim of fully exploiting the material benefits and create
economic advantages. According to Eurocode 3, the maximum rotational capacity is limited
and linked to the definition of cross-sectional classes. For class 1 the rotation θ is assumed to
be “infinite”, while it drops significantly for class 2 and 3 to a maximum rotation of θpl and
θel, respectively. In reality, in spite of their lower hardening capacity and ultimate strains,
high-strength steel sections display a non-negligible rotational capacity that exceeds these
code predictions, which were developed for mild steel and with a level of analysis in mind that
is suitable for hand calculations. For an increased use of high-strength steel sections, it is thus
very important to understand and to be able to predict a more realistic deformation and rota-
tion capacity, with the aim of implementing the findings in tools for advanced, FEM-based
Design by Analysis (DbA) approaches. As an initial step in this direction, the proposed paper
shows the results from numerical calculations on the rotational capacity of HSS rectangular
hollow sections. The numerical results are calibrated against laboratory tests. Consequently,
different rectangular cross section dimensions but also variations of the steel grade and the
thickness are chosen and studied using advanced FEM modelling techniques.

1 INTRODUCTION

Structural hollow sections have a wide range in engineering applications, especially when talking
about structural efficiency, self-weight optimization or structural design with a certain architec-
tural demand. They are basically separated in two different classifications depending on the
manufacturing process, being produced by hot-rolling or cold forming, whereby hot rolling can
be further divided in hollow sections formed hot and formed cold with additional heat treatment
(EN 10210 2006, EN 10219 2006). Although the latter case is not investigated in this paper it has
an influence on ductility in terms of elongation and tensile ration compared to hot rolled sec-
tions (Zhang et al. 2010). Hot rolled sections in general has homogeneous material properties
around their sections, due to the manufacturing process relatively low residual stresses combined
with a good ductility and tight corner radii. On the other side cold-formed sections undergo high
plastic deformations at ambient temperatures during the forming process which lead to
a variation of hardness and non-homogeneity of the material properties between the flat and the
corner regions of the sections – resulting in enhanced strength and a corresponding loss of duc-
tility (Yun et al. 2018, Gardner et al. 2010, Amouzegar et al. 2016). A recent experimental inves-
tigation on high strength cold formed hollow sections resulted in strength enhancements up to
34% for the corner regions (Ma et al. 2015). A significant difference between hot-rolled and
cold-formed steel sections is the presence and magnitude of residual stresses. In general, residual
stresses are not assumed for hot-rolled SHS profiles, as they are considered to be residual stress
free, at least at a negligible level as shown by Ma et al. (2015), without significant changes in the
properties of the base material (Kato 1982). It is the cold forming process that leads to locked-in
residual stresses in transverse and longitudinal direction (Tong et al. 2012, Abvabi et al. 2015,
Zhang et al. 2010, Gardner et al. 2010).

839
This paper shows influences on the rotation capacity, that arise by the choice of different
thicknesses and the corresponding changing geometry due to code provisions (EN 10210 2006,
EN 10219 2006). Different steel grades and material laws (see Section 3.2) were chosen to take
strain hardening into account and replicate a precise material stress-strain response. The influ-
ence of initial imperfection, based on 3D scan date and additional GMNIA calculations, was
investigated exemplary to show differences in the development of the plastic hinge location
and its impact on the post-buckling curves.

2 ROTATION CAPACITY

For the use of inelastic analysis methods, the structural members must have a certain amount
of ductility to undergo deformation after reaching its initial yield without any significant loss
in its ultimate strength. This rotation capacity Rcap is defined through Eq. (1). It is a measure
of how much the plastic hinge can rotate before failure occurs.

’pl;2  ’pl ’pl;2


Rcap ¼ ¼ 1 ð1Þ
’pl ’pl

φ represents the beam end sections rotation and its limit values φpl, φu, φpl,2. φpl,2 is the limit rota-
tion at which the moment drops below Mpl. EC3-1-1 (2005) provides rotation demands Rcap
depending on the cross-section classifications (Figure 1). According to the definition class 2 cross-
sections can reach its plastic moment, but fail to reach a rotation capacity of 3. Class 1 cross-
sections fulfill this demand by reaching a rotation capacity that is larger than 3, see Figure 1a.
Although the code requires to make use of a minimum permitted rotation capacity for the
sections (Rcap = 3), the rotation capacity of a structural member could be determined more
accurate by means of FE based plastic analysis. Figure 2a is exemplarily showing the evalu-
ation of a cold-formed SHS 200 profile with the thickness of t = 8 mm (Class 1) and the length
of l = 825 mm. Strain hardening effects on the rotation demand are clearly visible, even
though the profile is almost reaching Class 2.

Figure 1. Rotation capacity Rcap according to EC3-1-1 (2005); b) Definition of the rotation capacity
Rcap.

840
Figure 2. a) Comparison of test date with further methods of numerical calculation; b) Test set-up.

3 NUMERICAL MODELING

3.1 Chosen modelling techniques


A series of individual calculation runs of increasing sophistication were performed in the
numerical studies presented in this paper. Linear bifurcation analyses (LBA) were carried out
in order to identify the elastic critical buckling resistance. Geometrically and materially non-
linear analyses (GMNA) served the purpose of identifying the elastic-plastic buckling resist-
ance of the perfect structure. Geometrically and materially nonlinear analyses with
imperfections (GMNIA) were performed to determine the elasto-plastic buckling load, i.e. the
realistic buckling resistance that considers both material and geometric nonlinearities. For the
calculations based on the scanned real geometry data, no further analysis types – like LBA or
GMNA – had to be performed beforehand and a GMNIA could directly be performed on the
as-fabricated structural geometry. This calculation served as a means to calibrate and validate
the numerical model against the conducted full-scale tests.

3.2 Applied stress-strain curves for hot-rolled and cold-formed steels


For the numerical analysis, two recently published material models for hot-rolled and cold-formed
steels are used to describe a more accurate structural behavior. The material model for hot-rolled
carbon steel is based on a bilinear plus non-linear hardening description (Yun 2017). The stress-
strain response for cold-formed steels can be described by a two-stage Ramberg-Osgood model
(Yun 2018). The fact, that cold-formed steel shows a different stress-strain relation in the corner
and flat areas, as well as the overall presence of residual stresses, were not taken into account
during the preparations for this paper, as this is going to be part of ongoing investigations.

3.3 Description of the FE-model


The recalculation of the bending tests and additional numerical study on the influences of the rota-
tion capacity was performed with help of the commercial FEM programme Simulia Abaqus™
(Abaqus 2016). The numerical models made use of three-dimensional shell elements of type S4R
and the material models based on the description of section 3.2. Geometrical constraints were set

841
by DIN EN 10210-2 and DIN EN 10219-2. To ensure local buckling, the models were all set to
a maximum length of l = 700 mm resulting from the condition of λglob.  0,15.
For the evaluation of the numerical models recently conducted tests on SHS200 profiles,
from an intern project RFCS No. 709892 Hollosstab, were used to compare the moment-
rotation-relations. The full-scale specimens were first 3D-scanned by using an optical 3D digi-
tizing system to reproduce the real geometry with all the initial imperfections in order to
implement it into a FE based simulation (Abaqus 2016). This reverse engineering procedure
was already tested and performed successfully in a previous project and presented partially in
(Müller 2018). The specimens were subsequently tested by applying an eccentric normal force
to achieve almost pure bending, see Figure 2b. Additional GMNIA calculations with an
imperfection amplitude of w0 = B/600 were conducted and compared with a sufficient accord-
ance concerning strength and additional elongation.

4 PARAMETRIC STUDY ON THE ROTATIONAL CAPACITY OF SQUARED


HOLLOW SECTIONS

4.1 Studied geometries and load cases


The parametric study investigates the rotation capacity of SHS 200 profiles out of cold-
formed and hot-rolled material and two different steel grades, S355 and S700, in each case.
Also different thickness values are applied, varying from 2 to 16 mm (Table 1). The length
was set constant by a value of l = 700 mm. Throughout the study, the models were all loaded
by a moment M at both ends of the specimens to achieve a constant moment gradient.

4.2 Effects of initial imperfections on the rotation capacity


A considerable effort was made through the decades to define parameters that effect the rota-
tion capacity. Saloumi (2016) gives an overview of quantification and sensibility factors
regarding the rotation capacity.
To show the impacts of initial imperfection on the rotation capacity, GMNIA calculations with
different eigenmodes were conducted. The imperfection amplitude for all calculated models was
equal to the value of B/200. Figure 3a & 3b are showing the moment-rotation relation for a hot-
finished SHS 200 profile with the thickness of t = 6,3 mm (Class 2) and t = 12,5 mm (Class 1). By
looking at Figure 3a one can identify a high scatter in the post buckling branch, while the rising
branch remains for all imperfection shapes mostly the same until the point of reaching the max-
imum moment capacity. Figure 3b shows a significantly different behavior. The post buckling

Table 1. Overview of geometric and material parameters.


Steel grades

Thickness S355 S700 Cold-formed Hot-rolled

2,0 mm X X X X
2,5 mm X X X X
3,0 mm X X X X
4,0 mm X X X X
5,0 mm X X X X
6,3 mm X X X X
8,0 mm X X X X
10,0 mm X X X X
12,5 mm X X X X
16,0 mm X X X X

842
Figure 3. a) SHS200, t = 6,3 mm, Class 2; b) SHS200, t = 12,5 mm, Class1; c) Value for the coefficient
of variation.

brunch remains unchanged, even at higher rotations leading to an overall lower imperfection
sensitivity.
To get a better concept of the imperfection sensitivity throughout the slenderness, the coeffi-
cient of variation (cov) was calculated for every examined thickness and plotted against the slen-
derness. Figure 3c is showing the results of the calculated cov’s as well as the corresponding
cross-section classes. As one would expect the cov for the thickest cross-section with t = 16 mm
is also the lowest, cov16mm = 0,40. Starting from this value towards a higher slenderness, where
it reaches its maximum for t = 2 mm, with cov2mm = 29,7. The values in-between rise degres-
sively for class 1 sections in the beginning, reaching a kink around class 2 sections and then rise
continuously with a lower inclination. The biggest difference between individual values can be
identified for class 1 sections. That means that with rising slenderness of compact SHS sections
the scattering degree of post buckling curves for different imperfection shapes increases

843
disproportionally. Wilkonson (1999) also found out that the magnitude of the imperfection had
an unexpectedly significant impact on the rotation capacity, especially for stockier sections.
To show that the assumption towards different imperfections and amplitudes within one speci-
men is not only a theoretical topic, an imperfection evaluation was done on SHS profiles. The
effective imperfections of the real specimen geometries were determined by using the previously
described scanning and reverse engineering procedure, see Section 3.3 and also (Müller 2018) An
exemplary outcome of the comparison is shown in Figure 4. Calculations according the real test
setup were done for each side of the two specimens. The outcome of the moment-rotation-
relations of the post buckling paths are shown in Figure 5a & 5b below. The paths are grouped
from z+) to y-), corresponding to the presentation in Figure 4. The rotation is read in each case at
the top (black lines) and the bottom (red lines) of the specimen, as the post buckling path is
highly dependent on the location of the local buckling area. If the local buckling field is assumed
at the top, the corresponding rotation of the top boundary will also be higher compared to the
rotation at the bottom. By looking at side z- of Figure 5a, buckling appeared in the upper area.
Therefore, rotation at the top boundary leads to a pronounced post buckling path with
a soft subside. However, plotting the rotation of the bottom boundary leads to a sharp kink in

Figure 4. Imperfection analysis of cold-formed SHS profiles based on 3D-scan data.

Figure 5. Moment-rotation-relations for each side the SHS200*8mm (S355) and SHS200*5mm (S550).

844
the post buckling path. The opposite is valid for side y+, where buckling appeared at the
bottom of the specimen.
Figure 5b is showing the moment-rotation-paths of a SHS200 (S550) profile with a thickness
of 5mm, again for every side of the specimen. Even though the shown profile isn´t reaching the
plastic rotation limit and thus less relevant for the analysis of rotations, it can be taken as
a good example for the analysis of initial imperfections on the post-buckling behavior. In com-
parison to the SHS200 * 8 (Figure 5a) the deviations are relatively low, due to the fact that
buckling occurred always around the middle area. This buckling case produces considerable
equal rotations at both ends of the boundaries. Therefore, initial imperfections are highly influ-
encing the location of the plastic hinge and the moment-rotation curve, respectively.

5 EVALUATION OF THE PARAMETRIC STUDY

Figure 6 is showing the evaluation plots of the moment-rotation-relations for hot-finished and
cold-formed SHS200 profiles, where the thickness t was varied as described in Section 4.1.

Figure 6. Slenderness dependent moment-rotation-relation for SHS200; a) S355 Hot-finished; b) S355


Cold-formed; c) S690 Hot-finished; d) S700 Cold-formed.

845
The normalized Moment M/Mpl, where Mpl = fy * Wpl, is plotted against the slenderness in
the x-axis and the normalized rotation φ / φpl in the y-axis.

6 SUMMARY, CONLUSIONS AND OUTLOOK

In the framework of this paper a numerical study on the influence of imperfections, derived
on the one hand from real geometry scans of specimens tested at the laboratory of the Bundes-
wehr University and on the other hand from GMNIA calculations based on different eigen-
mode imperfection shapes, was conducted. It has been shown that initial imperfections have
an influence on the post buckling path and also on the location of the plastic hinge. Therefore,
a constant moment gradient will not always produce the same rotations at the boundaries of
the specimen. An influence on the rising brunch could not be determined.
An additional parametric study was conducted to derive 3D-plots, which show the slenderness
dependent rotation of hot-finished and cold-formed SHS200 profiles, taking different thicknesses,
steel grades and corner geometries, dependent on the manufacturing process, into account.
Part of further investigations will be the development of equations which describe the full
moment-rotation-relation depending on the slenderness of squared, rectangular and cylin-
drical hollow sections. Including additional studies on the rotation capacity affected by
residual stresses (manufacturing process), the length of the specimens and its influence on
imperfections and also further load cases to achieve a variety in the moment gradients.

REFERENCES

Abaqus. Refernce manual, version 6.16. 2016. Simulia, Dassault Systéms. France.
Abvabi, A. & Rolfe, B. & Hodgson, P.D. & Weiss, M. 2015. The influence of residual stresses on a roll
forming process. International Journal of Mechanical Sciences 101–102: 124–136.
Amouzegar, H. & Schafer, B.W. & Tootkaboni, M. 2016. An incremental numerical method for calcula-
tion of residual stresses and strains in cold-formed steel members. Thin Walled Struct. 106: 61–74.
DIN EN 1993-1-1. 2005. Eurocode 3. Design of steel structures – Part 1-1: General rules and rules for
buildings. CEN – European Committee for Standardization. Brussels.
DIN EN 10210-2. German version EN 10210-2:2006. Hot finished structural hollow sections of non-
alloy and fine grain steels – Part 2: Tolerance, dimensions and sectional properties.
DIN EN 10219-2. German version EN 10219-2:2006. Cold formed welded structural hollow sections of
non-alloy and fine grain steels – Part 2: Tolerances, dimensions and sectional properties.
Gardner, L. & Saari, N. & Wang, F. 2010. Comparative experimental study of hot-rolled and
cold-formed rectangular hollow sections. J. of Thin-Walled Structures 48: 495–507.
Kato, B. 1982. Cold-formed Welded Steel Tubular Members. Applied Science Publishers, London and
New York.
Ma, J.-L. & Chan, T.-M. & Young, B. 2015. Material properties and residual stresses of cold-formed
high strength steel hollow sections. J. of Constructional Steel Research 109: 152–165.
Müller, A. & Hausmann, B. & Taras, A. 2018. Study on the influence of imperfections and strain harden-
ing on the buckling strength of spiral –welded aluminum circular hollow sections. Conference paper,
Eighth ICTWS 2018. Losbon.
Saloumi, E. 2016. Development of a new design method to define the rotation capacity of steel hollow
sections. Phd-Thesis. Liége University, Saint-Joseph University.
Tong, L. & Hou, G. & Chen, Y. & Zhou, F. & Shen, K. & Yang, A. 2012. Experimental investigation on
longitudinal residual stresses for cold-formed thick-walled square hollow sections. J. Constr. Steel Res.
73: 105–116.
Wilkonson, T. 1999. The Plastic Behaviour of Cold-Formed Rectangular Hollow Sections.
Yun, X. & Gardner, L. 2017. Stress-Strain curves for hot-rolled steels. Journal of Constructional Steel
Research.
Yun, X. & Gardner, L. 2018. Description of stress-strain curves for cold-formed steels. Construction and
Building Material 189: 527–538.
Zhang, X.Z. & Liu, S. & Zhang, M.S. & Chiew, S.P. 2010. Comparative experimental study of
hot-formed, hot-finished and cold-formed rectangular hollow sections. J. of Case Studies in Structural
Engineering 6: 115–129.

846
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

An analytical solution for the compressed simply-supported plate


with initial geometric imperfections

M. Nedelcu
Technical University of Cluj-Napoca, Romania

ABSTRACT: The analytical solution of the uniformly compressed simply-supported plate with
large deflections was developed almost a century ago by various authors using the condition of
minimum of the strain energy. The resulted ultimate load was not in satisfactory agreement with
the experiments. Von Karman’s semi-empirical approach considers that the ultimate load is
reached when the critical stress of the effective width is equal with the yield stress. Winter modified
the formula in order to match the experimental results, and his expression is currently used in the
majority of the CFS design codes. The paper presents an accurate analytical solution using the
initial approach but introducing different displacement functions. The initial geometric imperfec-
tions are taken into account and their effect on the effective width is properly quantified.

1 INTRODUCTION

It is a well-known fact that the thin plates have large post-buckling load-carrying capacity. To
estimate this prost-critical reserve and eventually to find the ultimate load-carrying capacity,
numerous analytical and experimental studies were conducted, starting almost a century ago.
A classic analytical solution of the uniformly compressed simply-supported plate with large
deflections (see Figure 1a) is presented in Timoshenko&Gere (1961) and apparently attributed
to Lahde&Wagner (1936). The given approximate solution determines the deflection of the
buckled plate from the condition of minimum of the strain energy.
The solution is given with or without lateral restraints and the distribution of the longitudinal
normal stresses presented in Figure 1b clearly reflects the stress concentrations near the edges.
This complex stress distribution can be simplified by replacing it with constant longitudinal
normal stress on an effective width formed by 2 strips along the longitudinal edges (Figure 1c).
Everything seems clear, but there are a few drawbacks. Even if other variations of this analytical
solution were developed, none of them provided ultimate loads in satisfactory agreement with
experiments and also, none of them took into consideration the initial geometric imperfections,
even if all researchers agree that their effect is extremely important for thin plates. Von Karman
et al. (1932) previously had a different approach: one takes the same equivalent system made by
the two strips of the effective width and then a new concept is introduced - that the ultimate load
of the plate is equal with the critical load of this equivalent system. By considering the yield value
fy for the critical longitudinal normal stress σcr,eff, the effective width beff is calculated:
rffiffiffiffiffiffi
beff σcr 1
¼ ¼ ð1Þ
b fy λp

where b is the plate width, σcr is the critical stress of the entire plate without imperfections, and λp
is the normalized slenderness. Even if this is approximation is quite “rough” (as described by
Timoshenko and Gere) it provided better results than the analytical solutions. Winter (1947)
improved the formula obtained by von Karman presumably considering geometric and mechanic

847
Figure 1. (a) Simply-supported plate; (b) Post-buckling normal stresses; (c) Effective width
(Timoshenko&Gere 1961).

imperfections on the base of a large series of tests on Cold-Formed steel beams. Winter’s equation
is given below:
!
beff 1 0:22
¼ 1 ð2Þ
b λp λp

This semi-empirical formula has no input regarding the geometric imperfections of the plate,
yet it is used by the majority of national codes for the design of CF thin-walled steel members
(e.g. EC3-Part 1.3, AISI Specification, etc.).

2 ORIGINAL ANALYTICAL SOLUTION

The analytical solution presented in Timoshenko&Gere (1961) starts with the following
approximate expressions for the displacements of the middle surface:

πx πy
w ¼ fcos cos
2b 2a
πy πx
v ¼ C1 sin cos  ey ð3Þ
a 2b
πx πy
u ¼ C2 sin cos þ αx
b 2a

where f, C1, C2, e and α are constants, and u, v, w are the displacements in x, y, z directions,
respectively. For the case of lateral rigid frame α = 0, for free lateral displacements α was found
from the condition that the sum of normal stresses σx along vertical edges is equal to zero. Using
the Kirchhoff-Love plate theory, the only non-zero strains and stresses are in the in-plane direc-
tions, and the following kinematic equations are considered together with the constitutive
relations:
 
∂u 1 ∂w 2
εx ¼ þ
∂x 2 ∂x

848
2 3 2 32 3
  σx 1 μ 0 εx
∂v 1 ∂w 2 4 σy 5 ¼ E 4μ 1
εy ¼ þ 0 5 4 εy 5 ð4Þ
∂y 2 ∂y τxy 1μ 0 0
2 1μ
γxy
2

∂v ∂u ∂w ∂w
γxy ¼ þ þ
∂x ∂y ∂x ∂y

The strain energy of the plate is calculated, having Membrane and Bending components:
Z Z 
h
U ¼ UM þ UB ¼ σx εx þ σy εy þ τxy γxy dxdy
2
Z Z " 2 2  2 2  2 2 # ð5Þ
1 ∂ w ∂w ∂2 w ∂2 w ∂w
þ D þ þ 2μ 2 2 þ 2ð1  μÞ dxdy
2 ∂x2 ∂y2 ∂x ∂y ∂x∂y

where h is the plate thickness, μ is the Poisson’s ratio and D is the plate bending stiffness.
For a given compression level of the plate (namely varying e) the constants C1, C2 and f are
found from the condition of minimum of the strain energy, namely:
∂U ∂U ∂U
¼0 ¼0 ¼0 ð6Þ
∂C1 ∂C2 ∂f

The ultimate load was calculated using the strength theory of the maximum shear stress. It
was assumed that the complete failure of the plate occurs when the maximum shear stress in
the middle plane of the plate reaches the yield stress.
As stated before, this analytical solution is not in satisfactory agreement with the experi-
ments nor with the numerical results provided by shell Finite Element Analyses (FEA). There
could be only one problem: the displacement field given in Equation 3. The author found that
for plate width/thickness ratios smaller than 100, the w expression agrees with the deflection
given by common shell FEA. But the u and v expressions are not in agreement (no matter the
width/thickness ratio).

3 NEW ANALYTICAL SOLUTION

Figure 2a presents the considered plate, the coordinate system, the load and boundary condi-
tions. For comparisons with the original solution one must keep in mind that the coordinate
system is now rotated with 90º, and its origin has been translated.
The new displacement field is given as follows:
   
πy 2πx Lx
u ¼ C1 þ α1 sin sin e x
Ly Lx 2
   
2πy Ly 2πx Ly 2πy
v ¼ C2 sin  α2 y  cos  α3 y  þ C3 sin
Ly 2 Lx 2 Ly ð7Þ
πx πy
w ¼ fsin sin
Lx Ly
πx πy
w0 ¼ f0 sin sin
Lx Ly

Figure 2. Simply-supported plate.

849
Figure 3. (a) v displacement (original solution); (b) v displacement (new solution).

where w0 is the initial geometric imperfection having the same shape as w but with the known
amplitude f0. Because of the change of the coordinate system, w seems different compared
with the original solution, but in fact it is the same. Figures 3a and 3b present the
v displacement in the original and present formulation, respectively. Only the present one
agrees with the displacements obtained from shell FEA.
Considering the geometric imperfection, the kinematic expressions become:
 
∂u 1 ∂w 2 ∂w0 ∂w
εx ¼ þ þ
∂x 2 ∂x ∂x ∂x
 
∂v 1 ∂w 2 ∂w0 ∂w
εy ¼ þ þ ð8Þ
∂y 2 ∂y ∂y ∂y

∂v ∂u ∂w ∂w ∂w0 ∂w ∂w0 ∂w
γxy ¼ þ þ þ þ
∂x ∂y ∂x ∂y ∂x ∂y ∂y ∂x

Returning to Equation 7, the constant e stands for the unit compression as in the original for-
mulation (actually it is the linear component of the normal strain εx). The constants α1, α2, α3
are found from the conditions that the normal and shear stresses σx and τxy are null on the
lateral edges having y = 0 and y = Ly. Their expressions are given below:

Ly
α1 ¼ α2
Lx 2

2C1 μπ 2C2 π f π2 ðf þ 2f0 Þ


α2 ¼ þ  ð9Þ
Lx Ly 4Ly 2

2C3 π f π2 ðf þ 2f0 Þ
α3 ¼  eμ þ
Ly 4Ly 2

The remaining unknows are the constants C1, C2, C3 and f. The condition of minimum of the
strain energy is next considered:

∂U ∂U ∂U ∂U
¼0 ¼0 ¼0 ¼0 ð10Þ
∂C1 ∂C2 ∂C3 ∂f

The first 3 equations provide the expressions of C1, C2, C3 as functions of f- the w displacement
amplitude. The last equation provides a cubic equation of the form:
 
a1 f 3 þ a2 f 2 f0 þ f a3 e þ a4 þ a5 f02 þ a6 f0 e ¼ 0 ð11Þ

850
where ai are the coefficients of the cubic equation which have rather complicated (yet analytical)
expressions depending on Lx, Ly, E, μ and h. The varying input parameter in the above equation is
e (the linear component of the normal strain εx). One can start with zero value for e and gradually
increase this compression parameter. For each value of e, the smallest root of the cubic equation
provides the w displacement amplitude f.
The critical point for the plate without imperfections can be found if one considers f0 = 0,
then the cubic equation becomes:
 
f a1 f 2 þ a3 e þ a4 ¼ 0 ð12Þ

In order to have real and non-zero solutions for f, the following condition must be satisfied:

a3 e þ a4 50 ð13Þ

At the limit, the above condition provides the critical point and the corresponding critical
stress:

a4 E
ecr ¼  σx;cr ¼ ecr ð14Þ
a3 1  μ2

As stated before, an incremental procedure is used varying the compression parameter e, and
the state of stress is checked at each increment. The von Mises strength theory is used to find
the “maximum” load, assuming as in the previous solutions (von Karman’s as well) that the
failure is associated with the reach of the yield point. This is an approximation that generally
underestimates the strength capacity, the shell FEA proved that the ultimate load could be
significantly higher than the “first yield” load, when the amplitude f0 of the geometric imper-
fections is above 50% of the thickness.
The new analytical solution was validated by comparing the 1st yield load with the one pro-
vided by shell FEA. The differences are satisfactory (less than 5% in almost all cases) if the
following restrictions are considered:
– the plate slenderness (width/thickness ratio) is less than 100. For higher values, the assumed
displacement field is no longer similar with the one detected in shell FEA. The use of trig-
onometric series with more than one term seem to be the appropriate approach, and it’s
currently under work.
– the ratio between the edges lengths Lx/Ly is in the range (1, √2). If it is higher than √2, two or
more halfwaves of the w displacement may appear in x direction (see again Timoshenko&-
Gere (1961)), unless a large value of the geometric imperfection will counteract this well-
known buckling behavior. The same ratio can be less than 1 but for small values, again
a higher number of buckling halfwaves may be triggered in y directions and it must be men-
tioned that this last case has no practical use for the study of thin-walled bars.

4 ILLUSTRATIVE EXAMPLE

The new analytical formulation is now applied to the Geometric Nonlinear Analysis with Imper-
fections (GNIA) of a square plate with the following characteristics: dimensions Lx=Ly=100 mm,
modulus of elasticity E = 210 GPa, Poisson ratio μ = 0.3, the yield stress fy = 350 MPa. Three
values are taken for the plate thickness h: 2, 1.5, 1 mm. For each case, 7 values of the geometric
imperfection amplitude f0 were considered in the range 1. . .100% of the plate thickness h. The
analytical formulation was applied using Matlab (2005) and a public software application is avail-
able at http://users.utcluj.ro/~mnedelcu/AnSolBuckledPlates.htm.
For validation, numerical analyses using S4 shell elements were performed in Abaqus
(2002). The in-plane size of the shell elements was set as twice the plate thickness. Figure 4

851
presents the state of stress in the middle plane at 1st yield for the case h = f0 =1mm: von Mises
stresses, normal stresses in x and y directions, and the shear stress, all based on the proposed
analytical formulation. The same state of stress was found using shell FEA.
Table 1 presents the GNIA results and interesting conclusions can be drawn.
1. The new analytical formulation is in very good agreement with the shell FEA regarding the
axial load at the 1st yield point. The only case when the differences are above 5% is for rela-
tively stiff plates (Ly/h = 50) with very small geometric imperfections. In both formulations
the 1st yield point is considered via the von Mises stresses in the middle plane of the plate.
But the yield first appears at the exterior plate surface which is most compressed, and it
spreads gradually in all directions, decreasing the strength capacity. Usually this phenom-
enon doesn’t have a significant effect for slender plates, but it appears that for this case it
can not be neglected. This phenomenon also caused for the first 2 cases presented in Table 1,
the reach of the maximum load before the appearance of the 1st yield in the middle plane.
2. The geometric imperfections have a strong influence in reducing the strength capacity espe-
cially for the stiff plates Ly/h = 50: almost 50% in strength reduction with regard the 1st
yield point if the imperfection amplitude is equal with the plate thickness f0 = h. The ero-
sion is not so large for slender plates Ly/h = 100 (around 15%) but in some cases the imper-
fection amplitude could be even larger than the plate thickness.
3. Table 1 presents the maximum compression load given by shell FEA and the post 1st yield
strength reserve. It can be concluded that there is a favorable effect of the geometric imper-
fections -larger they are, the post 1st yield strength reserve increases. For slender plates h =
1mm, this effect leads to a reduction of only 6% for the maximum load when we have f0 =
h in comparison with f0 = 0.01h. Since the maximum load is usually detected in experi-
ments, this could explain the good agreement for slender plates between the Winter’s for-
mula and the testing results, regardless the inherent geometric imperfections.
4. In the last three columns, Table 1 presents a comparison with respect the effective width cal-
culated using the Winter’s formula (see Equation 2) and the new analytical formulation.
Large differences are encountered, only 30% of cases present differences less than 5%. For
the stiff plate case (Ly/h = 50) with large imperfections, Winter’s formula overestimates the
plate strength capacity up to 41%, demanding a fundamental reevaluation of the application

Figure 4. State of stress in the middle plane at 1st yield for h = f0 =1mm.

852
Table 1. GNIA results.
h f0 P1st yield Pmax Post 1st yield Effective width
reserve
shell
analytical FEA Difference Winter analytical Difference
[mm] [%h] [kN] [kN] [%] [kN] [%] [mm] [mm] [%]

1 62.09 58.30 6.51 58.78 0.82 88.71 19.84


5 58.76 56.16 4.63 56.16 0.01 83.94 13.40
10 55.48 53.73 3.25 54.02 0.54 79.25 7.06
2.0 25 48.11 48.02 0.18 49.82 3.75 74.02 68.73 7.15
50 39.84 41.48 3.95 45.40 9.44 56.92 23.11
75 34.33 34.83 1.43 42.51 22.06 49.04 33.74
100 30.52 30.86 1.19 40.25 30.32 43.60 41.10
1 32.44 31.29 3.66 31.92 2.00 61.79 4.57
5 31.72 30.79 3.03 31.61 2.68 60.42 2.25
10 30.80 30.15 2.15 31.26 3.67 58.67 0.71
1.5 25 28.50 28.55 0.18 30.30 6.13 59.09 54.28 8.13
50 25.27 25.66 1.49 28.91 12.70 48.14 18.53
75 22.84 23.86 4.29 27.72 16.19 43.50 26.38
100 20.97 20.94 0.13 26.72 27.58 39.94 32.40
1 14.59 14.01 4.12 15.40 9.94 41.68 0.22
5 14.50 13.97 3.79 15.37 10.02 41.43 0.84
10 14.34 14.04 2.12 15.32 9.12 40.96 1.95
1.0 25 13.91 13.65 1.87 15.17 11.13 41.78 39.74 4.88
50 13.24 13.39 1.10 14.90 11.25 37.84 9.41
75 12.64 12.76 1.01 14.70 15.15 36.10 13.58
100 12.14 11.89 2.08 14.46 21.65 34.68 16.99

of the Winter’s formula in design. One could argue that such large imperfections are not
allowed in practice, yet even for f0 = 0.01h, the Winter’s formula underestimates the plate
strength capacity with almost 20%.

5 CONCLUSIONS

A new analytical solution is presented for the compressed simply-supported plate with large
deflections. The formulation considers a new displacement field, different to the one initially
proposed by Lahde&Wagner (1936). The initial geometric imperfections are taken into consid-
eration and their effect on the 1st yield load, the maximum load and the effective width is quan-
tified. The illustrative examples proved the good agreement between the proposed
formulation and the shell FEA. Also these examples reveal the limitations of the Winter’s
semi-empirical formula and the need of a reevaluation of its application in practical design.

REFERENCES

Timoshenko, S.P. and Gere, J.M. 1961. Theory of elastic stability. 2nd Ed. New York, McGraw-Hill.
Lahde, R. and Wagner, H. 1936. Versuche zur Ermittlung der mittragenden Breite von verbeulten
Blechen. Luftfahrt-Forsch, 13: 214–223.
von Karman, T., Sechler, E.E., Donnell, L.H. 1932. Strength of thin plates in compression, ASCE Trans-
actions, vol. 54, APM-54-5.
Winter, G. 1947. Strength of thin steel compression flanges. ASCE Transactions, Paper No. 2305, Trans.,
112, 1.
Matlab 2005. Version 7. 1.0246 Documentation, The Mathwork Inc.
ABAQUS Standard (Version 6.3), 2002, Hibbit Karlsson and Sorensen Inc.

853
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Experimental study on the shear connections of composite girders


with trapezoidally corrugated web

G. Németh, B. Jáger, N. Kovács & B. Kövesdi


Department of Structural Engineering, Budapest University of Technology and Economics (BME), Budapest,
Hungary

ABSTRACT: Composite bridges made by steel beams and reinforced concrete slabs are the
most economical choice for medium spans (40-60 m). Most commonly, headed studs welded to
the upper flange of the beam are used as shear connector. Although the structural behavior of this
type of shear connector is well known, the research of the last decades showed that connections
made by concrete dowels can be more economical and have more favorable behavior. The applica-
tion of corrugated web beams is also popular in composite structures, they can be applied with
and without steel flanges beside the concrete slab. If no steel flange is used, the web is directly
embedded into the concrete slab. This layout is applicable in both composite (upper concrete slab)
and hybrid (upper and lower concrete slabs) structures. Experimental research program having
large number (57 specimens) of full-scale push-out tests were performed in the Structural Labora-
tory of the BME Department of Structural Engineering. The test program contained specimens
with embedded web and specimens having steel flanges as well. The series of experimental tests
made it possible to compare specimens with embedded corrugated web and specimens with upper
steel flange with the aim to investigate the influence of the steel flange on the structural behavior
of the connection resistance and deformation capacity. The current research program has the aim
to experimentally investigate the structural behavior of shear connectors having concrete dowels
or horizontal headed studs in steel-concrete composite girders with trapezoidally corrugated webs.

1 INTRODUCTION

Corrugated steel profiles are widely used structural elements for a long time, but it has only been
used worldwide since two decades as beam webs. The corrugated profile has numerous advanta-
geous properties borrowing positive effects on the structural behavior and resulting in savings in
material, production and construction costs. One of its great advantage is that the corrugation has
a low stiffness in longitudinal direction, so in case of pre-stressing the forces generated by creep and
shrinkage do not cause significant increase in tension forces. The shear buckling resistance of the
web against global and local buckling increases, so the number of stiffeners can be significantly
reduced. In addition, compared to flat web girders, high bending stiffness is obtained in transverse
direction, which makes possible to reduce the number of cross girders (Kuhlmann & Raichle 2011).
The most commonly used shear connectors are the well-known headed studs, however, the
application of corrugated web requires new type of shear connectors, which raises a number of
issues to be clarified in structural design. Several bridges have already been built with new, innova-
tive shear connections, where the design is mainly executed by using experimental methods. Japan-
ese Nakano Bridge is an excellent example for bridges using corrugated web and top steel flange,
where a single perfobond connection was welded to the flange (Raichle J. 2015). Hondani Bridge
located also in Japan, is an example for hybrid structures where transverse rebars are used guided
through the holes cut on the corrugated web. There are also longitudinal rebars applied welded to
both sides of the corrugated web (Raichle J. 2015). In addition to the examples described above,
there are many other different shear connector types used between the corrugated web and con-
crete slab. Research of embedded type connectors was started in 2000 by the work of Nakasu. The

854
out-of-plane bending resistance of embedded connections was studied by experiments and finite
element analysis (Nakasu et al. 2000). The investigation of corrugated hybrid structures started in
2006 by the work of Kosa. Their experiments revealed that the failure of the connection occurs
due to opening between the concrete slabs and the steel web (Kosa 2006). In 2010, Röhm and
Novák tested the out-of-plane bending moment resistance of concrete dowels on an experimental
basis (Röhm and Novák 2010). In 2011, a study was carried out to investigate the corrugated per-
fobond type connectors with transverse rebars by Kim et al. According to their results design pro-
posal was developed for the analyzed shear connector type (Kim et al. 2011).
The current research program has the aim to study the structural behavior of shear connectors
having concrete dowels or horizontal headed studs for composite structures using experimental
tests. This article deals with the so-called concrete dowel shear connectors, which consist of open
cutouts on the corrugated steel web embedded in the concrete slab. In the paper shear connectors
with horizontally positioned headed studs are also studied with the aim to investigate the influ-
ence of parameters such as stud diameter on the structural behavior and shear resistance.

2 DESIGN PROPOSALS FOR SHEAR CONNECTORS USING CORRUGATED WEBS

EN 1994-1-1 only proposes design methods and construction layouts for shear connections
with vertically positioned headed shear studs welded on steel flanges. Annex C of EN 1994-2
provides design formula for embedded type, horizontally positioned headed studs, but only
for flat web girders. Currently, Eurocodes do not provide design recommendations and pro-
posed layouts for embedded shear connections with innovative corrugated web. Therefore,
design of these types of shear connectors require additional research work. In 2009, a large
series of laboratory tests are carried out on 22 test specimens (Figure 1) at the University of
Stuttgart to study the shear connections for corrugated web girders. Four test specimens with
no additional connections, 11 specimens with horizontally placed headed shear studs and 7
specimens with concrete dowels with transvers rebars were tested (Kuhlmann et al. 2011).
Based on the results of the experimental and additional numerical calculations, two design
recommendations were developed to determine the shear capacity of embedded type shear
connectors using corrugated webs. According to (Novák & Röhm 2009) the shear capacity of
embedded connection with concrete dowels can be calculated from Equation 1.

PRd;L;NR ¼ 2L1w γ ðh  a3 Þ0;15 t0;55 0;35 0;45 0;10


w tE fck as ð1 þ μÞ
1;50
ð1Þ
v

where Lw = projection length of the corrugation profile (2a1+2a4, see Figure 2), γv = partial
factor (equal by 1.25), h = thickness of the concrete slab, a3 = horizontal projection of the
angled plate, tw = thickness of the corrugated web, tE = embedding depth, fck = characteristic
compression strength of concrete, as = cross-sectional area of rebars guided through the
dowels and μ = coefficient of friction between concrete and steel.

Figure 1. Experimental layout of push-out tests at University of Stuttgart (Kuhlmann & Raichle 2011).

855
Figure 2. Notations of the corrugated web according to the Eurocode.

Raichle and Kuhlmann developed a design equation (Raichle & Kuhlmann 2015) for the
estimation of the shear capacity of embedded shear connections with horizontal headed studs
given by Equation 2.

11500 0;6 0;35 0;2 0;05


PRd;L;RK ¼ kb ks kα ka1 fck tE tw asw ð1 þ asa Þ0;05 ð2Þ
L w  γv

where kb = coefficient depending on the support conditions, ks = coefficient depending on the


position of headed studs, kα = coefficient depending on bending angle of the corrugated pro-
file, ka1 = coefficient depending on the plate width (see Figure 2), asw = cross-sectional area of
stirrups and transverse reinforcement, asa = cross-sectional area of additional transverse
reinforcement above the corrugated web. The exact determination of the coefficients can be
found in (Raichle & Kuhlmann 2015). It has to noted that the experimental layout applied by
Raichle and Kuhlmann is unique because the standard does not contain instructions for
embedded type specimens. The strategy of the test program follows the main features of trad-
itional push-out test extended by several specialties coming from the structural behavior of the
corrugated profile. However, the number of experiments is relatively small compared to the
large number of parameters investigated in the test program, therefore, additional experimen-
tal results are required for better understanding of the structural behavior of these shear con-
nection layouts. The formulas above were developed for embedded type connections and the
effect of optional steel flanges was not tested. The experimental results presented in this article
will be compared to the values obtained with the formulas given by Equations 1 and 2.

3 THE PUSH-OUT TESTS

3.1 Test specimens


Experiments are performed in the Structural Laboratory of BME Department of Structural
Engineering in 2017-2018. A total of 57 specimens were tested, of which 14 had steel flanges
and 43 specimens were manufactured using embedded type shear connectors. The general aim
of the project was to investigate the structural behavior of different shear connectors, their
resistances and specialties coming from the corrugated web. The current article presents the
results of fifteen specimens; nine of them are used with embedded steel web (T), and six with
top flanges (H). In the experimental research program the trapezoidal profile (30, 45, 60 are
the angles in degree) and the shear connector type (R - reference, HS - headed stud, CD - con-
crete dowel) has been changed. The investigated corrugated steel profiles and the shear con-
nectors are shown in Figure 3; the geometry of the test specimens are shown in Figure 4.
Table 1 shows the analyzed parameters of specimens. The ‘type’ column next to the speci-
mens’ sign shows that the trapezoidal profile is embedded (no flange) or welded to the top
flange. The ‘angle of corrugation’ shows the angle between the corrugated plates and the axis
of the beam in degrees. The ‘additional shear connector’ shows whether the specimen is pro-
vided with other mechanical connectors besides the trapezoidal profile embedded in the con-
crete, and if so, it shows the type of connectors.
In case of specimens with flanges (H) the corrugated profile was welded to a HEM300 beam.
The headed studs and concrete dowels are placed to the center of the web folds parallel to the
edges of the concrete slab, 50 mm from the edge of the corrugated profile (see Figures 3 and 5).
Some specimens are prepared which, apart from the embedded corrugation, did not contain

856
Figure 3. Web profile and geometry of headed studs and cutouts.

Figure 4. Geometry of test specimens (T45-R and H45-R).

any other shear connector (R) and was used as a reference to compare the results using shear
connectors. The surfaces of the steel element that come into contact with the concrete have
been greased in order to allow the separation of the different materials. The concrete slabs were
made in two casting times, placed in horizontal position with two days apart. In the concrete
slab rebar with 10 mm diameter are used. There is a distance of 70 mm between the edges of the
concrete slab and the corrugated steel profile. At least 3 test cubes were made for each casting
to determine the concrete strength. Tensile coupon tests are also performed to determine the
tensile strength of the corrugated steel profile. The average compression strength of the concrete
is about 40 MPa, the average tensile strength of the steel web is 367 MPa.

3.2 Experimental layout and procedure


The experimental layout and the test specimens are shown in Figure 6. During the experiments,
the applied load and the slip between the corrugated steel web and the concrete slab are deter-
mined to analyze the initial stiffness, ductility and shear capacity of the connections. In case of
embedded specimens, three linear variable displacement transducers (LVDT) are placed to the
web, in case of specimens with flanges, two LVDTs are used. The load is produced by a hydraulic

857
Table 1. Analyzed parameters of specimens.
Specimen Type Angle of corrugation [°] Additional shear connector

T30-R embedded 30 -
H30-R with flange 30 -
T30-HS embedded 30 headed stud
H30-HS with flange 30 headed stud
T30-CD embedded 30 concrete dowel
T45-R embedded 45 -
H45-R with flange 45 -
T45-HS embedded 45 headed stud
H45-HS with flange 45 headed stud
T45-CD embedded 45 concrete dowel
T60-R embedded 60 -
H60-R with flange 60 -
T60-HS embedded 60 headed stud
H60-HS with flange 60 headed stud
T60-CD embedded 60 concrete dowel

Figure 5. Concrete dowels cut to the center of the plates parallel to the edges.

jack with a maximum loading capability of 6000 kN. The jack transferred the load to the rigid
steel plate at the top of the test specimens, or in case of specimens with flanges, directly to the end
of the HEM300 beam. In addition, three steel rods are placed at the bottom of the concrete slabs
to carry the horizontal force due to the eccentricity of the load.
Based on the load protocol of EN 1994-1-1 Annex B, the test specimens are cyclically
loaded to a level corresponding to 40% of their estimated load bearing capacity for breaking
the adhesion between concrete and steel surfaces, then the load is applied statically until reach-
ing the ultimate load and failure of the specimen.

4 TEST RESULTS

4.1 Failure modes


Two typical failure modes are observed during the tests. In the case of the reference specimens
(T-R and H-R types), due to the absence of shear connectors suitable for carrying transversal

858
Figure 6. Experimental layout and geometry of test specimens.

forces, the failure is caused by transverse bending of the concrete slab. Figure 7 shows that the
test specimens are opened since the inclined folds of the web generate horizontal forces to the
concrete slab. In the case of specimens with additional shear strengthening, the horizontal
forces are carried by the transverse rebars and the concrete slab could not open. Thus, the
concrete slabs crumbled as a result of vertical shearing transferred from the web. Figure 8
illustrates this failure mode, which also shows the crack of a headed stud welded to the web.
It has been observed that, in the case of the reference specimens, the amount of concrete crum-
bling increases with the increase of the angle of trapezoidal plate, since the horizontal projection of
the contact surface is larger. Thus, the ratio of vertical shear and horizontal forces are larger, too.

4.2 Force-slip relationship


Figure 9 shows the typical force-slip diagrams of specimens having corrugation angle of 45°
and Figure 10 presents the force-slip curves of specimens having corrugation angle of 30° and
60 (also it shows T45-R for comparison).

Figure 7. Failure mode of test specimens – transverse bending.

859
Figure 8. Failure of T-HS type test specimens – concrete crushing.

Results presented in Figure 9 show that the initial stiffness of specimens with steel flange
(H-R and H-HS) is larger than for the embedded type specimens (T-R, T-HS and T-CD). Test
results also proved that the shear connectors have no significant effect on the initial stiffness (com-
pared H-R and H-HS type specimens), which means that the initial stiffness is mainly influenced
by the corrugation profile and connection type (embedded or joint using flange). In addition, spe-
cimens with headed studs (T-HS and H-HS) are less ductile, with more significant resistance deg-
radation (steeper slope after the maximum point) than embedded type connections. Concrete
dowel shows ductile behavior. Sudden jumps on T-HS and T-CD curves are traces of headed stud
and rebar cracks. The figure clearly shows how much the flange increases the capacity; the refer-
ence specimen (H-R) has higher shear capacity than the embedded one with headed studs (T-HS).
Results presented in Figure 10 proves that the curves of the 30 and 60 degree specimens
show similar characteristics, shear connectors used on corrugated web generally show ductile
behavior (see HS and CD) and corrugation angle does not result in obvious change in the
measured force-slip curves.

4.3 Comparison of experimental and analytical results


Table 2 summarizes the test results and the calculated shear capacities based on the proposed
design methods (Equations 1 and 2). It should be noted that dimensions of test specimens
were out of the range of applicability of the formulas, so the calculated results merely illustrate
that the formulas can actually be developed to be applicable in the analyzed range.
Comparing the results of reference specimens, the supporting effect of the top flange increased
the shear capacity by 31-36% (T-R and H-R, 30°-45°-60°). Use of headed studs in the case of

Figure 9. Experimental force-slip curves of specimens with 45° angle corrugation.

860
Figure 10. Experimental force-slip curves of specimens with 30° and 60° angle corrugation.

embedded connection increased the capacity by ~51-25-8% (T-R and T-HS, order: 30°-45°-60°),
in the case of test specimens having top flange increased by ~33-32-9% (H-R and H-HS, order:
30-45-60). The concrete dowel increased the shear capacity by ~46-21-14% for the embedded cor-
rugated connection (T-R and T-CD, order: 30°-45°-60°). It can be seen that the effect of the top
flange is roughly constant (~33%). The increasing effect in the ultimate capacity of the additional
shear connectors decreases by increasing the angle of trapezoidal corrugation. This can be
explained by the increase in the ratio of vertical shear and horizontal forces transmitted by the
inclined panels, since the smaller the opening of the concrete slab, the smaller the extent to which
the additional connectors can contribute to the ultimate capacity. The results of the embedded
specimens show that headed studs and concrete dowels shown in Figure 3 increased the load bear-
ing capacity to a similar extent. For embedded reference specimens, the increase of corrugation
angle increased the load bearing capacity by ~40% (T30-R and T45-R) and ~60% (T30-R and
T60-R). For the reference specimens with top flange, the capacity increased with ~39% (H30-R
and H45-R) and ~63% (H30-R and H60-R), respectively. For test specimens using headed studs,
the ultimate load increase is ~9% (T30-HS and T45-HS), ~26% (T30-HS and T60-HS), ~37%
(H30-HS and H45-HS) and ~49% (H30-HS and H60-HS), respectively, while for concrete dowels,
the increase is ~12% (T30-CD and T45-CD) and ~37% (T30-CD and T60-CD). Comparing the

Table 2. Ultimate shear capacities.


Specimen* Ptest [kN] PRk.L.NR [kN] PRk.L.RK [kN] Ptest/PRk.L

T30-R 819 - 807 1.01


H30-R 1199 - - -
T30-HS 1656 - 1684 0.98
H30-HS 1798 - - -
T30-CD 1512 2398 - 0.63
T45-R 1364 - 811 1.68
H45-R 1963 - - -
T45-HS 1814 - 1692 1.07
H45-HS 2875 - - -
T45-CD 1726 2525 - 0.68
T60-R 2064 - 726 2.84
H60-R 3236 - - -
T60-HS 2250 - 1515 1.49
H60-HS 3538 - - -
T60-CD 2392 2604 - 0.92

* H: with flange, T: embedded web, 30-45-60: angle of trapezoidal corrugation in degree, R: reference,
HS: headed studs, CD: concrete dowels.

861
results of specimens having corrugation angle of 30°, 45° and 60°, it is clear that with increasing
angles, the load bearing capacity increases as well. This effect can obviously be explained by the
increased horizontal projection of the load-bearing surface (inclined plates). For embedded and
top flange reference specimens, the rate of increment is similar, from 30 to 45 degrees ~40%, from
30 to 60 degrees ~62%. The effect of increasing the angle of corrugation on the ultimate shear
capacity can also be observed in the case of test specimens with additional shear connectors,
although it is clearly smaller than in the case of reference specimens.

5 CONCLUSIONS

The article presents the results of an experimental research program investigating the struc-
tural behavior of shear connections of composite girders using trapezoidally corrugated web.
The research aim is to determine the structural behavior of shear connections and their impact
on the ultimate shear capacity by push-out tests. The results of fifteen specimens are presented
in the current paper; all specimens had similar parameters, differences are only in the angle of
trapezoidal corrugation, and in the presence of top flange and shear connectors. Based on the
test results, the following conclusions are drawn:
(i) supporting effect of steel flange has considerable increasing effect on shear capacity,
(ii) the higher the angle of inclination of the trapezoidal profile, the higher the shear capacity,
(iii) both horizontal headed studs and concrete dowels increase the ultimate shear capacity,
around the same extent, however, concrete dowel connectors show more ductile behavior,
(iv) additional shear connectors – which can carry horizontal loads – influences the failure
mode and ultimate shear capacity,
(v) without additional shear connectors failure is caused by transverse bending, with add-
itional connectors the failure mode is concrete crumbling resulting higher shear capacity.

ACKNOWLEDGEMENT

The research program is part of the “BridgeBeam” R&D project No. GINOP-2.1.1-15-2015-
00659, the financial support is gratefully acknowledged. Through the second and fourth authors
the paper was also supported by the ÚNKP -18-3-III. and ÚNKP-18-4 New National Excel-
lence Program of the Ministry of Human Capacities and by the János Bolyai Research Scholar-
ship of the Hungarian Academy of Sciences; the financial supports are gratefully acknowledged.

REFERENCES

Kim, S.H., Ahn, J.H., Choi, K.T., Jung, C.Y. 2011. Experimental evaluation of the shear resistance of
corrugated perfobond rib shear connections. Advances in Structural Engineering: 14(2),249–263.
Kosa, K, Awane, S, Uchino, H, Fujibayashi, K. 2006. Ultimate behavior of prestressed concrete bridge
with corrugated steel webs using embedded connection. Proc. JSCE: 62(1),202–220.
Kuhlmann, U. et al. 2011. New Developments of Steel and Composite Bridges. Proceedings of the 7th
National Conference of Steel Structures, (18–38). Volos, Greece.
Kuhlmann, U. & Raichle, J. 2011. Headed studs close to the concrete surface - fatigue behaviour and
application. Composite construction in steel and concrete VI (26–38). Tabernash, Colorado: ASCE.
Nakasu, K, Yoda, T, Sato, K. 2000. Study on out-of-plane bending of concrete dowels in a composite
girder with corrugated steel web. Proc. JSCE: 647(I–51), 267–279.
Novák, B. & Röhm, J. 2009. Anwendung von Trapezblechstegen im Brückenbau - Längsschubtragverhal-
ten von Betondübeln in Kombination mit Trapezblechstegen. Beton- und Stahlbetonbau: 104(9),562–569.
Raichle, J. 2015. Randnahe Kopfbolzen im Brückenbau (PhD dissertation). University of Stuttgart.
Raichle, J. & Kuhlmann, U. 2015. Trapezblechstege im Verbundbau – Längsschub und Querbiegung der
Verbundfuge mit randnahen Kopfbolzen. Stahlbau: 84(10),763–770.
Röhm, J. & Novák, B. 2010. Querbiegetragverhalten von Betondübeln bei Verbundtragwerken mit Tra-
pezblechstegen. Beton- und Stahlbetonbau 105(3),176–185.

862
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Numerical study of end-plate steel connections with two and four


bolts-per-row

D.L. Nunes & A. Ciutina


Politehnica University of Timisoara, Romania

ABSTRACT: The classic technical solution of beam-to-column connections with extended


end-plate with two bolts-per-row has a well-established behaviour and its design is finely
defined in the Eurocodes. It has, however, several limitations in situations where there is
a depth limit for flooring or in case of extreme actions. As a solution to these particular cases,
a configuration with four bolts-per-bolt may provide the necessary characteristics in order to
improve the performance of the connection and therefore, of the entire structure. Several stud-
ies already have shown that the classic T-stub approach may not accurately describe the
behaviour of such connections, as the failure modes do not follow a row-by-row progression.
In this context, a set of relevant numerical models of the tensioned area of the connection (the
first two bolt-rows, representing the tensile macro-component) was developed for both solu-
tions with two and four bolts-per-row, including parametric variation for end-plate thickness,
bolt diameter and flange width. The FEM analysis aims to reveal the change in the connec-
tion’s behaviour with the variation of the different parameters.

1 INTRODUCTION

The strength, stiffness and ductility of steel moment-resisting structures, are greatly influenced
by the configuration and geometry of the beam-to-column connections. Although such type
of connection is well depicted and exhaustively studied in their classic configuration with two
bolts-per-row, it also brings several limitations especially regarding its post-flexural behaviour,
for which the development of adequate ductility can be hard to achieve. Connections with
four-bolts-per-row instead of two could prove to be a feasible solution for these issues, where
high robustness is a requirement. However, the modern steel norms do not provide guidelines
for the design of connections with this configuration.
Although several researches have investigated this alternative configuration, the approach
has been greatly similar to the classic two-bolt per row configuration (Demonceau et al., 2010,
Couchaux et al., 2015, Kozolowski and Pisarek, 2008) and (Massimo et al., 2014), for which
each bolt-row is compared to a simple T-stub element. This approach could be inefficient in
explaining the four bolts-per-row configurations, for which there exist an interaction between
bolts of different rows (Nunes et al., 2018). In a general extended steel end-plate configur-
ation, the T-stub should integrate both the top flange and the beam web, thus the tensile
macro-component playing the essential role in the end-plate connection.

1.1 Macro-component
The approach of compiling the two top tension bolt-rows into a single macro-component has
been applied in previous studies with positive results (Ciutina, 2003). The advantage of these
elements is in taking into consideration both bolt-rows by considering the influence of the
flange and web, and the interaction between bolt-rows.
In an extended study on beam-to-column connections, Ciutina (2003) has reported the
results of several tensile macro-component tests, from a two bolts-per-row configuration, by
considering the tension area of a larger connection (Figure 1). The portion where problems of

863
Figure 1. Two-bolts-per-row macro-component extracted from a beam-to-column connection.

brittle failure may appear was isolated and a macro-component could be identified. This was
retrieved from the larger connection in order to study uniquely the part of the connection in
tension. The macro-component is composed by the flange of the beam and a portion of the
web, adjacent to the interior bolt-row. The tensioned macro-components were subjected to
monotonic and cyclic testing with various parameters and the results showed a good correl-
ation between experimental testing and the analytic results. In these tests, it was observed the
inverse proportionality between strength and ductility, and the combined effect of flexure and
traction (Figure 1).
Another study (Nunes and Ciutina, 2019), which investigated different configurations of
four-bolts-per-row connections, emphasizes that that the supplementary bolts can play an
important role in the robustness of a structure, by being able to regain an important percent-
age of the ultimate force after the first failure thus enhancing the ductility and post-flexural
behaviour in relation to the classic two-bolts-per-row configurations. The same study also sug-
gests that the failure mechanism for this type of connection may not be well described in
a row-by-row failure, but there is an interaction between different rows which can lead to
more complex patterns of failure, in which multiple rows are affected simultaneously due to
bolt coupling between rows.

2 REFERENCE STUDY

The structural robustness is recently highlighted in many studies and describes the capacity of
a structure to sustain localized damage, imposed by extreme actions, whilst avoiding the pro-
gressive collapse of the entire building. In this context, the research program FRAMEBLAST
(Dinu et al., 2017) was set up in order to assess the structural behaviour of a real-size steel
structure to an explosive blast. The study consists of a sequence of tests on structures and sub-
structures exposed to blast force in different layouts, which are then studied in order to assess
the blast effects. The results obtained until now show that an element such as a column can be
easily damaged to the point where it cannot carry loads, leaving the still-standing structure to
cope with what can be considered the complete removal of a column.
The reference building is a steel-frame structure with two 2,5m stories, two 5m spans and
two 3m bays (Figure 2). The stability is assured in both directions: transversally by moment
resisting frames and longitudinally by a centric bracing system. The beam-to-column connec-
tions are assured by bolted extended end-plate (rigid) in the transversal directions and fin
plate connection (pinned) in the longitudinal direction. Also, the connection between the col-
umns and the foundations is fully rigid. The columns are made using HEB260 profile, and the
main and secondary beams are made with IPE300 and IPE200, respectively. All steel elements
are made from S275JR.

864
Figure 2. Section and horizontal plan of the structure.

The structure test consists in exposing one column (outer transversal frame, central column)
to a blast designed to render the element unable of carrying loads. Thus, the affected column
is removed and the assembly is analysed for its capacity to rearrange the unloading paths in
order to avoid collapse, and for the advantages that a four-bolts-per-row connection bring to
the behaviour of the damaged structure.

3 NUMERICAL STUDY

3.1 Connection configuration


The initial configuration of the beam-to-column rigid connection consists of an end-plate of
20 mm thickness with 5 bolt-rows of M20 bolts, HR 10.9 in a two-bolt per row configuration.
Their geometric disposition is observed in Figure 3.
In order to study the influence of the additional the exterior bolts, an alternative connection
was conceived, considering the same structural elements and maintaining the other original
geometric features but widening the end plate from 180mm to 260. The top part of the connec-
tion was retrieved as a macro-component in order to study the behaviour of the tensile part of
the connection (Figure 4).

Figure 3. Isolation of the tensile macro-component.

865
Figure 4. Macro-component (2BR and 4BR).

The parametric numerical study covers a total of 40 FE models, where, based on the original
configurations, four parameters were varied: (i) configuration 2BR/4BR; (ii) bolt diameter; (iii)
beam profile and (iv) end-plate thickness. In this regard, the name of the models used in the next
sections consider the described parameters. For instance, the model “2B_M12_IPE300_EP12”
considers two bolts per row of 12 mm in diameter, connecting an IPE 300 beam with an end-
plate of 12mm in thickness. All the numerical models considered an assembly of two symmetric
macro-components connected back-to-back by bolts. All the models were built using 3D elements
on the basis of the measured geometry of the plates or the effective diameter of the bolts (accord-
ing to ISO 898-1).
In order to conduct the finite element analysis, the ABAQUS FEA software was used. The
different parts composing each model were meshed using the same type of solid finite element,
an 8-node brick C3D8R with enhanced hourglass control and reduced integration. For the
elements in bending, the mesh was intentionally left asymmetrical in order to better capture
the non-linear geometry of the assembly.
The macro-components were positioned symmetrically and secured by the bolts (using gen-
eral contact settings). While one element is pinned, the opposed one is subjected to a traction
force applied in the centre of gravity of the section.

3.2 Materials
The behaviour curve of the material used in the execution of the experimental structure (Dinu
et al., 2017) was obtained via tensile testing of coupons in laboratory. These results were then
compiled and formatted into a true-stress true-strain behaviour curve - Table 1). The bolt material
was tested and modelled following the same procedure. The results are shown in Table 2.

Table 1. Nominal values for steel S275JR.


fy fu Agt At

Element N/mm2 N/mm2 % %

Bolted T-stub end-plate t = 10 mm 310 408 22.5 34.7


t = 12 mm 305 445 23.3 32.7

Table 2. Nominal values for bolts group 10.9.


fy fu Agt At

Element N/mm2 N/mm2 % %

Bolt, M16 965 1080 5.0 6.5

866
Figure 5. Plastic behaviour for S275JR and bolt 10.9.

3.3 Numerical model


In order to calibrate the material settings, coupons of identical geometry were modelled in
finite elements and assigned the material elastic and plastic characteristics obtained in the
coupon tests (Figure 5).
There are, however, limits to the software’s capacity of interpretation of the full curve as is,
and consequently the degradation phase was re-calibrated according to specific parameters.
The degradation phase is modelled by considering the Damage Evolution parameter for duc-
tile materials. The parameter is defined using various coefficients (Fracture Strain, Stress
Triaxiability, Strain Rate) which can describe analytically the variability of the damage which
depends on the initial section of the material and the effective section after loading.
The fine material calibration is then carried by tuning the mentioned coefficients in order to
obtain a numerical behaviour matching the true experimental behaviour.

4 RESULTS

4.1 2BR-4BR comparison


Illustrating the global behaviour differences between the 2B and 4B configurations, two repre-
sentative Force – Displacement curves are superposed in Figure 6.
Generally, there can be observed a significant increase in the maximum force and in the
total dissipated energy. In order to assess the behaviour of the two different configurations
2BR and 4BR, the levels of strength and ductility were measured for each model.
Regarding the strength- Figure 7, all the 4BR models have registered a better performance
at the level of the yield strength, by proving an increase in the range of minimum of +5.4%
and a maximum of +64,0% with an overall average of +25,2%. It worth mentioning that the
highest values were obtained in models with stiff end-plates and small diameter bolts
(M12_20) which was prone to mode 3 failure.

Figure 6. Behaviour curves for models 2B and 4B M16_IPE300_EP20 and M16_IPE300_EP12.

867
Figure 7. Variation of yield and maximum force values between 2BR and 4BR.

Figure 8. Variation of displacement at maximum force between 2BR and 4BR.

At the level of the maximum strength, the 4BR configuration led to a largely general
increase with a small exception of one model (IPE300_M20_15), for which the maximum
force slightly decreased by −1,4%. The average increase was of 24,2% and the maximum
recorded increase was of +58,2%, again in the models which tended towards mode 3.
Concerning the ductility (represented by the displacement at the point of maximum force)-
Figure 8, although the variation between 2BR and 4BR was broadly positive, the observed per-
formance was not uniform and some of the models presented a negative evolution. The lowest
value observed was −24,7% and the highest was 380,2%, the average value being 76,1%. Also,
the narrower beam (IPE300) faired generally better than the broader counterpart (IPE360).

4.2 Flange width variation


In a classic 2BR connection, the bolts are invariably positioned within the width of the beam’s
flange. However, in the case of 4BR, the exterior bolt-rows are often placed outside the span of
the beam’s flange, especially in the case of beam IPE profiles. This changes the overall behav-
iour of the connection as it affects the initial rigidity of the macro-component and ultimately
the outer bolt participation rate (OBPR) (Nunes and Ciutina, 2019) (Nunes and Ciutina, 2019).
Tables 3 and 4 prove that the values of the yield and maximum force present a variation
between beam profiles for both 2BR and 4BR models. Although the variation for 2BR is
quite modest (avg. +4,7% for yield force and avg. +3,9% for maximum force), the variation

868
Table 3. Yield force comparison between beam profiles for 2BR and 4BR models.
Fy

2BR 4BR

End Plate
Bolts [mm] IPE300 IPE360 ΔFy (kN) ΔFy (%) IPE300 IPE360 ΔFy (kN) ΔFy (%)

M12 12 278,1 275,8 −2,3 −0,8% 323,8 363,4 +39,6 +12,2%


M12 15 341,6 352,0 +10,4 +3,0% 430,3 497,7 +67,4 +15,7%
M12 20 322,2 406,0 +83,8 +26,0% 528,4 613,2 +84,8 +16,0%
M16 12 300,3 296,7 −3,6 −1,2% 339,9 382,4 +42,5 +12,5%
M16 15 419,0 424,3 +5,3 +1,3% 461,7 525,9 +64,2 +13,9%
M16 20 585,7 623,2 +37,5 +6,4% 617,5 725,7 +108,2 +17,5%
M20 12 329,8 333,4 +3,6 +1,1% 379,5 426,8 +47,3 +12,5%
M20 15 461,4 469,3 +7,9 +1,7% 495,4 579,9 +84,5 +17,1%
Average ΔFy (%) +4,7% Average ΔFy (%) +14,7%

Table 4. Maximum force comparison between beam profiles for 2BR and 4BR models.
Fmax

2BR 4BR

End Plate ΔFmax ΔFmax ΔFmax ΔFmax


Bolts [mm] IPE300 IPE360 (kN) (%) IPE300 IPE360 (kN) (%)

M12 12 418,8 430,4 +11,6 +2,8% 518,3 577,3 +59,0 +11,4%


M12 15 431,3 447,0 +15,7 +3,6% 560,9 622,9 +61,9 +11,0%
M12 20 424,4 461,7 +37,3 +8,8% 652,5 730,5 +78,0 +12,0%
M16 12 587,6 615,9 +28,3 +4,8% 627,0 736,1 +109,1 +17,4%
M16 15 705,1 725,0 +19,9 +2,8% 789,3 900,9 +111,5 +14,1%
M16 20 763,0 778,3 +15,3 +2,0% 921,7 1014,3 +92,6 +10,0%
M20 12 723,4 741,6 +18,3 +2,5% 802,1 860,5 +58,4 +7,3%
M20 15 909,0 941,5 +32,6 +3,6% 896,1 1026,4 +130,4 +14,5%
Average ΔFmax (%) +3,9% Average ΔFmax (%) +12,2%

values for 4BR models is more expressive: avg. +14,7% for yield force and avg. +12,2% for
maximum force. This indicates that the beam’s flange dimensions play a more relevant role in
the 4BR models than in classic 2BR configurations.
Regarding the ductility aspect (Table 5), the observed values sustain this assumption, show-
ing a significant increase of the values for the 4BR models (avg. -10,4% for yield force and
avg. -26,1% for maximum force). The trend is, however, negative as a broader plate tends to
reduce the ductility levels of the connection.
In the conception of the 4BR models a third beam profile (HEA260) was introduced in
order to further study the influence of the width of the flange in the connection’s behaviour.
Figure 9 presents both yield and maximum force values.
Thus, considering the variability of beam profiles, the same variation of the yield and max-
imum force values can be observed. In all the cases there is a steady increase of both values
with the increase in the width of the profile. Although less uniformly, the ductility varies with
the beam’s width, by decreasing sharply in the models with a higher ratio between bolt and
end-plate rigidity - Figure 10.

869
Table 5. Displacement at maximum force comparison between beam profiles for 2BR and 4BR models.
Disp_max

2B 4B

End Plate ΔDisp ΔDisp ΔDisp ΔDisp


Bolts [mm] IPE300 IPE360 (mm) (%) IPE300 IPE360 (mm) (%)

M12 12 15,80 15,57 −0,2 −1,5% 21,76 18,86 −2,9 −13,4%


M12 15 7,58 6,36 −1,2 −16,0% 10,90 9,40 −1,5 −13,8%
M12 20 3,99 3,85 −0,1 −3,6% 7,36 6,75 −0,6 −8,4%
M16 12 39,75 31,11 −8,6 −21,7% 29,93 30,30 +0,4 +1,2%
M16 15 24,92 23,28 −1,6 −6,6% 46,61 33,44 −13,2 −28,3%
M16 20 12,82 10,38 −2,4 −19,0% 61,57 29,86 −31,7 −51,5%
M20 12 38,63 36,71 −1,9 −5,0% 86,87 33,09 −53,8 −61,9%
M20 15 38,65 34,80 −3,9 −10,0% 71,21 47,59 −23,6 −33,2%
Average ΔDisp (%) −10,4% Average ΔDisp (%) −26,1%

Figure 9. Yield and maximum force values for different beam profiles, for 4BR models.

Figure 10. Displacement at maximum force for different beam profiles.

870
5 CONCLUSION

In this study, connections with four-bolts-per-row have been studied using macro-components
in tension, in order to evaluate their behaviour and depict their characteristics in regard with
the classic two bolts per row configuration. When evaluating the impact of the added outer
bolts in the characteristics of the macro-component, several tendencies were observed by chan-
ging from a 2BR configuration to 4BR:
• in all the studied cases, the strength performance of the 4BR macro-component increased
both at the level of yield force and at the level of maximum force;
• on average, the displacement at the point of maximum force also increased, in some cases
due to partial plasticisation of the beam;
• the flange width plays a significant impact on the connection’s performance in the 4BR con-
figurations than for classic 2BR configurations;
• where the connection tends to fail towards mode 1, the levels of ductility decrease with the
widening of the flange. This is due to the stiffening effect of the flange on the end-plate,
which loses its deformation capacity.

REFERENCES

Ciutina, A. 2003. Assemblages et comportement sismique de portiques en acier et mixtes acier-béton:


expérimentation et simulation numérique, Rennes, Institut National des Sciences Appliquées.
Couchaux, M., Demonceau, J.-F. & Weynand, K. 2015. Calcul d’un assemblage par platine comportant
quatre boulons par rangée. Revue Construction métallique. Liège: Université de Liège, CTICM, Feld-
mann et Weynand Ingénierie.
Demonceau, J.-F., Jaspart, J.-P., Weynand, K. & Muller, C. Application of Eurocode 3 to steel connec-
tions with four bolts per horizontal row. Stability and ductility of Steel Structures 2010, 2010 Rio de
Janeiro.
Dinu, F., Marginean, I., Dubina, D., Kovacs, A. & Ghicioi, E. Testing of a full-scale building under
external blast. 2017 European Federation of Explosives Engineers, 2017 Stockholm.
Kozolowski, A. & Pisarek, Z. 2008. Effective length of end plate in moment connections with four bolts
in the row. Collection of scientific works VM Shimanovsky.
Massimo, L., Gianvittorio, R., Aldina, S. & da Silva, L.S. 2014. Experimental analysis and mechanical model-
ing of T-stubs with four bolts per row. Journal of Constructional Steel Research, 158–174.
Nunes, D.L. & Ciutina, A. Behaviour of end-plate steel connections with 4 bolts per row under large
deformations. 9th International Conferene on Steel and Aluminium Structures (ICSAS19), 2019
Bradford.
Nunes, D.L., Marginen, I., Ciutina, A. & Dinu, F. Influence of four bolts per row connections on a steel
frame building subjected to column. Integrity – Reliability – Failure 2018, 2018 Lisbon.

871
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Experimental study on square steel tubular columns under


compressive force with biaxial bending moment
In case of end bending moment ratio equal to 0.0 and load direction
equal to 45 degrees

T. Onogi & A. Sato


Nagoya Institute of Technology, Nagoya, Aichi, Japan

ABSTRACT: Tubular hollow square steel (HSS) columns are widely used for the steel struc-
tural buildings in Japan. The columns will essentially resist horizontal force in arbitrary direction
subjected to the spacial moment resisting frame. As a result, biaxial moment and axial force are
applied to the columns. The Architectural Institute of Japan (AIJ) has a publication that can be
used to design the column under compressive axial force and biaxial bending. However, the
design rule which is used to design the column in combined loading condition is not shown clearly
with the sufficient test data. In this paper, tubular hollow square steel columns are studied by the
full-scale testing to verify the ultimate strength and plastic deformation capacity. The parameters
selected for testing are the direction of bending moment, the compressive axial force, and the slen-
derness ratio. From the test results, local buckling at the plate elements were observed in the spe-
cimens where the applied axial force was relatively small. However, compared with the results
from the uniaxial bending moment where the bending moment was applied to the principal axis,
it was found that the column under 45 degrees loading direction was not sensitive to the local
buckling due to the stress distribution in the plate elements. Finally, it was found that the formula
recommended by AIJ can guarantee sufficient strength and plastic deformation capacity,
although the design rule in the recommendation provided overly conservative results.

1 INTRODUCTION

Tubular hollow square steel (HSS) columns are widely used for the steel structural build-
ings in Japan same as Wide-flange column. The horizontal force in arbitrary direction
subjects to the space moment resisting frame will be essentially resisted by the flexural
manner of the columns. As a result, biaxial bending moment and axial force are applied
to the columns. Design criteria (AIJ 2005) and recommendations (AIJ 2010, 2017) pro-
vides the design method for the members of the plane frame (2D frame) which horizontal
force is subjected. The design criteria for the column is mainly based on the axial force
with the uniaxial bending moment. The limit state design formulas for the columns under
the axial force with the biaxial bending moment are shown in the Recommendation for
Limit State Design of Steel Structures (LSD) (AIJ 2010) and the Recommendation for
Plastic Design of Steel Structures (PD) (AIJ 2017). In these recommendations, the design
formulas for the flexural and the compressive limit state are defined as a linear interpol-
ation between the formula of uniaxial bending moment and axial force. However, the
validity of these formulas are not explained sufficiently.
Testing of HSS columns that are under axial force with three kinds of uniaxial bending
moments (end bending moment ratio equal to 0.0, 0.5, 1.0) were conducted by our research
group, and the column that can guarantee the full plastic bending moment considering axial force
and sufficient plastic deformation capacity were evaluated (Sato & Mitsui 2018). Wide-flange col-
umns under the axial force with the biaxial bending moment are studied by Matsui (1986). In this

872
paper, the design formulas that are shown in SSRC (Structural Stability Research Council 1976),
Chen (1986), ESSC (European Convention Constructional steelwork) and PD (First edition) (AIJ
1975) were evaluated with the test results. As a result, the formulas shown in PD provided the
most conservative results. While, the formula proposed by Chen provided the appropriate results.
However, as for as our survey, the research dealing with the HSS column formulas that can guar-
antee the plastic hinge under the axial force with the biaxial bending moment can not be found.
In this study, the testing of HSS columns under the axial force with the biaxial bending
moment are conducted to verify the deformation, maximum bending moment and plastic
deformation capacity. The test parameters are axial force ratio and slenderness ratio. The
design formulas specified in PD, LSD, and recommended by Chen are used to verify the max-
imum bending moment.

2 RECOMMENDATION FOR STEEL COLUMN DESIGN IN JAPAN

2.1 Requirements for column which will form plastic hinge


PD (AIJ 2017) specifies the following formula when a plastic hinge is expected to be formed in
the column. This formula can be applied if the moment distribution (i.e. End moment ratio :
M2/M1=κ) is in the range of -0.5 to 1.0 (see Figure 1).

ny λ2c0 ≦ 0:10ð1 þ M2 =M1 Þ ð1Þ

where ny is the axial force ratio (=N/Ny), N is the axial force, Ny is the axial yield strength
(=σy・Ag), σy is the yield stress, Ag is the gross area of the cross-section, and λc0 is the non-
dimensional flexural slenderness ratio based on measured yield stress defined as Equation (2).
rffiffiffiffiffi
1 σy L
λc0 ¼ ð2Þ
π E ix

where E is the Young’s modulus, L is the length of the member, and ix is the radius of gyr-
ation. The end moment equal to the maximum moment in the elastic member can be derived
by the theory and following Equation (3) can be obtained. This equation guarantees that first
yielding occurs at end of the column.

Figure 1. Load condition.

873
ny λ2c0 ≦ 0:25ð1 þ M2 =M1 Þ ð3Þ

Moreover, column should satisfy the axial force limitation ny ≦ 0.75.

2.2 Strength for column under axial force with biaxial bending moment
Following formulas are specified for the column under axial force with biaxial bending moment.

N MX MY
þ ’X  0:85 þ ’Y  0:85 ¼ 1:0 ð4:aÞ
Ncr MpX MpY

MX MY
 1:0; ≦ 1:0 ð4:b; cÞ
MpcX MpcY

where Ncr is strength of flexural buckling, MX, MY are the bending moment around X axis and
Y axis, MpX, MpY are the full plastic bending moment around X axis and Y axis, MpcX, MpcY
are the full plastic bending moment considering axial force around X axis and Y axis, and φX,
φY are the coefficient of moment amplification around X axis and Y axis (see Fig. 2).

3 OUTLINE OF EXPERIMENT

3.1 Specimens
Cold-formed square steel columns (HSS) are used in this testing. The cross-section is BOX-
125×125×6 (mm), and the nominal width-thickness ratio B/t is equal to 20.8. The parameters are
the axial force ratio and the slenderness ratio. Figure 3 shows the plan of the specimen. End-
plates are welded on both ends of HSS column to setup the specimen in the testing device (section
angle is 45 degrees). In this paper, the angle is called “Loading angle” and the symbol αs is used.
Table 1 shows the material properties from coupon test, and Table 2 shows the geometrical
properties of the cross-section measured from the specimens. Where I is moment of inertia, Zp
is plastic section modulus, Mp is full plastic bending moment (= Zp・σy). I, Zp and Mp are
corresponding to loading angle. The moment of inertia, where the loading angles are 0 degree
and 45 degrees, is equal. The coupon test specimens were cut out from the center of the flat
part of the plates except for the seamed part. Total three coupon tests were conducted and the
average results are shown in Table 1. The yield stress is defined as 0.2% off-set strength. The
length of specimens are three types (1800 mm, 2100 mm, 2400 mm) and the end bending
moment ratio κ is equal to 0.0. The specimens where the welding angle is 0 degree is prepared
to compare with 45 degrees. Non-dimensional flexural slenderness ratios λc0 calculated by

Figure 2. Cross-section of column.

874
Figure 3. Test specimen (αs=45°).

Table 1. Material properties from the coupon test.


tb E σy σu Elong Y.R.

6 206800 385.7 443.2 40.4 87.0

tb : Nominal thickness [mm], E : Young’s modulus [N/mm2], σy : Yield stress [N/mm2], σu : Ultimate
stress [N/mm2], Elong : Elongation [%], Y.R. : Yield ratio = σy/σu×100 [%].

Table 2. Geometrical properties of cross section.


Cross section B t B/t Rout/t A αs I Zp Mp

□-125×125×6 124.8 5.77 21.63 1.67 2681 0 625.3 118.7 45.78


45 625.3 118.7 43.50

B : Nominal thickness [mm], t : Thickness [mm], Rout : Outer radius [mm], A : Area [mm2], αs : Loading angle [°], I :
Moment of inertia [×104 mm4], Zp : Plastic section modulus [×103 mm3], Mp : Full plastic bending moment [kN-m].

Table 3. Geometrical dimensions in longitudinal direction.


L λ λc0

1800 37.27 0.515


2100 43.48 0.600
2400 49.69 0.686

L : Length [mm], λ: Slenderness ratio, λc0 : Non-dimensional flexural slenderness ratio.

using yield stress are shown in Table 3. In this paper, the axial force ratio (N/Ny = ny) and the
flexural slenderness ratio are based on measured yield stress from the coupon test.
Figure 4 shows the relationship between the specimens and design formulas in LSD. The
numbers in Figure 4 corresponds to the specimens tabulated in Table 4. Table 4 shows the list
of specimens. The target compressive axial force Nt is based on the yield strength Ny calcu-
lated with yield stress σy measured from coupon test and the gross area of the cross-section.
θpc is elastic end rotation angle corresponding to the full plastic bending moment Mpc.

3.2 Testing
In order to observe the deflection of the member, measurement devices are set at section S-A to
S-G (see Figure 3). To measure the rotation at both ends of the column, two sets of displace-
ment transducers are set. Strain gauges are attached at section G-A to G-F to measure the
strain distribution in the longitudinal direction. Figures 5 and 6 shows the boundary condition
and the testing device used in this testing. Firstly, the axial force was applied to the specimen by
the hydraulic jack located on the west side where the testing jig is supported by the linear guide

875
Figure 4. Comparison between specimens and LSD limitations.

Table 4. List of specimens.


name No. L ny Nt αs Mpc θpc

B125bt21ny0375L1800a0 1 1800 0.375 387.8 0 36.90 0.0195


B125bt21ny0375L1800a45 2 45 35.78 0.0178
B125bt21ny050L1800a45 3 0.5 517.0 45 30.73 0.0157
B125bt21ny0625L1800a45 4 0.625 646.3 45 25.08 0.0131
B125bt21ny075L1800a0 5 0.75 775.5 0 15.02 0.0081
B125bt21ny075L1800a45 6 45 19.18 0.0123
B125bt21ny02L2100a45 7 2100 0.2 206.8 45 41.10 0.0224
B125bt21ny03L2100a45 8 0.3 310.2 45 38.36 0.0209
B125bt21ny04L2100a45 9 0.4 413.6 45 34.84 0.0190
B125bt21ny02L2400a45 10 2400 0.2 206.8 45 41.10 0.0257
B125bt21ny03L2400a45 11 0.3 310.2 45 38.36 0.0239

ny : Compressive axial force ratio, Nt : Target compressive axial force [kN], Mpc : Full plastic bending moment
considering the effect of compressive axial force [kN-m], θpc : Elastic rotation corresponding to Mpc [rad].

Figure 5. Boundary condition.

876
Figure 6. Testing device (Plan view).

Photo 1. Overview of testing device.

to accommodate the column shortening. After the axial force reached the target compressive
axial force Nt, bending moment was applied around the x axis from the east side of the column
end that is shown in Figure 6. The axial force is measured by the load cell. The bending
moment is applied by the couple force P through PC steel bar which is equipped with the arm
using the center hole jack. The bending moment is calculated from tensile force measured by
the load cell attached at the PC steel bars and the distance of couple force (1.0 m). The support-
ing conditions at the ends are pin (east side) and pin-roller (west side) that are free around
x axis. Out of plane displacement (x-direction) is restrained in both ends.

4 TEST RESULTS

Table 5 shows the test results. Maximum bending moment Mmax is the bending moment meas-
ured at the loading point M1. S-i is the location where the deflection was the lagest
at Mmax. Mi/Mmax is ratio of second-order moment. Rotation angles θmax and θu are the
values corresponding to the maximum bending moment and the bending moment at the
deformation evaluation point Mu (defined as Eq. 5).
 
Mu ¼ max 0:90Mmax ; Mpc ð5Þ

θu can be obtained from the specimens where M1 reached Mpc. Plastic deformation capacity
R is the value corresponding to θu. R is defined as the equation shown in below.

R ¼ θu =θpc  1 ð6Þ

877
Table 5. Test results.
name No. Mmax Mmax/Mpc S-i Mi/Mmax θmax θu R

B125bt21ny0375L1800a0 1 42.45 1.112 S-A 0.989 0.0809 0.1106 5.03


B125bt21ny0375L1800a45 2 40.70 1.138 S-A 0.970 0.0739 0.1270 6.14
B125bt21ny050L1800a45 3 34.52 1.123 S-A 1.041 0.0642 0.0952 5.08
B125bt21ny0625L1800a45 4 25.80 1.029 S-A 1.097 0.0434 0.0550 3.19
B125bt21ny075L1800a0 5 14.63 0.974 S-B 1.199 0.0264 - -
B125bt21ny075L1800a45 6 14.70 0.766 S-C 1.270 0.0311 - -
B125bt21ny02L2100a45 7 47.45 1.154 S-A 0.940 0.1004 0.2084 7.92
B125bt21ny03L2100a45 8 42.18 1.100 S-A 0.991 0.0830 0.1403 5.27
B125bt21ny04L2100a45 9 36.57 1.050 S-A 1.033 0.0709 0.0989 3.73
B125bt21ny02L2400a45 10 47.51 1.156 S-A 0.961 0.0945 0.1862 5.86
B125bt21ny03L2400a45 11 43.26 1.128 S-A 1.007 0.0796 0.1368 4.21

Figure 7 shows the relationship between bending moment and rotation angle. The triangle
legend in these figures corresponds to the point where the local buckling were observed. Photo 1
shows the deformed shape of each specimen after the testing that are taken from the west side.
The circle in these photos illustrates the location where the local buckling was observed. Photo 2
shows the deformed shape at the section where the local buckling occurred. Strength degradation
after the maximum bending moment was gradual when the local buckling was formed in the spe-
cimen. However, the strength degradation after the maximum bending moment was rapid when
the local buckling was not formed in the specimen. The strength degradation was significant
when the subjected axial force was high. This rapid strength degradation was delivered by
the second-order effect, and the Pδ moment made a significant effect. It can be observed that
the second-order effect will be significant when the axial force is high, and the dominant failure
mode will be determined by Pδ moment. When the failure mode is determined by local buckling,
the local buckling will appear close to the end when the axial force is low. The maximum bending
moment reduced when the axial force ratio increased due to the effects of Pδ moment. The loca-
tion of the local buckling will be close to the loading point (i.e. end) when the axial force is smal-
ler. The maximum bending moment and θmax measured from specimen No.7 and No.10 are
almost the same. In Figures 7(b), (c), specimen No.10 formed local buckling earlier than specimen
No.7. The length of specimen No.10 is longer than specimen No.7 and the deflection is also
larger. As a result, the formation of the local buckling was accelerated by Pδ moment.
Figure 8 shows the comparison of M-θ relationship between αs=0 degree and αs=45 degrees
in L=1800 mm. The maximum bending moments are equal in each axial force ratio. As can be
seen from Figure 8, strength degradation of the specimen No.1 due to the local buckling
occurred earlier than specimen No.2. As a result, the plastic deformation capacity R of speci-
men No.1 is smaller than specimen No.2. As can be recognized from Photos 2(a) and (b),
local buckling point of specimen No.1 is closer to the end than specimen No.2. The specimens
No.5 and No.6 where the axial force level is equal to the design limit, the M-θ relationships
showed similar characteristics, and also the Pδ effect dominated the failure mode.

Figure 7. Relationship between M1 and θ. (a) 1800 series. (b) 2100 series. (c) 2400 series.

878
Photo 2. Deformation after loading.

Photo 3. Local buckling deformation. (a) No.7. (b) No.10.

5 EVALUATION OF STRUCTURAL PERFORMANCE

5.1 Comparison between maximum bending moment and formula


In this section, the maximum bending moment Mmax is compared with the design formulas recom-
mended in PD (AIJ 2017), LSD (Second edition) (AIJ 2002), and the formula proposed by Chen
(1986). In this study, the bending moments around X axis and Y axis are defined as below.
MX ¼ M1  cos αs ; MY ¼ M1  sin αs ð7Þ

Figure 8. Comparison of loading angle.

879
where MX and MY are the bending moments around X axis and Y axis.
LSD (Second edition) specifies the design formula for flexural and compressive strength of
HSS columns. The design formula is shown below.
   
N MX MY
þ 0:85 þ 0:85 ¼ 1:0 ð8Þ
Ny MpX MpY

Figure 9(a) shows the comparison between testing results and Equation (4), (8). PD, and LSD
(Second edition). It can be confirmed that conservative results are obtained.
The design formula for the column under axial force with biaxial bending moment that is
proposed by Chen is shown below.
   
CmX MX η CmY MY η
þ ¼ 1:0 ð9:aÞ
MpnX MpnY
  
N N
MpnX ¼ 1   1 MpX ð9:bÞ
Ncr N0X
  
N N
MpnY ¼ 1   1 MpY ð9:cÞ
Ncr N0Y

N B
η ¼ 0:4 þ þ ≧1:0 ð9:dÞ
Ny D

where CmX and CmY are reduction factor around X axis and Y axis. CmX and CmY are calcu-
lated as 0.6 (Chen W.F. 1986). N0X and N0Y are Euler buckling strength around X axis and
Y axis. N*cr is flexural buckling strength (AISC 2010). B and D are depth and width of the
column. In this study, B/D is 1.0. Figure 9(b) shows the comparison between Equation 9 and
test results. It can be confirmed that reasonable results are obtained.

5.2 Plastic deformation capacity


Figure 10 shows the relationship between the plastic deformation capacity R and ny・λc02.
The columns which satisfy Equation 1 can guarantee R larger than 5.0. In terms of αs=45
degrees (legend is “〇”), the regression line can be derived as Equation (10). The correlation
coefficient of this equation is equal to -0.965.
 
R ¼ 47:1 ny  λ2c0 þ 10:8 ð10Þ

Figure 9. Comparison of formulas.

880
Figure 10. Plastic deformation capacity.

6 CONCLUSIONS

In this paper, the full-scale testing of HSS columns were conducted to clarify the maximum
bending moment and the plastic deformation capacity.
The main conclusions are:
1. The design formula recommended in PD for the column under the axial force with the
biaxial bending moment provided conservative results with a large safety margin in this
study. On the other hand, even if the column did not satisfy the formula to guarantee
forming the local buckling, the maximum bending moment reached the full plastic bend-
ing moment considering axial force except the column that axial force ratio ny was 0.75.
2. The formula recommended by Chen evaluated the test results in a reasonable accuracy.
3. In the columns where the loading angle is equal to 45 degrees, negative linear correlation
between the plastic deformation capacity R and ny・λc02 was found. The correlation coef-
ficient was equal to -0.965. As a result, the plastic deformation capacity R can be evalu-
ated when the ultimate limit state is not determined by the local buckling.

REFERENCES

American Institute of Steel Construction 2010. : Specification for Structural Steel Buildings, ANIS/AISC
360-10
Architectural Institute of Japan 1975 : Recommendations for Plastic Design of Steel Structures (in Japanese)
Architectural Institute of Japan 2002 : Recommendations for Limit State Design of Steel Structures
(in Japanese)
Architectural Institute of Japan 2005 : Design Standard of Steel Structures -Based on Allowable Stress
Concept- (in Japanese)
Architectural Institute of Japan 2010 : Recommendations for Limit State Design of Steel Structures
(in Japanese)
Architectural Institute of Japan 2017 : Recommendations for Plastic Design of Steel Structures
(in Japanese)
Chen W.F., Lui E.M. 1986: Structural Stability-Theory and Implementation-, Preatice-Hall Inc.
European Convention Constructional Steelwork : European Recommendation for Steel Construction,
THE CONSTEUCRION PRESS, ECCS, 1978
Matsui, C., Morino, S., Tsuda K. 1986 : An Experimental Study on Inelastic Behavior of Wide-Flange
Steel Beam-Columns under Constant Vertical and Two-Dimensional Horizontal Loads , Journal of
Structural and Construction Engineering (Transactions of AIJ), No.361, pp.113–122 (in Japanese)
Sato, A. and Mitsui, K. 2018 : Structural Performance Evaluation of Square Steel Tubular Columns
under Compressive Axial Force with Bending Moment, Journal of Structural and Construction Engin-
eering (Transactions of AIJ), Vol.83, No.751, pp.1365–1372 (in Japanese)
Structural Stability Research Council 1976 : Guide to Stability Design Criteria for Metal Structures, John
Wiley & Sons, SSRC

881
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Considering realistic weld imperfections in load bearing capacity


calculations of ring-stiffened shells using the analytical numerical
hybrid model

H. Pasternak, Z. Li, C. Stapelfeld & B. Launert


Chair of Steel and Timber Structures, Brandenburg University of Technology, Cottbus, Germany

A. Jäger-Cañás
Bauwesen GmbH, Lohfelden, Germany

ABSTRACT: The ring-stiffener is widely used to prevent the local and global buckling in
silos and tanks. For their assembly, welding is one of the most common techniques in steel
structures. However, the welding process inevitably also produces residual stresses and distor-
tions. These structural and geometrical imperfections influence the buckling capacity of stiff-
ened silos. Since they are in most cases thin walled structures, they are known to be very
sensitive to imperfect initial geometries. Nowadays, for the calculation of residual stresses and
weld distortions, welding simulation can provide satisfactory results. At the same time, due to
the huge requirement of calculation time and storage capacity, the 3-D welding simulation is
difficult to apply directly to the analysis of real large structures. The analytical numerical
hybrid model provides an alternative fast and simply applicable approach for the calculation
of weld imperfections. The approach is subdivided into an analytical calculation of mechanical
loads representing the thermomechanical effects caused by welds and then a subsequent
numerical calculation of the stresses and deformations with these loads. In this paper, the
application of this model is demonstrated on an exemplary buckling capacity calculation of
a silo structure with ring-stiffeners. First, by means of a parametric study applying different
heat inputs per unit length and different welding sequences, recommendations for
a subsequently planned experimental study are proposed targeting the lowest imperfections.
Finally, numerical calculations of the axial load bearing capacity of the welded silo specimen
are carried out to show the results under consideration of these results with the analytical
numerical hybrid model.

1 INTRODUCTION

Ring-stiffened silos are wildly used in industrial structures (Jäger & Pasternak, 2016). However,
until today the load-increasing effect of ring-stiffeners under axial pressure has received a limited
attention. Considering the structural safety, the ring-stiffened shell needs to be estimated on the
conservative side in practice. This is not desirable in terms of modern lightweight steel structures
(Jäger-Cañás & Pasternak, 2017). For this purpose, a research project (Pasternak, 2018) is
recently carried out to close this gap in the current design standard with regard to the axial buck-
ling behavior of ring-stiffened cylindrical shells. In this project, a series of experiments is per-
formed to calibrate numerical models, which are then used in further parameter studies. All
experimental specimen sizes are scaled down to a lab-operable level by a determined ratio of
radius and thickness. This means that the thickness of the shell is very thin. Therefore, the load
bearing capacity of the specimen is generally very sensitive to geometric imperfections. In par-
ticular, the weld distortions generated during the welding process are not negligible. Researches
on the influence of the welding process on the load bearing capacity of silos have started decades
ago (Rotter & Teng, 1989; Pircher & Bridge, 1998).

882
Meanwhile, the residual stresses and distortions of welds can be relatively exactly simulated
with commercial finite element (FE) code (Pasternak et al. 2015). Some references (Banke
et al. 2003; Loose, 2008) have already studied the influence of welding residual stresses and
welding distortions on the buckling behavior of shells. Even though a complete numerical
simulation can predict the weld imperfections, the welding simulation is, on the other hand,
very resource intensive. Very fine FE meshes around the welds and a huge number of time
steps, especially for long welds, are the main reasons for this. Therefore, welding simulations
are currently impossible to apply for the design of the large-scale shell structures in practice.
Nevertheless, simplified numerical approaches are available and allow to remedy this situ-
ation (Stapelfeld et al. 2009). However, the application of some of these models partly demands
more expertise than a conventional thermomechanical FE calculation (Duan et al. 2007) or the
simplifications are so extensive that the weld imperfections calculated by the approach partially
loses their validity (Thikomirov et al. 2008). Hence, fast but still sufficiently accurate procedures
that, at the same time, are easy in its application are required. In this paper, the so-called analyt-
ical numerical hybrid model will be employed to analyze the influence of the welding process.
By these investigations, effects of the welding sequence, the welding energy and the welding
speed on the load bearing capacity of the shell are considered. Accordingly, the welding process
can be also optimized to reduce the accompanying weld imperfections based on the results.

2 ANALYTICAL NUMERICAL HYBRID MODEL

The basic idea of the coupled analytical numerical shrinkage force model is the linking of the
major advantages of both, analytical and numerical procedures (Stapelfeld, 2016). On the one
hand, the matchless marginal calculation time of the analytical shrinkage force model and its
simple application, and on the other hand, the possibility to conduct a FE simulation to calcu-
late stresses and distortions at any location of complex welded structures. According to this, all
the determining factors on quality and quantity of weld imperfections are passed to an analytical
calculation program, capturing the mathematical approach of the shrinkage force model. The
output is a mechanical load and the point of action in longitudinal and transversal direction,
equivalent to the heat effect of welding. The loads are then applied to the FE model of the weld
structure and the distortions and stresses are calculated by an elastic, geometrically nonlinear
calculation. The influence of the weld sequence on the arising weld imperfections is captured by
a back coupling. The numerically calculated stresses in the regarded weld caused from
a previous weld are submitted to the analytical calculation. The complete calculation flow is
shown in Figure 1 in principle.

2.1 The analytical shrinkage force model


Weld imperfections depend significantly on the maximum temperatures that every point verti-
cal to the weld direction is exposed to and the stiffness of the structure. Equations for the cal-
culation of the maximum temperatures were derived by Rykalin (Rykalin, 1957) constituting
the basis of the shrinkage force model. Two border cases are considered: a line source in
a thin plate with isotherms penetrating the plate and a point source on a semi-infinite body
with circular isotherms around the weld. The proportion of each of the two border cases to
the resulting temperature field in the investigated weld structure depends on the heat input per
unit length, the geometrical properties as well as the heat exchange with the environment and

Figure 1. Calculation flow of the coupled analytical numerical hybrid model (Stapelfeld, 2016).

883
is calculated iteratively. Considering further influencing factors, an axial force Fx is calculated
equalling to the heat effect of welding (Kuzminov, 1974):

α
Fx ¼ 0:355 qs EKχδ Kk Kσ , ð1Þ

with the thermal expansion coefficient α, the specific heat capacity c, the density ρ and the
Young’s modulus E. qs denotes the heat source power. Kχδ , Kk and Kσ are capturing the tem-
perature field in welded plates, the stiffness of the weld structure and the effect of existing
stresses in the weld.
Furthermore, the transversal shrinkage force Fy is calculated as follows:

α  
Fy ¼ qs E ½0:255 þ 0:745Kσ ð0:04 þ 0:96Kav Þ
 1 þ Kμ Kδ

εF ð2Þ
 ð1 þ KC ð1 þ Kδ ÞÞKW Kδ þ qs ð1  Kδ Þ,
θ

where Kδ captures the degree of heating through the thickness, Kav captures the influence of
stiffening cross-beams, Kμ determines the effect of longitudinal strains on the plastic transver-
sal strains, KC is the degree of excessive heat and KW captures the effect of forced heat
exchange. εF is the yield strain and θ is the proportionality factor between the heat input per
unit length and the cross section of the weld.
The axial shrinkage force Fx , Equation 1, is proportional to the width of the plastic zone:

Fx
bPZ ¼ ð3Þ
εm Eδ_

Here, εm is the averaged yield strain and δ is the plate thickness.


The appropriate points of action zc are equal to the centre of the zone of plastic deform-
ations. They are significantly influenced by the material and its properties as well as the heat
input per unit length and the plate thickness. Depending on the points of action, an equivalent
linear strain distribution or respectively stress distribution over the plate thickness can be cal-
culated. Considering the point of origin in the center of gravity of the plate’s cross section, the
strain distribution ε(z) follows as:

12εm zc
εðzÞ ¼ εm þ z_ ð4Þ
δ2

2.2 The coupling procedure


For the coupling of the analytical shrinkage force model with the FE simulation, different
mechanical loads are available. The deformation state calculated by applying loads and
appropriate points of actions or alternatively eccentric pressures matches well with experimen-
tal results. However, the calculated stress state in the structure is qualitative wrong. Instead,
specifying strains or stresses linearly distributed over the plate thicknesses, Figure 2, leads to
correct stresses with tension in the weld and balancing compressive stresses in the nearby
regions. The procedure is already validated and verified with Ansys®, LS Dyna®, Sysweld®
and Abaqus®. In the case of loading the FE model with stresses σ, the Poisson’s ratio v must
be considered:

εx;y E þ εy;x Ev
σx;y ¼ _ ð5Þ
1  v2

884
Figure 2. Coupling by means of linearly distributed strains (Stapelfeld, 2016).

3 NUMERICAL STUDY

In this paper, the hybrid model is employed to analyse influence of the weld distortions and
residual stresses, which are produced by the necessary welding process for the production of the
test specimens. To study the enhancement of the axial load bearing capacity of the silo by the
application of ring-stiffeners, three cylindrical shells with different stiffener distances (without
ring-stiffener, with 70 mm and 140 mm distance of ring-stiffeners) are provided in this investiga-
tion. The cylindrical shell is manufactured by welding two half-cylindrical shells. Then the ring-
stiffeners are welded to the cylindrical shell. The thickness of shell is defined as 0.5 mm, and the
ratio of radius and thickness is 1600, i.e. r/t = 1600. An L-profile (15  25  1 mm) is used as
the ring-stiffener. The steel grade for shells is S460 and for the ring-stiffeners S235. Figure 3
shows the test set-up and corresponding specimen geometries.
In this preliminary study, the initial geometric imperfections of the half-cylindrical shells
and ring-stiffeners by themselves are ignored and only the effect of welding is considered as
a source of imperfections. Two different welding geometries, namely butt joint for joining the
two half-cylindrical shells and T-joint for the connection of stiffener and shell, are considered,
as shown in Figure 3. Metal active gas (MAG) welding with short-arc can be used to weld
also shells smaller than 1 mm. The energy inputs per unit length are assumed as shown in
Table 1.
The initial geometry of the silo is first defined without imperfections and a geometrically
and materially nonlinear analysis (GMNA) is employed for the axial bearing capacity calcula-
tion. In a second step, the weld effects are calculated with the analytical shrinkage force model
as longitudinal and transversal strains. Then, these strains are applied to the numerical model
and the residual stresses and distortions due to welding are calculated. In this calculation, the
stiffness of all stiffeners is considered in all weld steps since the stiffeners are initially tack-
welded to the structure. The results are considered in the load bearing capacity calculations
and a geometrically and materially nonlinear imperfection analysis (GMNIA) is carried out.

Figure 3. Set-Up of cylindrical shell specimen with ring-stiffeners.

885
Table 1. Energy inputs per unit length for different
joint types.
Joint type Energy input per unit length [J/mm]

T-joint 27.8
Butt joint 29.8

All calculations are run in Abaqus®. The shell element type S8R is used for all analyses and
the element size is defined as 10 mm. The material properties are taken from the material
library of specialized software (Simufact®.Welding, V. 5.0.0). The bottom end of the structure
is restrained in each translation direction, and the top end is also restrained except for vertical
direction due the applied constant line load.
The different collapse shapes for the shell with five stiffener-rings are shown in Figure 4.
From Figure 4 a), it is obvious that a very typical collapse shape for shells with a medium
strength ring-stiffener appeared. In this case, the wavelength of longitudinal buckling is equal to
half of the distance between two ring-stiffeners, wherep the
ffiffiffiffi distance between ring-stiffeners is two
times of the maximum of a buckling wave, i.e., 2 3:46 rt according to (Pasternak, 2018). When
distortions and residual stresses caused by welding are considered, the corresponding collapse
shape is different. According to the Figure 4 b), it is seen that the maximal deformation appears
on both sides of the vertical weld seam and this deformation is gradually reduced in the circum-
ferential direction as the distance from the vertical weld increases. This implies that the welding
process may cause local buckling of the shell structure, when weld imperfections are considered
in the numeral model.
To quantify the influence of the number of ring-stiffeners and the welding process on the
axial load bearing capacity, the corresponding simulation results are evaluated in Figure 5.
The results emphasize the strong differences between considering and ignoring the welding
process. In particular, for the axial load bearing capacity of the silo without ring-stiffeners,
the reduction due to welding is enormous. For the cases with ring-stiffeners and without weld-
ing, the axial load bearing capacity increases as the number of ring-stiffeners increases. If the
welding process is then taken into account in these numerical calculations, the maximum axial
load bearing capacity is in any case much higher than with no stiffening and welding, but the
axial load bearing capacity of the case with nine ring-stiffeners is now less than that of the
same shell with only five ring-stiffeners. Generally, welding is known to cause a reduction in
stiffness as well as load bearing capacity. In this particular example, this negative effect, which
increases with a higher number of welds, outweighs the positive effect of an increased number
of ring-stiffeners. Obviously, this can be no general conclusion and may be different for other

Figure 4. Collapse shapes for specimens with five ring-stiffeners, a) without weld imperfections,
b) considering the welding process.

886
Figure 5. Axial load bearing capacity of the different silo specimens.

cases or component sizes, but it clearly show the complexity of the problem when considering
welding.
The example calculations also point out a strong need to implement models that can realis-
tically capture the welding effect in corresponding numerical simulations. In this context, the
hybrid model can be a very useful tool for future calculations as it can be applied to different
geometries and for different weld processes and weld joints as well as different material
grades. Besides, it can also consider the sequencing of welds.
Finally, it should be mentioned that initial geometric imperfections before the welding pro-
cess may change the simulation results. For this, measurements of the initial geometries prior
to welding will be conduced in the experiments.

4 PARAMETER STUDY

To investigate also the influence of a variation of the welding parameters on the axial load
bearing capacity of the test specimen, the energy input per unit length is varied. Exemplary,
this variation is only considered for welds related to T-joints. The energy per unit length is
defined in a range of [14 J/mm, 34 J/mm], which equals 0.5 times to 1.2 times of the value
already given in Table 1. The welding sequence for these cases is given by welding successively
the welds from the bottom to the top. Corresponding numerical results are presented in
Figure 6 a).

Figure 6. a) Relationship of axial load bearing capacity and the applied energy input per unit length,
b) Influence of the welding sequence.

887
According to Figure 6 a), the axial load bearing capacity for both investigated cases
increases with an increased energy input per unit length. However, a stronger effect is only
noticed with five ring-stiffeners. With nine ring-stiffeners, the axial load bearing capacity is
more constant in the investigated range. Since these results are in contrast to what one would
expect at first glance, it again shows how important a computational consideration of such
effects can be.
Besides the energy input per unit length, three different welding sequences (WS) are also
studied. Exemplary, this investigation is shown only for the specimen with nine ring-stiffeners.
The sequences and the results are given in Figure 6 b) showing that these WS do not affect the
axial load bearing capacity of the shell in the investigated case.

5 CONCLUSIONS AND RECOMMENDATIONS

For the study of cylindrical shell specimens with welded ring-stiffeners, the analytical numer-
ical hybrid model can be used to take into account distortions and residual stresses for thin-
walled steel structures due to welding directly in numerical load capacity calculations. The
coupling of an analytical model and a FE simulation in this model meets also practical
requirements as it provides reliable values for the weld imperfections at very low calculation
times. The approach has particular potential in optimizing the welded shell structure accord-
ing to, e.g., the welding parameters or the weld sequence in cases where the shell structure
reacts sensitive to such imperfections.
The results show that the distortions and residual stresses due to welding cannot be ignored
for the specimen with welded ring-stiffeners. Due to its complex effects shown, it is necessary
to consider the welding effects directly in the numerical model, especially for the very thin
shell structures with welded ring-stiffeners dealt in this paper.
Concerning the planned experimental investigations in (Pasternak, 2018) and based on the
results of this preliminary numerical study carried out here, the following further recom-
mendations can be given:

– The influence of the welding sequence on the axial load bearing capacity of the investi-
gated specimen is shown to be negligible.
– Reducing the weld heat input does not necessarily reduce the welding effect to the accept-
ance range for the investigated cases.
– Even though welding is applied in practice, it might be necessary to avoid it in the manu-
facturing process of specimens in laboratory scale due to scale effects; instead, for
example, spot welding or adhesive bonding may be alternative processes (Pasternak &
Ciupack, 2014).
– If it is unavoidable to use welded joints, a straightening of the welded specimens is recom-
mended to reduce the effect of weld distortions.

ACKNOWLEDGMENTS

The authors would like to thank the German Research Foundation (DFG) for its financial
support on the projects DFG-No. 408366689.

REFERENCES

Banke, F. Schmied, J. Schulz, U. 2003. Der Einfluß von Schweißeigenspannungen und Schweißverformun-
gen auf das Beulverhalten von axialgedrückten Zylinderschalen. Stahlbau (77): 91–101.
Duan, Y. G. Vincent, Y. Boilot, F. Leblond, J. B. Bergheau, J. M. 2007. Prediction of welding residual
distortions of large structures using a local/global approach. Journal of Mechanical Science and Tech-
nology 21 (10):1700–1706.

888
Jäger, A. Pasternak, H. 2016. Studien zum Beulverhalten von eng ringversteiften Kreiszylinderschalen
unter Axialdruck. Bauingenieur (91):401–409.
Jäger-Cañás, A. Pasternak, H., 2017. Influence of closely spaced ring-stiffeners on the axial buckling
behavior of cylindrical shells. Eurosteel 2017, Copenhagen. Volume 1, 2-3, 928–937.
Kuzminov, S. A. 1974. Svarochnie deformazii sudovich korpusnich konstrukzii. Verlag Sudostroenie
Leningrad, Leningrad.
Loose, T. 2008. Schweißverzug, Schweißeigenspannungen und Schalenstabilität. Stahlbau (77): 111–119.
Pasternak, H. 2018. Versuche und grundlegende Studien zum Beulverhalten von eng ringversteiften
Kreiszylinderschalen unter Axialdruck, DFG-Project: 408366689.
Pasternak, H. Ciupack Y. 2014. Development of Eurocode-based design rules for adhesive bonded
joints. International Journal of Adhesion and Adhesives (53): 97–106.
Pasternak, H. Launert, B. Krausche, T. 2015. Welding of girders with thick plates - Fabrication, meas-
urement and simulation. Journal of Constructional Steel Research (115): 407–416.
Pircher, M. Bridge R. 1998. The Influence of Weld-Induced Residual Stresses on the Buckling of Cylin-
drical Thin-Walled Shells .Proc., 2nd Int. Conf. on Thin-Walled Structures (eds. N. E. Shanmugam et al),
671–678.
Rotter, J. Teng, J. 1989. Elastic Stability of Cylindrical Shells with Weld Depressions. Journal of Struc-
tural Engineering, American Society of Civil Engineers, vol. 115(5): 1244–1263.
Rykalin, N. N. 1957. Berechnung von Wärmevorgängen beim Schweißen, VEB Verlag Technik, Berlin.
Stapelfeld, C. 2016. Vereinfachte Modelle zur Schweißverzugsberechnung. Dissertation, Brandenburgische
Technische Universität Cottbus-Senftenberg, Aachen, Shaker Verlag.
Stapelfeld, C. Doynov, N. Michailov, V. 2009. Hybride Berechnungsansätze zur Prognostizierung und
Minimierung des Verzugs komplexer Schweißkonstruktionen. Sysweld Forum, 91–105.
Thikomirov, D. Rietman, B. Kose, K. Makkink, M. 2008. Computing Welding Distortion: Comparison
of Different Industrially Applicable Methods. SHEMET (11):195–202.

889
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Seismic response of steel dual eccentrically braced frames with


equal-strength joints

Č. Penelov & N. Rangelov


Department of Steel and Timber Structures, UACEG, Sofia, Bulgaria

ABSTRACT: Case studies on steel dual seismic resistant eccentrically braced frames (EBF)
for six- and ten-storey office buildings are presented. The beam-to-column joints are designed
as equal-strength joints according to the classification for performance objectives recently
introduced by the European EQUALJOINTS project. The seismic assessment of the dual
structures is performed by adequate nonlinear models for the potential dissipative zones: the
beam-to-column joints (end-plate connections and web panels) of the moment resisting frames
(MRF) as recommended by the EQUALJOINTS project, and the link elements of the EBF
subsystem, using hysteretic constitutive multi-linear models with kinematic strain-hardening.
Both nonlinear static pushover analysis and incremental dynamic analysis (IDA) are carried
out to assess the seismic response of the dual systems. The results are compared to those
obtained by the simple design procedures of EN 1998-1 based on elastic analysis.

1 INTRODUCTION

Along with the typical structural types of primary earthquake resisting structures, dual sys-
tems combining moment resisting frames (MRF) with concentrically (CBF) or eccentrically
braced frames (EBF) are quite promising for seismic applications. The main features of dual
systems are: high stiffness (due to CBF or EBF) plus high ductility (due to both subsystems),
prevention of a soft storey formation and recentring after strong earthquakes (due to the
MRF). On the other hand, their response to seismic actions is complex and additional
research is needed in this context.
In this paper, a study on two steel dual eccentrically braced frames for six- and ten-storey
office buildings is presented. Both subsystems (EBF and MRF) are designed to resist simul-
taneously the seismic action in the considered direction. The steel structures are designed to
EN 1998-1 (2004) with high ductility class (DCH). The moment resisting beam-to-column
joints are designed as equal strength extended stiffened (ES) end-plate bolted joints, using the
recommendations of the EQUALJOINTS European project (Landolfo et al. 2018) according
to which distribution of plastic deformations in the beams and the joints is permitted. The
design moment at the expected plastic hinge in the beam (at the tip of the rib stiffeners) is
equal to the nominal plastic moment resistance of the beam section.
Nonlinear static pushover analysis (NSA) is used to obtain the capacity curves of the frames.
Dynamic nonlinear analysis is carried out for a set of seismic records to estimate the structural
response of the dual systems to seismic actions. The elastic analysis is performed with SAP2000
software (CSI Inc.). The nonlinear analyses are performed with the OpenSees computational
framework (Mazzoni et al. 2007). The results are compared to those obtained by the simple design
procedure based on the elastic analysis and the capacity design methodology of EN 1998-1.

2 DESIGN OF THE EXAMPLE DUAL SYSTEMS

The example six- and ten-storey office buildings are located in a site with reference ground
acceleration agR = 0.35 g, ground type B, characterised by response spectrum type 1 according

890
to EN 1998-1. The building dimensions are 24 × 24 m with 6 m grid spacing in both directions.
The lateral load resisting systems are located in the building perimeter only, two dual frames
in each direction. Additional interior gravity columns carry only the vertical loads, which are
transferred by a floor framing system with simple primary and secondary beams. The con-
sidered dual frames are perpendicular to the secondary beams (spaced at 2 m). For example,
the considered six-storey dual frame is shown in Figure 1.
It is worth noting that in case of such a structural arrangement, for the considered dual sys-
tems the tributary width for gravity loads is 6/2 = 3 m, while the tributary area for the seismic
masses is half of the total floor area. Thus, from the seismic combination (Gk + ψ2Qk) the
concentrated force from a secondary beam is only F1 = 27.4 kN, while the storey mass for one
dual frame is 110 t.
The preliminary member design is based on the lateral force method (LFM) using the sim-
plified formula for the fundamental period of vibration in EN 1998-1, giving T1 = 0.74 s and
T1 = 1.08 s for the six- and ten-storey frame, respectively, and assuming the upper limit for
high ductility class (DCH) for the behaviour factor q = 6. The adopted cross-sections and the
influence of the P-Δ effects are then verified by modal response spectrum analysis (RSA).
As expected, the obtained by modal analysis fundamental periods are substantially differ-
ent: T1 = 1.02 s and T1 = 1.71 s for the two frames, respectively, therefore the behaviour fac-
tors in RSA are accordingly reduced to q = 4 and q = 3.5 to provide approximately equal
design base shear forces from LFM and RSA. The example structures appear non-sway. The
finally adopted member cross-sections are summarized in Table 1 and Table 2. EBF seismic
links are parts of the EBF beams. Since clause 6.10.2 (1) of EN 1998-1 is found somewhat
unsafe for the MRF subsystems, these are designed according to ASCE/SEI 7-16 (2017) to
resist 25% of the design seismic forces acting on the whole dual system.

3 INCREMENTAL DYNAMIC ANALYSIS

Several sets of nonlinear dynamic analysis (NDA) have been performed with variation of the
seismic intensity level (i.e. incremental dynamic analysis, IDA). A set of seven accelerograms
is adopted, particularly: Hollister, Imperial Valley, Kobe, Kocaeli, Loma Prieta, Northridge,
Trinidad, taken from SeismoMatch (Seismosoft 2016) library. The seismic records are
matched using a wavelet algorithm so that their elastic response spectra are compatible with
the elastic response spectrum type 1 of EN 1998-1. According to EN 1998-1 the results of the
NDA should be definеd as averaged results obtained from the seven accelerograms used.
The seismic assessment of the dual structures is performed by adequate nonlinear models for
the potential dissipative zones: the beam-to-column joints (end-plate connections and web panels)
of MRF as recommended by the EQUALJOINTS project (Landolfo et al. 2018); the link

Figure 1. Scheme of the example six-storey dual EBF-MRF system with ES beam-to-column joint.

891
Table 1. Member cross-sections of the six-storey dual system.
EBF beams MRF beams
Storey (B2) Diagonals (B1, B3) Columns

1 HEA 260 SHS 160×12 IPE 240 HEB 320


2 HEA 260 SHS 160×12 IPE 240 HEB 320
3 HEA 260 SHS 160×12 IPE 240 HEA 320
4 HEA 220 SHS 140×10 IPE 240 HEA 320
5 HEA 180 SHS 140×10 IPE 240 HEA 320
6 HEA 140 SHS 120×8 IPE 240 HEA 320

Table 2. Member cross-sections of the ten-storey dual system.


EBF beams MRF beams
Storey (B2) Diagonals (B1, B3) Columns

1 HEA 280 SHS 180×12 IPE O 240 HEM 320


2 HEA 300 SHS 180×12 IPE O 240 HEM 320
3 HEA 300 SHS 180×12 IPE O 240 HEB 320
4 HEA 280 SHS 160×12 IPE O 240 HEB 320
5 HEA 260 SHS 160×12 IPE O 240 HEB 320
6 HEA 240 SHS 160×12 IPE O 240 HEB 320
7 HEA 220 SHS 140×10 IPE O 240 HEA 320
8 HEA 200 SHS 140×10 IPE O 240 HEA 320
9 HEA 180 SHS 140×10 IPE O 240 HEA 320
10 HEA 140 SHS 120×8 IPE O 240 HEA 320

elements of the EBF subsystem using hysteretic constitutive multi-linear model with kinematic
strain-hardening according to Richards & Uang (2006) and the plastic hinges in the other steel
members (beams, columns and diagonals) following the recommendations of ASCE/SEI 41-06
(2007). More details on modelling of the plastic zones can be found in Penelov et al. (2018).

4 RESULTS FROM NONLINEAR DYNAMIC ANALYSIS

4.1 Incremental and pushover capacity curves


The ‘maximum base shear force – maximum roof displacement’ relationships (the incremental
capacity curves) obtained from IDA (using material overstrength factor γov = 1.25), are shown
in Figure 2. The markers of these curves correspond to the accelerogram scale factors (SF)
varying from 0 to 1.72 (up to maximum considered earthquake MCE with return period TR =
2475 years, associated with near collapse (NC) limit state).
For the six-storey frame the obtained incremental capacity curves lay between the capacity
curves obtained by pushover analysis with ‘modal’ and ‘uniform’ lateral load pattern, respect-
ively. For the ten-storey frame, however, the pushover analysis even with ‘uniform’ pattern
appears to underestimate the base shear capacity for some of the records, due to the more
pronounced influence of the higher vibration modes. This effect leads to an increase of the
IDA base shears compared to those obtained by pushover analysis with ‘modal’ lateral force
pattern. At the same time, the ratio between the maximum base shear obtained from IDA and
the base shear obtained from the RSA is about 2.5, which should be accounted for in the
design of the column bases and the foundations.

4.2 Behaviour of the MRF subsystem and the ES beam-to-column joints


The nominal bending resistances of the MRF beams and the EBF columns are reached only
at seismic action intensities close to MCE. This means that the MRF component subsystem
will remain almost elastic after a strong earthquake and will provide recentring action to the

892
Figure 2. Incremental and pushover capacity curves for the studied frames.

whole structure. The elastic behaviour of the MRF beams leads to almost elastic behaviour of
the ES bolted joints. The magnitude of the shear forces in the panel zones of the inner columns
of the example structures never exceeds the design shear forces obtained for the panel zones of
the single sided joints. Therefore, it seems more appropriate to design the internal joints of the
dual system according to the design procedure for single sided beam-to-column joint configur-
ation, using the properties of the MRF beam cross-section.

4.3 Maximum interstorey drift ratios (IDR)


As it can be observed in Figures 3, 4 the results from the elastic response spectrum analysis
(RSA) according to EN 1998-1 for seismic action with return period TR = 95 years (damage
limitation limit state, DL; SF = 0,5) are similar to the mean interstorey drift ratios from NDA.
The common practice to calculate the interstorey drifts from RSA as the difference between the
maximum storey displacements (as adopted in EN 1998-1) seems not justifiable from the point
of view of the structural dynamics (Newmark & Hall 1982) and may lead to underestimation of
the interstorey drifts at the upper storeys, when the response of the structure is elastic or with
limited plastic deformations. This is more pronounced for the braced frames. For DL limit
state the estimating of IDRs based on the so called ‘generalised displacements’ (shown as EC8
RSA GD in Figures 3, 4), an option available in SAP2000 software, gives a more realistic IDR
values for the storeys in the upper one third of the building height. Anyway, the maximum IDR
appears below the assumed design DL interstorey limiting drift ratio of 1%.
RSA based on EN 1998-1 and the ‘equal displacement rule’ overestimates IDR for the
design earthquake (DE) with return period TR = 475 years (associated with significant
damage limit state, SD; SF = 1) and for MCE for the upper storeys due to overestimation of
the drift components resulting from the elastic column axial deformations. This seems more
pronounced for the ten-storey frame.

4.4 Plastic rotations of the seismic links


Firstly, the link plastic rotations for DE and MCE are estimated based on the plastic compo-
nents of the expected interstorey drifts obtained from the elastic analysis (based on the ‘equal
displacement rule’), and using the kinematics of the EBF storey plastic collapse mechanisms
as illustrated in Figure 5. The displacement amplification factor is assumed equal to the
behaviour factor q.

893
Figure 3. Maximum interstorey drift ratios (IDRs) of the six-storey dual frame obtained from NDA
(with γov = 1.25) for the three seismic intensity levels: related to DL criteria; design earthquake (DE) and
maximum considered earthquake (MCE).

Figure 4. Maximum interstorey drift ratios (IDRs) of the ten-storey dual frame obtained from NDA
(with γov = 1.25) for the three seismic intensity levels: DL; DE and MCE.

Figure 5. Kinematical scheme for determination of link rotation angle.

It appears that for mid and high-rise EBFs this method overestimates the link plastic rota-
tions in the upper stories compared to the mean results from the nonlinear dynamic analysis.
For this reason, it is recommended (Richards & Thompson 2009) to calculate the plastic com-
ponents of the interstorey drifts considering ‘shear-only’ storey behaviour (neglecting the con-
tribution from the column axial deformations).

894
The plastic rotations of the link elements obtained from NDA (with the conservative
assumption for a lack of material overstrength, i.e. γov = 1) for two seismic intensity levels,
DE and MCE, are shown in Figures 6, 7. The structural behaviour for DE appears quite
adequate because the mean and even the maximum plastic rotations from the dynamic ana-
lysis appear smaller than the EN 1998-1 limit for DE for short links (0.08 rad). Apparently,
the maximum plastic rotations for MCE are close to the ultimate rotation of 0.12 rad accord-
ing to ASCE 41-06 (for links with two intermediate stiffeners), which still can be considered
acceptable. The simplified procedure based on the elastic analysis gives a good preliminary
estimate of the mean link plastic rotations. According to EN 1998-1 the mean values of the
plastic rotations for the DE should be used for determination of the required number of the
intermediate stiffeners of the links, see Figure 8. In the same figure the hysteretic behaviour of
the 3rd storey seismic link of the six-storey dual frame under the matched seismic record Loma
Prieta scaled for the MCE intensity is shown.

4.5 Dual system recentring capability


The maximum residual roof displacements obtained for the MCE are under 0,3% of the build-
ing height. Additionally, the maximum residual IDRs for the MCE are around 0,5% (Figure 9),
and, referring to the criteria of McCormick et al. (2008), these results confirm a very good
recentring capability of the example dual systems and thus prove the possibility for cost-
effective repair even after the seismic action corresponding to MCE.

Figure 6. Link plastic rotations for the six-storey frame from DE (SF = 1) and MCE (SF = 1.72).

Figure 7. Link plastic rotations for the ten-storey frame from DE (SF = 1) and MCE (SF = 1.72).

895
Figure 8. Structural detail and hysteretic behaviour of the 3rd storey link of the six-storey dual frame.

Figure 9. Time history plots of IDRs of the six-storey frame from Hollister matched accelerogram
(SF = 1.72).

Table 3. Comparison between the maximum bending moments (kNm) at the column bases.
Obtained from RSA and EN 1998-1
capacity design procedure for Mean values from NDA for

Dual MRF NC limit


Frame EBF systems systems SD limit state state

Six-storey 160 314 284 483


Ten-storey 380 549 556 902

4.6 Internal forces in diagonals and columns


As suggested in Vayas (2017), the determination of these internal forces should be based on
the EN 1998-1 methodology for capacity design of non-dissipative members of EBF systems,
using ΩEBF factor. The IDA conducted herein shows that this approach predicts well the
column axial forces for the NC limit state but underestimates the maximum column bending
moments even for the SD limit state. The reliable prediction of the maximum bending
moments in the first storey columns is essential for the design of the column bases and for
ensuring their elastic behaviour which contributes also to the recentring capability of the
structure. The methodology for capacity design of the columns as a part of MRF system given
in EN 1998-1 covers better column bending moment results obtained from NDA for the SD
limit state but also underestimates those bending moments for the NC limit state (Table 3).

5 CONCLUSIONS

In this paper case study on two dual eccentrically braced frames for a six- and ten-storey
office buildings is presented. The seismic assessment is performed by static nonlinear pushover

896
analysis and incremental dynamic analysis. The latter is carried out for a set of selected
matched seismic records to estimate the structural response to seismic actions of the dual
system. The beam-to-column joints of the MRF are designed as equal strength joints, accord-
ing to the classification of the EQUALJOINTS European project (Landolfo et al. 2018) with
extended stiffened end plate bolted connections.
Recentring capability of the dual systems is observed even under maximum considered
earthquake due to the elastic behaviour of the MRF subsystem and the significant strain-
hardening of the seismic links. As a result of the MRF elastic behaviour under strong earth-
quake and the large stiffness of EBF subsystem, the actual ‘moment-rotation’ behaviour of
the equal strength extended stiffened joints of the MRF subsystem has a very limited effect on
the global structural behaviour. The response spectrum analysis performed according to EN
1998-1 predicts well the interstorey drift ratios (IDR) for the damage limitation (DL) limit
state but overestimates IDRs for the upper stories for design earthquake and maximum con-
sidered earthquake, due to the overestimation of the drift components resulting from the
column axial deformations. This is more pronounced for the ten-storey frame. The simplified
procedure using the ‘shear-only’ components of the interstorey drifts and the kinematics of
the plastic collapse storey mechanisms (Figure 5), gives a good preliminary estimate of the
seismic link mean plastic rotations. To determine the column bending moments for NC limit
state it is advisable to apply nonlinear dynamic analysis.
The research work presented herein raises also some interesting points for discussion. It
seems that the concept of dual systems and the proper design of the two component subsystems
are not clarified in details in the current EN 1998-1. The design of columns that are common
for both subsystems may be also ambiguous. In EN1998 there are no codified hysteretic models
and acceptance criteria for some types of dissipative elements (EBF links, for example).

ACKNOWLEDGEMENTS

The research has been carried out within the European EQUALJOINTS PLUS Project,
funded by the Research Fund for Coal and Steel (RFCS) of the European Commission under
grant agreement № 754048. This support is gratefully acknowledged.

REFERENCES

ASCE/SEI 41-06 2007. Seismic rehabilitation of existing buildings.


ASCE/SEI 7-16 2017. Minimum design loads for buildings and other structures.
EN 1998-1: 2004/AC: 2009. Eurocode 8: Design of structures for earthquake resistance. Part 1: General
rules, seismic actions and rules for buildings.
Landolfo, R., D’Aniello, M., Costanzo, S., Tartaglia, R., Demonceau, J.-F., Jaspart, J.-P., Stratan, A.,
Jakab, D., Dubina, D., Elghazouli, A., Bompa, D. 2018. Volume with information brochures for 4 seis-
mically qualified joints (EQUALJOINTS), ECCS.
Mazzoni, S., McKenna, F., Scott, M.H., Fenves, G.L. et al. 2007. OpenSees command language manual.
McCormick et al. 2008. Permissible residual deformation levels for building structures considering both
safety and human elements. 14WCEE, Beijing, China.
Newmark, N.M. & Hall, W.J. 1982. Earthquake spectra and design. Berkeley: EERI.
Penelov, Č., Hadzhiyaneva, I., Rangelov, N. 2018. A numerical study on the seismic response of a
six-storey steel eccentrically braced dual system with equal strength joints. Proceedings of the XVIII
Anniversary International Scientific Conference “Construction and Architecture” VSU’2018, Sofia.
Richards, P.W., Uang, C.-M. 2006. Testing protocol for short links in EBFs, ASCE Journal of structural
engineering, 132:8, 2006.
Richards, P.W. & Thompson, B. 2009. Estimating inelastic drifts and link rotation demands in EBFs,
Engineering journal, AISC, vol. 46, pp 123–136.
SAP 2000, CSI, Computers and Structures Inc, www.csiberkeley.com.
SeismoMatch, version 2016, Seismosoft, www.seismosoft.com.
Vayas I. (ed.) 2017. INNOSEIS: Innovative anti-seismic devices and systems. ECCS.

897
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Tests and design of built-up section columns

Dang Khoa Phan & Kim J.R. Rasmussen


The University of Sydney, Sydney, New South Wales, Australia

ABSTRACT: Built-up sections are increasing in popularity as structural elements in the


cold-formed steel (CFS) industry. They are composed of two or more component sections
connected by discrete fasteners, most commonly screws or bolts, typically spaced evenly along
the length. Conventionally, two singly symmetric sections are connected to form a doubly
symmetric cross-section. However, this study focuses on sections formed by three or more
lipped channel C-sections, which are being used more frequently in recent years. The paper
details an experimental investigation of the strength and behaviour of screw-fastened built-up
columns with variable lengths. The cross-section configurations include singly-symmetric col-
umns composed of three channel sections (3C), for which the failure mode was either local,
distortional and/or flexural-torsional modes, and doubly-symmetric columns formed by four
sections (4C) which failed in local and/or flexural buckling modes.

1 INTRODUCTION

The recommendation for the modified slenderness (kL/r)m of columns composed of two sec-
tions connected together at discrete locations is specified in Section I1.2 of AISI S100 (2016).
Because the buckling involves relative deformations that induce shear forces in the fasteners
between component members, the modified slenderness is calculated as follows:
  s ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
  2ffi
kL kL 2 a
¼ þ ð1Þ
r m r o ri

where (kL/r)o is the overall slenderness of the built-up section about the built-up member axis
assuming full composite action, a is the longitudinal centre-to-centre distance between two
adjacent rows of fasteners, and ri is the minimum radius of gyration of full unreduced cross-
sectional area of an individual section.
Equation (1) is derived from studies of welded hot-rolled members (Zandonini, 1985, Aslani
and Goel, 1991) with relatively high shear stiffness of the weld fasteners. In this paper, screw
fasteners are considered for which the connection shear stiffness depends on the type and size
of fastener as well as the thickness and material of connected components. Section I1.2 of
AISI S100 (2016) also specifies the provision for maximum fastener spacing along the length
as shown in Equation (2). This limitation on the fastener spacing is imposed to ensure that
flexural buckling of any individual member between fastener points does not occur before
buckling of the built-up member. The one-half factor (1/2) is a reduction factor which
accounts for any one of the fasteners being loose or ineffective (AISI S100, 2016).
 
a 1 kL
 ð2Þ
ri 2 r o
The North American Specification (AISI S100, 2016) recommends two methods for the basic
design of compression members, viz. the Effective Width Method and the Direct Strength
Method. While the former method has been utilized by engineers for more than two decades,
providing a comprehensive and reliable design solution, the latter is a more robust and

898
contemporary method (AISI S100, 2016). Using the Effective Width Method as specified in
Section E3.1 of AISI S100 (2016), the nominal axial strength, Pnl, for local buckling interact-
ing with yielding and global buckling can be calculated as follows:

Pnl ¼ Ae fn ð3Þ

where Ae is the effective area (determined in accordance with Sections E3.1.1 and E3.1.2 of
AISI S100 (2016)), calculated at the global column stress fn, which is:
8
< 0:658λc fy for λc  1:5
2

fn ¼  ð4Þ
: 0:877 fy for λ c >1:5
λ 2
c

pffiffiffiffiffiffiffiffiffiffiffiffiffi
where λc ¼ fy =fcre , fy is the yield stress, and fcre is the minimum of the elastic torsional, flex-
ural and flexural-torsional buckling stresses.
Authors, including Reyes and Guzmán (2011), Whittle and Ramseyer (2009) and Stone and
LaBoube (2005), validated their experimental results against the design strength obtained by
using the Effective Width Method and the modified slenderness ratio for built-up members.
Besides the Effective Width Method, a number of authors have compared test strengths against
strengths predicted by the Direct Strength Method (DSM) with the aim of proposing modifica-
tions to the current DSM such that the method is applicable to built-up sections, such as Young
and Chen (2008), Georgieva et al. (2012a), Li et al. (2014), Zhang and Young (2018a), Zhang and
Young (2018b), Zhang and Young (2015), Zhang and Young (2012), Fratamico et al. (2018a),
Fratamico et al. (2018b). In the DSM, The nominal axial strength (Pn), specified in Sections E2
and E3.2 of AISI S100 (2016) and Section 7 of AS/NZS 4600 (2018), is as follows:

Pn ¼ minimum ofðPne ; Pnl ; Pnd Þ ð5Þ

where Pne, Pnl and Pnd are the nominal member strengths for global (flexural, torsional or flex-
ural-torsional), local and distortional buckling in compression, respectively, specified as follows:

Pne ¼ Ag fn ð6Þ

where Ag is the gross area and fn is specified in Equation (4);


8
< Pne for λl  0:776
 0:4  0:4
Pnl ¼ ð7Þ
: 1  0:15 Pcrlne
P Pcrl
Pne Pne for λl 40:776

pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
where λl ¼ Pne =Pcrl , Pcrl ¼ Ag fol , and fol is the elastic local buckling stress; and
8
< Py for λd  0:561
 0:6  0:6
Pnd ¼ ð8Þ
: 1  0:25 PPcrdy Pcrd
Py Py for λd 40:561

pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
where λd ¼ Py =Pcrd ,Py ¼ Ag fy , Pcrd ¼ Ag fod , and fod is the elastic distorsional buckling stress.
Although the effect of fastener looseness on the member strength has been reported in
a number of studies (Fratamico et al., 2018b, Niu et al., 2015, Georgieva et al., 2012a, Geor-
gieva et al., 2012b, Young and Chen, 2008), it has not been quantified and studied thoroughly.
Instead of testing conventional built-up sections formed by two single sections, the built-up
CFS sections in this paper were composed of three or more lipped channel sections. The test
columns were compressed under different end conditions (i.e. pinned and fixed) for different
lengths and fastener configurations. The test strengths are reported and compared with DSM
strength predictions.

899
2 EXPERIMENTAL STUDY

Before the compression tests, the connection stiffness and the geometric imperfections of each
built-up column were measured for the later use in finite element modelling and column design
calculations. The specimen ends were milled flat to ensure full contact with the loading plates of
the column test rig. Figure 1(a) shows a schematic drawing of the test set-up. The tests were con-
ducted in J.W. Roderick Materials and Structures Laboratory at the University of Sydney. A 16-
mm layer of pattern stone was poured enclosing the cross-section at either end in order to create
a rigid interaction between the specimen end and the loading plate. Figures 1(b) and 1(c) show the
geometry, orientation and position of the two cross-section configurations in the test setup. While
3C was a singly-symmetric section formed by connecting three single C-sections back-to-back, 4C
was doubly-symmetric consisting of four single sections and composing a square shape in the
middle. The component C-sections were C10010 lipped channel high-strength zinc-coated G550
steel with a nominal yield stress of 550 MPa. The measured values for the elastic modulus, the
static yield strength and ultimate strength were obtained from coupon tests, i.e. E = 213,300 MPa,
fy = 608 MPa and fu = 621 MPa, respectively. The nominal cross-section geometry of the C-section
is shown in Table 1. The nominal ratio of web depth to thickness (d/t) was 102. By using the com-
puter program Thin-Wall (Centre for Advanced Structural Engineering, 2019), the elastic local
(fol) and distortional (fod) buckling stresses for the C10010 section were found to be 111.1 MPa
and 205.1 MPa, respectively.
As shown in Figure 1(a), a 500 kN hydraulic actuator was used for applying monotonic
compression load on 500 mm, 2100 mm and 6000 mm long specimens, with displacement con-
trol rates of 0.1 mm/min, 0.3 mm/min and 0.6 mm/min, respectively. With the hinge assemblies
being enabled for the tests of the 2100 mm and 6000 mm columns, the loading plates were free

Figure 1. (a) Schematic drawing of test set-up, and cross-sections: (b) 3C and (c) 4C (elevation).

Table 1. Nominal geometry of a single C10010 section.


Ix Iy
Thickness Web Flange Lip Area J Iw
3 4
Section (mm) (mm) (mm) (mm) (mm) (10 mm ) (mm4) (106 mm6)

C10010 1.0 102 51 12.5 215.3 363 75.5 71.78 160.2

900
to rotate about the vertical axis, which created the pinned end condition. While the 3C built-
up columns were free to rotate about the major axis, the 4C specimens were oriented to be
free to rotate about the minor axis. The effective length (Le) is defined as the distance between
the two hinges for the pinned end condition, viz. Le = L + 150 mm, where L is the specimen
length (see Figure 1(a)) and the total dimension of the end platens is 150 mm. In the tests of
2100 mm and 6000 mm long specimens, the columns were positioned in the rig with a nominal
horizontal eccentricity (e) of Le/1500. The direction of eccentricity is specified in Figure 1(b)
and also labelled in Figures 3(c) to 3(f). On the other hand, the stub column tests of the
500 mm long specimens were conducted under a fixed-ended condition. For these tests, the
effective length was taken as half the specimen length, i.e. Le = L/2.
Table 2 summarizes the information about the test series, viz. cross-section configuration,
column length, boundary condition and screw configuration. Gauge 14 screws with a nominal
diameter of 6.3 mm (Buildex, 2019) were used to connect individual sections intermittently.
Most test columns had screws evenly spaced along the member length at distance (a) ranging
from 100 mm to 983 mm as shown in Figure 2 (Top). Two special columns with even screw
spacing of 492 mm also had an extra set of screws at each end to form an end fastener group
(EFG) as shown in Figure 2 (Bottom). The two columns were 6000 mm long and labelled as
3C-L6000-S492-EFG-1 and 4C-L6000-S492-EFG-1 in Tables 3 and 4 in the results section.
Each EFG had a length of 160 mm (i.e. 1.5 times the web depth) and screw spacing of 20 mm
(i.e. not more than four diameters apart), following Section I1.2(b) of AISI S100 (2016). The
EFGs were utilized to minimize the relative slip among component sections, which could
improve the global (flexural) capacity of the columns. Note all the test specimens had screw
spacing satisfying Equation (2) as required in Section I1.2 of AISI S100, and the screw end
distance of 50 mm satisfying Section J4.2 of AISI S100 (2016).
The experimental ultimate strengths (Ptest) were compared with DSM strength predictions
(PDSM) as per AISI S100 (2016) and AS/NZS 4600 (2018). For the calculations of the DSM
strengths (PDSM), the actual geometry of the 3C and 4C sections was transcribed to paper to
determine the actual area (A) and the actual second moments of area about the major and
minor axes, respectively, assuming full composite action. The modified slenderness and PDSM
values were then determined (see Tables 3 and 4 for the values of PDSM).

Table 2. Test matrix.


End Approximate screw
conditions spacing (mm)

Specimen Effective Rotation


Section length length freedom Eccentricity 500 &
configuration (L, mm) (Le, mm) Type about (e) 100 500 1000 EFG

500 250 Fixed Nil Nil ✓


3C 2100 2250 Pinned Major axis Le/1500 ✓ ✓
6000 6150 Pinned Major axis Le/1500 ✓ ✓ ✓

500 250 Fixed Nil Nil ✓


4C 2100 2250 Pinned Minor axis Le/1500 ✓ ✓
6000 6150 Pinned Minor axis Le/1500 ✓ ✓ ✓

Figure 2. Evenly spaced screws (Top), and evenly spaced screws plus end fastener groups (Bottom).

901
Table 3. Test results for 3C columns.
Length a Ptest Buckling PDSM Ptest/ Ptest/
Specimen (mm) (mm) (kN) mode (kN) PDSM Py

3C-L500-S100-1 500 100 180.4 L 178.2 1.01 0.46


3C-L500-S100-2 500 100 178.2 L 178.0 1.00 0.45
3C-L2100-S500-1 2100 500 160.3 L+FT 140.4 1.14 0.41
3C-L2100-S500-2 2100 500 156.9 L+FT 141.0 1.11 0.40
3C-L2100-S100-1 2100 100 166.7 L+FT 150.4 1.11 0.42
3C-L2100-S100-2 2100 100 166.2 L+FT 150.0 1.11 0.42
3C-L6000-S983-1 6000 983 62.4 L+D+FT 51.9 1.20 0.16
3C-L6000-S492-1 6000 492 56.6 L+D+FT 55.5 1.02 0.14
3C-L6000-S492-2 6000 492 56.8 L+D+FT 55.6 1.02 0.14
3C-L6000-S492EFG-1 6000 492+EFG 62.4 L+D+FT 55.6 1.12 0.16
Mean 1.08
CoV 0.06

* L = local, D = distortional, FT = flexural-torsional, CoV = coefficient of variation.

Table 4. Test results for 4C columns.


Length a Ptest Buckling PDSM Ptest/
Specimen (mm) (mm) (kN) mode (kN) PDSM Ptest/Py

4C-L500-S100-1 500 100 239.8 L 237.4 1.01 0.46


4C-L500-S100-2 500 100 238.8 L 237.5 1.01 0.46
4C-L2100-S500-1 2100 500 198.6 L+F 183.6 1.08 0.38
4C-L2100-S500-2 2100 500 197.1 L+F 184.6 1.07 0.38
4C-L2100-S100-1 2100 100 207.0 L+F 194.6 1.06 0.40
4C-L6000-S983-1 6000 983 66.6 L+F 60.8 1.10 0.13
4C-L6000-S492-1 6000 492 64.8 L+F 64.8 1.00 0.12
4C-L6000-S492-2 6000 492 65.1 L+F 63.9 1.02 0.12
4C-L6000-S492EFG-1 6000 492+EFG 63.2 L+F 64.1 0.99 0.12
Mean 1.03
CoV 0.04

* F = flexural.

3 EXPERIMENTAL RESULTS

The test series is aimed at studying the effect of varying cross-sectional configurations, specimen
lengths, screw arrangements and end conditions on the member strength and failure behaviour.
The test results are reported for the two cross-sectional configurations, viz. 3C and 4C, in
Tables 3 and 4, respectively. The general labelling convention of the tests is “Section configur-
ation - Column length - Screw spacing - Serial number”. For example, “3C-L6000-S492-2”
implies a built-up section formed by three single C-sections, a length of 6000 mm, screws evenly
spaced at 492 mm, and the 2nd test of the configuration. A special label “S492EFG” implies
that the end fastener groups are installed in addition to the existing screws which are evenly
spaced at 492 mm. In general, the average ratio of test strength to the strength predicted by the
DSM is 1.08 for the 3C specimens and 1.03 for the 4C specimens. The ratios of the test strength
to squash load (Ptest/Py) is smaller for longer (more slender) columns as expected. For the stub
column tests under the fixed end condition, all 500 mm specimens experienced local buckling
only (see Figures 3(a) and 3(b)). From Tables 3 and 4, the average test strength ratio (Ptest/Py)
for the 3C to 4C stub (500 mm) columns is exactly three quarters (3/4).
Since the 2100 mm and 6000 mm 3C specimens were free to rotate about the major axis, flex-
ural-torsional and local buckling was observed for 2100 mm columns (Figure 3(c)) while flexural-
torsional buckling interacting with minor local and distortional mode was observed for 6000 mm

902
Figure 3. Post-peak behaviour.

specimens (Figure 3(e)). The twisting about the member axis for the 2100 mm and 6000 mm col-
umns was pronounced and clearly observable near and after the peak load. For the 6000 mm 3C
specimens, the observed lip localised failures, distortional buckling (Figure 3(e)) and web elastic
local buckling (not visible in Figure 3(e)) in the middle third of the specimen length could possibly
be due to the flexural compressive stresses resulting from global (flexural-torsional) buckling.
Interestingly, the lip failures occurred at or very close to the longitudinal positions of screw rows,
which constituted the troughs of the wave shape of the deformed lip (Figure 3(e)).Under the
pinned end condition which allowed the rotation about the minor axis, both the 2100 mm and
6000 mm 4C columns experienced the interaction of local and flexural buckling (see Figures 3(d)
and 3(f)). Even with substantially higher member slenderness for the 6000 mm 4C specimens, the
presence of web local buckling (see Figure 3(f)) was because this web sustained large compressive
flexural stresses due to flexural buckling (it is located furthest away from the minor axis and on
the compressive side). Figure 4 shows typical test graphs for the flexural-torsional and flexural
behaviour for specimens 3C-L6000 and 4C-L6000, respectively. The use of two transducers for

903
Figure 4. Test graphs for 3C and 4C 6000 mm specimens.

measuring the lateral displacement and twist rotation at mid-length of 3C-L6000 failed to obtain
the last part of the post-peak curves because of the abrupt failure triggered by sectional buckling
during which the load dropped steeply and the transducers lost contact with the specimen.
The effect of using more screw rows, including end fastener groups, on the member strength
shall be discussed in the context of the cross-section slenderness, the rigidity of the end condition
and the screw size in comparison with the findings by Fratamico et al. (2018a), Fratamico et al.
(2018b), Fratamico et al. (2016). As shown in Tables 3 and 4, the increase in the number of fas-
teners does not clearly improve the capacity of the test columns (only increases of 4.9% and 5.2%
for 3C-L2100 and 4C-L2100 specimens, respectively, and no sign of improvement for L6000 spe-
cimens). This behaviour can be attributed to the rigid end platens used in the tests. The test speci-
men ends were milled flat to ensure full contact with loading plates. This, together with the use of
pattern stone, simulated an end condition at which the sections could barely move relative to
each other, thus assisting in minimizing the relative longitudinal slips among the component mem-
bers of built-up sections. In comparison, the end condition in the tests by Fratamico et al. (2018b)
behaved like a semi-rigid condition as the end cross-sections were not originally flush with the
webs of the loading tracks. Therefore, the role of the end fastener groups was more prominent in
preventing the shear slips under the semi-rigid end condition.
The second reason could be related to the slenderness of the component C10010 sections in
this study (d/t = 102). All the tests observably experienced local buckling as the single mode of
buckling or in combination with other modes. According to Fratamico et al. (2018b), the use of
EFGs only increases the member capacity of built-up columns formed by relatively stocky com-
ponent sections, but does not increase the capacity of slender columns when local buckling is
pronounced. Fratamico et al. (2018a), Fratamico et al. (2016) also observed that the use of
more screw rows (even at spacing as small as 19 mm) does not meaningfully increase the cap-
acity of built-up columns failing in local and/or distortional buckling. Further work is required
to validate the second point, viz. conducting finite element analysis with more compact sections
(which experience isolated global buckling, e.g. commercially available C10019 sections).

4 CONCLUSIONS

This study focuses on testing screw-fastened built-up cold-formed steel sections with innovative
shapes, various lengths, fastener configurations and boundary conditions. The cross-sections
include singly-symmetric columns constructed by three sections buckling in local, distortional
and/or flexural-torsional modes, and doubly-symmetric columns composed of four channel sec-
tions buckling in local and/or flexural modes. The Direct Strength Method was found to give reli-
able results for the test specimens. With the rigid end platens used to load the test specimens, the
increase in the number of screw rows and the use of end fastener groups did not boost the
observed member capacity. Referring to the study by Fratamico et al. (2018b) with semi-rigid end
conditions, this study provides a supplement to proving the significance of rigid end platens
coupled with fasteners in preventing relative slip among component members and thus improving

904
the built-up member flexural buckling strength. However, the significance of fasteners in increas-
ing member capacity was compromised when local buckling was pronounced. Following this
experimental study, a finite element analysis will be conducted for the purpose of validating
experimental results and proposing design curves. The finite element modelling will then be
expanded to study the effect of fastener configurations on the strength of more compact built-up
sections which fail in isolated global buckling when loaded between rigid end platens. This paper
forms part of a broader research program on built-up sections, including columns, beams and
sections failing in local, distortional and global modes to determine the buckling strength of sec-
tions with discrete fasteners subject to contact constraints between abutting surfaces.

REFERENCES

AISI S100 (2016) North American specification for the design of cold-formed steel structural members,
AISI S100-16. Washington, D.C, American Iron and Steel Institute.
AS/NZS 4600 (2018) Cold-formed Steel Structures, AS/NZS 4600: 2018Sydney, Australia, Standards
Australia/Standards New Zealand.
Aslani, F. & Goel, S.C. (1991) An analytical criterion for buckling strength of built-up compression
members. Engineering Journal, AISC, 28, 159–168.
BUILDEX (2019) Metal Teks. Australia, Buildex.
Centre For Advanced Structural Engineering (2019) THIN-WALL. 2. 1.48ed. Sydney, The University of
Sydney.
Fratamico, D.C., Torabian, S., Rasmussen, K.J. & Schafer, B.W. (2016) Experimental investigation of
the effect of screw fastener spacing on the local and distortional buckling behavior of built-up
cold-formed steel columns. Proceedings of the Wei-Wen Yu international specialty conference on cold-
formed steel structure.
Fratamico, D.C., Torabian, S., ZHAO, X., Rasmussen, K.J. & Schafer, B. W. (2018a) Experimental
study on the composite action in sheathed and bare built-up cold-formed steel columns. Thin-Walled
Structures, 127, 290–305.
Fratamico, D.C., Torabian, S., Zhao, X., Rasmussen, K.J. & Schafer, B.W. (2018b) Experiments on the
global buckling and collapse of built-up cold-formed steel columns. Journal of Constructional Steel
Research, 144, 65–80.
Georgieva, I., Schueremans, L. & PYL, L. (2012a) Composed columns from cold-formed steel Z-profiles:
Experiments and code-based predictions of the overall compression capacity. Engineering Structures, 37,
125–134.
Georgieva, I., Schueremans, L., Pyl, L. & Vandewalle, L. (2012b) Experimental investigation of built-up
double-Z members in bending and compression. Thin-Walled Structures, 53, 48–57.
Li, Y., Li, Y., Wang, S. & Shen, Z. (2014) Ultimate load-carrying capacity of cold-formed thin-walled col-
umns with built-up box and I section under axial compression. Thin-Walled Structures, 79, 202–217.
Niu, S., Rasmussen, K.J.R. & Fan, F. (2015) Local–Global Interaction Buckling of Stainless Steel
I-Beams. I: Experimental Investigation. Journal of Structural Engineering, 141, 04014194.
Reyes, W. & Guzmán, A. (2011) Evaluation of the slenderness ratio in built-up cold-formed box sections.
Journal of Constructional Steel Research, 67, 929–935.
STONE, T.A. & LABOUBE, R.A. (2005) Behavior of cold-formed steel built-up I-sections. Thin-Walled
Structures, 43, 1805–1817.
Whittle, J. & Ramseyer, C. (2009) Buckling capacities of axially loaded, cold-formed, built-up
C-channels. Thin-Walled Structures, 47, 190–201.
Young, B. & Chen, J. (2008) Design of Cold-Formed Steel Built-Up Closed Sections with Intermediate
Stiffeners. Journal of Structural Engineering, 134, 727–737.
Zandonini, R. (1985) Stability of compact built-up struts: experimental investigation and numerical simu-
lation. Construzioni metalliche, 288.
Zhang, J.-H. & Young, B. (2012) Compression tests of cold-formed steel I-shaped open sections with
edge and web stiffeners. Thin-Walled Structures, 52, 1–11.
Zhang, J.-H. & YOUNG, B. (2015) Numerical investigation and design of cold-formed steel built-up
open section columns with longitudinal stiffeners. Thin-Walled Structures, 89, 178–191.
Zhang, J.-H. & Young, B. (2018a) Experimental investigation of cold-formed steel built-up closed section
columns with web stiffeners. Journal of Constructional Steel Research, 147, 380–392.
Zhang, J.-H. & Young, B. (2018b) Finite element analysis and design of cold-formed steel built-up closed
section columns with web stiffeners. Thin-walled Structures, 131, 223–237.

905
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Buckling and strength of prestressed steel stayed columns

R. Pichal & J. Machacek


Faculty of Civil Engineering, Czech Technical University in Prague, Prague, Czech Republic

ABSTRACT: Prestressed steel stayed columns have been used since fiftieth of the last century
to enhance their critical and maximal capacities in compression. This research deals with central
steel tube columns, one and two tube crossarms and spatial prestressed rod or cable stays. Firstly
four columns made of stainless steel 1.4301 with one central crossarm and cable stays were tested
up to extreme buckling deflections. The following numerical analysis employed geometrically
and materially nonlinear analysis with imperfections (3D GMNIA) using ANSYS software.
Careful analysis was successfully validated and used to obtain optimal prestressing of stays
giving maximal critical loads and maximal collapse loads of the tested columns. All the following
parametrical studies cover columns of the same geometrical and nonlinear material characteris-
tics as given or resulted from testing. This enables comparison and evaluation of the significance
concerning mode and value of initial deflections, value of prestressing, number of crossarms,
material characteristics, sliding of stays at crossarms and ratios of maximal collapse to critical
loads. The main results concerning critical loads of an “ideal column” (with amplitudes of initial
deflections L/500000) and maximal loads of an “imperfect column” (with amplitudes of initial
deflections L/200) under various prestressings of stays are discussed in a detail. In the conclusions
a necessity of GNIA/GMNIA considering respective mode and value of initial deflections is
emphasized (linear buckling analysis is not sufficient in such prestressed elements). Also the sub-
stantial significance of activating of stays during buckling and radical increase of both critical/
maximal loads by adding the second crossarm is presented.

1 INTRODUCTION

With the innovative technologies and high strength materials available the prestressed steel
structures are becoming always more affordable and required by both engineers and archi-
tects. Prestressed stayed columns with cable or rod stays enable design of extremely slender
structural elements (Figure 1). Nevertheless, design and realization of these elements require
careful techniques, because both buckling and collapse strength under prestressing require
nonlinear analysis and sophisticated erection procedures.

2 EXPERIMENTS AND THEORY

Four columns with one central crossarm made of stainless steel 1.4301 and stainless steel cable
stays were tested up to extreme buckling deformations (Servitova & Machacek 2011). Careful
GMNIA using ANSYS software covering measured initial deflections was successfully valid-
ated and later used in a study to obtain optimal prestressing of stays giving maximal critical
loads and maximal collapse loads of the columns (Macháček & Pichal 2018).
The following parametric study covers columns of the same geometrical dimensions as in the
tests (i.e. central tube Ø 50 × 2 [mm] with total length 5000 mm, crossarms Ø 25 × 1.5 [mm] with
total length 2 × 250 ¼ 500 mm, but with rod stays Ø 4 mm) and the same nonlinear material
characteristics of stainless steel concerning columns and crossarms as resulted from testing
(Figure 2), while for rod stays in accord with Eurocode 3 (E ¼ 200 GPa). Together with columns

906
Figure 1. Examples of structures with stayed elements.

Figure 2. Tested columns and material characteristics.

having one central crossarm also the same columns with two crossarms located in the thirds of
the column length were investigated.
Linear buckling analysis (LBA) applied to prestressed stayed columns provides reasonable
critical loads only for a limited values of prestressings (in the so called prestressed “zones” 1
and 3 according to investigations of Hafez et al. 1979). Due to a sudden change of the internal
energy in a distinctive buckling of the prestressed column and the following restoring of the
equilibrium the LBA (Macháček & Pichal 2018, Saito & Wadee 2008) can’t be used in mid
prestressings (“zone” 2).
Therefore, geometrically and materially nonlinear analysis with imperfections (3D GMNIA)
using ANSYS software was employed. The shapes of initial deflections (symmetrical and antisym-
metrical, see Figure 3) were introduced in accordance with the first two modes resulting from 2D
LBA of columns without any prestress, using SCIA Engineer software (Machacek & Pichal). To
model “ideal columns” the negligible values of amplitudes w0 ¼ L/500000 (i.e. approaching to
zero) were introduced, while for “imperfect columns” the amplitudes w0 ¼ L/200 (corresponding
to cold-formed steel tubes in accordance with EN 1993-1-1) were employed.
The 3D finite element (FE) model in ANSYS used elements BEAM188 for the column and
crossarm, while LINK188 (with no-compression option) for stays. The mesh optimization
resulted into meshing L/250 and a/25 (for designation see Figure 2). First initial deflections
were introduced, followed by the relevant prestress through a thermal change (cooling) and
finally imposing an axial loads to the column. Newton-Raphson (N-R) iteration was used in
the numerical solution.

907
Figure 3. Initial deflections introduced in the studies (LBA for the stays fixed to crossarms).

3 PRINCIPAL RESULTS OF STUDIES

3.1 Columns with 1 and 2 crossarms using GMNIA and GNIA


The principal results of 3D GMNIA concerning stayed columns with one central crossarm
and columns with two crossarms located in the thirds of the column length are shown in
Figure 4. The critical loads of “ideal columns” and maximal loads of “imperfect columns”
under various prestress are demonstrated.
The maximal loads in Figure 4 resulted from antisymmetrical initial deflections for columns
both with 1 and 2 crossarms. However, the critical loads of “ideal columns” with one central cross-
arm are determined by symmetric initial deflection, while in the case of two crossarms by antisym-
metric initial deflections. The decisive initial deflection mode generally depends on the column
geometry, initial deflection amplitudes and level of prestress. When the first two critical loads
resulting from the LBA corresponding to the antisymmetric and symmetric modes get closer (as in
the studied geometry, see Figure 2), the interactive mode for “imperfect columns” with low initial
deflection amplitudes (lesser than required by Eurocode) may be decisive (Saito & Wadee 2008).
Material nonlinearity of stainless steel (Figure 2) decreases the critical and maximal values.
As shown in Figure 5, the 3D GNIA using constant Young’s modulus E ¼ 200 GPa gives for
the same columns significantly higher critical and maximal values.
Note that even in the “ideal columns” with low prestress the critical load may be increased
to corresponding “maximal load” due to activating of the stays on convex side of buckled
shape after buckling (shown for “ideal columns” with 2 crossarms only). This is quite import-
ant in cases of no stays prestress in practice.

Figure 4. 3D GMNIA: Column with 1 crossarm (left), the same stayed column with 2 crossarms (right).

908
Figure 5. 3D GNIA: Column with 1 crossarm (left), the same stayed column with 2 crossarms (right).

The activating of convex side stays is explained in Figure 6 (3D GMNIA, for symmetrical
initial deflections). In a low prestress the buckling starts by slackening of stays (giving “critical
load”), followed by immediate activating of stays at convex side of buckling, leading to “max-
imal load”. On the other side in a high prestress the buckling invokes sudden changes of all
stays forces – a drop in concave stays and an increase in convex ones, while both “critical
load” and “maximal load” coincide.

3.2 Influence of initial deflection amplitude values


The cardinal influence of the value of the initial deflection amplitude follows from Figures 4
and 5. More detailed investigation in this field was performed in GNIA for the column with
one central crossarm and antisymmetric initial deflection mode, resulting in maximal loads for
various amplitude values as follows: w0 ¼ L/500 000 (nearly zero) gives Nmax ¼ 36 180 N;
w0 ¼ L/100 000 gives Nmax ¼ 35 770; w0 ¼ L/50 000 gives Nmax ¼ 35 430; w0 ¼ L/200 (amplitude
required by Eurocode EN 1993-1-1) gives Nmax ¼ 24 840 N.

3.3 Support at crossarms


Another study concerned influence of the stays support at the crossarms. In case of rod stays
(usually made of Macalloy or Detan rods) the connection is formed by rod end forks, creating
a fixed/hinge support. In case of cable stays the support may be either fixed or sliding (with

Figure 6. Activating of stays at convex side during buckling.

909
Figure 7. Column with one central crossarm – cable support variations.

cable continuing over a saddle). 3D GMNIA comparison of columns with one central cross-
arm having either fixed or sliding cable stays is shown in Figure 7. The column under investi-
gation had initial deflection w0 ¼ L/500 000 (i.e. representing “ideal columns” and critical
loads were evaluated).

4 CONCLUSIONS

• Determination of both critical and collapse loadings require GNIA/GMNIA considering


respective mode and value of initial deflections (LBA is not sufficient in such prestressed
elements).
• Activating of stays at convex sides of buckling may increase maximal loading in com-
parison to critical one, especially in low prestressings (see Figures 4 and 5).
• The Eurocode prescribed initial deflections (L/200) substantially decrease maximal loadings
of “imperfect columns” in comparison to critical loadings of “ideal columns” (in Figure 4
to 62%, in Figure 5 to 61%).
• Sliding stays at crossarms decrease both critical/maximal loadings only in cases of anti-
symmetrical buckling.
• Nonlinear behaviour of stainless steel material significantly decreases the maximal loading
(the decrease resulting from Figures 4 and 5 is to 78%).
• Adding the second crossarm to the otherwise same column with just one central crossarm
leads to substantial increase of both critical/maximal loadings (in the investigated col-
umns up to 28%).

ACKNOWLEDGEMENT

The research was supported by the Czech Grant Agency (GACR), grant No. 17-24769S.

REFERENCES

Hafez, H.H., Temple, M.C. & Ellis, J.S. 1979. Pretensioning of single-crossarm stayed columns.
J. Structural Division ASCE 14362(ST2): 359–375.
Macháček, J. & Pichal, R. 2018. Buckling and collapse capacity of prestressed steel tube stayed columns
with one and two crossarms. Thin-Walled Structures, 132: 58–68.
Saito, D. & Wadee, M.A. 2008. Post-buckling behaviour of prestressed steel stayed columns. Eng. Struc-
tures, 30: 1224–1239.
Saito, D. & Wadee, M. A. 2009. Numerical studies of interactive buckling in prestressed steel stayed
columns. Eng. Structures, 31: 432–443.
Servitova, K. & Machacek, J. 2011. Prestressed Stainless Steel Stayed Columns. In Dunai, L. et al. (eds),
EUROSTEEL 6th European Conf. on Steel and Composite Structures; Proc. intern. Conf., Budapest,
31 Aug. – 2 Oct. 2011: 1767–1772, Budapest: ECCS.

910
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Numerical simulation and analysis of axially restrained stainless


steel beams in fire

A. Pournaghshband
University of West London, London, UK

S. Afshan
University of Southampton, Southampton, UK

M. Theofanous
University of Birmingham, Birmingham, UK

ABSTRACT: This paper reports the results of a numerical investigation into the response of
restrained stainless steel beams in fire, where in addition to the degradation of strength and
stiffness at elevated temperatures, the influence of thermally induced stresses, are also
included. The finite element (FE) programme ABAQUS has been used to model stainless steel
welded I-section beams of different axial end restraint stiffness subjected to fire. The FE
models are firstly validated against a selection of literature test data, and then used to perform
parametric studies. The generated models capture the effects of restrained thermal deform-
ations with a high degree of accuracy, and thereby allow the influence of frame continuity to
be explicitly considered in design of stainless steel members in fire to quantify the required
strength and ductility demands on connections for catenary action to develop. Comparisons
with carbon steel beams demonstrate that while stainless steel beams show similar stages of
behaviour in fire, they are capable of withstanding higher temperatures prior to the onset of
catenary action, while developing similar levels of maximum tensile catenary force to carbon
steel beams, despite the higher thermal expansion of the material.

1 INTRODUCTION

Stainless steel members are often used in buildings in load-bearing applications. Due to their
superior behavior in terms of strength and stiffness retention at elevated temperatures com-
pared to carbon steel, stainless steel members are often used unprotected, thus reducing over-
all building costs. The performance of unprotected stainless steel structures in fire has been
studied by a number of researchers over the past years. At material level, the behaviour of
stainless steel elevated temperature has been studied extensively by means of isothermal and
anisothermal test, where it has been shown that stainless steel generally offers better retention
of strength and stiffness than carbon steel, especially at the important temperature range of
500-800 °C, owing to the beneficial effects of the alloying elements. At member level, experi-
mental and numerical modelling studies of the response of unprotected stainless steel struc-
tural members exposed to fire have been performed, which provided a valuable insight into
the effects of instability, temperature gradients and full cross-sectional behaviour on fire per-
formance. However, most previous studies have been mainly focused on understanding the
fire performance of individual elements such as statically determinate compression and flex-
ural elements, leading to the development of component based fire design guidelines such as
those in EN 1993-1-2 (2005). In structural frames, due to continuity, each structural member
is restrained axially and/or rotationally by adjacent members. Therefore, in addition to the
degradation of material strength and stiffness at elevated temperature, the effects of thermally

911
induced stresses, due to restrained thermal expansion and thermal bowing effects, on the
structural response have to be considered. Hence, to obtain a better understanding of the
response of stainless steel structures in fire, the behaviour of axially restrained stainless steel
beams exposed to uniform temperature distribution across the cross-section is investigated
herein.

2 FIRE BEHAVIOUR OF RESTRAINED STEEL BEAMS

The behaviour of a restrained beam in fire is considerably different from that of an isolated
beam without restraints. Figure 1 depicts the stages of the behaviour for the general case of
a restrained steel beam with axial and rotational end restraints subjected to a standard fire
with no decay phase, and giving rise to a non-uniform temperature distribution across the
beam section. Assuming that the beam is laterally restrained (i.e. no lateral torsional buckling)
and is made of a compact cross-section (i.e. no local buckling), there are three typical stages of
behaviour as described hereafter.
• In Stage I, the response is predominantly elastic. The axial restraint partially prevents the
thermal expansion of the beam, thereby leading to the development of a compressive axial
force accompanied by an increase in beam length due to axial expansion. Similarly, the rota-
tional restraint in conjunction with the thermal curvature resulting from the temperature gra-
dient in the cross-section of the beam, leads to the development of hogging bending moment
at the beam ends accompanied by thermally induced curving and a corresponding increase in
beam deflections. The fire induced internal forces and deflections continue to increase with
temperature until yielding of the most highly stressed beam cross-section occurs.
• In Stage II, the response is elasto-plastic. As plasticity spreads throughout the beam sec-
tion, the compressive axial force and the hogging bending moment will start to be relieved,
which is accompanied by an increase in the beam deflection and a reduction in the axial
expansion. This continues until the internal forces return back to zero.
• In Stage III, the beam enters the catenary phase where the deflection in the beam becomes
sufficiently large for the load bearing mechanism to change from flexure to catenary action
until failure occurs by tensile fracture of the beam or the connection.

3 VALIDATION OF NUMERICAL MODELS

3.1 Summary of literature test results


All fire tests on stainless steel beams reported in the literature have been performed on simply
supported beams without axial restraint. Hence, the results of the fire tests on restrained

Figure 1. The stages of behaviour of an axially and rotationally restrained beam in fire.

912
Table 1. Summary of the restrained beam tests reported in Liu et al. (2002).
Specimen Connection type kA (kN/mm) kR (kNm/rad) P (kN) Load Level

FUR09 Simply supported 0 0 40 0.5


FUR20 Simply supported 0 0 56 0.7
FUR14 Web cleat 8 0 40 0.5
FUR16 Web cleat 8 0 56 0.7
FUR21 End plate 8 14000 16 0.2
FUR13 End plate 8 14000 40 0.5
FUR15 End plate 8 14000 56 0.7
FUR17 End plate 8 14000 72 0.9
FUR27 End plate 35 14000 25 0.3
FUR25 End plate 35 14000 40 0.5
FUR26 End plate 35 14000 56 0.7
FUR32 Web cleat 62 0 40 0.5
FUR29 End plate 62 14000 24 0.3
FUR31 End plate 62 14000 40 0.5
FUR30 End plate 62 14000 56 0.7

carbon steel beams carried out by Liu et al. (2002) were used to validate the numerical models
developed herein. The beams were UB 178 × 102 × 19 in Grade S275 and were restrained
between two S275 UC 152 × 152 × 30 columns in the form of a ‘rugby goalpost’ frame, which
was arranged in a reaction frame placed in a furnace box. Two types of beam-to-column con-
nections, namely flush end-plate and double angle web-cleat, were employed. The beams were
loaded in the four-point bending configuration where two transverse loads, of magnitude P,
were applied at a distance of 0.6 m from each beam ends. Table 1 presents a summary of these
tests, where kA is the employed axial restraint stiffness, kR is the employed rotational restraint
stiffness, P is the applied point load per loading jack and the load level is defined as the ratio
of the applied maximum bending moment to the plastic bending moment capacity of the
beam at ambient temperature, assuming a design yield strength of 275 MPa. All fire tests were
performed anisothermally, where the loads were applied at room temperature and then kept
at a constant level, while the furnace temperature was set to increase following the ISO834
standard fire curve.

3.2 Development of numerical models

3.2.1 Thermal and stress analysis models


A sequentially coupled thermal-stress analysis using the FE analysis package ABAQUS (2015)
was carried out, where a nonlinear thermal analysis was first conducted to compute the tempera-
ture development in the beams, which was followed by a geometrically and materially nonlinear
stress analysis to determine the structural response under the application of load and temperature.
The measured furnace temperature-time curve was applied to the exposed surfaces of the
beams in the heat transfer model, and temperature development was simulated through the con-
vection, radiation and conduction heat transfer mechanisms. Radiation was modelled as surface
radiation by means of the *SRADIATE command in ABAQUS with the emissivity coefficient
taken as 0.7 from EN 1993-1-2 (2005). Convection was modelled as a film condition using the
*SFILM command in ABAQUS with the convective heat transfer coefficient taken as 25 W/m²K
from EN 1993-1-2. Other thermal properties including specific heat, thermal conductivity and
thermal expansion from EN 1993-1-2 were adopted. The beams were modeled with the DS4 shell
elements with a uniform mesh size of 10 mm × 10 mm, which from the mesh sensitivity analysis
gave the best compromise between accuracy and computational efficiency.
In order to simulate the finite translational and rotational end restraint conditions of the
beams in the stress analysis model, a structural model with axial and rotational springs as

913
Figure 2. Structural model of axially and rotationally restrained heated beam.

shown in Figure 2 was employed. Axial and rotational springs were modelled by means of
SPRING elements in ABAQUS using the *Element, type = Spring2 and *Element, type =
Spring1 commands for the axial and rotational springs, respectively. Both spring types were
assumed to be linear elastic with axial stiffness kA and rotational stiffness kR set equal to
those reported in Table 1.

3.2.2 Material modelling, boundary conditions and analysis steps


The elevated temperature stress-strain relationship for carbon steel provided in EN 1993-1-2,
along with the measured room temperature material properties from the experimental pro-
gramme and the elevated temperature strength and stiffness reduction factors set out in EN
1993-1-2, were utilised to construct a series of temperature dependent stress-strain curves. In
ABAQUS, the material behaviour was modelled as elastic–plastic with a von Mises yield cri-
terion and isotropic hardening. ABAQUS requires that for shell elements the input material
stress-strain curves are in the form of multi-linear true stress and logarithmic plastic strain
responses; these were obtained from the constructed engineering stress-strain relationships,
and incorporated into the FE models. The boundary conditions were carefully defined
through reference points RP-1 and RP-2 as shown in Figure 3 to simulate the test conditions.
The transverse loads were applied to the model at 0.6 m from each beam end through refer-
ence points RP-3 and RP-4 as shown in Figure 3. The models were laterally restrained against
lateral-torsional buckling. The geometrically and materially nonlinear stress analysis was car-
ried out in two steps; in Step 1, the transverse loads P were applied at room temperature and
maintained constant throughout Step 2, where the temperature was increased following the
time-temperature relationships stored from the thermal analysis model. To measure the axial
force and bending moment induced by the restrained axial and rotational displacements, the
axial and rotational springs were activated at the start of step 2. The general static solver in
ABAQUS was used for both Steps 1 and 2.

3.2.3 Geometric imperfections and residual stresses


Initial local and global geometric imperfections in the form of the lowest global and local
buckling modes were obtained from a linear eigenvalue buckling analysis and assigned to the
numerical models as the starting geometry. The global imperfection amplitude was taken as
L/1000, where L is the beam length, in accordance with the permitted out-of-straightness tol-
erance in EN 1090-2 (2018), while the local imperfection amplitude was taken as that pre-
dicted from the modified Dawson and Walker model from Gardner et al. (2010). The residual
stress pattern for hot-rolled steel sections recommended in (ECCS, 1984) for cross-sections

Figure 3. Definition of boundary conditions for the stress analysis model.

914
with height-to-width ratio greater than 1.2 was applied to the FE models. The four-node
doubly curved shell element with reduced integration S4R, which is compatible with the DS4
elements, with the same mesh element size as for the thermal models 10 mm×10 mm, was
used to discretise the FE models. The models included web stiffeners with the same arrange-
ment as that in the tests in order to avoid premature local bearing failure of the web.

3.3 Validation
The sequentially coupled thermal-stress analysis models described were validated against the
15 axially and rotationally restrained steel beam tests presented in Table 1. Figures 4-6 com-
pare temperature-deflection and temperature-axial force curves obtained from the FE models
with the corresponding experimental ones for the beams with flush end-plate connection and
an axial restraint level of 8 kN/mm, 35 kN/mm and 62 kN/mm axial restraints, respectively.
The same tests were also replicated numerically by Yin & Wang (2004) and Liu & Davies
(2001), the results of which are also depicted in Figures 4-6, where it is shown that comparable
results are obtained by all numerical models. The axial force and deflection characteristics typ-
ical of axially and rotationally restrained beams in fire as exhibited by the tested beams were
accurately captured by the FE models. The unavoidable discrepancies that exist between the
test and FE results are due to the variations in the actual and simulated temperature distribu-
tions and the generalised elevated temperature stress-strain relationships assumed. The failure
modes of the tested steel beams were reported by Liu et al. (2002) to mainly include (1) forma-
tion of two plastic hinges under the loading points and two others close to end supports as

Figure 4. Test and FE results for FUR13 specimen – Flush end-plate, kA = 8 kN/mm, load ratio = 0.5.
(a) Temperature-axial force. (b) Temperature-deflection.

Figure 5. Test and FE results for FUR25 specimen – Flush end-plate, kA = 35 kN/mm, load ratio = 0.5.

915
Figure 6. Test and FE results for FUR31 specimen – Flush end-plate, kA = 62 kN/mm, load ratio = 0.5.

Figure 7. Replication of experimentally observed failure mode by numerical models.

well as (2) local buckling of the bottom flange close to the end supports. The deformed shapes
obtained by the FE models closely replicated the experimentally observed failure modes as
shown in Figure 7. Following successful validation, the developed finite element models were
employed to perform parametric numerical investigations on axially restrained stainless steel
beams in fire, detailed descriptions of which are presented in the following section.

4 PARAMETRIC STUDY

Parametric studies were performed to investigate the structural response and failure pattern of
axially restrained stainless steel beams in fire, with a focus on the influence of catenary action and
restraint axial forces. The analyses were performed on conventionally welded IEP400 beams with
a length of 8 m loaded in the three-point bending configuration. The selected cross-section is clas-
sified as Class 1 according to EN 1993-1-4 (2015). For a building subjected to a compartment
fire, the action of cooler adjacent structures to the heated structural members can be represented
by suitable end restrain conditions which are usually modelled by end springs with axial and rota-
tional stiffness. Table 2 presents a summary of the examined axially restrained beams. The axial
restraint stiffness ratio αA is defined as the ratio of the axial restraint stiffness to the axial stiffness
of the beam at room temperature EA/L = 202 kN/mm, where E is the Young’s modulus at room
temperature, A is the cross-sectional area and L is the beam length and the load level is as defined
in Section 3.1. All beams were free to rotate at both ends. The temperature distribution was

Table 2. Summary of examined parametric study models.


Beam configuration Axial restraint condition Load level (SS) Load level (CS)

No axial restraint 30%, 50%, 70% 50%


Full axial restraint 30%, 50%, 70% 50%
αA = 0.02, 0.05, 0.15, 0.3, 0.5, 1 30%, 50%, 70% 50%

916
uniform across the depth of the modelled cross-sections. For comparison purposes, carbon steel
beams with the same levels of axial restraint stiffness as those for the stainless steel beams and
with 50% load level were also modelled.
The same modelling procedure as in Section 3.2 was used with the input parameters for
carbon steel taken as those described previously and those for stainless steel taken as those
described hereafter. For room temperature, the material properties recommended by Afshan
et al. (2019) for hot-rolled austenitic stainless steel plates were adopted. The two-stage Ram-
berg-Osgood model and the elevated temperature strength and stiffness reduction factors for
grade EN 1.4571 recommended in the Design Manual for Structural Stainless Steel (2017) were
employed. The standard ISO 834 time-temperature curve was used for the thermal model. The
convective heat transfer coefficient factor and the emissivity were taken as 25 W/m2K and 0.4,
respectively (EN 1993-1-2, 2005). Specific heat, thermal expansion and thermal conductivity as
recommended in Design Manual for Structural Stainless Steel (2017) for austenitic stainless
steels were used. The global imperfection amplitude was set to L/1000. The local imperfection
amplitude was determined by the modified Dawson and Walker model from Gardner & Neth-
ercot (2004). The residual stresses for conventionally welded stainless steel I-sections proposed
by Yuan et al. (2014) were incorporated into the FE models. The models were laterally
restrained to avoid lateral-torsional buckling.

5 RESULTS AND DISCUSSIONS

5.1 Temperature-deflection and temperature-axial force responses of stainless steel beams


Figure 8 compares the temperature-deflection curves from the FE models of stainless steel
beams with varying axial restraint stiffness ratio αA and subjected to 30%, 50% and 70% load
levels. For each of the considered load levels, there is little difference in the temperature-
deflection behaviour of the modelled beams during Stage I response. The temperature at
which the beam deflection starts to runaway is a function of the αA, and, as expected, is lower
for the beams with higher αA. During Stage II, the deterioration of strength and stiffness of
stainless steel with increasing temperatures dominates the structural response and causes fur-
ther deflections, which are higher for higher applied load levels. Additional deflections arise
due to the P-δ effects which are also higher for beams with higher axial restraint stiffness αA
and under higher load levels. As a result, the onset of catenary action, which is triggered by
the occurrence of large deflections, occurs at lower temperatures for beams under higher
applied load levels. However, for a given load level, catenary action began at almost the same
temperature for all different levels of αA. The temperature-axial force curves from the FE
models of stainless steel beams with varying axial restraint stiffness ratios αA subjected to
30%, 50% and 70% load levels are shown in Figure 9. It can be observed that the axial com-
pressive force generated as a result of the restrained thermal expansion strains grows steadily
with increasing temperature for all beams until a maximum compressive axial force is reached,
beyond which the axial force starts to reduce as the deflections start to runaway. The max-
imum axial compressive force reached is higher for beams with higher αA, and results in earlier
onset of runaway deflections. During the catenary stage, the axial force changes to tension
and the applied vertical load on the beam is supported by the vertical component of the tensile
catenary force. The magnitude of the maximum catenary axial tensile force developed depends
on the residual tensile strength of the material at elevated temperatures and the magnitude of
the applied vertical load, which is shown in Figure 9 to be larger for higher load levels.

5.2 Comparison of stainless steel and carbon steel beam responses


Figure 10(a) compares the temperature-deflection relationships of the axially restrained
stainless steel and carbon steel beams with varying levels of αA for 50% load level. Prior
to the start of the runaway deflections, stainless steel beams deflect slightly more than
carbon steel beams for all αA. This is due to the higher axial compressive forces that

917
Figure 8. Parametric temperature-deflection results for axially restrained stainless steel beams.

develop as a result of the higher thermal expansion of stainless steel, which in turn give
rise to higher P-δ deflections. For low levels of axial restraint stiffness (αA ≤ 0.05), the
runaway deflections start at higher temperatures for stainless steel beams, as the superior
strength and stiffness properties of stainless steel outperform the effects of higher restraint
forces developed. As a result, the start of catenary action also takes place at higher tem-
peratures. In stainless steel beams with higher axial restraint stiffnesses (αA ≥ 0.15), while
the higher axial compressive forces developed cause early runaway deflections than for
carbon steel beams, the onset of catenary action takes place at higher temperatures. The
delay in reaching catenary action is due to the lower deflections at high temperatures due
to the better stiffness retention of stainless steel over these higher temperatures. Figure 10
(b) compares the temperature-axial force responses of the stainless steel and carbon steel
beams with different levels of αA and with 50% load level. For all αA, the axial compres-
sive force grows at a higher rate in stainless steel beams than carbon steel beams, demon-
strating the higher thermal expansion of the material. For low levels of axial restraint
stiffness (αA ≤ 0.05), the stainless steel beams are capable of reaching their peak axial
compressive loads, prior to the onset of runaway deflections, at higher temperatures, while
for higher levels of axial restraint stiffness stiffnesses (αA ≥ 0.15), the maximum axial com-
pressive load is reached at increasingly lower temperatures with increasing levels of αA.
For all stainless steel beams, the catenary stage begins at higher temperatures (750-850 °C)

918
Figure 9. Parametric temperature-axial force results for axially restrained stainless steel beams.

Figure 10. Comparison of (a) temperature-deflection and (b) temperature-axial force responses of stain-
less steel and carbon steel beams under 50% load level.

919
than the carbon steel beams (550-650 °C). The maximum catenary force reached for stain-
less steel and carbon steel beams are however very similar; this is due to their comparable
residual ultimate tensile force at their respective catenary stage temperature ranges.

6 CONCLUSIONS

The large deflection behaviour of axially restrained stainless steel I-sections beams in fire has
been examined through a numerical investigation. FE models, validated against the results of
physical tests, were used to conduct a parametric study. The results showed that the axially
restrained stainless steel and carbon steel beams show similar stages of behaviour in fire. How-
ever, owing to the more favorable strength and stiffness retention factor of stainless steel at
high temperatures, stainless steel beams are capable of withstanding comparatively higher
temperatures before the onset of catenary action, while developing similar levels of maximum
catenary force to carbon steel beams, despite the higher thermal expansion of the material.

REFERENCES

ABAQUS. Version 2015. 2016. Dassault Systmes Simulia Corp. USA.


Afshan, S., Zhao, O. & Gardner, L. 2019. Standardised material properties for numerical parametric
studies of stainless steel structures and buckling curves for tubular columns. Journal of Constructional
Steel Research 152: 2–11.
Steel Construction Institute (SCI). 2017. Stainless Steel Design Manual. fourth edition.
European Committee for Standardization. 2005. “Design of steel structures – Part 1-2: General rules –
Structural fire design.” Eurocode 3, EN 1993-1-2, Brussels, Belgium.
European Committee for Standardization. 2015. “Design of steel structures – Part 1.4: General rules sup-
plementary rules for stainless steels.” Eurocode 3, EN 1993-1-4, Brussels, Belgium.
European Committee for Standardization. 2018. “Execution of steel structures and aluminium structures.
Technical requirements for steel structures.” EN 1090-2, Brussels, Belgium.
ECCS. 1984. Ultimate limit state calculation of sway framed with rigid joints. Technical Committee 8 of
European Convention for Constructional Steelwork (ECCS), Tech. Rep., No. 33.
Gardner, L. & D. Nethercot. 2006. Numerical modelling of stainless steel structural components –
a consistent approach. Journal of Structural Engineering (ASCE) 130(10): 1586–1061.
Gardner, L. Saari, N. & Wang, F. 2010. Comparative experimental study of hot-rolled and cold-formed
rectangular hollow sections. Thin-Walled Structures 48(7): 495–507.
Liu, T.C.H., Fahad M.K. & Davis J.M. 2002. Experimental investigation of behaviour of axially
restrained steel beams in fire. Journal of Constructional Steel Research 58: 1211–1230.
Liu, T.C.H. & Davies, J.M. 2001. Performance of steel beams at elevated temperatures under the effect
of axial restraints. Steel and Composite Structures 1 (4): 427–440.
Yin, Y.Z. & Wang, Y.C. 2004. A numerical study of large deflection behaviour of restrained steel beams
at elevated temperatures. Journal of Constructional Steel Research 60 (7): 1029–1047.
Yuan, H.X., Wang, Y.Q., Shi, Y.J. & Gardner, L. 2014. Residual stress distributions in welded stainless
steel sections. Thin-Walled Structures 79: 38–51.

920
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Experimental and numerical investigations of unstiffened steel


girders with non-rectangular panels subjected to bending and shear

V. Pourostad
The Institute of Structural Design, University of Stuttgart, Germany

U. Kuhlmann
Steel, Timber and Composite Construction, The Institute of Structural Design,
University of Stuttgart, Germany

ABSTRACT: Slender structures of thin plates are usually designed according to EN 1993-
1-5. In order to adapt to an optimized overall curved shape in the longitudinal direction of
bridges with plated girder the development of buckling rules for non-rectangular steel panels
is addressed. This paper aims to show experimental as well as numerical analyses on the buck-
ling behaviour of non-rectangular panels, which are conducted in the frame of a European
Research Project OUTBURST (OpTimization of Steel Plated BRidges in Shape and
Strength). To investigate the behavior of the non-rectangular panels, six large tests composed
of unstiffened and stiffened girders have been conducted at the laboratory at the University of
Stuttgart. The numerical simulations are validated by recalculation of the tests. In addition, to
investigate the effect of angle, aspect ratio and slenderness of panels on the buckling behavior
of tapered panels subjected to the interaction of bending and shear, a parametric study is con-
ducted for unstiffened panels. Different resistance models are developed and statistically
evaluated. A resistance model and design procedure on basis of the EN 1993-1-5 in case of
non-rectangular panels is proposed to achieve a safe and economic design.

1 INTRODUCTION

1.1 Type area


Non-rectangular steel panels are increasingly used in the design of new bridges due to architec-
tural and/or structural demands. At large spans, in order to save material and consequently to
decrease the impact on the environment, the girders are curved in elevation with a maximum
depth at intermediate support and minimum depth at the mid-span. Steel bridges built up of
slender panels, which tend to buckle may be designed based on EN 1993-1-5 which offers dif-
ferent methods such as effective width method, reduced stress method and verification based
on finite element methods of analysis. Among the mentioned methods, the effective width
method is based on the reduction of the cross-section area considering the local buckling of
the sub-panels between the stiffeners and the global buckling of the stiffened panel. As
a consequence of the optimized shape of new bridges, there also are non-rectangular panels,
most commonly as web panels of girders with a lower flange curved in the longitudinal direc-
tion. However, for an application of the effective width method, only rectangular panels with
parallel flanges are assumed. This may also be applied to non-rectangular panels with an
angle up to 10 according to the existing rules. However, in case of the larger angles, the sta-
bility of panels has to be checked with the higher width which leads to conservative results. At
the moment many interpretations of the design rules in EN 1993-1-5 is not clear for non-
rectangular panels are possible. In order to avoid confusing the angle α with the aspect ratio
α, the angle of panels is indicated by  in the following.

921
In this paper at the first, a short summary of the test program is introduced and then
the recalculation of tests and validation of the numerical model is shown. Finally, the
results of the parametric study are compared with the results of different resistance
models of non-rectangular panels to develop appropriate design rules for non-
rectangular panels.

1.2 Experimental investigations of non-rectangular panels


As part of a European research project OUTBURST, in the frame of the testing program on
the buckling behaviour of non-rectangular panels, tests were conducted on six welded plate
girders within the field of bending-shear interaction.
The tests explicitly focused on the angle of panels under the interaction of bending moment
and shear. The structural system of a cantilever, the flanges (350 x 20 mm) and length of the
test panels remained the same for all tested girders, so only the plate thickness and the angle
of a test panel were varied within the testing program. Figure 1 shows a schematic representa-
tion of a non-rectangular panel. The test setup is shown in Figure 2 and the parameters of test
panel are summarized in Table 1.
Table 2 shows the comparison of the load capacity for the tested girders together with
the angle and the plate thickness of the tested panel. The highest capacity is found for
girder T5 followed by T3. In Table 2 the maximum load of the regarded girder is given
in relation to the maximum load of the test T5 to show the tendency. Comparison of
tests T3 and T5 shows that with decreasing the angle from 15 to 11 the maximum
load slightly increases. As the thickness of the web increases, obviously the maximum
load increases.

Figure 1. Definition of dimensions of a non-rectangular panel.

Figure 2. Test setup of test T3.

922
Table 1. Dimensions of the tested panels.
Test a b2 b1 tw  Stiffener
[mm] [mm] [mm] [mm] [ ]

T1 1500 1500 1100 8 15 Without stiffener


T2 1500 1500 1100 6 15 Without stiffener
T3 1500 1500 1100 8 15 Parallel to lower flange
T4 1500 1500 1100 6 15 Parallel to lower flange
T5 1500 1500 1200 8 11 Parallel to lower flange
T6 1500 1500 1200 8 11 Parallel to upper flange

Table 2. Maximum loads of tested girders.


Test T1 T2 T3 T4 T5 T6

t [mm] 8 6 8 6 8 8
 [ ] 15 15 15 8 11 11
Fgirder =F5 0.848 0.812 0.962 0.859 1.000 0.898

2 NUMERICAL INVESTIGATIONS

2.1 Recalculation of tests


The numerical models were developed in ABAQUS. In order to reduce calculation time and
use the model for the parametric study, the model has been simplified. The simplified model is
composed of only the tested panel. The applied load consists of a shear force and a bending
moment which were determined to gain the appropriate shear force and bending moment in
the tested panel. In Figure 3 the complete model and the simplified model are shown. For the
recalculation of tests, the σ  ε curve was obtained from the tensile material tests and modified
according to EN 1993-1-5 into true σ  ε relation. During the tests, initial imperfections and
deformations were measured with photogrammetric methods. The measured initial imperfec-
tions were modelled in the numerical model. The plates were meshed with 4 and 3 node shell
elements with reduced integration. The ultimate loads of the numerical simulations were veri-
fied with respect to the results of the experimental tests. The deviation of numerical and test
results may be defined with the Equation (1).
FFE  FTest
Dev: ¼ ð1Þ
FTest

The deviation of test and numerical results are given in Table 3. A detailed description of the
numerical recalculation of tests and simplification of numerical models are discussed in

Figure 3. Numerical model of test T1 and T2.

923
Table 3. Deviation of numerical results.
Model T1 [%] T2 [%] T3 [%] T4 [%] T5 [%] T6 [%]

Simplified Model 2.4 -0.3 -3.8 -0.3 -4.6 0.4


Complete Model 3.5 -0.3 -2.0 -0.6 -3.4 -0.5

Pourostad & Kuhlmann. Table 3 shows that the ultimate load obtained by both numerical
models are very close to the test results. Therefore, the model is verified and can be used for
the parametric study.

2.2 Parametric study


To evaluate the current design rules acc. to EN 1993-1-5 and especially the effective width
method, a parametric study has been conducted using the validated model. Following parameters
were considered and varied in order to investigate the buckling behaviour of non-rectangular
panel: b2 = 2100 [mm]; a = 2100, 4200 [mm]; tw = 6, 8, 10, 12, 15, 20 [mm]; bf  tf ¼ 315 × 20,
350 × 36, 490 × 36, 560 × 36 [mm];  = 0, 10, 12.5, 15 [ ]. The application of a suitable imperfec-
tion shape, which reproduces the real behaviour of tapered panel, is sophisticated. In this investi-
gation, the first positive eigenmode shape form LBA with an amplitude of b2 =200 was applied.
For each configuration, two types of numerical analyses were carried out: 1- Linear Buckling
Analysis (LBA) to obtain the eigenmode shap 2- Geometrically and Materially Nonlinear Ana-
lysis with Imperfections (GMNIA) to determine the resistance of panel. As material characteris-
tics, nominal values for structural steel S355 with elastic modulus E = 210000 MPa and yield
strength f y ¼ 355 MPa (on the safe side without strain hardening) were modelled. To avoid
numerical problems, the nominal plateau slope of E/10000 acc. to EN 1993-1-5 was assumed.
The panels were subjected to bending, shear and interaction of bending and shear.

3 RESISTANCE MODELS

Four theoretical resistance models have been developed and evaluated to be compared to the
numerical database in order to determine the reliability. For bending analysis acc. to EN
1993-1-5 the plate buckling verification of the panel should be carried out for the stress result-
ants at a minimum distance 0,4a or 0,5b, from the panel end where the stresses are greater.
That means the acting stresses may be reduced. Additionally, the gross sectional resistance
should be checked at the end of the panel subjected to actual acting forces without any
reduction. M-V interaction for this system should be verified at the sections located at
a distance b/2 from the support. Since the design rules in EN 1993-1-5 have been written for
the rectangular panels, the design procedure of tapered panels is not clearly explained. So far,
current design rules have not been thoroughly investigated and evaluated in case of non-
rectangular panels. Due to the shape of tapered panels, in this study, the mentioned distances
are obtained using the larger width b2 instead of b.
Figure 4 shows the position of sections of verifications acc. to EN 1993-1-5. The gross cross
section should be checked with respect to the corresponding acting forces at section 1 and 2.
For bending, section 3 and 4 should be verified. Section 3 is also relevant for M-V interaction.
In this paper, the results of four resistance models are presented. To obtain the ultimate
loads, the verifications are transformed as a function of shear force V and compared with cal-
culated V from the numerical results.
The first resistance model consider the verifications of the bending, shear
and M-V interaction acc. to EN 1993-1-5.
In addition to the first resistance model, the second resistance model considers the check of
the gross sectional resistance.
Due to the shape of the inclined compression flange, the acting shear forces may be modi-
fied. The force distribution due to inclined compression flange is shown in Figure 5. Nx; f is

924
Figure 4. Position of sections of verification.

Figure 5. Influence of inclined compression flange on shear force.

the horizontal force resulting from the bending moment. This force may be split into the force
NFlange and the vertical force Vz; f . The vertical force Vz; f may modify the shear forces. The
modification of the shear forces are calculated by Equations (3) and (4).

Vz; f ¼ Nx; f  tanðÞ ð2Þ

VRed ¼ VEd  Vz; f ð3Þ

The third resistance model considers the mentioned verifications, additionally, the influence of the
inclined flange is taken into account. Since at the moment the authors do not know any literature
or free software to calculate the buckling coefficient of non-rectangular panels, the buckling coef-
ficient is calculated on the safe side assuming the panel as a rectangular with the larger width (b2 ).
However, advanced commercial FE software enables users to determine the buckling coeffi-
cient or elastic critical buckling load for practically arbitrary geometry, stiffener shapes, and
stiffeners arrangement. Fourth resistance model uses this possibility and considers the exact
shape of the panel. The buckling coefficients for bending and shear are calculated using linear
buckling analysis with ABAQUS considering the non-rectangular shape of the panel. The
influence of flange on the shear is considered in the fourth resistance model. It should be men-
tioned for all resistance models the partial factors γM0 and γM1 are assumed equal to 1:0.

4 STATISTICAL EVALUATION OF RESULTS

In order to determine the reliability level of each of the four resistance models, the results of
numerical database are compared with the results of the resistance models. Following the
method in EN1990, Annex D is applied for the statistical evaluation.

925
In Figure 6 numerical results are compared with the results of the resistance models for all
calculated angles of panels ( ¼ 0  15). The values obtained by the resistance models are indi-
cated on the x-axis and the values derived from the numerical simulations on the y-axis. The
red line shows re ¼ rt , which means the resistance model is equal to the numerical resistance.
It can be observed that the results of the first resistance model rt;1 are partly on the unsafe
side. The second resistance model shows better results than the first resistance model. Hence,
an additional check of cross section is needed. However, the distribution of results is wide.
The results of the third and fourth resistance models have a good agreement with numerical
results. The reduction of the shear forces due to the contribution of the flange obviously
improves the results of the third and fourth resistance models.
To investigate the influence of the angle of panel  on the resistance models, the re  rt dia-
gram of the second and third resistance model are compared separately for each angle, see
Figure 7.
More conservative results are reached from the second resistance model with increasing the
angle, while the results of the third resistance model show very good agreement with the
numerical results. Furthermore, the third resistance model leads to a more narrow distribution
and to more economic results in comparison to the second resistance model.
Finally the partial factors for each resistance model and angle  are calculated and summar-
ised in Table 4. It is obvious that the check of the gross sectional resistance is necessary for the
verification. Due to the missing cross section check, the first resistance model leads to the
highest partial factor. The best results can be found by the third and fourth resistance model,
which consider the influence of inclined flange on shear forces. Both resistance models have
similar results. The third resistance model it does not need a advanced software to calculate
the buckling coefficients, which is a benefit, and the panel can be assumed as rectangular. It
can be seen that the partial factor increases with increasing the angle. The maximum value of
1.06 can be found for the angle of 15 , which is the less than required 1.1 for shear forces. It
should be mentioned that the partial factor in EN 1993-1-5 for shear and direct stresses are
different. For shear forces a partial factor 1.1 and for direct stresses 1.0 is given in EN 1993-
1-5. In the last row, the partial factors for all angles are given. The partial factor for the third
resistance model has a value of 1.04. Therefore, the third resistance model can be used design-
ing of unstiffened panels with non-rectangular shapes leading to safe and economic results.

Figure 6. Comparison of the results of the resistance models with the numerical results
( ¼ 0; 10; 12:5; 15 )

926
Figure 7. Influence of inclined compression flange on shear forces.

927
Table 4. Partial factor γM for evaluated resistance models.
 rt;1 rt;2 rt;3 rt;4

0 1.07 1.00 1.00 1.00


10 1.10 1.07 1.03 1.03
12:5 1.14 1.10 1.05 1.05
15 1.21 1.11 1.06 1.07
0  15 1.46 1.07 1.04 1.05

5 CONCLUSIONS

Experimental investigations have been conducted on six welded girders. The tests focused
explicitly on the influence of the angle of panels subjected to the interaction of bending
moment and shear. The results show that with decreasing angle from 15 to 11 the maximum
load increases.
For further numerical investigations, the tests have been recalculated for the verification of
the FE-model. The numerical models were simplified. The results of the numerical simulations
were validated against the results of the experimental tests. The numerical simulations show
a good agreement with test results. Then, a numerical parametric study has been performed to
cover different configurations. Based on numerical data the current design EN 1993-1-5 rules
are evaluated, enhanced, and also supported by on a statistical approach.
These investigations show that the EN 1993-1-5 can be used for non-rectangular panels and
that the given the limitation of 10 can be increased up to 15 . Additionally, it is observed for
a tapered panel with inclined compression flange that the effect of flange inclination can be taken
in to account. The obtained results with consideration of the flange force show more economic
results.
In the next step, the stiffened panels will be investigated and the introduced resistance
models will be evaluated and enhanced for stiffened panels.

ACKNOWLEDGEMENT

The financial support of the European Commission Research Program of the Research Fund
for Coal and Steel (RFCS) for the investigation of non-rectangular panels is greatly appreci-
ated. The girders for the experimental research were donated and fabricated by MCE GmbH.
Acknowledgment is also given to the Institute for Photogrammetry in Stuttgart, especially to
Dipl.-Ing. Alessandro Cefalu for the photogrammetric measurements and to MPA, the mater-
ial testing institute at the University of Stuttgart. Our thanks also refer to our colleagues of
the project OUTBURST for their cooperation.

REFERENCES

EN 1990. Eurocode: Basis of structural design, European committee for standardization, 2002.
EN 1993-1-5.Eurocode 3 - Design of steel structures - Part 1-5: Plated structural elements, European com-
mittee for standardization, 2006.
Simões da Silva L., Kuhlmann U., Može P., Pedro J.O., Hendy Chr. et al., Optimal and aesthetic design
of curved steel bridges (OUTBURST). Research Fund for Coal and Steel, Grant Agreement No.
RFCS-2015-709782, 2016–2019.
Simulia. Abaqus 6.11 Documentation. Dassault Systèmes, 2011.
Pourostad, V. & Kuhlmann, U. Experimental investigation on girders with nonrectangular panels,
Eighth International Conference on Thin-Walled Structures- ICTWS 2018, July 2018. Lisbon, Portugal.

928
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Studying bolt force distribution in ultra-large capacity


end-plate connections

A.A. Ramzi, I.M. El Aghoury, S.M. Ibrahim & A.I. El-Serwi


Department of Structural Engineering, Ain Shams University, Cairo, Egypt

ABSTRACT: Bolted connections are classified according to their flexural resistance and
rotational stiffness. The moment capacity of any connection is controlled by its different com-
ponents (end-plate thickness, bolt diameter, bolt grade and stiffeners (if any)). Increasing the
number of bolts in the tension zone achieves a higher moment capacity for the connection. In
this paper, the ultra-large capacity connection with 12 bolts in tension (4 bolts per row) is
studied. A proposed finite element model was built and verified using experimental results.
A parametric study is carried out using different parameters: end-plate thickness, bolt diam-
eter, effect of horizontal stiffeners and their thickness. Results are presented in the form of
curves using the relation between end-plate thickness and the moment capacity for different
parameters. Moreover, bolt force distribution is studied at yielding and ultimate limit of con-
nection. Results showed that the chosen parameters are effective on moment capacity and bolt
force distribution among bolts.

Keywords: Moment Connections, Ultra High Capacity Connections, Bolt Force Distribution

1 INTRODUCTION

The bolted end-plate moment connection is one of the most widely used connection in steel
structures since 1960. Due to the ease of its fabrication, overall performance and cost effect-
iveness compared to welded connection types. Researchers started investigating its behaviour
since the early 1950s up till now has resulted in refined design procedures for both flush and
extended end-plate connections. Conventional known configurations of end-plate connections
are two bolts per row. Many researchers studied the performance of conventional (two bolts
per row) (Samaan 2010; Shi et al. 2008). Also, several design methods have been proposed in
different design codes or guides (AISC 2010; EN1993-1-8:2005 2005).
As the moment capacity required for beam-to-column connections is increased in steel frames,
conventional end-plate connections might not meet the required moment resistance, being limited
by the number of the bolts in tension. Improved configurations have been developed by using four
bolts per row, both classified as large capacity end-plate connections with eight bolts in total
arranged in tension zone. The large capacity end-plate connection has been proposed and investi-
gated by several researchers (Demonceau et al., 2010; Mohamed Eldemerdash et al., 2012; Prinz
et al., 2014). Ultra-large capacity end-plate connections with 12 bolts to 16 bolts around tension
flange was investigated using experimental programs. The study showed that in such type is
characterized by non-uniform bolt force distribution depending on bolt location, and end-plate is
subjected to biaxial bending (Gang Shi et al., 2017). Figure 1 illustrates different types of end-plate
connections.
In this paper, ultra-large capacity end-plate connection with different end-plate thickness,
bolt diameter, horizontal stiffeners effect and their thickness to investigate the moment cap-
acity and bolt force distribution among bolts using verified finite element model.

929
Figure 1. Different types of end-plate connections.

2 FINITE ELEMENT MODEL

2.1 Element type and material modeling


The study was carried out using the Finite element software package (ANSYS) program. 3D
SOLID186 element is used in modelling all components: flanges, webs, end-plate, stiffeners
and bolt. This element is characterized by 20-node solid element that exhibits quadratic dis-
placement behaviour having three degrees of freedom per node: translations in the nodal x, y,
and z directions. CONTA178 and TARGE170 are used to define contact surfaces between
end-plate, column flange and bolt shank. PRETS179 is used for defining pretension forces
inside bolt. A tri-linear stress-strain curve is used in modelling the steel plates while bilinear
stress-strain curve is used in modelling bolts as shown in Figure 2.

2.2 Model validation


(Samaan et al. 2017) conducted an experimental study to investigate the large-capacity end-
plate beam-to-column connection. Five different specimens were examined with different con-
figurations: two flushed connections, two extended unstiffened connection and one stiffened
extended connection. Steel used was st37 with yield strength 2.4 t/cm2 and M20 grade 10.9
bolts with an ultimate tensile strength of 10 t/cm2. The test setup is shown in Figure 3. The
finite element model created for calibration with the experimental work is shown in Figure 4.

2.2.1 Comparison between FEM and experimental results


Table 1 lists the ultimate loads for both the experimental program and those obtained from
the finite element model. It is evident form Table 1 that the finite element model results show
good agreement with the experimental results with average ratio reaching 100% and within
a standard deviation ratio of 6%.

Figure 2. Stress-strain curves for different components of the studied connections.

930
Figure 3. Test setup by (Samaan, El-Serwi and El-Hadary, 2017).

Figure 4. Meshing of the entire model.

Table 1. Comparison between FEM and experimental results.


Specimen Experimental test (kN) Finite element (kN) FEM/Test

ENS20 156.5 148.2 0.95


ENS32 177.1 176 0.99
ES20 174 176 1.01
F20 95 88 0.93
F32 90 101 1.12
Average 1.00
Standard deviation 0.06

3 PARAMETERIC STUDY

A parametric study is conducted to study the behaviour of three connection configurations as


shown in Figure 5. A finite element model was developed to represent a typical beam-to-column
connection with fixed supports on both ends of column and a lateral support for beam to prevent
LTB as illustrated in Figure 6. End-plate thickness, bolt diameter and adding horizontal stiffeners
are the main studied parameters to study their effect on the behaviour and the moment capacity
of the connection as well as the force distribution among bolts.. For all studied configurations,
the effect of different bolt diameters {M16, M20, M22, M24 and M27} and end-plate thicknesses
{14mm, 20mm, 25mm, 30mm, 35mm, 40mm and 50mm} on the behaviour of connection is stud-
ied. The following parameters are kept constant throughout the parametric study:
– Constant cross section and length is used for the column, with web dimensions 400 × 20mm
and flange dimensions (16 × Dia. of bolt) × 50mm.

931
Figure 5. The three studied configurations.

Figure 6. Studied parametric model.

– Constant cross section and length is used for the beam, with web dimensions 600 × 16mm
and flange dimensions (16 × Dia. of bolt) × 40mm.
– Stiffeners with thickness 40mm are used in tension, compression zone and at load applied
position to prevent local deformation of the column and beam elements.
– Pitch distance is taken four times the diameter of bolt. While edge distance is two times the
bolt diameter.
– The main type of steel used for plates is mild steel 24/37 with yield stress 2.4 t/cm2 and
ultimate tensile stress 3.6 t/cm2.
– The main type of high strength bolts used in this study is bolts of grade 10.9 with an ultim-
ate tensile stress of 10 t/cm2 and yield stress equal to 9.0 t/cm2.

4 RESULTS

4.1 Effect of end-plate thickness and bolt diameter


Figure 7 plots the moment capacity of different connections versus the thickness of the end
plate. By inspecting Figure 7, it is divided into two parts; first part represents end-plate rup-
ture failure mode while the second part represents bolt rupture failure mode. End-plate failure
happens at relatively small end-plate thicknesses with an obvious increase in moment capacity.
With the increase in end-plate thickness bolt rupture occurs with a slight increase in moment
resistance.
From the results shown, it is found that there is a slight increase in moment capacity in the
bolt rupture zone. Figure 8 highlights the difference in behaviour, a comparison is made
between two specimens having the same bolt diameter with different end-plate thickness of 22
and 50mm. For large end-plate thickness the plate can rotate as one unit leading to failure in
bolts due to pure tension, while in specimens with smaller end-plate thickness a deformation
happens leading to the bolts to fail by both tension and local bending which in turn decreases
the overall moment capacity of the connection.

932
Figure 7. Connection moment capacity - end-plate thickness curve.

Figure 8. End-plate deformations in two different bolt rupture specimens.

4.2 Effect of adding horizontal stiffeners


To increase the moment capacity of the connection and redistribute the force distribution
among bolts, a horizontal stiffener is used to enhance the behaviour. As referred in Figures 9
and 10, three curves is plotted representing specimens before and after adding horizontal stiff-
eners between second and third bolts row for M22 and M24 as shown in parametric study
configurations. It is found that adding one horizontal stiffener has great effect on moment
capacity with low (tep/d) ratios reaching 28%. This increase slightly decreases with increasing
(tep/d) ratios until no increase is noted. Since, at bolt rupture zone the end-plate has its own
initial stiffness to redistribute the force among bolts. On the other hand, adding two horizon-
tal stiffeners hasn’t achieve more moment capacity more to the connection. Due to second
stiffener exists more far from the tension flange to redistribute forces among the bolts. Finally,
adding one horizontal stiffener enhance the connection in end-plate rupture zone only.

Figure 9. Connection moment capacity – (end-plate thickness/bolt diameter) for bolt diameter M22.

933
Figure 10. Connection moment capacity – (end-plate thickness/bolt diameter) for bolt diameter M24.

4.3 Bolt force distribution at yielding and fracture


The yielding moment for connection is determined as 2/3 of the ultimate moment according to
the recommendation of (Faella et., 2000). In this study, bolt force distribution among tension
bolts has been studied at both yielding and ultimate limits for the connections. Two specimens
were studied, one exhibited end-plate failure mode (M16-14) and other exhibited bolt rupture
mode (M16-50). By investigating the force distribution on bolts For M16-14 shown in Figure 11,
it is concluded that:

Figure 11. Bolt force distribution among tension bolts for specimen M16-14 at yielding and fracture limit.

934
1- Adding one stiffener has nearly no effect on bolt force distribution among bolts in yield-
ing limit.
2- Adding one stiffener redistributed forces and increased bolt tension force in bolt B32 by
13% in ultimate limit.
3- Adding two stiffeners hasn’t any enhancement on the bolt force distribution in both
yielding and ultimate limit.
While in M16-50 as shown in Figure 12, it can be concluded that adding stiffeners (one
or two) has no effect on the bolt force distribution in yielding and ultimate limit. As in
bolt rupture controlled specimens, the end-plates are characterized by thick thicknesses
which achieve its own stiffness to redistribute the force among bolts without any need for
stiffeners.

4.4 Effect of stiffener thickness


As concluded in section 4.3, adding one horizontal stiffener enhance moment capacity of end-
plate connection especially in end-plate rupture zone. Thickness of horizontal stiffeners effect is
studied for three different bolt diameters with variable stiffener thickness and constant end-plate
thickness tep = 16 mm. As shown in Figure 13, a curve is plotted representing the connection
moment capacity versus ratio between the stiffener to end-plate thickness. It is concluded that
increasing stiffener thickness effects the moment resistance until a certain limit (ts = 0.75tep), then
any increase has no effect.

Figure 12. Bolt force distribution among tension bolts for specimen M16-50 at yielding and fracture limit.

935
Figure 13. Moment capacity – (stiffener thickness/end-plate thickness) for 16mm end-plate.

5 CONCLUSIONS

The following conclusions can be obtained from this study:-


– The relation between end-plate thickness and connection moment capacity is divided into
two zones; the first zone represents end-plate rupture with a nearly linear relation while
the second part represents bolt rupture and the curve is nearly horizontal.
– Adding one horizontal stiffener enhances the moment capacity by 26% in end-plate failure
zone and vanishes in bolt rupture zone. While, adding the second stiffener has no extra
benefit to neither the overall moment capacity of the connection nor for the bolt force
distributions.
– Redistribution of bolt force among bolts is affected by adding one stiffener especially at
ultimate limit to specimens share end-plate rupture. While, at yielding limit of connection
stiffeners hasn’t any effect.
– Increasing the thickness of the horizontal stiffeners increases the moment capacity of con-
nection until a certain thickness. It is recommended to be (0.75) thickness of end-plate.

REFERENCES

AISC, 2010. ANSI/AISC 360-10 Specification for Structural Steel Buildings[S], Chicago.
Demonceau, J. et al., 2010. Application of Eurocode 3 to steel connections with four bolts per horizontal
row. In SDSS’Rio 2010 STABILITY AND DUCTILITY OF STEEL STRUCTURES.
EN1993-1-8:2005, 2005. Eurocode 3: Design of Steel Structures, Part 1-8: Design of Joints[S], British
Standards Institution.
Faella, C., Piluso, V. & Rizzano, G., 2000. Structural steel semirigid connections: theory, design, and soft-
ware, Boca Raton, Florida: CRC Press LLC.
Gang Shi, Xuesen Chen, D.W., 2017. Experimental study of ultra-large capacity end-plate joints. Journal
of Constructional Steel Research, pp.354–361.
Mohamed Eldemerdash, Taher Abu-Lebdeh, M.A.N., 2012. Finite element analysis of large capacity
endplate steel connections. Journal of Computer Science 8 (4), pp.482–493.
Prinz, G. et al., 2014. Experimental testing and simulation of bolted beam-column connections having
thick extended endplates and multiple bolts per row. Engineering Structures, pp.434–447.
Samaan, R., 2010. Behavior and design of steel I-beam-to-column rigid bolted connections. PhD diss.,
Ph. D. thesis, Ain Shams University, Cairo, Egypt, 2010.
Samaan, R.A., El-Serwi, A.A.I. & El-Hadary, R.A., 2017. 01.05: Experimental and theoretical study of
large capacity extended end-plate moment connection. EUROSTEEL 2017, 1(2–3),pp. 205–214.
Shi, G. et al., 2008. Numerical simulation of steel pretensioned bolted end-plate connections of different
types and details. Engineering structures, pp. 2677–2686.

936
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Sensitivity of the stiffness reduction model used to analyze the


ultimate load condition of steel frames

B. Rosson, T. Villalon-Camacho & H. Gurneian


Florida Atlantic University, Boca Raton, Florida, USA

R. Ziemian
Bucknell University, Lewisburg, Pennsylvania, USA

ABSTRACT: The stiffness reduction of compact wide-flange steel shapes with an ECCS
residual stress pattern was studied using a detailed fiber element model. For a given moment,
axial load and residual stress ratio, the reduced stiffness conditions were evaluated and used to
develop a new inelastic material model. Ultimate load analyses were conducted on three bench-
mark frames using the new model and other material models from the literature, and the results
were compared with those from advanced nonlinear finite element models. Discussion is given
regarding the material models and their ability to match the ultimate load capacity results.

1 STIFFNESS REDUCTION MATERIAL MODELS

1.1 Introduction
Appendix 1 of the AISC Specification for Structural Steel Buildings (2016) allows for the use
any method that uses inelastic analysis to proportion members with localized yielding provided
it meets specified strength, ductility and analysis requirements. Researchers over the past couple
of decades have proposed different methodologies and stiffness reduction models to account for
the spread of plasticity in steel frames (Ziemian & McGuire 2002, Surovek-Maleck & White
2004, Zubydan 2010, Kucuckler et al. 2016). A new inelastic material model is proposed that
allows for the input of parameters to adjust the stiffness reduction based on the W-Shape cross-
section dimensional properties, axis of bending and residual stress ratio. To assess the accuracy
and sensitivity of material models used to determine the limit load conditions of steel frames,
the new inelastic material model results are compared to those obtained when using two other
inelastic material models currently available in MASTAN2 (2015).

1.2 New proposed material model


The stiffness reduction τ that results from yielding of the cross-section due to major- or minor-
axis bending and axial compression was studied in detail using a fiber element model for
W-Shapes with an ECCS (1984) residual stress pattern (Rosson 2017). The model used 2,046
fiber elements over the cross-section (400 fiber elements in each flange and 1,246 fiber elem-
ents in the web). For a given normalized moment m = M/Mp, axial load p = P/Py, and residual
stress ratio cr = σr/σy, the stiffness reduction was carefully assessed for a W8x31 with cr = 0.3.
Using the m and p results with loading increments of 0.01, over 7,000 data points were used to
produce the m-p-τ plot in Figure 1(a) for major-axis bending and axial compression.
The equation to determine the extent of τ = 1 in Figure 1(b) is found in the literature
(Attalla et al. 1994). For a given residual stress ratio cr and axial compression load condition

937
Figure 1. (a) major-axis bending m-p-τ surface plot, (b) new inelastic material model.

p, the maximum normalized moment m1 at which τ = 1 is maintained for major-axis bending


is given as
Sx
m1 ¼ ð1  cr  pÞ ð1Þ
Zx

where Sx is the major-axis elastic section modulus and Zx is the major-axis plastic section modu-
lus. The maximum normalized moment m1 at which τ = 1 is maintained for the minor-axis bend-
ing and axial compression condition is determined in a similar manner and is found to be
Sy
m1 ¼ ð1  cr  pÞ ð2Þ
Zy

where Sy is the minor-axis elastic section modulus and Zy is the minor-axis plastic section
modulus.
The stiffness reduction when m = 0 and p > 1- cr is given as τp in Figure 1(b) and is determined
by dividing the major-axis second moment of the area of the remaining cross-section that has not
yielded by the original major-axis second moment of the area Ix. The relationship for τp assuming
an ECCS (1984) residual stress pattern is found to be
2 sffiffiffiffiffiffiffiffiffiffiffi!3 3 sffiffiffiffiffiffiffiffiffiffiffi
1p 5 1  ph i
λλ21 41  1 þ 2 þ 6ð1 þ λ1 Þ2
cr cr
τp ¼ ð3Þ
λλ21 þ 2 þ 6ð1 þ λ1 Þ2

where λ = Aw/Af and λ1 = dw/tf (Rosson 2017). The stiffness reduction τp for the minor-axis
condition is determined in a similar manner and is found to be

sffiffiffiffiffiffiffiffiffiffiffi!3 sffiffiffiffiffiffiffiffiffiffiffi
1p 1p
2 þ λλo
2
cr cr
τp ¼ ð4Þ
2 þ λλ2o

938
where λo = tw/bf (Rosson 2016). For W-Shapes in which λλo2 is very small compared to 2,
a close approximation to Equation 4 that excludes the web effect is given as
 
1  p 3=2
τp ¼ ð5Þ
cr
For a given p condition, two equations are needed to determine m0 when τ = 0 in Figure 1(b).
Closed-form equations are given by Chen and Sohal (1995); however, the same results can be
obtained with fewer computations using the constants λ, λo and λ1 (Rosson 2016). For the
major-axis bending with axial compression condition, one equation is needed when the plastic
neutral axis is in the web, and the other equation is needed when it is inside the flange thickness.

p2 ð2 þ λÞ2 λ
m0 ¼ 1  when p < ð6Þ
4λo þ λð4 þ λÞ 2þλ

ð2 þ λ1 Þ2  ½pð2 þ λÞ  λ þ λ1
2 λ
m0 ¼ when p  ð7Þ
4 þ λ1 ð4 þ λÞ 2þλ

For the minor-axis bending with axial compression condition, one equation is needed when
the plastic neutral axis is inside the web thickness, and the other equation is needed when it is
outside the web thickness.

p2 ð2 þ λÞ2 2λo þ λ
m0 ¼ 1  when p < ð8Þ
ð2 þ λλo Þð2 þ λ1 Þ 2þλ

4  ½pð2 þ λÞ  λ
2 2λo þ λ
m0 ¼ when p  ð9Þ
2ð2 þ λλo Þ 2þλ

The new proposed inelastic material model takes advantage of the closed-form equations for the
perimeter conditions given by m1, τp and m0. The 3D surface in Figure 1(b) was developed from
Equations 10 and 11. An appropriate value for the exponent n is selected based on fiber
element m-p-τ results for a W-Shape with its given cross-section dimensional properties and the
axis about which bending occurs. In the absence of any other effort to determine an appropriate
n value for a given W-Shape, it is recommended to use n = 4 for major-axis bending and n = 2 for
minor-axis bending.
 
m  m1 n
τ ¼1 when p < 1  cr ð10Þ
m0  m1
 n
m
τ ¼ τp 1  when p  1  cr ð11Þ
m0

For a given axial compression p condition, and a W-Shape with its λ, λo, λ1 and cr constants,
the m1, τp and m0 values are evaluated from Equations 1, 3, 6 and 7 for major-axis bending, and
Equations 2, 5, 8 and 9 for minor-axis bending. As illustrated in Figure 1(b), for a given p and
its corresponding m1, τp and m0 values, τ is evaluated based on the magnitude of the normalized
moment m. If p < 1 – cr and m  m1, there is no stiffness reduction and τ = 1. If p < 1 – cr
and m > m1, there is stiffness reduction between 1 and 0 using Equation 10 depending on the
magnitude of m. If p  1 – cr, then stiffness reduction is between τp and 0 using Equation 11
depending on the magnitude of m. If m  m0 for any given p condition, then τ = 0.

1.3 Et material model


To approximate the residual stress effects on member behavior, the tangent modulus in Equa-
tion 12 is used in combination with the yield surface in Equation 13 to account for full cross-

939
section yielding at the end of the element (Ziemian & McGuire 2002). Equation 12 is derived
from the Column Research Council column equation and has frequently been used in the sta-
bility analysis of steel framed structures (Galambos 1998). If p ≤ 0.5, then τ = 1.

τ ¼ 4pð1  pÞ when 0:5 < p  1:0 ð12Þ

F ¼ p2 þ m2x þ m4y þ 3:5p2 m2x þ 3p6 m2y þ 4:5m2x m2y ¼ 1 ð13Þ

where mx = Mx/Mp is the normalized major-axis moment and my = My/Mp is the normalized
minor-axis moment.

1.4 Etm material model


To better capture the inelastic behavior of W-Shapes due to yielding from both bending and
compression, Ziemian and McGuire (2002) modified the tangent modulus approach in Equa-
tion 12 to the following expression.
 

1
τ ¼ min ð14Þ
ð1 þ 2pÞ 1  p  0:65my  m2x

2 ANALYSIS RESULTS

2.1 Column strength study


Column limit load analyses were conducted on a W8x31 pinned-pinned column with minor-axis
bending to compare the results using the new proposed model with those using the Et and Etm
material models. All of the member cross-sections were assumed to be fully-compact and their
out-of-plane behavior fully restrained. All column analyses were conducted using a second-order
inelastic analysis with E = 200 GPa and Fy = 345 MPa. The initial geometric imperfections were
directly modeled using the first eigen-mode (equivalent to a half sine wave) normalized to L/1000
at mid-height. The column was modeled with MASTAN2 (2015), and in all cases using 10 line
elements. Limit load analysis results of columns with L/r values up to 200 are given in Figure 2.
Equations (E3-2) and (E3-3) from the Specification for Structural Steel Buildings (2016) are

Figure 2. Column strength comparison of W8x31 with minor-axis bending.

940
included in the figure as a curve of reference to the modeled results. The limit load results are con-
sistently higher for the Et model and generally lower for the Etm model. A comparison of the
curves indicates that the new proposed model generally gives results between the Et and Etm
results and more closely approximates the AISC (E3-2) and (E3-3) curve. This simple column
example illustrates that even when using the same structural model and analysis conditions, fairly
significant discrepancies exist in the limit load results based solely on the stiffness reduction model
used in the analysis.

2.2 Test frame studies


System limit load analyses were conducted on three benchmark frames using the new proposed
model and were compared with the results obtained from advanced finite element models and
from models using Et and Etm. All of the member cross-sections were assumed to be fully-
compact and their out-of-plane behavior fully restrained. Frames were analyzed using a second-
order inelastic analysis in accordance with Appendix 1 of the Specification for Structural Steel
Buildings (2016). All given loads were defined to result in a system limit load condition of the
finite element models using FE++ in MASTAN2 (2015) at an applied load ratio of 1. The
modulus of elasticity and yield stress were reduced by 0.9, and the initial geometric imperfec-
tions were directly modeled using L/500 in the compounding direction. Frame analyses with the
three different material models had load increments of 0.005 of the total load, and either 10 or
20 line elements per member as indicated in Figures 3, 6 and 8. All members were oriented such

Figure 3. Test frame 1 (Miller 1995).

Figure 4. Test frame 1 location and extent of stiffness reduction (a) new proposed model, (b) Etm model.

941
Figure 5. Applied load ratio vs. lateral displacement comparison for test frame 1.

Figure 6. Test frame 2 (Martinez-Garcia 2002).

that major-axis bending occurred, except for the columns in Figure 3. The shaded regions in
Figure 4 illustrate the location and relative extent of stiffness reduction at the limit load. For
each member where there is no shaded portion τ = 1, and where the shaded portion touches in
the middle τ = 0. Comparison of the shaded regions in Figure 4(a) and 4(b) indicate a fairly
significant difference in the modeled stiffness reduction between the new proposed model and
the Etm model at the limit load.
The results in Figure 5 for test frame 1 illustrate the differences in the applied load ratio at
the limit load and the final lateral displacements at the top right corner of the frame. The

942
Figure 7. Applied load ratio vs. lateral displacement comparison for test frame 2.

Figure 8. Test frame 3 (Schimizze 2001, Martinez-Garcia & Ziemian 2006).

results in Figure 7 for test frame 2 illustrate a closer convergence of final results. Due to the
relatively low axial forces in test frame 2 at the limit load, there is no loss of stiffness when
using the Et model. The nonlinear behavior observed for the Et model is the result of only
elastic second-order effects and yield surface control defined by Equation 13.

943
Figure 9. Applied load ratio vs. lateral displacement comparison for test frame 3.

Figure 10. Test frame 3 relative percent difference in ultimate lateral load conditions, cr = 0.18 and 0.42.

Test frame 3 is given in Figure 8. The results in Figure 9 for test frame 3 illustrate very simi-
lar results between the new proposed model and the Et model with yield surface control. The
Etm model gives the lowest applied load ratio results for all three test frames, and this
response is consistent with the column strength results in Figure 2 and with the more extensive
stiffness reduction throughout the frame as illustrated in Figure 4(b).
Test frame 3 was studied further by first applying the vertical loads to a given normalized
magnitude pi, then the W lateral loads were applied in increments up to their maximum value
at which instability occurred. Since the new proposed model can accommodate any residual
stress ratio, limit load results were compared between the frame modeled with all members

944
having cr = 0.18 versus with cr = 0.42 (Shayan et al. 2014, Rosson 2018). Figure 10 reveals a
significant residual stress effect when pi > 0.35. Since the Et and Etm models cannot consider
a given residual stress ratio, they are not able to detect this effect.

3 CONCLUSIONS

This study investigated the degree to which several inelastic material models provide limit
load results that are consistent with one another. Given that Section 1.3 of Appendix 1 of the
Specification for Structural Steel Buildings (2016) allows for any method that uses inelastic
analysis to proportion members with localized yielding provided it meets specified strength,
ductility and analysis requirements, it is important for the designer to recognize that material
models can give significantly different limit load results and to be aware of their limitations.
The new proposed material model provides consistently closer limit load results when com-
pared to the advanced finite element results, and it allows for direct input of paramters to
adjust the stiffness reduction based on the W-shape’s dimensional properties, axis of bending
and the residual stress ratio.

REFERENCES

American Institute of Steel Construction, 2016. Spec. for Struct. Steel Bldgs. AISC 360-16. Chicago, IL.
Attalla, M.R., Deierlein, G.G. & McGuire W. 1994. Spread of plasticity: quasi-plastic-hinge approach.
Journal of Structural Engineering 120: 2451–2473.
Chen, W. & Sohal, I. 1995. Plastic design and second-order anal. of streel frames. New York: Springer.
European Convention for Constructional Steelwork, 1984. Ultimate limit state calculation of sway
frames with rigid joints. ECCS TC 8 Report No. 33. Brussels.
Galambos, T.V. (ed.) 1998. Guide to Stability Design Criteria for Metal Structures. New York: Wiley.
Kucuckler, M., Gardner, L. & Macaroni, L. 2016. Development and assessment of a practical stiffness reduc-
tion method for the in-plane design of steel frames. Journal of Constructional Steel Research 126: 187–200.
Martinez-Garcia, J.M. 2002. Benchmark studies to evaluate new provisions for frame stability using second-
order analysis. MS Thesis, Department of Civil and Environmental Engineering, Bucknell University,
Lewisburg, PA.
Martinez-Garcia, J.M. & Ziemian, R.D. 2006. Benchmark studies to compare frame stability provisions.
Proc. 2006 Annual Stability Conference, SSRC, San Antonio, TX.
Miller, A.R. 1995. Advanced second-order inelastic analysis of steel structures with columns experiencing
minor axis bending subject to strength limit state requirements. MS Thesis, Department of Civil Engin-
eering, Bucknell University, Lewisburg, PA.
Rosson, B.T. 2016. Elasto-plastic stress states and reduced flexural stiffness of steel beam-columns. Proc.
2016 SSRC Annual Stability Conference, SSRC, Orlando, FL.
Rosson, B.T. 2017. Major and minor axis stiffness reduction of steel beam-columns under axial compres-
sion and tension conditions. Proc. 2017 SSRC Annual Stability Conference, SSRC, San Antonio, TX.
Rosson, B.T. 2018. Modeling the influence of residual stress on the ultimate load conditions of steel
frames. Proc. 2018 SSRC Annual Stability Conference, SSRC, Baltimore, MD.
Schimizze, A. 2001. Comparison of p-delta analyses of plane frames using commercial structural analysis
programs and the AISC specifications. MS Thesis, Department of Civil and Environmental Engineer-
ing, Virginia Tech, Blacksburg, VA.
Shayan, S., Rasmussen, K.J.R. & Zhang, H. 2014. Probabilistic modeling of residual stress in advanced
analysis of steel structures. Journal of Constructional Steel Research 101: 407–414.
Surovek-Maleck, A.E. & White, D.W. 2004. Alternative approaches for elastic analysis and design of
steel frames. I: Overview. Journal of Structural Engineering 130(8): 1186–1196.
Surovek-Maleck, A.E. & White, D.W. 2004. Alternative approaches for elastic analysis and design of
steel frames. II: Verification studies. Journal of Structural Engineering 130(8): 1197–1205.
Ziemian, R.D. & McGuire, W. 2002. Modified tangent modulus approach, a contribution to plastic
hinge analysis. Journal of Structural Engineering 128: 1301–1307.
Ziemian, R.D. & McGuire, W. 2015. MASTAN2, Version 3.5.
Zubydan, A.H. 2010. A simplified model for inelastic second order analysis of planar frames. Engineering
Structures 32(10): 3258–3268.

945
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Statistical evaluation of the bearing capacity of short polygonal


columns

G. Sabau, E. Koltsakis, & O. Lagerqvist


Luleå University of Technology, Luleå, Sweden

P. Manoleas
Norwegian University of Science and Technology, Trondheim, Norway

ABSTRACT: Regular convex polygon sections (RCPS) are commonly used as towers sup-
porting transmission lines, stadium lightning and street lamps. Their use provides advantages
in the bearing capacity and can simplify erection. Over the last 50 years experimental studies
have been conducted to check the applicability of the plate theory to stocky polygonal col-
umns. The paper presents the processed data gathered from compression tests found in the
literature. Results from 70 specimens tested under pure compression were statistically ana-
lysed. Specimens with yield strength varying from 235 to 700 MPa and angles varying from
144 to 175.5 (5 to 40 sides) were investigated. The local non-dimensional slenderness was cal-
culated using buckling lengths according to EN 1993-1-3 and EN 1993-1-5 with values ranging
from 0.55 to 4.52. The objective of the paper was to compare the plate buckling resistance
predictions to the experimental results. The paper concludes with a buckling width recommen-
dation for evaluating the critical stress as calculated according to EN 1993-1-3 or EN
1993-1-5.

1 INTRODUCTION

High-rise towers are traditionally realized as lattice structures or monopoles hollow sections.
In the latter case the preferred section is the cylinder. As an alternative, a regular convex pol-
ygonal sections (RCPS) can be used, offering several advantages. In contrast to the cylindrical
shells, limited studies can be found for the structural behaviour of RCPS. Even though experi-
ments on RCPS profiles can be found in the literature, no systematic statistical evaluation has
been performed so far. The objective of this paper is to evaluate the current rules available
against experimental data.

1.1 General
The popularity of the RCPS increased significantly once cold-formed elements were intro-
duced in design codes. Cold-forming allowed manufacturing with minimum welding, which is
highly important for fatigue, posing a major threat for such structures.
More importantly, the number of circumferential welds needed is greatly reduced compared
to cylinders. The length of the individual pieces is limited by the capacity of the press-brake
and transportation conditions. Circumferential welds can be entirely avoided if the parts are
connected with gravity slip joints, a method described in EN 50341-1.
Monopole towers demand an increased moment of inertia at the base, driving the design to
a conical shape. Achieving this shape imposes manufacturing difficulties, both for the trad-
itional can-welding technique and the innovative spiral-welded towers. In contrast, it is very
simple to taper an RCPS tower. Examples of tapered towers can be seen in Figures 1 and 2 as
supports for lighting and electrical lines.

946
Figure 1. Example of RCPS light tower.

Figure 2. Example of RCPS electricity support tower.

This study investigates the local behaviour of RCPS, with the expected failure mechanism
due to the high stresses near the base. For higher slenderness members under pure compres-
sion (λ ≥ 0.6), Sabau (2018) indicates that RCPS can be safely designed using the buckling
curve “c”, provided that the profile is not in class 4.

1.2 Geometric and structural properties


To describe the application of the alternative design methods, it is necessary to name some
basic geometric properties. The cross-section of a hexagonal RCPS is shown in Figure 3.
The RCPS resistance is calculated based on plate theory according to EN 1993-1-1, EN
1993-1-5 using the width of the flat plate bf and the cross-section area Ap.
 
bf ¼ 2 rp sin θ  rb tan θ ð1Þ
 
Ap ¼ 2t nv rp sin θ  nv ab tan θ þ πab t ð2Þ

where nv = number of sides; and ab ¼ rb =t


To apply the calculations of EN 1993-1-6, an equivalent cylinder must be defined. The most
intuitive way is to consider the cylinder with equal wall thickness and cross-section area to
a RCPS. By doing so, the radius of the equivalent cylinder is expressed as follows.
 
rc ¼ nπv rp sin θ  ab t tan θ þ ab t ð3Þ

947
Figure 3. Basic geometric properties of a RCPS.

1.3 Description of experiments


In this study, experiments from 7 different sources are gathered, with low flexural slenderness,
λ ≤ 0.2. Bulson (1969) performed the first and most extended experimental study (39 tests with
nv varying from 4 to 40 and high local slenderness (56 ≤ b/t ≤ 790).
The second study including experiments on RCPS was performed by Aoki et al. (1991). In
their paper they presented results from 15 experiments with a much lower local slenderness
and fewer sides (b/t ≤ 66.7 and nv ≤ 8). Soon after, Migita et al. (1992) added 9 more tests
(similar nv and b/t) with flexural slenderness reaching up to λ = 0.5.
Further on, 8 tests on octagonal profiles were performed by Harraq (1997), 4 of which were
regular (RCPS). Only 2 of them were tested under pure compression.
A more systematic study was performed by Godat et al. (2012) who reported previous
experiments and performed 6 new. Apart from Bulson’s slender profiles, this study provided
the first results for nv > 8. Specimens of 8, 12 and 16 sides with local slenderness in the region
of interest to civil engineering (31 ≤ b/t ≤ 56) were tested.
A numerical investigation coupled with 9 laboratory tests was performed by Manoleas
(2018). The 9 additional tests covered higher-order polygons having 16, 20 and 20 sides. These
are the only tests in high strength steel (S700).
More recently, 3 additional tests on octagonal specimens were performed by Jiong-Yi et al.
(2019) providing experimental results for thicker profiles (b/t = 10.8).

2 CALCULATION OF COMPRESSIVE RESISTANCE

Considering European standards, RCPS can be designed based on EN 1993-1-1 in conjunc-


tion with either EN 1993-1-5, EN 1993-1-3 or EN 1993-1-6. In the first two cases, the facets of
the RCPS are assumed as simply supported plates, adopting two definitions for the buckling
width. In the third case, the RCPS is assumed to be a cylindrical shell.
Additionally, some design rules specifically for RCPS are given in EN 50341-1. This stand-
ard extends the classification definitions of EN 1993-1-1 to RCPS of 6 to 18 sides.

2.1 Resistance according to EN 1993-1-5


By employing EN 1993-1-5, the RCPS is perceived as an extended rectangular hollow section
(RHS). As such, the cross-section resistance is the sum of its individual facets.

948
This definition implies two assumptions:
– The buckling width is equal to the flat portion of the facet.
– Each facet acts as a simply supported plate.
The resistance is calculated for the reduced effective width of the plate as follows:

NRk ¼ Aeff  fy ð4Þ


 
Aeff ¼ nv A0;eff þ Acorner ð5Þ

where A0;eff ¼ ρbf t and ρ = plate buckling reduction factor, EN 1993-1-5 4.4.

2.2 Resistance according to EN 1993-1-3


The application of EN 1993-1-3 for the RCPS requires the same basic assumption of the
simply-supported plate. Oriented towards cold-formed members, EN 1993-1-3 provides
a more sophisticated approach for the plate buckling width, which is measured between the
midpoints of two adjacent corners on the mid-thickness line of the profile (Figure 3).
Note that EN 1993-1-3 regards to thin gauge members and sheeting and its scalability to
larger structures is questionable. In addition, the use of EN 1993-1-3 is limited to cold-formed
corners between 135° and 90°, which implies RCPS with 4 up to 8 sides.

2.3 Resistance according to EN 1993-1-6


An alternative approach to calculate the resistance of a RCPS is to assume the equivalent
cylinder and apply the hand calculation procedure of EN 1993-1-6. This assumption becomes
increasingly valid as the number of sides increases.
The local behaviour of an RCPS column under pure compression is equivalent to a short
cylinder with a uniform compressive meridional stress field. Eurocode defines four limit
states, only two of which are relevant to monotonic compression loading:
– Plastic limit state, LS1.
– Buckling, LS3.
Von Misses equivalent stress is considered for LS1. Thus, the case of simple meridional com-
pression this check is identical to the cross section check of EN 1993-1-1. On LS3, a reduction
is applied to the characteristic yield stress, leading to the following formula.

NRk;shell ¼ A σx;Rd ¼ χx Afyk ð6Þ

The reduction factor χx is calculated based on the critical stress, σx,Rcr, and the local imperfec-
tions. The critical stress is a function of the non-dimensional shell slenderness, ω = l/(r t)0.5.

3 STATISTICAL EVALUATION OF EXPERIMENTAL DATA

The ratio of the reported experimental resistance values Nu and the characteristic resistance
NRk, using EN1993-1-3 and EN1993-1-5 as per Section 2 are presented in Figures 4 and 5
along with the probability density function (PDF), assuming a Gaussian distribution around
the mean value. The kernel density estimate (KDE) was represented using a kernel size deter-
mined by means of a Gaussian estimator. The KDE gives a compromise between the histo-
gram and the normal distribution. A linear regression was performed having as parameter the
number of sides. The 90% confidence interval was represented for the regression estimate.
The characteristic value Xk was calculated as per Equation 7 and is represented as the
bottom limit (dashed line) of the standard deviation in Figures 4 to 10. In Equation 7, kn was
taken as 1.64, assuming that the result scatter would not change for higher number of test.

949
Figure 4. Scatter data and linear regression of the ratio between the experimental results and
EN1993-1-3 calculated resistance for RCPS

Figure 5. Scatter data and linear regression of the ratio between the experimental results and
EN1993-1-5 calculated resistance for RCPS.

  kn  s
Xk ¼ x ð7Þ

where: x = mean sample value; kn = is a coefficient considering the number of tests; and s =
sample mean standard deviation.
Figures 4 and 5 clearly show that the assumption of pure plate buckling does not stand with
increasing number of sides.
The applicability of the EN 1993-1-3 is limited to bending angles of 90° to 135° between
two adjacent facets. Thus, it should allow for rectangular and octagonal sections. Addition-
ally, the EN 1993-1-3 rules are not applicable to cold-formed sections that do not satisfy the
bending radius r criteria as per Equation 8.

r > 0:04tE=fy ð8Þ

where: t = plate thickness; E = elastic modulus; and fy = yield strength


Figure 6 presents a linear regression of the test data as a function of number of sides.
The statistical quantities were calculated for all the reported data. The information
found in the literature was not sufficient to verify the requirement of Equation 8. How-
ever, the limitation allows for very large bending radius seldom used in cold-forming
technology.
Several octagonal specimens that reached a lower resistance than the value predicted by the
EN1993-1-3 were reported to have plate imperfections e0/b of magnitude between 1/30 to 1/40.
The reported imperfections do not satisfy the minimum requirements imposed by the
EN1090-2 for cold-formed profiles.

950
Figure 6. Scatter data and linear regression of the ratio between the experimental results and
EN1993-1-3 calculated resistance, the KDE and the normal PDF for RCPS of 5 to 8 sides.

Figure 7. Scatter data and linear regression of the ratio between the experimental results and
EN1993-1-3 calculated resistance, the KDE and the normal PDF for RCPS of 5 to 18 sides.

EN 1993-1-3 provides a safer estimate than EN 1993-1-5 for RCPS of up to 18 sides. The
statistical quantities do not differ significantly between the sampled population consisting of
RCPS with up to 8 sides and 18 sides as seen in Figures 6 and 7. Thus, EN 1993-1-3 is better
for the estimating the plate buckling of RCPS up to 18 sides.
Annex K of EN 50341-1 recommends the use Table 5.3.2 of ENV 1993-1-1 (replaced
by table 4.1 of EN 1993-1-5) for effective widths of RCPS of 6 to 18 sides. The linear
regression of Figure 8 indicates a high uncertainty concerning this methodology. Thus,
the recommended safety factor γM1 ¼ 1:1 might be insufficient. EN1993-1-1 provides
rules for plates thicker than 3 mm. Figure 8 shows the results of the tested RCPS.

Figure 8. Scatter data and linear regression of the ratio between the experimental results and EN
1993-1-5 calculated resistance, the KDE and the normal PDF for RCPS of 6 to 18 sides and thickness
above 3 mm.

951
Figure 9. Scatter data and linear regression of the ratio between the experimental results and EN
1993-1-6 calculated resistance, the KDE and the normal PDF for RCPS of 20 to 40 sides.

Figure 10. Scatter data and linear regression of the ratio between the experimental results and EN
1993-1-6 calculated resistance, the KDE and the normal PDF for RCPS of 20 to 40 sides and λp < 1.75.

The limit of 18 sides of EN 50341-1 is justified by the failure mode change. The buckling
wavelength is no longer restricted between edges, but extends to adjacent facets. In Figure 9
the ratio between the experimental and EN 1993-1-6 calculated resistance shows a significant
uncertainty for the lower bound limit of sides. Results from RCPS between 20 and 22 fall
below the predicted values and show large scatter.
Figure 10 contains the results from the RCPS with plate slenderness λp < 1.75, equivalent to
50% effective area reduction. With that limitation, the sampled data show a more conservative
trend. This limitation seems counterintuitive at first since a higher plate slenderness should push
the buckling mode towards a plate buckling collapse. However, a higher plate slenderness ultim-
ately leads to a reduced in-plane stiffness of the bended corners. Buckling resistance is thus sensi-
tive to interaction between plate and distortional imperfections, causing premature opening of the
corners. It is unclear where the turning point between plate and meridional buckling is since it is
a combination of multiple factors. A formula to predict the transition point between the two
modes was suggested by Manoleas (2018). The formula is applicable even to high-strength steels.
The mean value, standard deviation, the coefficient of variation and the lower 5% fractile of
the sampled data has been summarized in Table 1.

4 CONCLUSIONS

A total of 70 centrically compressed tests performed on cold-formed RCPS specimens have


been collected from literature and compared against current design standards. The results
showed that the design standards are unsafe if simplified approaches are used. For 5-8 sides

952
Table 1. Statistical quantities of the analysed data.
Number of sides Model Mean Standard Coefficient 5% Characteristic
deviation of variation fractile value

EN 1993-1-3 0.934 0.142 0.152 0.679 0.701


All
70 specimens EN 1993-1-5 0.907 0.136 0.150 0.657 0.684
EN 1993-1-6 0.822 0.245 0.299 0.367 0.419
5 – 8 sides EN 1993-1-3 1.000 0.114 0.114 0.772 0.814
28 specimens EN 1993-1-5 0.962 0.116 0.121 0.739 0.771
EN 1993-1-3 0.985 0.133 0.135 0.707 0.767
6 – 18 sides
43 specimens EN 1993-1-5 0.953 0.131 0.137 0.693 0.739
EN 1993-1-6 0.729 0.247 0.340 0.328 0.323
EN 1993-1-61 0.989 0.141 0.142 0.798 0.758
20 – 40 sides EN 1993-1-62 1.023 0.129 0.126 0.885 0.812

1– includes all 25 specimens; 2 – 15 specimens with λp < 1.75

specimens, the EN 1993-1-3 gives safer results as opposed to the EN 1993-1-5. This shows that
the buckling length extends over the flat part of the facet.
The resistance calculated according EN 50341-1 showed good agreement with the test data.
The use of the safety factor brings the characteristic value close to unity for specimens with t > 3.
The methodology of EN 1993-1-6 could be applied to RCPS with more than 18 sides if
higher safety factors are implied. The experimental values of specimens with 20-40 sides came
closer to the EN 1993-1-6. However, there is an uncertainty regarding the specimens with 16-
24 sides since none of the standards provided good estimates for this interval.

REFERENCES

Aoki, T. Migita, Y. Fukumoto, Y. 1991. Local buckling strength of closed polygon folded section
columns. Journal of Constructional Steel Research 20: 259–270.
Bulson, P.S. 1969. The strength of thin-walled tubes formed from flat elements. International Journal of
Mechanical Sciences 11: 613–620.
EN 1090-2. Execution of steel structures and aluminium structures - Part 2: Technical requirements for
steel structures, 2008. CEN: Brussels.
EN 50341-1. Overhead electrical lines exceeding AC 45 kV - Part 1: General requirements - Common speci-
fications, 2001. CENELEC, Brussels.
Eurocode 3: Design of steel structures – Part 1-1: General rules and rules for buildings, 2005. CEN: Brussels.
Eurocode 3: Design of steel structures – Part 1-3: General rules - Supplementary rules for cold-formed
members and sheeting, 2004. CEN: Brussels.
Eurocode 3: Design of steel structures – Part 1-5: Plated structural elements, 2005. CEN: Brussels.
Eurocode 3: Design of steel structures – Part 1-6: Strength and stability of shell structures, 2006. CEN:
Brussels.
Godat, A. Legeron, F. & Bazonga, D., 2012. Stability investigation of local buckling behaviour of tubu-
lar polygon columns under concentric compression. Thin-Walled Structures 53: 131–140.
Harraq, A. 1997. Étude de la stabilité de poteaux d’acier formés à froid de section octogonale (MSc thesis).
Université de Sherbrooke: Québec.
Manoleas, P. 2018. Between square and circle: A study on the behaviour of polygonal steel profiles under
compression (PhD Thesis). Luleå Tekniska Universitet: Luleå.
Migita, Y. Aoki, T. & Fukumoto, Y. 1992. Local and interaction buckling of polygonal section steel
columns. Journal of Structural Engineering 118: 2659–2676.
Sabau, G. Koltsakis, E. & Lagerqvist, O. 2018. Stability analysis of newly developed polygonal
cross-sections for lattice wind towers. Wind Engineering 42: 353–363.
Zhu, J.-Y. Chan, T.-M. Young, B. 2019. Cross-sectional capacity of octagonal tubular steel stub columns
under uniaxial compression. Engineering Structures 184: 480–494.

953
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Direct Strength Method (DSM) design of end-bolted cold-formed


steel columns failing in distortional modes

W.S. Santos & A. Landesmann


Civil Engineering Program, COPPE, Federal University of Rio de Janeiro, Brazil

D. Camotim
CERIS, DECivil, Instituto Superior Técnico, Universidade de Lisboa, Portugal

ABSTRACT: This work addresses the behaviour, ultimate strength and DSM design of end-
bolted simply-supported cold-formed steel lipped channel columns failing in distortional
modes – the end supports consist of a pair of cylindrical high-strength bolts with hexagonal
heads and nuts, and inserted in standard-size holes located at the intersections of the flanges with
the principal centroidal axis parallel to the web. After selecting the column geometries (cross-
section dimensions and lengths) and yield stresses, ANSYS SFEA are used to acquire knowledge
about their non-linear behaviour, strength and collapse, as well as to gather numerical failure
loads. Together with the experimental failure loads previously reported by the authors, these
numerical failure loads are then employed (i) to assess the merits of the available DSM distor-
tional design curves and, in view of their inadequacy, (ii) to propose and validate new strength
curves, aimed at providing efficient predictions for the whole failure load set considered.

1 INTRODUCTION

A past numerical investigation by the authors (Landesmann & Camotim 2013) on the distor-
tional failure strength of simply-supported (end sections locally/globally pinned and allowed to
warp freely) cold-formed steel columns with several cross-section shapes provided solid evidence
that the currently codified Direct Strength Method (DSM – e.g., Schafer 2008 or Camotim et al.
2016) design curve is not able to predict adequately (safely and accurately) the failure loads of
such columns. Indeed, it was found that this design curve overestimates the failure load of most
cold-formed steel columns, namely the lipped channel ones. On the other hand, simple bolted
connections are generally viewed as perfectly pinned/hinged end supports (at least in practice,
for the purpose of simplified structural calculations). This type of end connections, which are
commonly employed in civil engineering applications, consists of pairs of high-strength cylin-
drical bolts with hexagonal heads and nuts, which are inserted in two standard-size holes located
at the intersections of the flanges with the centroidal axis parallel to the web. Figure 1 provides
an overall view of a lipped channel member with the bolted end support conditions and sche-
matic representations of the top and bottom pin-ended supports, each formed by (i) a pair of
M20 A490 high-strength steel bolts (ASTM-A490 2014) and (ii) two pairs of 3 mm thick M20
steel flat washers (ASTM-F436/F0436 2016), welded to the flange hole inner and outer faces –
these washers aim to preclude (or, at least, delay) the occurrence of a localised failure. It is
worth noting that, in order to avoid eccentric loading, the flange holes (nominal diameter df =
21.5 mm) and washer inside/outside circumferences (nominal diameters di = 21.5 mm and do =
37 mm, respectively) are aligned with the lipped channel cross-section minor centroidal axis.
This paper continues an experimental and numerical investigation carried out by Santos
(2017), who showed that bolted end supports influence the collapse mechanism and failure
load of cold-formed steel lipped channel columns buckling and failing in distortional modes –
note that, for practical purposes, these support conditions may deemed to allow free warping

954
of the column end cross-sections. The main purpose of this work is to obtain additional
numerical failure loads, propose a DSM-based design approach for these columns and to
assess its merits in predicting all the available failure loads.

2 NUMERICAL SIMULATIONS

2.1 Column selection


The first task of this work consists of selecting the end-bolted (EB) lipped channel columns to be
analysed, (i) exhibiting cross-section dimensions commonly used in practice with various wall
width proportions (mostly web-to-flange width ratios), (ii) buckling and failing in distortional
modes, and (iii) having lengths associated with single half-wave buckling modes. The selection
procedure was performed by means of ANSYS (2009) shell finite element buckling analyses – details
of the numerical simulations are given in Section 2.2. The output of this effort is the 15 lipped
channel dimensions bw, bf, bl, t (web-flange-lip widths and wall thickness – see Figure 1) and
lengths LD given in Table 1 – the columns are labelled as bw × bf (widths in mm) and
t = 2.65 mm and bl = 10.5 mm are the same for all the columns. Table 1 also provides the column
(i) nominal cross-section areas A, (ii) web-to-flange width ratios bw/bf (equal to ≈0.7, 1.0 and
≈1.43), (iii) critical distortional buckling loads PcrD, calculated with E = 205 GPa and ν = 0.3 (steel
Young’s modulus and Poisson’s ratio) and (iv) ratios Pb1.L/PcrD and Pb1.G/PcrD, where Pb1.L and
Pb1.G are the lowest local and global bifurcation loads – these ratios lie inside the intervals 1.51-
2.07 and 6.93-72. For illustrative purposes, Figure 2 shows the 100 × 70 column signature curve,
providing the variation of its critical buckling load Pcr with the length L (logarithmic scale) – also
shown are the selected column length LD and critical (distortional) buckling mode shape.

2.2 ANSYS shell finite element model


As mentioned before, the column linear buckling and geometrically non-linear elastic-plastic
analyses were carried out in code ANSYS (2009), employing shell finite element model involving
column discretisations into fine SHELL181 element meshes - 4-node shear deformable thin-shell
elements with six degrees of freedom per node and full integration. Convergence studies per-
formed by Santos (2017) showed that accurate results can be obtained, with a reasonable com-
putational effort, by (i) discretising the member into a 5 × 5 mm mesh (refined around the
holes), and (ii) refining this mesh in the close vicinity (around) the rounded corners and, mostly,

Figure 1. Overall view and geometric characteristics of a lipped channel column with bolted end supports.

955
Table 1. Selected columns dimensions, critical buckling loads and bifurcation-to-critical load ratios.
Column bw A LD PcrD Pb1:L Pb1:G
(bw × bf) bf (cm²) (mm) (kN) PcrD PcrD

100×70 1.43 6.52 350 322.9 1.51 6.93


100×100 1.00 8.11 450 215.6 1.81 8.23
100×142.9 0.70 10.38 550 142.5 1.93 10.8
130×91 1.43 8.43 450 235.5 1.54 12.13
130×130 1.00 10.5 550 157 1.83 16.13
130×185.7 0.70 13.45 650 103.0 1.97 24.9
150×105 1.43 9.7 500 198.6 1.59 17.59
150×150 1.00 12.09 600 132.4 1.87 24.43
150×214.3 0.70 15.49 700 86.10 2.03 35.8
180×126 1.43 11.61 550 159.6 1.69 30.77
180×180 1.00 14.47 650 106.4 1.96 44.26
180×257.1 0.70 18.56 800 69.03 2.07 58.6
200×140 1.43 12.88 600 140.9 1.72 39.9
200×200 1.00 18.06 750 93.8 1.89 51.55
200×285.7 0.70 20.6 900 60.76 2.04 72.0

Figure 2. 100 × 70 column signature curve and selected column length and distortional buckling mode
shapes.

the bolt holes – Figure 3(a) shows a general view of the SFE mesh adopted to perform the
column analyses in this work. In order to simulate as accurately as possible the interface regions
between the bolt shanks and hole surfaces, non-linear contact shell finite elements (CSFE) were
employed – Figure 3(b) illustrates the mesh adopted in one interface region of the 200 × 140
column. Although Santos (2017) showed that the CSFE model is able to simulate adequately
the column non-linear behaviour obtained from the experimentally tests (the tests concerning
columns 100 × 142.9 and 200 × 140 were successfully simulated), he also concluded that this
model (i) requires very large and time-consuming computational efforts and (ii) often entails
convergence difficulties, which may lead to a premature interruption of the analysis – similar
problems, associated with the use of CSFE models, have been reported in the literature (e.g.,
Stolarski et al. 2018). These problems prompted the search for an alternative approach to per-
form the simulations – inspired by the concept of surface stresses due to the contact between
two curved bodies, Santos (2017) devised and implemented a methodology not requiring CSFE
and involving the following assumptions, steps and procedures:
(i) On the basis of the available CSFE results, obtain an estimation of the circular arc angle/
length along which the contact between the bolt shank and the hole surface takes place –
a value of 97° was found (see the zoomed view at the right of Figure 3(b)).
(ii) Assume that the whole compressive load is applied to the column through the 97° contact
circular arc lengths of the two bolt hole surfaces ensuring each end support condition –
see Figure 3(a).

956
Figure 3. Column SFE discretisation: (a) general view, (b) detailed of an interface region between a bolt
shank and a hole surface (CSFE model), and (c) loading and support conditions concerning the SFE model.

(iii) Assume that, in each bolt, the compressive load is applied through radial surface stresses lin-
early distributed along the 97° contact circular arc length defined in item (i), as depicted in the
left side of Figure 3(c) – in each bolt, the transversal and longitudinal resultants of these radial
surface stresses are null and equal to half of the applied compressive load, respectively. The
value of the above radial surface stress linear distribution is determined by means of analytical
expressions to calculate the surface stresses caused by the contact between two infinite circular
elastic bodies, found in a book by Young & Budynas (2002). Although the authors are aware
that the use of these analytical expressions out of their domain of validity (the bolt shank and
hole surface are very far from infinite. . .) has no sound justification, it seems fair to argue that
they provide a reasonable approximation of the real surface stresses – this assertion will be
backed by the comparison between SFE and CSFE results, addressed a bit ahead in the paper.
(iv) Concerning the support conditions, the SFE model considers simple supports along the four
bolt 97° contact circular arc lengths, as illustrated on the right side of Figure 3(c). These
supports prevent the hole surface displacements normal to the web surface, while keeping
their in-plane and longitudinal counterparts free – these support conditions may be deemed
to constitute the best possible materialisation of the (“idealised”) simply supports with free
warping, even if there exists some unavoidable (but small) warping restraint. Moreover, to
preclude column rigid body motions, the (iv1) 97° contact circular arc length mid-point in-
plane and (iv2) mid-span mid-web axial displacements were prevented (note that the loading
consists of equal compressive forces applied at the two end bolt pairs).

In order to validate the above SFE model, Santos (2017) performed elastic linear buckling
and elastic-plastic non-linear analyses of columns 100 × 142.9 and 200 × 140 (those success-
fully analysed with CSFE). The comparison between the results obtained with the proposed
SFE and CFSE models showed that both the critical buckling and failure loads were virtually
identical (all differences below 0.5%) – moreover, the non-linear equilibrium paths practically
coincided up to the failure load (there were small differences in the descending branches).
The geometrically and materially non-linear ANSYS SFE analyses carried out to determine the
column failure loads were performed by means of an incremental-iterative technique that com-
bines Newton-Raphson’s method with an arc-length control strategy – the applied compressive
forces are increased in small increments, taking advantage of the ANSYS automatic load stepping
procedure. All the columns analysed (i) contain critical-mode (distortional) initial geometrical
imperfections with small amplitudes (maximum equal to 23% of the wall thickness t, which cor-
responds to the average of the values measured in the specimens tested), and (ii) exhibit an elastic-
perfectly plastic material behaviour (Prandtl-Reuss model: Von Mises yield criterion and associ-
ated flow rule), characterised by E = 205 GPa, ν = 0.3 and several yield stresses fy – no strain-hard-
ening, residual stresses or corner effects were included in the analyses. The incorporation of the
critical-mode initial geometrical imperfections is made automatically by (i) determining the

957
column critical buckling mode shape, through an ANSYS SFE linear buckling analysis (adopting
the same discretisation employed to carry out the subsequent non-linear analysis), (ii) scaling it to
exhibit a maximum displacement equal to 0.23 t and (iii) “transforming” this buckling output
into an input of the non-linear analysis. In agreement with existing column distortional post-
buckling asymmetry studies (e.g., Silvestre & Camotim 2006), the initial geometrical imperfections
involve outward flange-stiffener motions (those shown to lead to lower post-buckling strengths).

2.3 Failure loads – parametric study


This section presents and discusses the performance of a parametric study, intended to validate the
proposed SFE numerical model and gather numerical failure load data, involving a total of 210
columns, which correspond to all possible combination of the (i) 15 column geometries given in
Table 1 and (ii) the 14 yield stresses (fy) considered, selected to enable covering a wide distortional
slenderness (λD) range, varying between 0.4 and 3.5 – recall that λD = (Py/Pc.D)0.5, where Py=A∙fy
s the squash load (A is the cross-section area, given in Table 1). Figure 4(a) plots the numerical
failure load ratio Pu/Py against λD for all the columns analyzed in this work. One readily notices
a huge Pu/Py scatter (vertical dispersion) in the low and moderate slenderness ranges
(λD below around 1.75). In order to identify the source of this very surprising finding, the numer-
ical failure loads of the columns involved inspected in great detail. This inspection revealed that 77
stocky columns (all with fy below 277 MPa) collapse prematurely, due to the occurrence of local-
ised yielding in the flange regions located below the flat washers (see Figures 1 and 3(b)), and,
therefore, never exhibit significant distortional deformations – all the remaining columns fail in
clear distortional modes. This is illustrated in Figures 5(a)-(b), which display the failure modes and
von Mises stress (σVM) contours of 100 × 100 columns with slenderness λD = 0.965 and λD = 1.484,
respectively – the distinct failure mode natures are clearly visible: while the first column fails by
localised yielding, the collapse of the second one is purely distortional. Figure 4(b) shows the
Pu/Py vs. λD plot when all the columns failing by localised yielding are excluded – it is readily
observed from the that the Pu/Py “cloud” are fairly well aligned along a “Winter-type” strength
curve, with very little vertical dispersion. However, this figure also shows that the number of fail-
ure loads concerning columns with low or moderate slenderness is drastically reduced – e.g., there
are only 9 columns with λD < 1.5. In order to enable an appraisal of this reduction, and also of the
failure load drop stemming from localised yielding, the two left columns in Table 2 provide numer-
ical results concerning the 100 × 100 columns analysed, namely the distortional slenderness λD and
failure load Pu – the shaded cells identify the columns failing by localised yielding. Similar results
for the whole set of columns analysed have been reported by Santos (2017).
In order to enlarge the distortional failure load pool concerning columns with low and moderate
slenderness, the 72 columns found to fail by localised yielding are analysed again assuming
the presence of wider washers – their outside nominal diameter (see Figure 1) is increased to
do = 80 mm, instead of the original do = 37 mm – all the other column dimensions remain unaltered.

Figure 4. Plots Pu/Py vs. λD for (a) all the original columns, (b) all the original columns failing in distor-
tional modes and (c) all the original and new columns failing in distortional modes, together with the
tested specimens.

958
Figure 5. Collapse modes and von Mises stress contour of 100x100 columns: do = 37 mm and (a) localised
(λD = 0.965) or (b) distortional (λD = 1.484) failure, and (c) do = 80 mm and distortional failure (λD = 0.965).

It is worth noting that, naturally, a larger washer diameter leads to higher critical distortional
buckling loads – the increase varies between 5%, for the 200 × 285.7 columns, and 24%, for the
100 × 70 columns – largest and smallest cross-section areas, respectively. In order to keep the
original distortional slenderness values, the column yield stresses also increased proportionally.
The inspection of the new column collapse modes showed that 42 of them (out of 72) still fail by
localised yielding – Figures 5(a) and 5(c) compare the failure modes and von Mises stress with the
same slenderness (λD = 0.965) and different washer diameters, making it possible to assess the influ-
ence of this last parameter. The right column in Table 2 provides the failure loads Pu2 of the new
100 × 100 columns – the shaded cells identify again the columns failing by localised yielding
(naturally, their number decreases). The ultimate strengths of the 30 new columns failing in
distortional modes were added to the failure load data pool (25 of these columns are such that
λD < 1.5). Figure 4(c) jointly plots the Pu /Py values of all the original and new columns failing in
distortional modes, as well as the 10 specimens tested by Santos (2017) – it is observed that the
three failure load sets (two numerical and one experimental) follow the same “Winter-type”
strength curve trend, with a quite reasonable vertical dispersion (naturally, the experimental Pu/Py
values are more scattered than the numerical ones).

3 DSM DESIGN CONSIDERATIONS

This section addresses the DSM-based prediction of the lipped channel column experimental
and numerical failure load data gathered in this work, namely by the currently codified design
curve (AISI 2016) – the corresponding nominal failure load (PnD) is given by

PnD
1     for λD  0:561
¼ ð1Þ
Py 1  0:25 λD 1:2 λD 1:2 for λD 40:561

Figure 6(a1) compares this design curve with the (i) numerical (white dots) and (ii) experimental
(crosses) failure load ratios Pu/Py gathered in this work (already shown in Figure 4(c)). Moreover,
Figure 6(b1) plots, also against λD, the “exact”-to-predicted failure load ratios Pu/PnD, providing
a pictorial representation of the quality failure load estimation quality achieved by the current
design curve – also indicated are the Pu/PnD statistical indicators (averages, standard deviations
and maximum/minimum values of the numerical and experimental results). The observation of
these results prompts the following remarks:
(i) It is readily noticed that the numerical and experimental failure loads are grossly overesti-
mated by the current DSM design curve, as quantified by the corresponding Pu./PnD statis-
tical indicators, which read 0.64-0.11-1.00-0.45 – note that the amount of overestimation
increases with λD. At this stage, it is worth recalling that the calibration and validation of this
DSM design curve was based almost exclusively on failure loads of columns with rigid plates
attached to their end cross-sections, thus fully preventing their warping (Schafer 2008).

959
Table 2. 100 × 100 column distortional slenderness values and failure loads.
λD Pu (kN) Pu2 (kN)

0.446 25.4 45.8


0.530 35.8 64.4
0.545 37.7 67.9
0.705 63.3 112.9
0.965 117.5 181.2
1.225 170.8 170.8
1.484 190.9 190.9
1.744 201.5 201.5
2.003 210.2 210.2
2.263 217.5 217.5
2.523 224.0 224.0
2.782 229.0 229.0
3.042 235.1 235.1
3.301 240.7 240.7

Figure 6. (a) Comparison between the available or proposed DSM distortional design curves and the
column failure load ratios Pu/Py, and (b) associated “exact”-to-predicted failure load ratio plots against
the distortional slenderness.

(ii) Findings similar to those described in the previous item were reported by Landesmann &
Camotim (2013), on the basis of numerical failure loads of standard simply supported
columns. It was concluded that the significant differences in distortional post-critical
strength, stemming from the end cross-section warping restraint (warping completely free
or fully prevented), are not adequately reflected by the distortional critical buckling stres-
ses. Based on the numerical failure loads gathered, the authors preliminarily proposed an
alternative DSM design curve for simply supported columns – the corresponding nom-
inal failure load, denoted PnD.SS, is given by the expressions
8
>1 for λD  0:561
PnD:SS <
 1:2   1:2 
¼ 1  0:25 λD λD for 0:5615λD  1:133 ; ð2Þ
Py >
:
 1:5   1:5 
0:65 þ 0:2 λD λD for λD 41:133

only differing from the currently codified ones (see Eq. (1)) for λD > 1.133. Figure 6(a2) makes it
possible to compare this design curve (red line) with the failure load ratios Pu/Py considered in
this work – the corresponding Pu /PnD.SS values are plotted against λD in Figure 6(b2). These
plots clearly show the failure load prediction quality improvement achieved by the strength

960
curves proposed by Landesmann & Camotim (2013), spanning the whole distortional slender-
ness range, as attested by the Pu/PnD.SS indicators: 0.95-0.05-1.19-0.77. However, it must be rec-
ognised that these indicators remain inadequate and that there are a fairly high of unsafe failure
load predictions (Pu /PnD.SS < 1.00) – 153 out of 220 columns considered. Therefore, it is neces-
sary to look for a lower DSM-based strength curve to predict adequately the available experi-
mental and numerical failure loads within the whole column slenderness range considered.
In view of the above findings, a first attempt is made here to find modified DSM design
curves capable of providing adequate (safe, accurate and reliable) predictions for the failure
loads of the end-bolted columns considered in this work and collapsing in distortional modes.
Adopting the approach employed previously (e.g., Landesmann & Camotim 2013), the
authors carried out a “trial-and-error” curve fitting procedure – its output, denoted PnD.EB, is
the modified/lowered DSM-based strength curve
8
PnD:EB < 0:620λ1:85
D for λD  1:588
¼
    ; ð3Þ
Py :
0:55 þ 0:2 λD 1:5 λD 1:5 for λD 41:588

differing from the previous ones (Eqs. (1)-(2)) in the fact that (i) the flat yield plateau and ini-
tial portion of the descending curve (low and moderate slenderness ranges – λD below around
1.5) are replaced by a single curve cast in the form of Johnson’s parabola, (ii) the distortional
slenderness transition value becomes 1.588 and (iii) the coefficient 0.65, appearing in Eq. (2),
is replaced by 0.55. Figure 6(a3) displays the PnD.EB strength curve (blue line) and the Pu
/PnD.EB values are plotted against λD in Figure 6(b3). It is readily observed that the failure load
prediction quality improves quite significantly, as attested by the Pu/PnD.EB indicators: 1.11-
0.06-1.40-0.89. Moreover, note also that the number of unsafe failure load predictions falls
considerably from 153 to 12 (73% to 6% of the 220 columns considered).
Since the PcrD values used to determine the column slenderness were obtained by means of
rather involved numerical calculations, resorting to ANSYS CSFE or SFE models, an alternative
(simpler) approach was sought. After some failed attempts, it was decided to propose an approach
that combines the use of Eq. (2) with PcrD values calculated for “ideal pin-ended columns” with
lengths equal to the distances between bolt centre lines – distance LD in Figure 1. Note that the
new PcrD values, not shown here (due to space limitations), are lower than the “exact” ones by
about 20% (in average), which is most likely due the absence of the warping restraint/continuity
provided by the bolts and the column lengths extending beyond the bolt centre lines. The ensuing
(higher) slenderness values are denoted λD.SS, while the nominal failure loads retain the designa-
tion PnD.SS. Finally, Figure 6(a4), which differs from Figure 6(a2) only in the fact that the circles
and crosses are shifted to the right, compares the PnD.SS design curve (expressed as a function of
λD.SS) with the experimental and numerical Pu/Py values gathered in this work – the corresponding
Pu/PnD.SS values are also plotted against λD.SS in Figure 6(b4). It is observed that this DSM-based
design proposal (i) clearly outperforms the original PnD.SS design curve, proposed by Landesmann
& Camotim (2013) and expressed in terms of λD (see Figures 6(a2) and 6(b2)), and (ii) performs
similarly to the PnD.EB design curve (also expressed in terms of λD) – this assertion can be attested
through the comparison between (i) Figures 6(a4) + 6(b4) and 6(a3) + 6(b3), (ii) the associated stat-
istical indicators, which are practically identical (both quite good), and (i) the fairly small number
of unsafe failure load predictions (now only 14 out of 220). The key advantage of this last design
approach, which is highly relevant for practitioners, is the fact that the PcrD values can be straight-
forwardly calculated by means of commonly available standard software tools – the use of sophis-
ticated CSFE or SFE models is not necessary.

4 CONCLUSION

This paper reported the results of an ongoing experimental and numerical investigation on the
buckling behaviour, ultimate strength and DSM design of end-bolted cold-formed steel columns
failing in distortional modes, which was initiated by Santos (2017). The numerical failure load

961
data obtained in this work concern 15 lipped channel end-bolted column geometries and yield
stresses that were carefully selected (i) to ensure, as much as possible, “pure” distortional buckling
and failure modes, and also (ii) to cover a wide (distortional) slenderness range. Since it was found
that almost all the non-slender (low or moderate slenderness) columns analysed failed prematurely
by localised yielding in the flange regions located below the flat washers, it was decided to consider
analyse identical columns with wider washers – this made it possible to obtain a reasonable
amount of non-slender column distortional failure loads. The above numerical distortional failure
loads, together with the experimental ones reported by Santos (2017), were then used to assess the
merits of the available DSM column distortional design curves in predicting them. It was shown
that practically all these failure loads are overestimated by both (i) the currently codified design
curve (AISI 2016) and (ii) the strength curve proposed by Landesmann & Camotim (2013), in the
context of standard simply supported columns – the overestimations are larger in the former case.
This finding led to the search for new DSM-based design approaches to provide better quality
failure load predictions for the end-bolted columns dealt with in this work. Two preliminary (fur-
ther validation is necessary) proposals were put forward and shown to yield quite reasonable and
very similar failure load predictions: (i) a modification of the design curve proposed by Landes-
mann & Camotim (2013), which has the non-negligible drawback of requiring a rigorous critical
distortional buckling load calculation, by means of sophisticated CSFE or SFE models, and (ii)
the use of this same design curve, but expressed in terms of a slenderness based on critical distor-
tional buckling loads of “ideal pin-ended columns” with lengths equal to the distance between the
bolt centre lines, which can be calculated straightforwardly. One last word to mention that further
work on this topic is currently under way and should be reported by the authors in the not too
distant future.

ACKNOWLEDGMENTS

The first and second authors gratefully acknowledge the financial support of the following
Brazilian institutions: CAPES (Finance Code 001), CNPq (303860/2016-2) and FAPERJ
(E-26/202.758/2017).

REFERENCES

AISI (American Iron and Steel Institute) (2016). North American Specification (NAS) for the Design of
Cold-Formed Steel Structural Members (AISI-S100-16), Washington DC.
ASTM A490-14a (2014), Standard Specification for Structural Bolts, Alloy Steel, Heat Treated, 150 ksi
Minimum Tensile Strength, ASTM International, West Conshohocken.
ASTM F436/F0436-16 (2016), Standard Specification for Hardened Steel Washers Inch and Metric Dimen-
sions, ASTM International, West Conshohocken.
Camotim, D., Dinis, P.B., Martin, A.D. (2016). Direct Strength Method (DSM) – a general approach for
the design of cold-formed steel structures, Recent Trends in Cold-Formed Steel Construction, C. Yu
(ed.), Woodhead Publishing (Series in Civil and Structural Engineering), Amsterdam, 69–105.
Landesmann, A., Camotim, D. (2013). On the Direct Strength (DSM) design of cold-formed steel col-
umns against distortional failure, Thin-Walled Structures, 67(January), 168–187.
Santos, W.S. (2017). On the Strength and DSM Design of End-Bolted Cold-Formed Steel Columns Buck-
ling in Distortional Modes, Ph.D. Thesis in Structural Engineering, COPPE, Federal University of Rio
de Janeiro, Brazil. (Portuguese)
Schafer, B.W. (2008). Review: the Direct Strength Method of cold-formed steel member design, Journal
of Constructional Steel Research, 64(7-8), 766–788.
Silvestre, N., Camotim, D. (2016). Local-plate and distortional post-buckling behavior of cold-formed
steel lipped channel columns with intermediate stiffeners. Journal of Structural Engineering (ASCE),
132(4), 529–540.
Stolarski, T., Nakasone, Y., Yoshimoto, Y. (2018). Engineering Analysis with ANSYS Software (2nd edition),
Butterworth-Einemann (Elsevier), Oxford.
Swanson Analysis Systems Inc. (SAS) (2009). Ansys Reference Manual (vrs. 12).
Young, W.C., Budynas, R.G. (2002). Roark’s Formulas for Stress and Strain (7th edition), McGraw-Hill,
New York.

962
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Design limitations for the steel beam-column to ensure full plastic


moment

A. Sato, M. Aoyama, K. Inden & T. Ono


Nagoya Institute of Technology, Nagoya, Japan

K. Mitsui
Nippon Steel and Sumitomo Metal Co., Japan

ABSTRACT: Under the seismic load action, the horizontal force will be significant, and the
column in the moment resisting frame (MRF) shall resist the axial force and the shear force (i.e.
This load condition is so called beam-column). In the Ultimate Limit State (ULS) design, the
inelastic characteristics are used; strong-column weak-beam philosophy is used to prevent the col-
lapse of the structure. The column overstrength factor (COF) is checked between the capacity of
the connected beams and columns; the full plastic moment considering the axial force is used.
Therefore, it is crucial to guaranty that the columns can resist the full plastic moment. The struc-
tural performance of the beam-column will be affected by the axial force level, and second-order
effects might be dominant in a relatively slender column. In this paper, the second-order effect
introduced by compressive axial force to the beam-column is studied by numerical simulation.

1 INTRODUCTION

Columns in the building structure are mainly supporting the gravity loads. Once horizontal forces
subjects to the moment resisting frame (MRF) structure, these external forces will be resisted by
the columns in a flexural manner. As a result, columns are in a combined loading condition (i.e.
This load condition is so-called beam-column). Under seismic action, horizontal forces will be sig-
nificant and combined loading subjected to the column can trigger the instability. In the Ultimate
Limit State (ULS) design, the structure is allowed to accommodate plastic deformation (ASCE
2016, EC8 2004, BSL 2016). The location of dissipative zones where accommodates the plastic
deformation are expected in the ends of the beams of the MRF structure. The strong-column
weak-beam philosophy is used to guarantee not forming the plastic hinge in the columns; the soft
story collapse is prevented (AISC 2016, EC8 2004, BSL 2016). The column overstrength factor
(COF) is checked between the capacity of the connected beams and columns in the connection;
the full plastic capacity that is considering the effect of axial force is used in the calculation. The
full plastic capacity of the beam-column can be computed from the section property when the
member is stocky. However, the capacity of the member can be dominated by the second-order
effect when the member gets relatively slender; the capacity of the member may not reach the
performance computed from the section property.
In this paper, the second-order effect to the beam-column which is introduced by axial force is
studied. To clarify the structural behaviour of the beam-column, numerical simulation where axial
force and bending moment are subjected to the member simultaneously is conducted. Parameters
selected for the analysis are the compressive axial force, the slenderness ratio, the bending moment
distribution, and the initial imperfection. Firstly, the members that can guarantee the full plastic
moment that is considering the axial force is discussed. Secondly, the results obtained from the
analysis is compared with the current design requirements specified in the recommendation com-
piled by the Architectural Institute of Japan (AIJ 2010, 2017). Finally, the formula where the
beam-column can guarantee the full-plastic moment with the admissible deformation is proposed.

963
2 DESIGN RECOMMENDATION FOR THE STEEL BEAM-COLUMN

2.1 Maximum bending moment equal to the end moment


Under the combined loading condition, the location of the maximum bending moment will place
somewhere in the middle of the member when the second-order effect due to the axial force gets
significant. From the force equilibrium in the elastic range, the following inequality equation can
be derived where the maximum bending moment is equal to the end moment (Chen & Lui 1987).
 pffiffiffiffiffiffiffiffiffiffiffiffiffi
M2 =M1 ≥  cos π N=N0 ð1Þ

where M1 and M2 are bending moment subjected at both ends (|M1|≥| M2|), N is the compres-
sive axial force, and N0 is the Euler’s buckling strength where the flexural buckling length is
equal to column length (see Eq. 2). End moment ratio, M2/M1, will be positive under double
curvature bending. Figure 1 summarizes the notations that are used in the equations; the dir-
ection of the arrows used in the force assigns the positive direction.

π2 EI
N0 ¼ ð2Þ
lc2

where EI is the bending stiffness around the bending axis (x-axis), and lc is the member length.
Appling the Taylor’s expansion and approximation of π2/8≈1.0, Equation 1 can be
expressed as follow (AIJ 2010):
   2
N Ny  λc0  0:25ð1 þ M2 =M1 Þ ð3Þ

where Ny is the axial yield strength (=σy·Ag), σy is the yield stress, Ag is the gross area of the cross-
section, and λc0 is the non-dimensional flexural buckling slenderness ratio defined as follow:
rffiffiffiffiffi
1 σy lc
λc0 ¼ ð4Þ
π E ix
where E is the Young’s modulus, and ix is the radius of gyration around the bending axis.

Figure 1. Beam-Column (under combined loading).

964
2.2 Requirements for beam-column to form the plastic hinge
The beam-column that is expected to form plastic hinge must fulfil the following conditions.
These requirements are stipulated in the Recommendation of Limit State Design for Steel
Structure (hereinafter LSD) (AIJ 2010) for the steel hollow structural section (HSS).
   2
N Ny  λc0  0:10ð1 þ M2 =M1 Þ ð5Þ

N  0:75Ny ð6Þ

λc  2:0 ð7Þ

where λc is the non-dimensional flexural buckling slenderness ratio computed based on the
effective flexural buckling length.

2.3 Flexural buckling resistance


Flexural buckling resistance, Nb,Rd, specified in LSD is as follows:

Nb;Rd ¼ Ny λc  p λc ð8Þ
 
λc  p λc
Nb;Rd ¼ 1:0  0:5 Ny p λc 5λc  e λc ð9Þ
e λc  p λc

1
Nb;Rd ¼ Ny e λ5λc ð10Þ
1:2  λ2c

where pλc is the plastic limit of slenderness


pffiffiffiffiffiffiratio
ffi (=0.15, i.e. length of the plateau), and eλc is
the elastic limit of slenderness ratio (=1= 0:6).

3 NUMERICAL SIMULATION

3.1 Analysis model of the beam-column


Figure 2 shows the analysis model used in this study. Compressive axial force and bending
moment are subjected to the member simultaneously. In-plane behaviour is only considered in
this study; three degrees of freedom (i.e. y and z displacement, and rotation around x-axis) are
provided at each node that is used to compose the column. The open source numerical soft-
ware “OpenSees” (PEER 2006) was used to conduct the analysis. The fiber element is used to
compose the member; the length of each element is fixed to 100mm with ten integration points
between the nodes (i.e. element “nonlinearBeamColumn” is used). The material and geometric
nonlinearities are considered in the analysis.
The shape of the cross-section used in the analysis is shown in Figure 3; bending moments
are applied around the x-axis (i.e. around the strong axis). During the analysis moment-
curvature based on the defined cross-section shape is computed at each integration points and
nodes, material nonlinearity is considered at each element that is composing the cross-section.
The flanges (i.e. parallel to the bending x-axis) are divided into ten layers along with the thick-
ness; the web (i.e. perpendicular to the bending x-axis) is also divided into ten elements
through the depth (De Souza 1999). The mild steel property obtained from the coupon test,
where the yield stress is σy = 300N/mm2, is used for the elements (Ono & Sato 2003).

3.2 Compressive axial force level and member length


The length of the member and the corresponding compressive axial force subjected to the
member is determined according to the formulae shown in chapter 2. The maximum bending
moment at the beam-column should be the end moment when it is expecting to form the

965
Figure 2. Loading and boundary conditions.

Figure 3. Wide flange section (W-flange).

plastic hinge; therefore, at least Equation 3 which is derived from elastic theory should be sat-
isfied. Moreover, LSD specifies Equation 5 for the beam-column that is expected to form the
plastic hinge. Specimens used for the analysis were determined from these equations. Figure 4
shows the sample relationships between specimens and design formulae. In the figures, flex-
ural buckling resistance which is defined by Equations 8-10 are also shown, and the specimens
that will overcome this resistance were not selected.
The compressive axial force was determined by the ratio between the axial yield strength
Ny. The axial force ratio (i.e. N/Ny) was ranged from 0.1 to 0.75 by 0.05 pitch. The maximum
ratio 0.75 followed the limitation shown in Equation 6. As can be recognized from the figures,
moment distribution (i.e. M2/M1) in the member will affect the limitation of column length,
a slender member is allowed in the same axial force level when the value of M2/M1 is larger
(i.e. when the moment distribution gets close to the antisymmetric bending moment).
Moment distributions used in the analysis are M2/M1 = 0.0, 0.25, 0.5, 0.75, and 1.0. M2/M1 = 0.0
is one end bending moment, and M2/M1 = 1.0 is antisymmetric bending moment conditions,
respectively.

3.3 Initial imperfection


Initial local bow imperfection that will exist in the actual structural member is modelled in the
analysis. In this study, a sinusoidal geometrical imperfection with a maximum value of u0,d at

966
Figure 4. Compressive axial force level vs. non-dimensional flexural buckling slenderness ratio.

the midspan in assumed (see Eq. 11). The amount of u0,d is assumed to be equal to lc/1000. See
Figure 2 for the coordinates.
 
π
u0 ðzÞ ¼ u0;d sin z ð11Þ
lc

4 RESULTS FROM THE NUMERICAL SIMULATION

4.1 End moment vs. rotation


Sample results computed from the analysis are shown in the Figure 5. The vertical axis M1 is the
moment applied at the member end; the horizontal axis is the end rotation measured at the load-
ing point of M1. Bending moment is normalized by full plastic moment Mpc which is considering
compressive axial force, and rotation is normalized by elastic rotation ϴpc corresponding to Mpc
force level. The parameters corresponding to the specimens are shown in Figure 4 with the gray
solid circular legend. Three results are show in each figure; one specimen is close to the border of
Equation 3, one is close to the border of Equation 5, and one is in between these two.
From one end bending moment (M2/M1= 0) results, as shown in Figure 5a, the member which
is satisfying the requirement Equation 5 could develop the full plastic moment Mpc at the max-
imum moment level. On the other hand, the column that is close to the border of Equation 3
could not develop the full plastic moment Mpc. On the contrary, the results obtained from the
antisymmetric bending moment (M2/M1= 1.0, see Fig. 5b) showed that the maximum bending
developed at the member end reached the full plastic moment Mpc even if the specimen is close to
the border of Equation 3.

4.2 Full plastic moment considering the compressive axial force


As mentioned in chapter 1, strong-column weak-beam philosophy is used in the structural
design to avoid story collapse. Following inequality should be fulfilled (AISC 2016, EC8 2004,
BSL 2016).
P P
MRc ≥ γCOF  MRb ð12Þ

where ΣMRc (=ΣMpc) is the sum of the design resistance of the columns framing the joint,
ΣMRb (=ΣMb) is the sum of the design resistance of the beams framing the joint, and the γCOF
is the column overstrength factor. γCOF is depending of the Standard or Specification and 1.3
(EC8 2004) or 1.5 (BSL 2016) is used.

967
Figure 5. End moment-rotation curves.

To check the inequality of Equation 12, the full-plastic moment of the column cannot be the
capacity corresponding to the maximum bending moment that can be resisted. As can be recog-
nized from Figure 5, is some members, the end bending moment can develop the full-plastic
moment at the maximum level. Moreover, some of the columns that can develop the full-plastic
moment reached the maximum value after significant plastic deformation (Sato & Ono 2017).
From the point of view to avoid significant inelastic deformation in the column (i.e. Eq. 12); the
in-elastic deformation in the column should be limited to a minor level (i.e. admissible deform-
ation). To form a plastic hinge in the column, Equation 5 is specified in the recommendation of
AIJ. However, this is the formula derived from the test results where the maximum bending
moment reached the full-plastic moment (Ono et al. 1989). Therefore, it is crucial to evaluate
the members that can form the full-plastic moment in the minor inelastic deformed level. The
full-plastic moment Mpc (or MRc) is based on the fictitious definition, and it is not available to
obtain this value from the results of the testing or the numerical simulation. In this study, the
full-plastic moment obtained from the analysis was defined when the value of tangent stiffness
of the moment-rotation relationships reached 5% of the elastic stiffness.
Figure 6 shows the results of the longest members at each axial force level that could
develop the full-plastic moment Mpc in the minor inelastic deformed level. It can be recognized
that the members under the moment distribution closer to the antisymmetric bending can
reach the full-plastic moment where the members are closer to the boundary of Equation 3.
Moreover, the members under less axial force are closer to the boundary of Equation 3.

Figure 6. Beam-column that can ensure the full-plastic moment with admissible deformation.

968
Table 1. Coefficients α and DM*.
M2/M1
Coef. 0.0 0.25 0.5 0.75 1.0

α 1.60 1.67 1.63 1.61 1.53


DM* 0.145 0.152 0.167 0.180 0.202

The results that can develop the full-plastic moment at the member end were analyzed by
regression analysis. The following formula was used in the regression analysis, and the result
is shown in each figure with the broken line.
   α
N Ny  λc0 ¼ DM ð1 þ M2 =M1 Þ ð13Þ

where α is the coefficient to evaluate the effect of slenderness ratio, DM* is the coefficient corres-
ponding to the moment distribution. The coefficients are tabulated in Table 1.

5 EVALUATION OF THE DESIGN FORMULA

As shown in Figure 6 and Table 1, a clear trend can be found; therefore, it seems to be pos-
sible to propose the evaluation formula for the beam-column that can guarantee the full-
plastic moment. The coefficient α ranged between 1.53 and 1.67. For simplicity, it can be
assumed that the coefficient of α is equal to the average value (i.e. α =1.61). The relationships
between the coefficient DM* and moment distribution (M2/M1) is shown in Figure 7. The
value of DM* has a positive relationship between the value of M2/M1. The dotted line shown
in the figure can be express by Equation 14.

0:516
DM ¼ ð14Þ
3:56  ðM2 =M1 Þ1:20

Therefore, Equation 13 can be expressed in the following formula.

Figure 7. Coefficient DM* vs. M2/M1.

969
Figure 8. Simplification of the formula (Eq. 16).

   1:61 ð1 þ M2 =M1 Þ
N Ny  λc0 ¼ 0:516  ð15Þ
3:56  ðM2 =M1 Þ1:20

The regression analysis can simplify the right-hand side of the formula; the result can be
express as follow.
(  1:30 )
 ð1 þ M2 =M1 Þ M2
dM ¼ 0:516  ) 0:145  1 þ 1:75 ð16Þ
3:56  ðM2 =M1 Þ1:20 M1

Figure 8 shows the result of the regression analysis of the right-hand side of the formula. Rea-
sonable accuracy can be found.
Finally, Equation 17 can be proposed to ensure the full plastic moment of the beam-column
in a minor inelastic deformed level (i.e. with admissible deformation).
(  1:3 )
   1:6 M2
N Ny  λc0  0:145  1 þ 1:75 ð17Þ
M1

6 CONCLUSION

Numerical simulation of the members that are under compressive axial force with bending
moment (i.e. beam-column) were conducted in this study. From the in-plane analysis, second-
order effect due to compressive axial force was studied; the limitation of the combined loading
condition that can ensure the full-plastic moment in a minor in-elastic deformed level (with
admissible deformation) was discussed. Following results were found in this paper:
1. The moment distribution in the beam-column will influence the member length that can
ensure full-plastic moment under combined loading condition. The moment distribution
closer to the antisymmetric bending moment can have a slender member on the same com-
pressive axial force level.
2. The member that can develop full-plastic moment at the minor in-elastic deformed level
(with admissible deformation) will have the length limit closer to the boundary of first
yielding at the member end (i.e. Eq. 3) when the axial force level is low.
3. The design limitation of the beam-column to ensure the full-plastic moment at the minor
in-elastic deformed level that can be used for the strong-column weak-beam check was pro-
posed from the regression analysis of the numerical results.

970
Regarding the research about beam-column, extensive projects are conducted in several
research groups analytically and experimentally. Lignos (Elkady & Lignos 2014, 2018, Suzuki
& Lignos 2015, 2017) conducted both steel wide flange and box columns full-scale testing;
evaluation of the structural performance of the beam-column based on testing and numerical
simulation are reported. Uang (Newell & Uang 2008) conducted steel wide-flange columns
full-scale testing, and the structural performance under cyclic loading is reported. Simões da
Silva (Tankova et al. 2017) studied the design formula of beam-column based on the theoret-
ical derivation, and the validity of the proposed formula is compared with the advanced
numerical simulation results (GMNIA). In the next step, the derived design formula in this
paper will be compared with the existing proposed results.

REFERENCES

American Institute of Steel Construction. 2016. Seismic Provisions for Structural Steel Buildings, ANSI/
AISC 341-16.
American Society of Civil Engineers. 2016. Minimum Design Loads and Associated Criteria for Buildings
and Other Structures, ASCE/SEI 7-16.
Architectural Institute of Japan. 2010. Recommendation for Limit State Design of Steel Structures, AIJ.
Architectural Institute of Japan. 2017. Recommendation for the Plastic Design of Steel Structures, AIJ.
Building Standard of Law of Japan. 2016. BSL, Ministry of Land, Infrastructure and Tourism.
Chen, W. F. & Lui, E. M. 1987. Structural Stability Theory and Implementation, Prentice Hall.
De Souza, R. M. 1999. OpenSees Command Language Manual, Patch Command.
Elkady, A. & Lignos, D. G. 2015. Analytical investigation of the cyclic behavior and plastic hinge forma-
tion in deep wide-flange steel beam-columns, Bulletin of Earthquake Engineering 13: 1097–1118.
Elkady, A. & Lignos, D. G. 2018. Full-scale testing of deep wide-flange steel columns under multiaxis
cyclic loading: loading sequence, boundary effects, and lateral stability bracing force demands, Journal
of Structural Engineering 144 (2):0401789-1-15.
European Committee for Standardization. 2004. Design of Structures for earthquake resistance – Part 1:
General rules, seismic actions and rules for buildings, EUROCODE 8, EN 1998-1, CEN.
Ishida, T., Ono, T. & Nomoto S. 1989. Study on Design Criteria of Steel Beam-Columns Based on
Experimental Data, Summaries of Technical Papers of Annual Meeting, Architectural Institute of
Japan (AIJ), C, Structures II, 1287–1288.
Newell, J. D. & Uang, C. M. 2008. Cyclic behaviour of steel wide-flange columns subjected to large drift,
Journal of Structural Engineering 134(8): 1334–1342.
Ono, T. & Sato, A. 2003. The influence of the notch, temperature and strain rate on the strength of metal-
lic materials, The 4th International conference on behaviour of steel structure in seismic areas, STESSA
2003: 99-105.
Pacific Earthquake Engineering Research Center. 2006. The Open System for Earthquake Engineering
Simulation (OpenSees), UC Berkeley.
Sato, A. & Tetsuro, Ono. 2017. Deformation Capacity of Steel Column under combined loading- Compres-
sive axial force with double curvature bending moment, EUROSTEEL 2017, Copenhagen 13-15 September,
Denmark.
Suzuki, Y. & Lignos, D. G. 2015. Large scale collapse experiments of wide flange steel beam-columns,
The 8th international conference on behaviour of steel structures in seismic areas, STESSA 2015: No. 57.
Suzuki, Y. & Lignos, D. G. 2017. Collapse behaviour of steel columns as part of steel frame buildings:
experiments and numerical models, The 16th world conference of Earthquake, 16WCEE 2017: No. 1032.
Tankova, T., Simões L. da Silva, & Marques, L. 2017, Ayrton-Perry formulation for the buckling resist-
ance of prismatic beam-columns, XI Conference on Steel and Composite Construction 1(4): 415–425.
Tankova, T., Marques, L., Simões L. da Silva, & Andrade, A. 2017. Development of a consistent meth-
odology for the out-of-plane buckling resistance of prismatic beam-columns, Journal of Constructional
Steel Research 128: 839–852.

971
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Experimental study on LTB behaviour and residual stresses of


welded I-section members

L. Schaper & R. Winkler


Chair of Steel, Lightweight and Composite Structures, Ruhr-Universität Bochum, Germany

F. Jörg
Institute of Structural Design, University of Stuttgart, Germany

U. Kuhlmann
Steel, Timber and Composite Construction, Institute of Structural Design, University of Stuttgart, Germany

M. Knobloch
Steel, Lightweight and Composite Structures, Ruhr-Universität Bochum, Germany

ABSTRACT: This paper presents an experimental study on the lateral torsional buckling
behaviour of members with doubly and mono-symmetric welded I-sections. The experimental
study comprised not only member buckling tests but also residual stress measurements. The
results of the structured member tests are used to assess simplified European design methods.
A comparative study confirms that these methods lead to conservative results and should be
amended to consider the effects of residual stresses more thoroughly.

1 INTRODUCTION

Residual stresses caused by welding have a marked effect on the structural stability behaviour,
in particular in case of lateral torsional buckling. Common models for residual stresses of
welded steel girders are based on old tests series and do not comply with current manufactur-
ing conditions. For both welded cross-sections with thick plates and mono-symmetric cross-
sections residual stress measurements and LTB test results are rare.
This paper presents a comprehensive experimental study on the lateral torsional buckling
behaviour of welded steel members. The study comprised residual stress measurements. The
tests results are used to perform a comparative study and to assess the simplified European
method.

2 EXPERIMENTAL INVESTIGATIONS

2.1 Background motivation


As a part of the AiF-/DASt-research project ‘Simplified method for lateral torsional buckling –
consistent model for welded beams at ambient and elevated temperatures’ experimental studies
have been carried out on welded steel girders at the Chair of Steel, Lightweight and Composite
Structures of Ruhr-Universität Bochum and the Institute of Structural Design of the University
of Stuttgart. The aim was to investigate the influence of residual stresses on the structural
member stability behaviour, especially the lateral torsional buckling behaviour of welded cross-
sections with thick flanges and mono-symmetric I-sections. The project comprised both residual
stress measurements and lateral-torsional buckling tests.

972
2.2 Test specimens
For the experimental investigations test specimens with doubly- and mono-symmetric
welded I-sections as well as I-sections with an additional lamella welded to one of the
flanges were used. The program facilitated analysing the effect of different parameters
on the stability behaviour and the residual stress distribution of the specimens. The pro-
gram comprised specimens of steel grades S355J2 and S460M. To characterize the actual
material behaviour, tensile material tests of the flange and web plates were carried out.
The web thickness of all test specimens was 8.0 mm. The thickness of the flanges varied
between 12 mm and 40 mm. The flanges were welded to the web with double fillet welds
of 5 mm. The dimensions of the test girders as well as investigated steel grades are sum-
marized in the testing program in Table 1. The test specimens of both the lateral-
torsional buckling tests and the residual stress measurements were prepared from the
same profile and separated by saw cutting.
The length of the test specimens of the residual stress measurements guaranteed that
the measuring points for the residual stress measurements were approx. 2.5 times the
flange width or 2.5 times the web height from the end of the girder. For the lateral tor-
sional buckling tests the specimen lengths corresponded to a relative slenderness ratio of
0.8 to 1.6.
Two welding procedures were used, submerge arc welding for specimens of steel grade S355
and MAG-welding for steel grade S460. The welding parameters of both welding procedures,
i.e. energy and speed were similar. To avoid influences of post treatment flame straightening
and surface sealing were not applied.
The specimens were produced by two different manufacturers, hereafter referred as ‘Manu-
facturer 1ʹ and ‘Manufacturer 2ʹ. In order to ensure comparability between the manufacturers,
steel specimens with the same dimensions, steel grade and welding procedures were also tested
(specimen no. 1 & 1*).

Table 1. Testing program.


Length [mm] Dimensions [mm]

Identification Steel grade RSMa LTBb hw bf tf

1 S355 J2 1800 10200 370 350 20


1* S355 J2 1800 - 370 350 20
1Bc S355 J2 1800 - 370 350 20
1fy S460 M 1800 10200 370 350 20
1Bfy S460 M 1800 - 370 350 20
2-2 S355 J2 4100 - 800 350 20
3-1 S355 J2 1800 6200 370 200 20
3-1fy S460 M 1800 6200 370 200 20
3-2 S355 J2 1800 10200 370 350 20
4-1 S355 J2 1800 10200 370 350 12
4-2 S355 J2 1800 - 370 350 40
5 S355 J2 1800 10200 370 200d, 350e 20d, 40e
5fy S460 M 1800 - 370 200d, 350e 20d, 40e
13 S355 J2 1800 - 370 350 25
14 S355 J2 1800 10200 370 200 40
a
Residual stress measurements at the Ruhr-Universität Bochum
b
Lateral torsional buckling tests at the University of Stuttgart
c
Test girder with an additional lamella
d
Dimensions of the compression flange
e
Dimensions of the tension flange

973
2.3 Residual stress measurements

2.3.1 Test setup


The sectioning method was used for the measurement of the residual stresses. Thereby, the
residual stresses in longitudinal direction were determined by sectioning the girder into small
parts. The residual stresses were determined using the released constraints and the material
properties from the tensile tests. For the measurement of the released constraints strain gauges
were used, the sectioning was carried out with saw cuts and with water jet cuts in the area of
the strain gauges.
Depending on the cross-section size, 33 to 82 strain gauges were applied on each specimen.
For mono-symmetric I-sections and cross-sections with a lamella, the girder was measured cir-
cumferentially. For double-symmetric cross-sections the top flange was measured and the
affiliated half of the web taking advantage of the symmetry. In this case, one half of the flange
was measured in a narrow grid and for the other half of the flange just significant points were
measured for control purposes. The strain gauges were varnished and sealed after application
for protection during water jet cutting (Figure 1).
Continuous measurements were carried out throughout the entire experiment. Through
this, numerous information were collected and the method verified, such as the influence of
the cutting sequence, influence of the distance of the measuring point to the end of the speci-
men and the development of residual stresses during the individual cuts. The final measure-
ment of residual stresses of the welded cross-section was carried out after the material had
settled with same temperature as at the beginning of the measurement (see Figure 2).

2.3.2 Test results


The residual stresses was computed with the results of the tensile tests. For small strains (≤ 0,15‰)
the residual stresses were determined as the product of the strain measurement on the Young’s

Figure 1. Applied strain gauges at the top flange (left). Test specimen with sealed strain gauges (right).

Figure 2. Specimen after sectioning (left). Strain gauges after sectioning of the left part of the top flange
(right).

974
Figure 3. Measured residual stresses. Influence of parameters: a) welding procedure and manufacturer;
b) steel grade and flange width; c) flange thickness.

modulus. For larger strains, the residual stresses were determined with the actual stress-strain-
relationship of the material. Figure 3 exemplarily shows the computed residual stresses of different
specimens. Markings indicate the mean values of the measured residual stresses of the respective
plate side. The distributions of the residual stresses were supplemented by symmetry and equilib-
rium conditions of the cross-sections. In addition, Figure 3 shows the residual stresses for welded
girders according to ECCS Publication No. 33 (gray).
The measurements in Figure 3 a) show that the welding procedure (submerge arc welding and
MAG-welding) and the manufacturer had a minor influence on the residual stress distribution
and the maximum stress value. The welding procedure caused tensile residual stresses at the mid-
part of the flange and top and bottom end of the web. The flame cutting caused residual tensile
stresses at the flange tips. Almost constant compressive residual stress distribution occurred at the
mid-parts of the web and the flange parts. Strain measurements in the area of the welds could only
be taken at the outside of the flange. The residual stresses at the inside of the flange, at the position
of the weld, were calculated by considering equilibrium conditions of the cross-section. The values
of the tensile stresses at the flange tips scatter. Due to the high slope of the distribution of the
residual stresses the position of the strain gauge significantly influences the results of the stress
values.
The influence of the steel grade and the flange width on the residual stresses is shown in
Figure 3 b). The residual stresses of grade S355 sections (triangle markings) and grade S460 sec-
tions (circular markings) with a flange width of 200 mm (dotted line) and 350 mm (dashed line)
are compared. The values and distributions of the residual stresses were similar for both steel
grades for the respective cross-sections types. The steel grade had no marked influence on the
residual stresses. In comparison of the cross-sections types it can be seen, that the compressive
stresses are slightly higher for the cross-sections with the smaller flange width. Their effect was

975
caused by the equilibrium of the cross-section. Due to the smaller area of the compressive stres-
ses the values increased to fulfill the equilibrium with the approximately consistent tensile stres-
ses at the welds and flange tips. By this a significant influence of the cross-section geometry on
the size of the residual stresses was given.
The influence of the flange thickness is presented in Figure 3 c). The results of the residual
stresses of cross-sections with varying flange thicknesses (12 mm, 20 mm and 40 mm) with
constant flange width, web height and thickness and steel grade are compared.
The ratio of flange-area-to-web-area significantly influenced both the size and the distribu-
tion of the residual stresses. This effect is presented for specimen Pos. 4.1 with a small flange
area, which leads to high tensile stresses at the flange and for specimen Pos. 4-2 with a large
flange area, which led to high tensile stresses at the top of the web. The stiffness of the affili-
ated plates caused this distribution of the residual stresses.

2.4 Lateral torsional buckling tests

2.4.1 Test setup


The test setup for the lateral torsional buckling tests corresponded to a three-point-bending
test. The load application occurred at mid span through a loading frame. In order to avoid
stabilizing effects due to a load introduction below the shear centre, the test girders were
loaded from bottom to top. Accordingly, the compression flange of the test girders was
always at the bottom.
Due to the lateral torsional buckling failure of the test girders a fork restraint had to be
arranged. This means that the displacements v and w of the cross-section axes as well as the
rotation ϑ about the longitudinal axis had to be prevented. At the same time the displacement
u in longitudinal direction and the rotations v’ and w’ of the cross-section axes should be pos-
sible. To achieve these fork restraints for the supports, a frame at each end was used, where the
flanges of the test girders were supported by threaded rods, see Figure 4. A frictionless rotation
was obtained through an additional teflon coating on the threaded rods.
The bending moment distribution of the specimen was achieved through a concentrated
loading at mid span. The loading construction contained a loading frame of HEM profiles,
which was connected to stiffeners of the test girders. In order to avoid significant influence on
the course of the residual stresses in the area of the flanges, the stiffeners were welded exclu-
sively to the webs. The load was applied with a hydraulic load jack, which was constructed as
a pendulum rod in order to avoid a tied axis of rotation.
To ensure a constant load direction the rotation of the hydraulic cylinder was observed
through an inclination sensor. The cylinder was placed in a kind of sledge and was moved
horizontally during loading on a roller rail system. Since the cylinder could only be moved in
one direction over the roller rail system, it had to be connected slightly eccentrically through
the loading frame. Due to the additional torsion of the beam, the direction of the stability
failure could be determined. The whole load introduction construction at mid span can be
seen in Figure 5.
The loading was applied at the beginning force controlled till approx. 60% of the previously
calculated numerical loading capacity. Then the load was applied path-controlled till reaching
the ultimate loading capacity of the test girder. Several different measuring devices were used
to observe the behaviour of the test specimens. The horizontal and vertical displacements were
recorded using several displacement sensors at the supports and through integrated sensors in
the hydraulic cylinder. Beside the vertical and horizontal displacements, the rotation ϑ of the
test girder about the longitudinal direction was measured through different inclination sensors
at mid span. Additional inclination sensors at both fork supports measured the rotation about
the strong axis.
Apart from the mechanical measuring devices, the three-dimensional deformations as
a result of the lateral torsional buckling failure were measured at mid span by photo-
grammetric means. With the photogrammetric method, photos of the measured panel,
which was marked with coded targets, could be taken with four special cameras. The

976
Figure 4. Fork restraint for the test specimens at the support.

deformations of each measuring point were generated afterwards by comparing their


coordinates in two consecutive photos.

2.4.2 Test results


Figure 6 shows the comparison of the load-displacement curves for different test girders.
Due to the different geometries of the regarded test specimens, they showed partly
a different lateral torsional buckling behaviour. Since the lateral torsional buckling fail-
ure is characterized by horizontal displacements v and rotations ϑ, the normalized cylin-
der force is shown in the diagram over the horizontal displacements of the hydraulic
cylinder at mid span.
Due to the slightly eccentric loading, the test specimens were subjected to an additional
small torsional moment right from the start. For that reason, small horizontal displacements
could be measured immediately after loading. From approx. 60 % of the ultimate load cap-
acity the measured horizontal displacements increased significantly, which can be traced back
to the lateral torsional buckling failure of the test specimens.
For the compact test specimen 14 with thick flanges (tf = 40 mm), the lowest horizon-
tal displacements were measured at the beginning compared to the other three test gir-
ders. Due to the high torsional stiffness of this specimen initially smaller rotations
occurred and therefore smaller horizontal displacements could be measured. However,
the displacements by reaching the ultimate load capacity were almost identical to those
of the test girders 1 and 1fy.
The test specimens 1 and 1fy show that the lateral torsional buckling behaviour depends
mainly on the cross-section geometry. Although the higher steel grade led to a higher load
bearing capacity, the load-displacement behaviour was not influenced by the different steel
grade.

977
Figure 5. Load introduction at mid span with hydraulic cylinder on a roller rail system.

Figure 6. Comparison of load-displacement curves for test girders 1, 1fy, 5 and 14 showing normalized
cylinder force over horizontal displacement at mid span.

Despite the larger torsional stiffness of the mono-symmetric test girder 5 in comparison to
the reference girder 1, the measured horizontal displacements for this test specimen were
larger. Obviously the stiffness of the compression flange, which was significantly smaller for

978
Table 2. Maximum loads of the tested girders.
Test 1 1fy 3-1 3-1fy 3-2 4-1 5 14

Steel grade S355 S460 S355 S460 S355 S355 S355 S355
Length [mm] 10200 10200 6200 6200 10200 10200 10200 10200
Flange width [mm] 350 350 200 200 350 350 200a, 350b 200
Flange thickness [mm] 20 20 20 20 20 12 20a, 40b 40
Fult [kN] 194.8 203.9 143.3 191.3 177.8 75.6 179.6 243.6
a
Dimensions of the compression flange.
b
Dimensions of the tension flange.

the mono-symmetric girder compared to girder 1, has the greatest influence on the horizontal
displacements.
Table 2 shows the load capacity of the tested girders. The maximum load of the
regarded test girders can be found for the test girder 14 with thick flanges due to the
huge torsional stiffness and therefore lower slenderness compared to the other test speci-
mens. Test girder 4-1 with a flange thickness of only 12 mm had the lowest loading cap-
acity due to the very small torsional stiffness. The influence of the different steel grades
for test specimens 1 and 1fy is not as pronounced as for 3 and 3-1fy even though the
slenderness was almost identical.

3 COMPARATIVE STUDY

In order to compare the experimental results with the current European design rules for the
lateral torsional buckling in EN 1993-1-1, the characteristic design resistance was calculated
with two different approaches. First the capacity was calculated with the lateral torsional
buckling verification for the ‘general case’, para. 6.3.2.2, and secondly with the ‘simplified
method’ as a verification of flexural buckling of the compression flange, para. 6.3.2.4. Results
of the lateral torsional buckling tests in comparison to the verifications are shown in Figure 7.
The experimental results fit well to the buckling curves of the ‘general case’. It can be seen
that the results of the experimental investigations are close to the associated buckling curve of
the analytical verification. It has to be considered that a small initial torsional moment, which
was caused by the slightly eccentric load application at mid span, is not taken into account.

Figure 7. Comparison of test results with LTB design methods. Left: ‘General case’ (EN 1993-1-1,
6.3.2.2). Right: ‘simplified method’ (EN 1993-1-1, 6.3.2.4).

979
In comparison to the ‘simplified method’, the experimental results scatter. There are devi-
ations to the non-conservative side which may be caused by the slightly eccentric load applica-
tion and by the fact that the load application point is not taken into account for this method.
Clarification and possibly improvements are needed.
The method led to conservative results for cross-sections with thick flanges. The method
does not consider the effect of the torsional stiffness of these sections on the buckling resist-
ance. Furthermore, the method applies buckling curve d for all welded cross-sections. Using
buckling curve c for these cross-sections instead leads to slightly increased load capacity but
cannot compensate the large difference and does not lead to an adequate design. An improve-
ment of the method, however, would facilitate an efficient design of welded bridge and
runway beam sections with thick flanges.

4 CONCLUSIONS

A comprehensive experimental study on the lateral torsional buckling behaviour of welded I-sec-
tions members have been conducted. The study comprised 15 different welded sections of steel
grades S355 J2 and S460 M. The study has addressed the influence of the dimensions of the sec-
tions on the course and the size of the residual stresses caused through welding. Torsional buck-
ling tests have been carried out to analyse the structural stability behaviour of the same girders.
The investigations of the residual stresses have shown, that the cross-sections geometry has
a significant influence on the distribution and size of the residual stresses. The residual tensile
stresses are markedly affected by the flange-to-web area of the sections.
The ultimate loads for lateral torsional buckling have been determined and compared to the
results of European design methods, i.e. the ‘general case’ (para. 6.3.2.2) and the simplified
method (para. 6.3.2.4). For welded girders with thick flanges, the simplified method leads to
conservative design results in comparison to the experimental results. Based on the results of
this study the simplified method could be amended to consider the effects of residual stresses
more thoroughly for welded cross-sections with thick flanges.

ACKNOWLEDGMENTS

Special thanks are given to the German Federation of Industrial Research Associations “Otto
von Guericke” e.V. (AiF) and the German Commission for Steel Construction (DASt) for the
financial support and to all the industry partners Astron and Goldbeck GmbH for the donation
and fabrication of the test specimens. Also, contribution and support by the Institute for Photo-
grammetry in Stuttgart for the photogrammetric measurements and by the material testing insti-
tute at the Universities of Bochum and Stuttgart (MPA) for carrying out the tests are greatly
appreciated.

REFERENCES

EN 10025-1: 2005, Hot rolled products of structural steels – Part 1: General technical delivery conditions.
EN 1993-1-1: 2005, Eurocode 3: Design of steel structures – Part 1-1: General rules and rules for
buildings.
EN 1993-1-2: 2005, Eurocode 3: Design of steel structures – Part 1-2: General rules – Structural fire
design.
EN ISO 6892-1: 2009, Metallic materials – Tensile testing – Part 1: Method of test at room temperature.
ECCS: 1984. Ultimate Limit State Calculations of Sway Frames with Rigid Joints. European Convention
for Constructional Steelwork – TC 8, ECCS-Publication No. 33, Brussels, Belgium.

980
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Behavior of column base plates under bi-axial bending moment

L. Seco, M. Couchaux & M. Hjiaj


Laboratoire de Génie Civil et Génie Mécanique, INSA, Rennes, France

L.C. Neves
Department of Civil Engineering, University of Coimbra, Coimbra, Portugal

ABSTRACT: Base plates of corner columns in buildings are usually subjected to biaxial
bending moment combined with axial force. This combination of forces is not covered by
most of the existing design methods and particularly of the component method adopted in
Eurocode 3 Part 1-8 that treats exclusively the case of a bending moment about the major
(strong) axis combined with an axial force.
This paper deals with a full-scale experimental test program of six column bases subjected
to bi-axial bending moment, focusing on the influence of the thickness of the base plates and
bending loading conditions on the overall behavior. The presented results are expected to be
useful for the validation of analytical models for the resistance, initial stiffness and rotational
capacity calculation to be developed in the future.

1 INTRODUCTION

Load transmission from the superstructure to the foundation rely on the column base plates
that, as a consequence, are one of the most critical elements of steel structures. Their efficiency
and performance strongly affects the overall behavior, namely the global stability, durability,
resistance and ductility.
Column base plates are probably the less studied type of connections, and the majority of the
studies on these elements are focused on the resistance, stiffness and rotational capacity when
subjected to major axis bending moment combined with compressive axial force. The first devel-
opments began very early in Europe in the 70’s and consisted in elastic models for the estimation
of the resistance. These advances were the basis for multiple other investigations that led to the
creation of calculation procedures for the resistance and stiffness commonly used by structural
designer engineers in Europe (EC3 Part 1-8) and the United States (AISC Steel Design Guide
One) for the design of column base plates under axial loads and in-plane bending moment.
Some research works deserve particular attention for its strong contribution in the comprehen-
sion of the behavior and performance of column base plates subjected to major axis bending
moment. Picard and Beaulieu (1985) carried out an experimental investigation on column base
plates subjected to axial force and bending moment aiming at estimating the influence of the load-
ing on the rigidity of the connection. Jaspart and Vandegans (1998) made progresses in the study
of the behavior of column base plates under major axis bending moment by developing
a mechanical model to accurately calculate the moment-rotation curve, which was validated
against experimental data from the tests performed at University of Liège. Later, Stamatopoulos
and Ermopoulos (2011), Abdollahzadeh and Ghobadi (2014) and Márai et al. (2014) proposed
refined models for the evaluation of the resistance of column base plates, considering the effect of
the non-linear deformation characteristics of the elements or the additional resistance provided by
the column web in compression. The analytical, numerical and experimental works of Lee et al.
(2008 a, b) stands out in the field of column bases subjected to minor axis bending moment.
More recently, Bajer et al. (2014), Amaral (2014) and Fasaee et al. (2018) contributed through

981
experimental tests and sophisticated numerical simulations for the development of analytical
models to predict the resistance and stiffness for bi-axial bending moment loading.
This paper presents the results of an experimental investigation performed on six column
bases subjected to bi-axial bending moment. Based on this study, the influence of the base
plate thickness and the orientation of the applied bending moment on the resistance, initial
stiffness and rotational capacity of the connections is analyzed.

2 EXPERIMENTAL PROGRAM

2.1 Tests set-up


Six experimental tests on column base plates connected to a concrete block have been performed
at INSA of Rennes. The key variables of the study were the influence of the orientation of the
applied bending moment and the thickness of the base plate. The tests set-up consisted in
a 1500 kN capacity load-jack attached to the column, applying a vertical force at 1085 mm
from the base plate. The connections were rotated at a convenient angle to create out-of-plane
and bi-axial bending moment (see Figure 1). In these cases, for technical reasons, the connec-
tions were linked to a beam attached to the load-jack by means of two bolted circular plates.
The column was a S275 HEA 200 steel profile with yield and ultimate tensile strengths meas-
ured of 350 MPa and 504.4 MPa, respectively. The S355 base plates shown in Figure 2 were
welded to the column by 7 mm throat fillet welds. Two thicknesses were considered for the base
plates: 10 mm for SPE1 test series and 20 mm for SPE2 test series, as this parameter largely
influences the behavior of this type of connection and the redistribution of forces on the anchor-
ing system. The system was composed of four plain round anchor bolts M16 class 5.6 with
a total length of 350 mm with corresponding nuts and 50 × 50 × 10 mm plate washers. The
embedment length was 300 mm. The yield and ultimate strength were respectively 520.4 MPa
and 626.8 MPa. Concrete blocks with dimensions of 1450 × 900 × 610 mm were designed to
avoid any cracking due to concrete cone failure or concrete blow-out. The concrete grade was
C25/30, typically used for concrete foundations. Tests were performed on concrete cylinders at
31, 73 and 80 days, corresponding approximately to the dates of the tests. Coupon tests were
also performed for the steel used in the column, base plates and anchor bolts (see Table 1).
Tests were performed under displacement control and subjected to monotonic loading con-
ditions. The reference tests SPE1-M0 and SPE2-M0 were the specimens tested under in-plane

Figure 1. Testing device and bending moment orientation.

982
Figure 2. Geometry of the tested column base plates.

Table 1. Mechanical properties of test specimens.


Loading Base plate Concrete
Base plate thick- Base plate yield
Test conditions ultimate compressive
ness tbp (mm) strength (MPa)
(bending) strength (MPa) strength* (MPa)

SPE1-M0 Major-axis 10 423.6 599.8 28.2 (41 days)


SPE2-M0 Major-axis 20 401.3 581.6 28.6 (43 days)
SPE1-M90 Minor-axis 10 423.6 599.8 29.5 (48 days)
SPE2-M90 Minor-axis 20 401.3 581.6 30.4 (51 days)
SPE1-M45 Bi-axial 10 423.6 599.8 32.3 (70 days)
SPE2-M45 Bi-axial 20 401.3 581.6 33.3 (79 days)

* The values were obtained by interpolation based on the standard cylinder tests.

bending moment. For SPE1-M90 and SPE2-M90, the connections were rotated by 90º, per-
forming out-of-plane bending tests. The case with bi-axial bending was also tested in speci-
mens SPE1-M45 and SPE2-M45 in which the connections were rotated by 45º. The loading
history for all tests was: three loading/unloading cycles at 50% of FRd, evaluated according to
EC3 Part 1-8, followed by a loading/unloading cycle applying the full design resistance FRd.
Then, the force was increased until failure. The block was supported in the front by two neo-
prene supports and at portal B, by rollers in the upper and bottom surfaces. Lateral supports
were placed at mid-length in the column to prevent lateral torsional buckling. To monitor the
displacements, seven Linear Variable Differential Transducers were used (see Figure 3). Hori-
zontal transducers Ut1, Ut2, Ub1 and Ub2 were used to evaluate the rotation of the connection.

2.2 Tests results


According to EC3 Part 1-8, the possible failure modes for column base plates are: Mode 1 – yield-
ing of the base plate with prying effects; Mode 2 – anchor bolts failure with yielding of the base
plate and prying effects; Mode 3 – failure of the anchor bolts without prying effects; Mode 1-2 –
similar to Mode 1 but without prying effects. The observed failure mode was bolt failure in ten-
sion without prying effects. However, this failure was preceded by yielding of the flange for the
case of 10 mm base plate thickness. For thicker base plates, the flanges remained rigid during all
the test. On the other hand, for thinner base plates, its behavior was characterized as flexible and
the deformation at the end of the tests was quite significant. As shown in Figure 4, this fact con-
tributed for a more ductile behavior of the connections, causing a later failure at higher values of
rotation. Figures 4 and 5 show the relationship between the bending moment, M, and the rotation
of the connection, θj, obtained from Equation (1). Test results are summarized in Table 2.

983
Figure 3. Instrumentation of the testing device.

Ut  Ub
θj ¼ ð1Þ
h

where Ut = average horizontal displacement of the base plate obtained from Ut1 and Ut2;
Ub = average horizontal displacement of the base plate obtained from Ub1 and Ub2; h = distance
between Ut1 and Ub1.
For tests SPE2-M0, SPE2-M90 and SPE2-M45 with a thicker base plate, as aforementioned
the yielding of the anchor bolts occurred before the base plate yielding, that for the latter was
smooth with the development of a yield line along the base plate in the compression side (con-
firmed by FEM). For SPE1-M0, SPE1-M90 and SPE1-M45, the deformation of the base
plate was quite significant, leading to the conclusion that the first yield lines developed in this
element, and the section of the activated anchor bolts also yielded until failure. As expected,
the rotation values for the tests with a thinner base plate are higher than the values of the
SPE2 test series. This is easily explained by the fact that the 10 mm base plate undergoes
greater deformations comparing to the rigid plate.
From the information depicted in Figures 4 and 5 for the series of tests SPE2, the curves of
cases M45 and M90 are quite different from that of M0, as a clear hardening phase is
observed that is due to the redistribution of the forces in the inner anchor bolts. The curve of
SPE2-M0 is directly proportional to the force displacement curve of the anchor bolts and thus
characterised by a quick loose of stiffness due to necking.
From Table 2, it was concluded that peak moment, initial rotational stiffness and rotation cap-
acity strongly depend on the applied loading case and the base plate thickness. An apparent trend
of the base plate thickness can be observed. For cases with thicker base plates, the yielding
moment, ultimate moment and initial stiffness are visibly higher. For minor axis bending, which
was the most severe case in terms of peak moment and initial stiffness for the two different thick-
nesses of the base plate, the difference between maximum bending moments was approximately

Figure 4. Comparison of the moment-rotation curves between SPE1 and SPE2 test series.

984
Figure 5. Evaluation of the influence of the bending moment orientation.

Table 2. Ultimate resistances, rotations and failure modes of experimental tests.


Myield θ
Test Mmax (kNm) Sj,ini (kNm/rad) u Failure mode
(kNm) (mrad)

SPE1-M0 40.2 47.0 5069.5 36.6 Anchor bolts failure and base plate yielding
SPE2-M0 44.8 52.7 11554.0 31.9 Anchor bolts failure
SPE1-M90 30.8 36.2 2610.1 46.6 Anchor bolts failure and base plate yielding
SPE2-M90 37.0 43.7 7707.7 34.3 Anchor bolts failure
SPE1-M45 36.1 42.8 4769.4 26.1 Anchor bolts failure and base plate yielding
SPE2-M45 41.2 51.4 8388.3 20.1 Anchor bolts failure

20%. As far as the initial stiffness is concerned, this variation is even more pronounced since the
initial stiffness for the 20 mm base plate configuration is approximately three times the 10 mm
base plate configuration initial stiffness. Test results also shown that the specimens subjected to
major axis bending revealed greater moment resistances and initial stiffness but lower rotation
capacity (see Table 2). On the opposite side, specimens under minor axis bending revealed the
weakest bending moment capacity and initial stiffness, but exhibiting higher rotation capacity.
Taking into account the fact that failure was always reached by the anchor bolts without prying
effects, the moment resistance is calculated by Equation (2). This fact is highly related to the
orientation of the applied bending moment since for the in-plane bending moment case, the
center of compression is located between the column flange and the edge of the extended part of
the base plate (see Figure 7). Thus, the lever arm of the binary system of forces Ft,Rd and Fc,Rd is
larger. For out-of-plane bending moment, the location of the centre of compression varies from
the tip of the column flanges to the edge of the base plate. Beyond the fact that this distance is
smaller, the lever arm of the binary of forces Fc,Rd e Ft,Rd is also shortened (see Figures 2 and 7).

Mj;Rd ¼ 2  Ft;Rd  z ð2Þ

where Mj,Rd = moment resistance; Ft,Rd = anchor bolt design resistance; z = lever arm between
Fc,Rd and Ft,Rd.
Another source of discrepancy between these results is associated to the base plate thickness.
As shown in Figure 8, the position of the resultant Fc,Rd of the stress distribution in the compres-
sion side varies according to the behavior of the base plate during the test. If the thickness is
higher and hence more rigid, the center of compression tends to move towards the edge of the
base plate and away from the footprint of the column flange. However, for thinner and more
deformable base plates, the yield line in the compression side tends to develop near the column
flange, leading to a smaller distance between Fc,Rd and Ft,Rd.
The typical sequence of the experimental tests is described below in more detail, emphasizing
some key aspects. Starting with the test series M0, SPE1-M0 was initially subjected to three load-
ing/unloading cycles up to 7.5 kN, followed by a cycle up to 15 kN, which corresponds to the
aforementioned design resistance FRd. As the column base plate was still in the elastic behavior

985
Figure 6. Definition of the experimental yield rotation.

Figure 7. Variation the position of Ft,Rd and Fc,Rd according to the applied bending moment orientation.

region, the registered deformations returned to its initial position after each loading/unloading
cycle. In the fifth cycle, in which the column base plate was loaded until failure, the first sign of
yield lines appearing on the base plate, at the level of the activated anchor bolts and the column
flange in compression occurred for a bending moment of 25 kNm. As soon as the anchor bolts
entered the necking region for M = 47 kNm, the column base plate resistance started to decrease,
with significant elongation of the anchor bolts. The maximum displacement recorded between the
base plate and the concrete before the failure of the anchor bolts was about 10 mm. No cracks
were observed in the concrete block, but some concrete detachment at the level of the yielded
anchor bolts was locally observed. For the test with a 20 mm plate, the sequence was similar,
except for the base plate that slightly yielded throughout the test, causing it to move as a rigid
plate, creating contact with the concrete surface only in the compression area. Thus, although the
observed failure mode for both tests was the same, it is easy to understand that the different base
plate thicknesses originated quite distinct redistribution of internal forces throughout the loading.
For the M90 series of tests, the sequence of events for the 10 mm and 20 mm base plate con-
figurations was quite similar to the SPE1-M0 and SPE2-M0 tests, with the development of
prying forces during the SPE1-M90 test, disappearing at the end of the test where the anchor
bolts elongation was notorious. However, given the orientation of the bending moment, it can
be observed from the curves plotted in Figure 5 that until failure, the loss of resistance capacity
was smoother compared to test series M0. In fact, before reaching the failure of the anchor
bolts, for tests SPE1-M90 and SPE2-M90, the resistance capacity remains nearly constant, the
decrease of force due to necking was not so evident comparatively to tests on SPE1-M0 and
SPE2-M0. This difference is probably due to the redistribution with the inner anchor bolts.

986
Figure 8. Variation the position of Ft,Rd and Fc,Rd according to the base plate thickness (rigid and flexible).

Figure 9. Deformed base plate at the end of tests SPE1-M0, SPE1-M90 and SPE1-M45 (side view).

This redistribution in the anchoring system was also quite distinct during SPE1-M45 and SPE2-
M45 tests. In fact, regardless the more or less deformation capacity of the base plate, in both
cases, the upper anchor bolt was the one that was subjected to a larger force concentration. How-
ever, at the time of failure, a slight elongation was observed at the level of the second and third
anchor bolts. However, although in both cases the three anchor bolts positioned at higher levels
contributed to the overall resistance of the connection, only one was subjected to a higher concen-
tration of forces before reaching its ultimate strength fu. Curves from Figure 5 show that the stric-
tion phase for the most requested anchor bolt was very short, with a small loss of resistance
before reaching failure. The specimens are presented in Figures 9 and 10 before or after anchor
bolt failure.
From the measurements taken during tests and from the base plate deformed shapes in the fig-
ures above, it can be observed that the variation of the position of the contact area for these two
cases is quite clear and in agreement with Figure 8. For SPE1-M0, bending line was located some-
where between 60 to 90 mm from the lower end of the base plate. For SPE2-M0 this distance was
reduced to the range between 30 and 60 mm from the lower end of the base plate. For the case of
SPE1-M90, it was concluded that the contact between the base plate and the concrete in the part
in compression was maintained from the bottom of the base plate until 60 mm height at the end
of the test. However during the beginning of the test, this contact length extend on 90 mm an also
on a part of the beam depth. The fact that the distance between the upper free end of the base
plate and the anchor bolts center line was higher caused the plate to fold at the level of the acti-
vated anchor bolts, crushing against the concrete surface. Similarly to SPE2-M0, the behavior of
the base plate in SPE2-M90 was rigid, being the anchor bolts the only source of deformation. The
maximum uplift was registered at the upper free edge of the base plate and equal to 12 mm. The

987
Figure 10. Deformed base plate at the end of tests SPE2-M0, SPE2-M90 and SPE2-M45 (side view).

displacements of the base plate in the lower part were essentially due to the initial gap, recovered
at the end of the test, resulting in a contact between the plate and the concrete surface.

3 CONCLUSIONS

Experimental tests have been carried out on column bases for two different base plate thicknesses
and loading conditions, showing that these two variables strongly affect the overall behavior of these
connections. It was shown that thicker base plates confer higher resistance and initial stiffness but
less rotation capacity to the connections. On the other hand, thinner base plates demonstrated lower
values for the bending moment resistance with greater deformability. The influence of the orienta-
tion of the applied bending moment on the mechanical properties of the connections was carried out
as well, proving that the greater the distance between the centers of compression and tension, the
higher the moment resistance, as in the case of in-plane bending moment. The presented studies
proved that the mechanical properties corresponding to the bi-axial bending cases were limited
superiorly and inferiorly by the in-plane and out-of-plane bending moment cases, respectively.
This paper aimed to assess and better understand the behavior of column base plates sub-
jected to biaxial bending moment. The presented results can constitute a useful data base for
the calibration and validation of forthcoming analytical models for the prediction of the prop-
erties of this type of connections.

REFERENCES

Abdollahzadeh, A. & Ghobadi, F. 2014. Mathematical modeling of column-base connections under


monotonic loading. Civil Engineering Infrastructures Journal 47(2): 255–272.
Amaral, P. 2014. Steel column bases under biaxial loading conditions (master degree dissertation). Porto:
University of Porto.
Bajer, M.Vild, M.Barnat, J. & Holomek, J. 2014. Influence of selected parameters on design optimiza-
tion of anchor joint. 12th International Conference on Steel, Space and Composite Structures, Prague,
28–30 May 2014: 149–158.
Fasaee, M. Banan, M. & Ghazizadeh, S. 2018. Capacity of exposed column base connections subjected
to uniaxial and biaxial bending moments. Journal of Constructional Steel Research 148: 361–370.
Jaspart, J.P. & Vandegans, D. 1998. Application of the component method to column bases. Journal of
Constructional Steel Research 48: 89–106.
Márai, P. Kövesdi, B. Oly, R. Belica, A. & Dunai, L. 2014. Resistance model for fixed column bases.
Eurosteel 2014, Naples, 10–12 September 2014.
Picard, A. & Beaulieu, D. 1984. Behaviour of a simple column base connection. Canadian Journal of Civil
Engineering 12: 126–136.
Stamatopoulos, G. & Ermopoulos, J. 2011. Experimental and analytical investigation of steel column
bases. Journal of Constructional Steel Research 67: 1341–1357.
Wald, F. Sokol, Z. Steenhuis, M. & Jaspart, J.P. 2008. Component method for steel column bases. Heron
Volume 53: 3–20.

988
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Quantifying the seismic ductility of lightweight steel lateral force


resisting systems through procedures of FEMA P695

S. Shakeel
Department of Structures for Engineering and Architecture, University of Naples “Federico II”, Italy

ABSTRACT: Seismic design regulations rely on certain factors to reduce the earthquake
actions due to inherent ductility and overstrength of the lateral force resisting systems. The
behavior factors given in Eurocodes are one such example, which when applied during the
design process will ensure the life safety performance level of a building. Lightweight steel
(LWS) structures fabricated with Cold formed steel (CFS) profiles have now been in use for
quite a sometime in earthquake prone European regions, however the current in practice edition
of Eurocodes fail to acknowledges their seismic performance by does not providing behavior
factors for them. The study presented herein addresses this issue by evaluating behavior factor
for the two most commonly used LWS systems: CFS strap braced studs walls and CFS shear
walls with gypsum board sheathing through FEMA P695 methodology. For each type of
system, a set of archetypes, which represent a range of design parameters and building configur-
ations are designed following the capacity design approach and their response is idealized by
nonlinear models. The performance of archetype models is evaluated systematically through the
static pushover and the incremental dynamic analysis under a suite of forty-four normalized
and scaled earthquake records, representing the probable seismic hazard to the buildings.
Finally, by calculating the collapse probability while also considering the uncertainties from
various sources, the suitability of trial value of behavior factor used in the design phase of arche-
types is evaluated. Based on the results, it is concluded that a behavior factor of 2.5 and 2.0 for
CFS strap braced stud walls and CFS shear walls with gypsum board sheathing is appropriate.

1 INTRODUCTION

Improved structural efficiency and superior environmental performance of Light Weight Steel
(LWS) constructions made with a Cold Formed Steel (CFS) frame and sheathing panels have
made them stand out as a preferable alternative in contrast to traditional masonry or concrete
constructions for low rise building applications. To resist the earthquake horizontal loads,
two different energy dissipative solutions are used in LWS constructions, all-steel and
sheathed-braced. Lateral resistance can be provided by thin steel straps acting as braces in an
X configuration or sheathing-braced solutions with steel sheets, wood or gypsum-based panels
can provide the lateral bracing effect to withstand horizontal loads.
Currently, only the North American standard for seismic design of CFS structures AISI
S400 (AISI, 2015) provides design guidelines for various typologies of lateral force resisting
systems (LFRS) used in CFS constructions. In contrast with AISI S400, the main European
earthquake standard EN 1998-1 (CEN, 2004) lacks guidelines for any type of LFRS used in
LWS constructions. In order to bridge this gap, a series of numerical and experimental investi-
gations (Fiorino et al, 2014; Macillo et al, 2014; Fiorino et al. 2016; Fiorino et al., 2017b; Ter-
racciano et al., 2018; Fiorino et al. 2018; Pali et al, 2018; Fiorino et al. 2019) were carried out
at University of Naples Federico II”, Italy in recent years to better understand the seismic
response of these constructions and to propose design guidelines within the European design
framework. These studies were part of a wider effort (Campiche et al., 2019; Costanzo et al.,

989
Figure 1. a) CFS gypsum sheathed shear wall b) CFS strap braced wall.

2017; Dell’Aglio et al., 2017; Tartaglia et al., 2018) carried out at University of Naples Feder-
ico II with its collaborators to better understand the seismic performance of steel structures.
Behavior factors are employed in EN 1998-1 for buildings equipped with LFRS to reduce
the effect of seismic actions due to building’s inherent structural ductility and overstrength.
Behavior factor for any structural system can be evaluated preliminary through either a test-
based or a numerical approach. In a test based approach, usually monotonic and cyclic tests
(Macillo et al., 2017) of the wall specimens are used to evaluate the response while numerical
approach make uses of several building archetypes to evaluate the collapse probability of
LFRS system and to make a judgement on reasonable value of behavior factor. A more elab-
orate methodology is outlined by FEMA document P695 (FEMA, 2009), which uses non-
linear analysis techniques, and explicitly considers uncertainties in ground motions,
modelling, design, and experimental data to achieve an acceptably low probability of collapse
of the seismic-force resisting system for a particular value of behavior factor. The work pre-
sented in this paper uses this methodology to evaluate behavior factor for CFS gypsum
sheathed shear and strap braced walls (Figure 1).
The methodology outlined by FEMA P695 (FEMA, 2009) considers an archetypes design
space to characterize the system behavior, which are designed using a trial value of behavior
factor and whose response is idealized by nonlinear models. A value of 2.5 and 2.0 is used as
a trial value for CFS strap braced and CFS gypsum sheathed shear walls, respectively. The
performance of archetypes is evaluated systematically through static pushover and incremen-
tal dynamic analysis techniques. By calculating the collapse probability, while also considering
the uncertainties from various sources, it evaluates the suitability of trial value of behavior
factor used in design phase of archetypes. The organization of this paper also follows the
arrangement of different steps in the FEMA methodology.

2 ARCHETYPE SELECTION AND DESIGN

A set of 14 archetypes (Table 1) are designed separately for each system. The archetypes dif-
fered with each other in terms of occupancies, building heights (H) and seismic hazards. The
complete design space is considered as one performance group due to the fact that the changes
in structural configuration and gravity loads among different archetypes will not alter the
behavior of seismic force resisting system, significantly. Three anonymous geographical sites
with low, medium and high seismic intensity levels representative of the type of seismic hazard
exposed to European continent are assumed as location of archetypes. The Low, Medium and
High seismic hazard are assigned with reference peak ground acceleration (ag,R) values of

990
Table 1. Archetype design space.
CFS gypsum sheathed shear walls CFS strap braced wall

Archetype Design base shear No of walls in Archetype Design base shear No of walls in
ID* [kN] each direction ID* [kN] each direction

R1L 129 4 R2L 292 4


R2L 175 6 R3L 470 6
R3L 322 12 R4L 644 8
R4L 469 16 R2M 486 6
R1M 258 10 R3M 783 8
R2M 350 12 R4M 1073 10
R3M 528 18 R2H 681 8
R1H 387 14 R3H 1096 10
R2H 525 18 O1L 212 6
O1L 101 4 O2L 475 6
O2L 332 12 O1M 352 6
O1M 203 8 O2M 792 6
O2M 504 18 O1H 494 6
O1H 304 12 O2H 1108 10

*Archetype ID: First letter: (Type of Occupancy R-residential or O-office), Second number: umber of storeys,
Third alphabet: Level of seismic hazard (L-low, M-medium or H-high).

0.1g, 0.2g and 0.3g, respectively in case of gypsum sheathed shear walls and with the values of
0.15g, 0.25g and 0.35g, respectively in case of strap braced walls. The reason for the choice of
a comparatively lower hazard for gypsum sheathed shear walls is due to their limited strength,
with which it was not possible to design a reasonable number of archetypes at the same
hazard level as that for strap braced walls. All of the archetypes are assumed to be built on
Soil class C and belong to an Importance class II according to EN 1998-1 (CEN, 2004).
Two type of occupancies are considered for the archetypes: residential and office. The
major differences among the two occupancies is different intensity of live loads they cause. In
particular, Eurocode EN 1990 (CEN, 2002) defines a live load of 2.0 kN/m2and 3.0 kN/m2 for
residential and office type of occupancies, respectively. The two occupancies also differ with
each other in terms of accessibility to roof. In case of residential occupancy, roof is considered
accessible with an expected live load equal to floors i.e. 2.0 kN/m2, while the roof of office
archetype is considered inaccessible except for normal maintenance and repair, which resulted
in 0.4 kN/m2 expected live load defined based on EN 1990 (CEN, 2002). Two different types
of floor are used: a lightweight steel floor sheathed with panels (light solution) and
a composite steel concrete deck (ordinary solution). The use of light floors is limited to only
those gypsum sheathed shear walls braced archetypes which resulted in an excessive design
base shear, if an ordinary floor is used. In case of strap braced archetypes, only ordinary
floors are used. The resulting dead loads on floors due to light or ordinary floor, ceilings and
vertical partitions are 1.30 or 1.80, 0.10 and 0.80 kN/m2, respectively, while the roof dead
loads due to light or ordinary roof and ceilings are 1.40 or 2.30 and 0.10 kN/m2, respectively.
In addition to dead loads, a snow load of 1.00 kN/m2 and a wind load of 0.35 kN/m2 is
assumed to be acting on the roof of archetypes. The height of the archetypes reflected the cur-
rent trend of LWS construction application in seismic areas, which is only limited to low rise
buildings. For residential building 3.0 to 12.0 m high archetypes are selected, while for office
archetype maximum height of archetype is limited to 6.0 m. Total height of each storey for
both type of occupancies is 3.0 m including a floor depth of 0.3 m.
The archetypes are structurally designed according to the relevant parts of Eurocodes. The
design of archetypes against the gravity load is carried out using EN 1993-1-3 (CEN, 2006),
which provides supplementary rules for CFS members and sheeting. The lateral loads acting
on the archetypes included a wind load of 0.70 KN/m2 and earthquake loads as explained

991
earlier. Lateral force method in EN 1998-1 (CEN, 2004) is used to compute the design base
shear, because all of the archetypes meet the criteria of regularity in structural plan and eleva-
tion. The seismic design base shear is distributed over height of structure corresponding to
seismic mass of each floor or the first modal shape. Both types of lateral force distribution are
used in pushover analysis of archetypes as explained in upcoming sections in order to obtain
the worst-case scenario response. A 2.40 m wide shear wall configuration which is already
tested by authors (Macillo et al., 2017) and has shown acceptable performance under both
cyclic and shake table testing (Fiorino et al., 2017a) is selected as the only configuration used
here for archetypes with gypsum sheathed shear walls as main seismic force system. On the
other hand, three wall configurations of low, medium and high shear force capacities were
used for archetypes with strap braced walls as main seismic force system. The low and
medium capacity walls were already tested experimentally (Iuorio et al., 2014), while high cap-
acity wall was designed theoretically based on formulations given in (Macillo et al., 2018).
Further details about its geometry, material properties and methods to predict theoretically
the design strength are presented in (Macillo et al., 2018b).

3 NUMERICAL MODELLING

Numerical modelling of the archetypes is needed for their seismic performance evaluation. It
is expected from the building using CFS gypsum sheathed shear walls as primary seismic force
resisting system, that it would mainly develop the nonlinear mechanism only at the sheathing
connections, while rest of the element would remain elastic. For the strap braced wall, the
main nonlinear mechanism only happens in steel straps due to their plastic yielding. Thus, the
simulation of response of the walls is the single most important factor in the modelling of
complete building response. OpenSees software (Mazzoni et al., 2009) is used for developing
the numerical models. A two-stage modelling approach is adopted. At first, numerical model
for single wall is developed and calibrated based on the experimental result and then in the
later stage, the model of single wall is incorporated in a complete building representing the
global response of archetype. The choice of the type of model for single wall and its validation
and different modelling options available in literature are explained in detail by authors in
(Fiorino et al., 2018; Macillo et al., 2018) for CFS gypsum sheathed shear walls and in
(Macillo et al., 2018) for CFS strap braced walls. A simplified model (Figure 2), which relies
on a pair of nonlinear diagonal truss elements is used to simulate the response of a wall.
Pinchinng4 material available in OpenSees is used to represent the nonlinearity in truss elem-
ents, which have a unit area and are the representative of the behavior of walls as a whole.

Figure 2. Simplified model for the walls.

992
Rest of the wall elements, including chord studs and hold down devices are also modelled as
elastic elements, which are necessary for the transfer of forces to the rest of building.
A three-dimensional model representing the most important structural elements that would
have significant implications on the global seismic response of building is created for each
archetype. The skeleton of a typical CFS building is made by joining the wall elements, such as
studs and wall tracks, to the floor elements, such as floor tracks and joists. The gravity studs
are modelled using truss elements. Floor elements (joist, tracks and composite floor) are mod-
elled as the rigid elastic beam column elements. To capture the lateral displacement arising
from P-delta effect, a rigid frame made of axially rigid elastic beam column elements with low
flexure stiffness having a co-rotational coordinate transformation is connected to the building
in the direction of seismic action. Moment releases are created at the ends of columns of P-delta
frame and are linked to the main building frame using the rigid truss elements. Moment releases
at the ends of columns are intended to avoid any transfer of moments from P-delta fame to
main frame and hence only transmitting the axial force, which induce an equivalent lateral dis-
placement in the main frame. Gravity load is applied to the P-delta columns at floor levels.
More information on modelling of CFS archetypes is give in (Fiorino et al., 2017b). For
dynamic analysis, a 2% damping ratio is used based on the Rayleigh damping model. This 2%
value reflected the measured damping values during the shaking table experiments (Fiorino
et al., 2017a) on a two storey CFS building.

4 SEISMIC ANALYSIS AND RESULTS

4.1 Non-linear static analysis


FEMA P695 (FEMA, 2009) methodology uses non-linear static analysis to estimate archetype
over-strength by calculating the ratio of maximum base shear resistance of archetype obtained
through non-linear static analysis to the design base shear. After achieving the maximum base
shear resistance, all archetype models have the concentration of damage at a ground storey in case
of gypsum sheathed shear wall archetypes, while in case of strap braced archetype the story with
the damage concertation varied from case to case depending on the distribution of shear resistance
over the height of archetype. Moreover, this particular difference is due to the difference in design
analogy of both systems. In case of gypsum sheathed shear walls, all stories had a same design
shear resistance since there was only one configuration of gypsum sheathed shear wall used in all
archetypes on all floors. While in case of strap braced walls, the design shear resistance descended
with ascending heights since there were three different types of wall configurations used in arche-
types design. The response in two planar directions do not differ much in terms of Vmax in most
cases, while the ultimate roof displacement (δu ) presented slight differences. Ultimate roof displace-
ment (δu ) is taken as the roof displacement at the point of 20% strength loss (0.8Vmax ) in pushover
analysis.
Additionally, the pushover analysis is also used to compute the period-based ductility
according to formulations in FEMA (FEMA, 2009), which is used to evaluate the effect of
spectral shape of different earthquake records, used for dynamic analysis, on the archetype
performance. Subsequently, the final value of over-strength and period-based ductility for
each archetype is computed based on its average response in both directions and under both
types of load distribution patterns based on FEMA P695 (FEMA, 2009) recommendations.

4.2 Non-linear Incremental Dynamic Analysis (IDA)


The collapse performance of archetypes of archetypes in terms of the Collapse Margin Ratio
(CMR) can be assessed through incremental dynamic analysis (IDA), which uses a set of ground
motions scaled with increasing scaling factors to reach a magnitude that will cause the collapse.
CMR is the ratio between median collapse intensity (SCT ) and maximum considered earthquake
intensity (SMT ). The 4% inter storey drift ratio is used as a threshold value causing the collapse at
global level for archetypes with CFS gypsum sheathed shear walls. The 4% value is selected based
on a similar study on shear walls with fiber cement board panels (Zeynalian et al., 2018), which

993
Figure 3. IDA curves: a) CFS Strap braced walls, b) CFS gypsum sheathed shear walls.

exhibit comparable behavior as walls with gypsum board panels and are classified under the same
category in AISI S400 (AISI, 2015). Contrarily, a 5% inter storey drift value is chosen for strap
braced archetypes as failure criterion. However, this value was chosen based on experiment results
(Iuorio et al., 2014), in which strap braced walls were able to sustain load carrying capacity until
at least 5% drift in all specimens.
Far-field record set provided by FEMA (FEMA, 2009) is used to conduct IDA in both planar
directions with a 20% increment in intensity starting from a 20% (Scaling factor SF = 0.2) to
600% (SF = 6.0). Before applying the records to archetypes, they were normalized to remove the
unwarranted variability between records due to inherent differences in event magnitude, distance
to source, source type and site conditions, without eliminating overall record-to-record variability
based on Peak Ground Velocity (PGV) and were matched to the archetype design response spec-
trum through the scaling of median response spectrum of all records within the range of funda-
mental vibration periods of archetypes. The SF = 1.0 is representative of design level earthquake,
while SF = 1.5 corresponds to MCE (Maximum Considered Earthquake). Figure 3 shows the
IDA curves for three storeys residential (R3M) archetype for both types of systems

5 SEISMIC PERFORMANCE EVALUATION

The acceptably low, yet reasonable, probability of collapse of the structural systems is established
by checking the suitability of trial value of behavior factor used in design as per performance
evaluation criteria of FEMA P695. The CMR’s obtained from the IDA analysis are adjusted to
take into the spectral shapes of different records used for the analysis to obtain the adjusted CMR
(ACMR), which is the product of CMR and a spectral shape factor (SSF). Two performance cri-
teria are defined by the methodology (FEMA, 2009): the average value of adjusted collapse
margin ratio, ACMRavg , for each performance group should be greater than ACMR10% , and the
individual values of adjusted collapse margin ratio, ACMRi , for each archetype within
a performance group should be greater than ACMR20% . The value of ACMR10% and ACMR20%
are the function of uncertainties, which are summed up as then the total collapse uncertainty
βTOT, that takes into account different sources of uncertainty which could contribute to variability
in collapse capacity. In particular, the main sources of uncertainty include: record to record uncer-
tainty (βRTR), design requirements uncertainty (βDR), test data uncertainty (βTD); and modelling
uncertainty (βMDL).
βRTR represents the variability in response of the archetypes model under different ground
motion records. FEMA P695 recommends to use βRTR equal to 0.40 for archetypes having
period-based ductility greater than 3, which is the case for all archetypes in this study. The other
three uncertainty parameters βDR, βTD and βMDL, are the qualitative measures. FEMA P695
(FEMA, 2009) provides the following qualitative scale to quantify them: (a) Superior: β = 0.10, (b)
Good: β = 0.20, (c) Fair: β = 0.35, and (d) Poor: β = 0.50. A quality rating of ‘Good’ (β = 0.20) is

994
Table 2. Performance evaluation of archetypes with CFS strap braced walls.
Pass ACM Pass
Archetype ID SCT SMT SSF CMR ACMR ACMR20% /Fail Ravg ACMR10% /Fail

R2L 3.40 1.50 1.14 2.27 2.58 1.57 Pass


R3L 3.67 1.50 1.14 2.45 2.79 1.57 Pass
R4L 3.70 1.50 1.13 2.47 2.79 1.57 Pass
O1L 2.87 1.50 1.14 1.91 2.18 1.57 Pass
O2L 3.40 1.50 1.14 2.27 2.58 1.57 Pass
R2M 2.27 1.50 1.14 1.51 1.73 1.57 Pass
R3M 2.18 1.50 1.13 1.45 1.64 1.57 Pass
R4M 2.53 1.50 1.12 1.69 1.89 1.57 Pass 2.05 1.97 Pass
O1M 2.40 1.50 1.14 1.60 1.82 1.57 Pass
O2M 2.50 1.50 1.14 1.67 1.90 1.57 Pass
R2H 1.95 1.50 1.30 1.30 1.69 1.57 Pass
R3H 1.94 1.50 1.31 1.29 1.69 1.57 Pass
O1H 1.72 1.50 1.33 1.15 1.53 1.57 Near Pass
O2H 2.16 1.50 1.30 1.44 1.87 1.57 Pass

Table 3. Performance evaluation of archetypes with CFS gypsum sheathed shear walls.
Pass ACM ACMR Pass
Archetype ID SCT SMT SSF CMR ACMR ACMR20% /Fail Ravg 10% /Fail

R1L 2.70 1.50 1.12 1.80 2.02 1.51 Pass


R2L 2.80 1.50 1.11 1.87 2.07 1.51 Pass
R3L 3.10 1.50 1.10 2.07 2.27 1.51 Pass
R4L 3.33 1.50 1.08 2.22 2.39 1.51 Pass
O1L 2.47 1.50 1.13 1.64 1.86 1.51 Pass
O2L 2.90 1.50 1.09 1.93 2.11 1.51 Pass
R1M 3.60 1.50 1.13 2.40 2.71 1.51 Pass
R2M 2.25 1.50 1.09 1.50 1.64 1.51 Pass 1.91 1.88 Pass
R3M 2.00 1.50 1.07 1.33 1.43 1.51 Near Pass
O1M 2.09 1.50 1.10 1.39 1.53 1.51 Pass
O2M 2.23 1.50 1.08 1.49 1.61 1.51 Pass
R1H 2.57 1.50 1.25 1.71 2.13 1.51 Pass
R2H 1.80 1.50 1.18 1.20 1.42 1.51 Near Pass
O1H 1.83 1.50 1.23 1.22 1.50 1.51 Near Pass

adopted for all of them in both cases except a ‘superior’ rating is given to test data available for
gypsum sheathed shear walls. In addition to quasi cyclic test data on the wall (Macillo et al.,
2017), there was also significant evidence on the exceptional performance of gypsum sheathed
shear walls from shaking table tests (Landolfo et al., 2018). Finally, the total collapse uncertainty
βTOT is computed to be 0.49 and 0.53 for CFS gypsum sheathed shear walls and strap braced
walls, respectively. Based on the value of βTOT, ACMR10% and ACMR20% are obtained from
Table 7-3 of FEMA P695 (FEMA, 2009). All of the archetypes in both cases passed the criterion,
as it can be seen from Tables 2 and 3 and hence this concludes that the value of 2.5 and 2.0 used in
design phase for CFS strap braced and gypsum sheathed shear walls, respectively is appropriate.

6 CONCLUSIONS

Behavior factor for CFS gypsum sheathed shear walls and strap braced walls, which are not yet
standardized as seismic force resisting system in EN 1998-1, is evaluated here through FEMA
P695 methodology. A total of fourteen residential and office archetypes representing heights up
to 12 m are designed following capacity design approach to withstand the low, medium and high

995
seismic load. Numerical models of archetypes with an ability to simulate nonlinear hysteretic
response of walls are developed in OpenSees software and analyzed under the action of 44 records
in FEMA P695 far-field record set. The analysis of models followed an incremental dynamic ana-
lysis approach until the collapse happens. Finally, by gauging the collapse performance of models
against the FEMA P695 acceptance criteria for adjusted collapse margin ratios, it is concluded
that a behavior factor value of 2.0 and 2.5 is appropriate for buildings using CFS gypsum
sheathed shear walls and strap braced walls as main seismic force resisting system, respectively.

REFERENCES

AISI. 2015. S400-15 North American Standard for Seismic Design of Cold formed Steel Structural Sys-
tems. American Iron and Steel Institute (AISI).
Campiche, A., Shakeel, S., Bucciero, B., Pali, T., Fiorino, L., & Landolfo, R. 2019. Seismic Behaviour of
Strap-Braced LWS Structures: Shake Table Testing and Numerical Modelling. IOP Conference Series:
Materials Science and Engineering, 473: 012032.
CEN. 2002. EN 1990 Eurocode 0: Basis of structural design. Brussels: European Committee for
Standardization.
CEN. 2004. EN 1998-1 Eurocode 8: Design of Structures for earthquake resistance-Part 1: General rules,
seismic actions and rules for buildings. Brussels: European Committee for Standardization.
CEN. 2006. EN 1993- 13Eurocode 3: Design of steel structures-Part 1-3: General rules-Supplementary
rules for cold-formed members and sheeting. Brussels: European Committee for Standardization.
Costanzo, S., D’Aniello, M., & Landolfo, R. 2017. Seismic design criteria for chevron CBFs: Proposals
for the next EC8 (part-2). Journal of Constructional Steel Research, 138: 17–37.
Dell’Aglio, G., Montuori, R., Nastri, E., & Piluso, V. 2017. A critical review of plastic design approaches
for failure mode control of steel moment resisting frames. Ingegneria Sismica, 34(4): 82–102.
FEMA. 2009. FEMA P695: Quantification of Building Seismic Performance Factors. Washigton, DC, USA.
Fiorino, L., Iuorio, O., & Landolfo, R. 2014. Designing CFS structures: The new school bfs in naples.
Thin-Walled Structures, 78: 37–47.
Fiorino, L., Macillo, V., & Landolfo, R. 2017a. Shake table tests of a full-scale two-story
sheathing-braced cold-formed steel building. Engineering Structures, 151: 633–647.
Fiorino, L., Pali, T., Bucciero, B., Macillo, V., Teresa Terracciano, M., & Landolfo, R. 2017b. Experi-
mental study on screwed connections for sheathed CFS structures with gypsum or cement based
panels. Thin-Walled Structures, 116(January): 234–249.
Fiorino, L., Shakeel, S., Macillo, V., & Landolfo, R. 2018. Seismic response of CFS shear walls sheathed
with nailed gypsum panels: Numerical modelling. Thin-Walled Structures, 122: 359–370.
Fiorino, L., Terracciano, M.T., & Landolfo, R. 2016. Experimental investigation of seismic behaviour of
low dissipative CFS strap-braced stud walls. Journal of Constructional Steel Research, 127: 92–107.
Iuorio, O., Macillo, V., Terracciano, M.T., Pali, T., Fiorino, L., & Landolfo, R. 2014. Seismic response
of Cfs strap-braced stud walls: Experimental investigation. Thin-Walled Structures, 85: 466–480.
Landolfo, R., Iuorio, O., & Fiorino, L. 2018. Experimental seismic performance evaluation of modular
lightweight steel buildings within the ELISSA project. Earthquake Engineering & Structural Dynamics.
Macillo, V., Fiorino, L., & Landolfo, R. 2017. Seismic response of cold-formed steel shear walls sheathed
with nailed gypsum panels: Experimental tests. Thin-Walled Structures, 120(August): 161–171.
Macillo, V., Iuorio, O., Terracciano, M.T., Fiorino, L., & Landolfo, R. 2014. Seismic response of Cfs
strap-braced stud walls: Theoretical study. Thin-Walled Structures, 85: 301–312.
Macillo, V., Shakeel, S., Fiorino, L., & Landolfo, R. 2018b. Development and calibration of a hysteretic
model for CFS strap braced walls. Advanced Steel Construction, 14(3): 337–360.
Mazzoni, S., McKenna, F., Scott, M.H., & Fenves, G.L. 2009. Open System for Earthquake Engineering
(OpenSees).
Tartaglia, R., D’Aniello, M., Gianmaria, D.L., & Attilio De Martino 2018. Influence of EC8 rules on
P-Delta effects on the design and response of steel MRF. Ingegneria Sismica: International Journal of
Earthquake Engineering, 13(3): 104–120.
Terracciano, M.T., Vincenzo, M., Pali, T., Bucciero, B., Luigi, F., & Landolfo, R. 2018. Seismic design
and performance of low energy dissipative CFS strap-braced stud walls. Bulletin of Earthquake Engin-
eering, Springer, 17: 1075–1098.
Zeynalian, M., Shahrasbi, A. Z., & Riahi, H.T. 2018. Seismic Response of Cold Formed Steel Frames
Sheathed by Fiber Cement Boards. International Journal of Civil Engineering, 16(11): 1643–1653.

996
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Numerical modelling of a two storey LWS building braced with


gypsum-based panels

S. Shakeel, A. Campiche & R. Landolfo


Department of Structures for Engineering and Architecture, University of Naples “Federico II”, Italy

ABSTRACT: The seismic behaviour of a full-scale two-storey Cold-Formed Steel (CFS)


building is explored via the shake-table tests on its bare structure and the complete construc-
tion phases. Starting from the shake-table test results, advanced numerical models have been
developed using the OpenSees software, which also consider both the structural and the non-
structural elements along with the contribution offered by the finishing components to the lat-
eral force resistance. Numerical models implemented for both construction phases are able to
simulate with good accuracy the experimental seismic behaviour of the building, in term of its
fundamental period, the drift peaks and the displacement time history.

1 INTRODUCTION

The recent seismic events have highlighted the need to deeply know the dynamic behaviour of
structures and their response to earthquake actions, in order to prevent and minimize damages
and losses. Subsequently, in the last decades, many studies have been started with the goal of
exploring the seismic response of structures (Dell’Aglio et al., 2017; Piluso et al., 2019; Tartaglia
et al., 2018a; Cassiano et al., 2018; Tartaglia et al., 2018b) and in particular CFS steel structures
at University of Naples Federico II, Italy (Landolfo, 2019). The structural elements in a typical
Lightweight Steel (LWS) building are the cold-formed steel (CFS) profiled members and the
main lateral force resistance components are the shear walls, which could be either sheathed by
flat sheets of steel or panels made of wood, gypsum or cement-based materials. Although the
capability of LWS constructions to exhibit high structural, technological and environmental
performances is well established by now, the lack of high-fidelity structural modelling tools,
which possess the ease of use, is quite evident. The state of the art of current numerical model-
ling of CFS structures has been extensively reviewed by the authors elsewhere (Fiorino et al.,
2018; Macillo et al., 2018). The most vital element for the simulation of dynamic response of
LWS structures is the shear walls. Its non-linear behaviour is often represented by a pair of
diagonal truss elements, calibrated on available experimental data. However, modelling only
shear walls cannot be enough for an accurate prediction of CFS building lateral response and it
can be necessary to develop 3D models, which include also other components often neglected.
In this perspective, the present paper provides the details of numerical modelling of a two-
storey CFS gypsum sheathing-braced building, Figure 1 (ELISSA mock-up) in OpenSees soft-
ware. The building was tested on the shake-table as a part of European project, named as
ELISSA (Energy efficient Lightweight-Sustainable-Safe steel construction. The proposed
numerical model includes all the structural and non-structural elements, while remaining quite
simple from the computational point of view.

2 SHAKE-TABLE TESTING

The ELISSA mock-up was designed for gravity loads as an “all-steel” solution and for seismic
loads as a sheathing-braced solution, in which the seismic resistant elements are made of CFS

997
Figure 1. ELISSA mockup.

998
gypsum sheathed shear walls. The dimensions of the mock-up were about 2.5×4.5 m in plan,
with an inter-storey height of about 2.3 m and a total height of about 5.4 m. The mock-up was
designed in accordance with current European building codes (CEN, 2004, 2006) and technical
approvals (Cocoon, Transformer, ETA-11/0105, 2011., in press). More details about the
research project are available in (Fiorino et al., 2018; Macillo et al., 2018). The mock-up was
built in two different stages: in the first phase, named as the bare structure (Configuration B), in
the second phase, named as complete construction (Configuration C).
The mock-up in Configuration B consisted mainly of shear walls and gravity walls without
finishing parts, internal partition walls, floors and roof without finishing parts. Shear, gravity
and partition walls were built with 147×50×1.5 mm C studs (steel grade S320GD+Z) with
a spacing of about 600 mm on the centre connected at their end with 150 × 40 × 1.5 mm
U wall tracks. Walls were sheathed with 15 mm thick impact resistant gypsum board panels
on both sides. Connections between studs and wall tracks were made by self-drilling screws,
whereas sheathing panels were connected to steel frame by ballistic nails. The main difference
between shear and gravity walls is the presence of hold-down at ends of the shear walls in
order to withstand the axial force due to overturning phenomena. Internal partition walls,
instead, were not designed to carry gravity loads. Floors and roof were made of back-to-back
lipped C 197×50×10×2.0 mm joists with a spacing of about 500 mm on the centre connected
at their end with 200×40×1.5 mm U floor tracks, sheathed on the top side with 28 mm thick
high-density gypsum fibre board panels. Connections between joists and floor tracks were
made by self-drilling screws. Bottom sides of second floor and roof were sheathed with 15 mm
thick impact resistant gypsum board panels connected to steel frames.
The mock-up in Configuration C was completed with non-structural elements, which mainly
consisted of internal counter walls and finishing parts of shear walls, gravity walls, and finishing
parts of floors and roof. In particular, internal counter walls of shear and gravity walls were
made of 50×50×0.6 C non-structural studs spaced at about 600 mm on the centre, connected at
their end with 50×40×0.6 U wall tracks and sheathed only on the internal side with a double
layer of 15 mm thick impact resistant gypsum board panels. Internal counter walls were com-
pleted with 20 mm thick vacuum insulated panel and 50 mm thick insulation mineral wool.
External faces of shear and gravity walls were completed with a ventilated façade. Top sides of
first and second floors were sheathed with additional gypsum fibre board panels, whereas
bottom sides of second floor and roof were completed with a ceiling. Further details about
mock-up are available in (Landolfo, 2018). Two shake-table test typologies were performed in
the shorter direction of mock-up i.e. X direction.: dynamic identification tests using a white
noise signal to evaluate the dynamic properties, such as fundamental period of vibration and
damping ratio, and earthquake tests for the evaluation of the seismic response of the building.
The dynamic identification tests were carried out at the both phases of construction of mock-up
while the earthquake tests were performed only at the complete construction phase
i.e. Configuration C.

3 DEVELOPMENT OF DETAILED ELISSA MOCK-UP MODEL

A two-stage modelling approach was adopted. In the first phase, the model of individual
shear walls in the mock-up were developed and calibrated. Later in the second phase, the com-
plete 3D model of the mock-up was developed.
Eight different shear walls were identified considering the two construction phases as shown
in Figure 1 (in red). Two wall specimens representative of the shear wall B-2650 in the mock-up
were also tested under monotonic and quasi static cyclic loading and a single wall specimen rep-
resentative of the shear wall C-2650 was tested under quasi static cyclic loading (Macillo et al.,
2017). The hysteretic response of the shear walls (Figure 2a) was concentrated in a pair of diag-
onal trusses elements with a pinching4 material, calibrated based on quasi static cyclic test
response of the walls. In particular, the pinching4 material requires a backbone envelope and set
of cyclic parameters to represent any hysteretic response. The calibration process of these walls
models is further explained by the authors in (Fiorino et al., 2018). An assumption that strength

999
Figure 2. a) shear wall w/intermediate studs, b) gravity and partition walls w/intermediate studs, c) internal
counter wall.

of wall per unit length remains constant properties per unit wall length is use to model the shear
walls without any experimental data on their in-plane response. In particular, the backbone
envelope of these walls was extrapolated from the backbone envelope of walls B-2650 and
C-2650 using a sailing L/L2650, where L represents the length of the shear wall without experi-
mental data and L2650 = 2650 mm. Zero Length elements were used to model the hold-downs
present at the ends of the shear walls. OpenSees uniaxial Material MinMax was used in conjunc-
tion with uniaxial Material Elastic Multilinear to explicitly model the different brittle collapse
mechanism in the hold-downs using the limiting values of strains associated to these failures.
More details are provided in (Fiorino et al., 2018). The tracks were modelled by rigid truss elem-
ents. Studs were also modelled as a truss element having a uniaxial Elastic Material wiht the
same properties of the profiles used in the mock-up. Similar to hold downs, uniaxial Material
MinMax was also used for chord studs to explicitly model buckling mechanisms and tensile rup-
ture. Since the floors behaved rigidly during experiments, a rigid Diaphragm constraint was
used at floor nodes to reproduce the rigid behaviour of the floor.
The behaviour of gravity walls and partition walls was also idealized as a pair of diagonal
truss elements with Pinching4 material (Figure 2b). A similar approach as used for the shear
walls without the reference experimental tests was also used to calibrate the partition and
gravity wall model. The main difference between gravity, and partition and the shear wall is
the absence of hold-downs in gravity and partition walls and the presence of hold down in
shear walls as shown in Figure 1. It must be noted that the lateral force contribution provided
by the double layer of impact resistant gypsum boards in counter walls was already lumped in
the model of shear and gravity walls with finishing parts. Therefore, only the studs of internal
counter walls (Figure 2c) were modelled as a truss element with their position in model and,
the physical and mechanical properties being the representative of profiles used in tested
mock-up (60×27×0.6 mm C sections). The end nodes of these studs were linked to the rigid
truss elements of floors and they were constrained by the same rigid Diaphragm used to repro-
duce the rigid behaviour of the floor.
Second floor and roof were modelled following two different approaches. In the first
approach, Model F1 (Figure 3a), the diaphragm was simulated by rigid in-plane X shaped
trusses applied at both floor intrados and extrados. In the second approach, Model F2
(Figure 3b), the X trusses applied at floor intrados was replaced by joist elements. Joists were
modelled as elastic beam column elements pin connected to the floor track elements, with the
same properties of joists used in the mock-up (197×50×10×2.0 mm C back-to-back section).
Where the F1 model was used for modelling the floors, the floor seismic mass was assigned as
lumped mass in the centre of the floor, whereas in the case of the F2 model the floor seismic
mass was applied in the four corner points of the floor. The roof and second floors had a mass
of 1.50 and 2.54 tons, respectively, in B configurations and 3.04 and 5.19 tons, respectively, in
C configuration. A Rayleigh damping ratio of 4% was used based on the experimental results
(Landolfo et al., 2018).

1000
Figure 3. Schematic drawings of modelling used for second floor and roof.
a) Model F1 b) Model F2.

Two classes of models were developed: B models, which simulate the dynamic characteris-
tics of mock up in B Configuration; C models, which represent dynamic characteristics and
seismic response of the building in C Configuration. Models were developed with an increas-
ing degree of detail, ranging from using a simple pair of diagonal trusses for simulating the
lateral stiffness of mock-up, e.g. only shear walls without intermediate studs and finishing
materials, to the use of more elements representing also other minor lateral stiffness contribut-
ing components of the mock-up, e.g. shear, gravity internal counter walls with intermediate
studs and finishing materials. Table 1 highlights the main difference between different models.

4 NUMERICAL VS EXPERIMENTAL RESULTS

Validation of numerical models was done via the comparison with shake-table results, in term
of fundamental vibration period, inter-storey drift peaks and time history. Dynamic identifica-
tion test results using the white noise signal showed that the fundamental period of vibration,
TEX, for the B Configuration was 0.126 s and it was greater than the period for the
C Configuration before earthquake test, which was 0.102 s. Note that the variation of TEX was
not only due to the variation of lateral stiffness, but was also affected by the variation of the
mass. As far as the inter-storey drift peaks during the earthquake tests are concerned, the posi-
tive inter-storey drift peaks were 0.80% for 1st level and 0.52% for 2nd level, whereas the nega-
tive inter-storey drift peaks were 0.35% for 1st level and 0.17% for 2nd level. All the peaks
corresponded to the earthquake with maximum intensity (Landolfo, 2018).
For each Model listed in Table 1, the fundamental period of vibration TM, was evaluated by
modal analysis. In general, numerical results follow the same trend of experiments results, TEX
with higher value of TM for B Configuration than the C Configuration, except for the B1 and

Table 1. Differences among different model types.


Model Shear walls (1) Gravity walls (1) Partition walls Floors Model Internal counter walls
B1 w/o intermediate studs No No F1 No
B2 w/intermediate studs No No F1 No
B3 w/intermediate studs Yes No F1 No
B4 w/intermediate studs Yes No F2 No
B5 w/intermediate studs Yes Yes F2 No
C1 w/o intermediate studs No No F1 No
C2 w/intermediate studs No No F1 No
C3 w/intermediate studs Yes No F1 No
C4 w/intermediate studs Yes No F2 No
C5 w/intermediate studs Yes Yes F2 No
C6 w/intermediate studs Yes Yes F2 Yes

(1) Shear walls and gravity walls with finishing for Models B1 to B5 and without finishing for Models
C1 to C6.

1001
Table 2. Comparison between test and modelling results in terms of fundamental period of vibration.
TM ΔTM TD TEX ΔTEM ΔTED
Model Features (1) [s] [%] [s] [s] [%] [%]
B1 Shear walls w/o intermediate studs, F1 0.143 - - 13 -
Floors
B2 As B1, but shear walls w/intermediate 0.123 17 0.123 3 3
studs 0.126
B3 As B2, plus Gravity walls 0.118 5 - 7 -
B4 As B3, but F2 Floors 0.117 1 - 8 -
B5 As B4, plus Partition walls 0.100 17 0.110 21 13
C1 Shear walls w/o intermediate studs, F1 0.150 - - 47 -
Floors
C2 As C1, but shear walls w/intermediate 0.111 36 - 8 -
studs
C3 As C2, plus Gravity walls 0.095 17 0.100 0.102 7 2
C4 As C3, but F2 Floors 0.093 2 - 9 -
C5 As C4, plus Partition walls 0.088 6 - 14 -
C6 As C5, plus internal counter walls 0.082 8 0.091 21 11

(1) Shear walls and gravity walls with finishing for Models B1 to B6 and without finishing for Models C1
to C7.

C1 Models, which have TM equal to 0.14 s and 0.15 s, respectively. The effect of different com-
ponents on the variation of the fundamental period is evaluated through the ratio ΔTM =
(TM,i+1 - TM,i)/TM,i, where TM,i is the fundamental period of the ith, Model TM,i+1 is the funda-
mental period of the Model i+1th. Results showed that, the highest decreasing of 17% in TM was
observed for the B Configuration (from 0.143 s to 0.123 s) and 36% for the C Configuration
(from 0.150 s to 0.111 s). These variations are due to the presence of intermediate structural
studs, which greatly affect the dynamic behaviour of the building when other components are
not explicitly considered in the model. The explicit modelling of the gravity walls changes the
TM from 0.123 s to 0.118 s (variation of 5%) and from 0.111 s to 0.095 s (variation of 17%), for
B and C Models, respectively. It is important to remember that there were two gravity walls in
the building (one for each storey), placed in the Y direction and they are bare (without finishes)
in B Models, while completed with non-structural finishing in C Models. Since the gravity walls
in B Configuration did not have any finishes, the effect in modelling is not very important. In
C Models, instead, gravity walls with finishing elements should be considered, because their
presence affects results enough. As it is shown, the effect of joists in modelling is negligible (1%
TM decreasing for B Models and 2% TM decreasing for C Models), because in the ELISSA
mock-up diaphragms behaved rigidly, according to experimental results. The partition walls
were placed in the X direction (loading direction) and they altered the dynamic response with
their contribution more significant in B5 Model (decreasing of TM equal to 17%) then in C5
(decreasing of TM equal to 6%). At the end, internal counter walls were added for the C Model
only, as in the mock-up. They affect the fundamental period by 7%, up to 0.082 s. Therefore,
elements which mostly affect the results are intermediate structural studs for both B and
C Models, partition walls for B Models and gravity walls for C Models, as demonstrated by
modal analysis. The experimental fundamental periods of vibration TEX were compared
through the ratio ΔTEM = (TEX-TM)/TM for all the models, as shown in Table 8, which shows
that the models with shear walls and intermediate studs for both B and C Configurations
matches the experimental fundamental period with good accuracy.
Fundamental period of vibration was also evaluated by a numerical dynamic identification
(TD), i.e. by performing the time history analysis under the same white noise signal used in the
experimental activity. TD was equal to 0.123 s and d 0.110 s for the B2 and B5 models, respect-
ively, and 0.100 s and 0.091 s for the C3 and C6 models, respectively. The comparison between
TD and TEX through the ratio ΔTED = (TEX-TD)/TD shows better match between experimental

1002
and numerical results, with ΔTED of about 3% and 14% for B2 and B5 Models, respectively,
and 2% and 11% for C3 and C6 Models, respectively.
To investigate the response of models under the seismic loads, the earthquakes were applied
in sequence to the mock-up with different scaling factors during the experiments. The same
earthquake input sequence was used as input to perform the time history analysis in OpenSees
for the dynamic analysis. The inter-storey drift peaks and displacement time history were
evaluated with the numerical analysis and compared with the experimental results. Compari-
son of the maximum inter-storey drift peaks are shown for the C2 and C7 Models against the
experimental results in Table 3. In C2 Model maximum drift peaks are over-estimated by 27%
and 87% for first and second storey, respectively. Instead, the C7 Model underestimates the
maximum drift peaks by 23% and 2% for first and second storey, respectively and hence is
more accurate. The numerical and experimental inter-storey drift time histories were also com-
pared for the first and second storey (Figures 4 and 5) and it is clear that the C7 Model is able
to capture the experimental response, while the behaviour predicted by C2 Model overesti-
mates the inter storey displacements.

Table 3. Experimental and Numerical maximum drifts.


Storey Experimental Numerical C2 Numerical C7
First 0.80 1.02 0.61
Second 0.52 0.87 0.53

Figure 4. Experimental and Numerical drift time history for first storey.

Figure 5. Experimental and Numerical drift time history for second storey.

1003
5 CONCLUSIONS AND FURTHER DEVELOPMENTS

The paper presents the numerical modelling of a full-scale two-storey CFS building, subjected
to shake-table tests carried out at University of Naples, as a part of the ELISSA project. The
shear walls were modelled as a pair of diagonal truss elements, calibrated on the basis of
experimental results. Subsequently, six models for the bare structure (B models) and seven for
the complete construction (C models) phases of the mock-up were developed, in which the
shear walls, floors, gravity walls, partition walls and internal counter walls were explicitly con-
sidered. The models differed with each other in terms of the different building component
being modelled on them, which were altered in consecutive models to better understand the
contribution of any individual component to the lateral force resistance. The dynamic charac-
teristics of model were investigated via the modal analysis and the time history analysis using
the white noise signal. The models with shear walls and intermediate studs for both B and
C Configurations matched the experimental fundamental period with good accuracy. Add-
itionally, the seismic response of C Models was explored through the dynamic analysis using
the experimental earthquake records, which were applied in a same sequence of the experi-
ments. Comparison of experimental and numerical results in terms of the maximum inter
storey drift peaks and time histories revealed that a model equipped with all the components
i.e. shear walls, gravity walls, partition walls, internal counter walls and joists of the floors
possess the ability to predict the experimental seismic response with the highest accuracy. In
future, the effect of damping ratio on the seismic response would be further investigated, as it
generally tends to decrease with the progress of damage in a building, while in the current
study a constant 4% value is used for it. Moreover, a comparison with the available design
guidelines is necessary in order to introduce a useful CFS modelling techniques for engineers,
which would require minimal amount of experimental data for calibration.

REFERENCES

Cassiano, David; D’Aniello, Mario; Rebelo, C. 2018. Seismic behaviour of gravity load designed flush
end-plate joints. Steel and Composite Structures, 26(5): 621–634.
CEN. 2004. EN 1998-1 Eurocode 8: Design of Structures for earthquake resistance-Part 1: General rules,
seismic actions and rules for buildings. Brussels: European Committee for Standardization.
CEN. 2006. EN 1993- 1-3Eurocode 3: Design of steel structures-Part 1-3: General rules-Supplementary
rules for cold-formed members and sheeting. Brussels: European Committee for Standardization.
Cocoon, Transformer, ETA-11/0105, 2011.
Dell’Aglio, G., Montuori, R., Nastri, E., & Piluso, V. 2017. A critical review of plastic design approaches
for failure mode control of steel moment resisting frames. Ingegneria Sismica, 34(4): 82–102.
Fiorino, L., Shakeel, S., Macillo, V., & Landolfo, R. 2018. Seismic response of CFS shear walls sheathed
with nailed gypsum panels: Numerical modelling. Thin-Walled Structures, 122: 359–370.
Landolfo, R. 2019. Lightweight steel framed systems in seismic areas: Current achievements and future
challenges. Thin-Walled Structures, 140: 114–131.
Landolfo, R., Iuorio, O., & Luigi, F. 2018. Experimental seismic performance evaluation of modular
lightweight steel buildings within the ELISSA project. Earthquake Engineering & Structural Dynamics.
Macillo, V., Fiorino, L., & Landolfo, R. 2017. Seismic response of CFS shear walls sheathed with nailed
gypsum panels: Experimental tests. Thin-Walled Structures, 120(August): 161–171.
Macillo, V., Shakeel, S., Fiorino, L., & Landolfo, R. 2018. Development and Calibration of a Hysteretic
Model for CFS Strap braced stud walls. Advanced Steel Construction, 14(3): 336–359.
Piluso, V., Pisapia, A., Castaldo, P., & Nastri, E. 2019. Probabilistic Theory of Plastic Mechanism Con-
trol for Steel Moment Resisting Frames. Structural Safety, 76: 95–107.
Tartaglia, R., D’Aniello, M., & De Martino, A. 2018a. Ultimate Performance of External End-plate
Bolted Joints Under Column Loss Scenario Accounting for the Influence of the Transverse Beam. The
Open Construction and Building Technology Journal, 12(1): 132–139.
Tartaglia, R., D’Aniello, M., Gianmaria, D.L., & Attilio De Martino, D. M. 2018b. Influence of EC8
rules on P-Delta effects on the design and response of steel MRF. Ingegneria Sismica: International
Journal of Earthquake Engineering, 13(3): 104–120.

1004
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Elastic buckling strength of H-shaped beams subjected to end


moment and uniformly distributed load

Daiki Shinohara & Kikuo Ikarashi


Tokyo Institute of Technology, Tokyo, Japan

ABSTRACT: The objective of this study is to evaluate the elastic buckling strength of
H-shaped steel beams subjected to end bending moments and uniformly distributed load. To
this end, the elastic lateral buckling strength is first calculated under various boundary condi-
tions. Based on this, approximation formulas to determine the moment gradient correction
factor which is generally used for evaluating lateral buckling strength are proposed. We also
consider the local buckling strength by conducting theoretical analysis. The contribution of
this study involves three main points: 1) Approximation formulas of the moment gradient cor-
rection factor are proposed under various boundary conditions. 2) Lateral buckling strength
is particularly affected at positions where the uniformly distributed load acts. 3) The effects of
the moment ratio on local buckling strength are small compared with the case of lateral
buckling.

1 INTRODUCTION

Important factors that influence the lateral buckling strength of H-shaped beams include the
boundary conditions at the ends of the beam and the moment gradient on the beam. The influ-
ence of the moment gradient on the lateral buckling strength is generally evaluated using the
moment gradient correction factor; this has been studied by many researchers. Approximation
formulas of the moment gradient correction factor have been proposed using both simple and
fixed supports at both ends, and the formulas are shown in the current Japanese steel design
standards [1]. Approximation formulas of the moment gradient correction factor under various
boundary conditions based on the elastic lateral buckling strength calculated from the energy
method have also been proposed [2]. In these studies, it was assumed that only the bending
moments acted on the beam; however, in reality, the vertical load due to dead load, live load, etc.,
also acts on the beam. Thus, proposed approximation formulas are insufficient. Approximation
formulas of the moment gradient correction factor have also been proposed when bending
moments and a uniformly distributed load act on the beam [3]. However, there are problems in
that the boundary conditions are insufficient and the accuracy of the approximation formulas
varies depending on the cross-section of the beam.
Studies on the local buckling of H-shaped beams have also been conducted. From
a theoretical analysis by the energy method, the elastic buckling strength of a simply supported
web plate subjected to bending shear stress was calculated, and evaluation methods for the
buckling strength have been proposed [4]. However, studies on the case of uniformly distributed
loads and bending shear are yet to be conducted.
Therefore, this study evaluates the elastic buckling strength when bending moments and
a uniformly distributed load act on H-shaped steel beams. For lateral buckling, approxima-
tion formulas of the moment gradient correction factor are calculated for six types of bound-
ary conditions, and an evaluation method for buckling strength is proposed. For local
buckling, we examine how the magnitude of the bending moment and uniformly distributed
load influences the elastic buckling strength under three kinds of boundary conditions.

1005
2 BUCKLING LENGTH AND MOMENT GRADIENT CORRECTION FACTOR
FOR CALCULATING ELASTIC LATERAL BUCKLING STRENGTH OF
H-SHAPED BEAMS

2.1 Calculation method for elastic lateral buckling strength of H-shaped beams
In this section, a calculation method for the elastic lateral buckling strength by a theoretical
analysis based on the energy method is described. Figure 1 shows the external force and
moment distribution acting on the beam. The moment acting on both the left and right ends
of the beam during buckling is Mcr, (β-1)Mcr. The range of the bending moment gradient β is
0 ≦ β ≦ 2. Moreover, the range of the coefficient α, which represents the magnitude of the
uniformly distributed load, is 0 ≦ α ≦ 2.
As shown in Figure 2, the lateral displacement and the angle of torsion during lateral buckling
are assumed to be μ and φ, respectively, and the displacement function is assumed according to
the boundary conditions at both ends of the beams. For each case with and without restraints at
the left and right ends of the beam, the lateral displacement μ is expressed by the formulas
(a1)–(a3) in Table 1, while the angle of torsion φ is expressed by the formulas (a4)–(a6) in
Table 2. In reality, three types of boundary conditions are considered: “simple supports” without
rotation and warping restraints; “warping restraint” with a warping restraint and without
a rotation restraint; and “fixed support” with rotation and warping restraints. Six combinations
of these boundary conditions are considered in this study, as shown in Table 3. Because it is
common for members to have a large moment at the end with the large restraint, this study does
not consider the case where the restraint of the right end becomes larger than the left end.
Here, Mcr is calculated from the strain energy ΔU of the member and the work ΔT done by
the external force using the displacement function, as shown in the following equations.

ð l "  2 2  2 2  2 #
1 d μ d ’ d’
ΔU ¼ EIy þ EIω þ GJ  dx ð1Þ
2 0 dx2 dx2 dx
ðl   ðl
4α þ β 4α 2 dμ2 qh 2
ΔT ¼ Mcr 1 xþ 2 x ’  dx þ ’  dx ð2Þ
0 l l dx 0 2

Mcr is expressed by Eq. (3) using the buckling coefficient Kcr, the cross-section index R is
expressed by Eq. (4) using the torsional rigidity GJ and warping rigidity EIω [2].

Figure 2. Lateral displacement μ and angle


Figure 1. Analysis model. of torsion φ.

1006
Table 1. Displacement function of lateral displacement μ.
Right Without rotation restraint With rotation restraint
Left

Without rota- P
M mπx
l am sin    ða1Þ
tion restraint m¼1 l

P
M h i P
M h i
With rotation ðm þ 1Þπx
l l  m þ 1 sin
am sin mπx m
l    ða3Þ l  am cos ðm l 1Þπx  cos ðm þl1Þπx    ða2Þ
restraint m¼1 m¼1

Table 2. Displacement function of angle of torsion φ.


Right Without warping restraint With warping restraint
Left
P
N
Without warping restraint l    ða4Þ
bn sin nπx
n¼1
PN h i P
N h i
ðn þ 1Þπx
With warping restraint l  n þ 1 sin
bn sin nπx n
l    ða6Þ bn cos ðn l1Þπx  cos ðn þl1Þπx    ða5Þ
n¼1 n¼1

Table 3. Displacement function by boundary condition.


Boundary condition μ φ

Simple support at both ends Eq.(a1) Eq.(a4)


Fixed support at both ends Eq.(a2) Eq.(a5)
Warping restraint at both ends Eq.(a1) Eq.(a5)
Left end fixed support – Eq.(a3) Eq.(a6)
Right end simple support
Left end fixed support – Eq.(a3) Eq.(a5)
Right end warping restraint
Left end warping restraint – Eq.(a1) Eq.(a5)
Right end simple support

sffiffiffiffiffi
π2 EIy Iω
Mcr ¼ Kcr 2 ð3Þ
l Iy

GJl 2
R¼ ð4Þ
π2 EIω

An eigenvalue analysis is performed to obtain Kcr by equating Eq. (1) and Eq. (2) with Eq. (3)
and Eq. (4). In this study, Mcr is calculated by substituting the obtained Kcr into Eq. (3).
In addition, in Japanese design guidelines, the elastic lateral buckling strength Mcr of
H-shaped beams is expressed by Eq. (5) [5]. Where C denotes the moment gradient correction
factor, and ku and kβ denote the buckling length coefficient for weak axial bending and bend-
ing twist, respectively. Kcr is expressed by the following equation using Eq. (3) to Eq. (5).
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u (  2 )
π2 EIy utIω ku GJ ðku l Þ2
Mcr ¼ C þ 2 ð5Þ
ðku l Þ2 Iy kβ π EIω
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 2 ffi
C ku
Kcr ¼ 2 þ ku R 2 ð6Þ
ku kβ

1007
2.2 Influence of boundary conditions on elastic lateral buckling strength of H-shaped beams
subjected to uniform bending
In this section, the influence of boundary conditions on the lateral buckling strength when
uniform bending (β = 0) is applied at the end of the beam is described. The buckling coeffi-
cient Kcr0 when the beam is subjected to this uniform bending is expressed by the following
equation with C = 1 in Eq. (6).
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 2 ffi
1 ku
Kcr0 ¼ 2 þ ku R 2 ð7Þ
ku kβ

Figure 3 shows plots of Kcr0 obtained by analyzing of the beam with warping restraint at both
ends. Considering the practical limits, the range of R is taken to be 0 < R ≦ 40 in this study.
These plots are evaluated by Eq. (7), and the coefficients ku and kβ are determined such that
the error calculated by the least squares method is minimal. The values of coefficients ku and
kβ under various conditions are determined and the results are listed in Table 4.

2.3 Approximate formulas of moment correction factor under various boundary conditions
The elastic lateral buckling strength of beams subjected to bending (except pure bending) is
determined by calculating the moment gradient correction factor C. The buckling coefficients
Kcr0 and Kcr in the case of β = 0 and β > 0, respectively, are obtained, and the ratio of these
coefficients with respect to C shown in Eq. (8). Furthermore, when the absolute value of the
maximum moment in the span Mmax is larger than the absolute value of the left end
moment Ml, C is corrected based on Mmax and is calculated by Eq. (9).
Kcr
C ¼ C1 ¼ ðMmax  Ml Þ ð8Þ
Kcr0
Mmax
C ¼ C1  ðMmax 4Ml Þ ð9Þ
Ml

First, C is calculated by only considering bending moments on both ends of the beam and with-
out considering the uniformly distributed load (α = 0). C is thus approximated by Eq. (10).

Figure 3. Relationship between Kcr0 and R (Warping restraint at both ends).

Table 4. Buckling length coefficient suggested value.


Boundary condition ku kβ Error from Kcr0 [%]

Simple support at both ends 1.00 1.00 0.0


Fixed support at both ends 0.500 0.500 0.0
Warping restraint at both ends 0.933 0.463 3.2
Left end fixed support – Right end simple support 0.696 0.696 0.0
Left end fixed support – Right end warping restraint 0.659 0.464 4.5
Left end warping restraint – Right end simple support 0.957 0.638 2.3

1008
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
  ffi
1 þ a4 β þ a5 β2 þ a6 β3 =ð1 þ a7 RÞ
C¼ ð10Þ
1 þ a1 β þ a2 β2 þ a3 β3

Table 5 lists the coefficients a1 to a7 of Eq. (10) obtained by the least squares method. C is
evaluated based on each boundary condition and the influence of R. Table 5 also shows the
difference between the value of C obtained from Eq. (10) and that obtained by the analysis. In
this case, the error is found to be less than 7%; hence, the accuracy is assumed to be good.
Next, C is calculated by considering a uniformly distributed load in addition to the bending
moments. First, it is assumed that the uniformly distributed load acts on the shear center of the
beam. This is investigated without considering the influence of twist. In this case, Reference [3]
proposed an approximation formula for C in the case where the beam was simply supported at
both ends, as shown Eq. (11). The equation does not include a factor indicating the cross-section
index. Therefore, the influence of the cross-section index R is evaluated, as shown Eq. (12).

1 1
C ¼ pffiffiffiffi ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð11Þ
B 0:283ð1  βÞ2 þ 0:434ð1  αÞð1  βÞ þ ð0:283  0:869α þ 0:780α2 Þ
1
C¼ pffiffiffi
ξ
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi


1þða14 αþa15 α2 þa16 α3 Þ=ð1þa20 RÞþ a4 f1þða17 αþa18 α2 Þ=ð1þa21 RÞgβþa5 f1þa19 α=ð1þa22 RÞgβ2 þa6 β3 =ð1þa7 RÞ
¼
ð1þα8 αþa9 α2 þa10 α3 Þþa1 ð1þa11 αþa12 α2 Þβþa2 ð1þa13 αÞβ2 þa3 β3
ð12Þ
The values of coefficients a1 to a7 listed in Table 5 are used. The coefficients a8 to a22 are
determined by the least squares method and shown in Table 6. Figure 4 shows the results of
the analysis together with Eq. (11) and Eq. (12).
Although the errors are described later, the proposed approximation formula shows
roughly good accuracy even though there are variations depending on the boundary condi-
tions and α. Moreover, the influence of R is evaluated.
Next, the approximation formula in the case where the beam is subjected to a uniformly
distributed load on positions except the shearing center is calculated. In this case, C proposed
in Reference [3] is given in Eq. (13). B in Eq. (13) is the one in Eq. (11), and S is expressed by
Eq. (14) using the distance h between the point of application of the uniformly distributed
load and the shear center, and bridge bw. In this study, C is evaluated by Eq. (15).
r
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2 ffi
4α pffiffiffiffiffiffiffi 4α pffiffiffiffiffiffiffi
S
π 2 1þR
þ π2 1þR
S
þB
C¼ ð13Þ
B

Table 5. Suggested value of a1 to a7.


Boundary condition a1 a2 a3 a4 a5 a6 a7 Error from C

Simple support at both ends −0.92 0.25 −0.0015 −0.0077 0.28 −0.12 0.22 1.9
Fixed support at both ends −0.90 0.21 0.020 0.11 0.00077 0.018 0.080 1.7
Warping restraint at both ends 0.89 0.20 0.013 0.17 −0.074 0.063 0.095 2.3
Left end fixed support – Right −1.1 0.29 0.048 0.031 0.52 −0.25 0.12 4.1
end simple support
Left end fixed support – Right −1.1 0.34 0.018 0.13 −0.26 0.14 0.21 6.5
end warping restraint
Left end warping restraint – −0.78 0.044 0.065 0.038 1.0 −0.52 0.15 4.8
Right end simple support
Units [%]

1009
Table 6. Suggested value of a8 to a22.
Boundary condition a8 a9 a10 a11 a12 a13 a14 a15 a16 a17 a18 a19 a20 a21 a22

Simple support at both ends 5.54 −11.7 5.54 6.69 −6.59 7.67 7.78 −1.17 0.453 −238 47.0 −4.07 0.00501 0.231 0.00782

Fixed support at both ends 1.92 −7.04 4.26 3.11 −4.37 5.00 4.66 −0.0116 −0.00500 6.51 −2.22 −871 0.0163 0.714 0.0948
Warping restraint at both ends 13.8 −27.1 12.7 15.9 −15.2 19.6 17.4 −2.64 1.10 15.2 −5.76 27.3 0.00944 0.391 0.0182

Left end fixed support –


1.82 −5.42 2.87 3.43 −3.55 6.47 3.48 −0.758 0.428 104 −28.4 −2.58 0.00558 0.142 0.305
Right end simple support
Left end fixed support –
2.69 −7.72 4.18 3.78 −4.38 5.41 5.30 −1.40 0.560 19.8 −4.40 4.49 0.0120 −0.0203 0.0449
Right end warping restrain

Left end warping restraint –


1.01 −3.98 2.13 3.37 −3.57 26.9 3.73 −1.36 0.406 −2.28 10.2 −0.395 0.0160 −0.0236 0.0576
Right end simple support

Figure 4. Relationship between C and β.

2h
S¼ ð14Þ
bw
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u0 12
u
u
a23 qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
αS
þ t@a23 qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
αS A þξ
ðku =kβ Þ þku2 R ðku =kβ Þ þku2 R
2 2

C¼ ð15Þ
ξ

The analysis results are shown in Figures 5 and 6, and Table 8 shows some of the errors in the
value of C in each boundary condition. The position at which the uniformly distributed load
acts has a great influence on the buckling strength. If the point of action is on the upper flange

Table 7. Suggested value of a23.


a23
Boundary condition S<0 S>0

Simple support at both ends 0.392 0.307


Fixed support at both ends 0.303 0.175
Warping restraint at both ends 0.338 0.245
Left end fixed support – Right end simple support 0.227 0.135
Left end fixed support – Right end warping restraint 0.331 0.188
Left end warping restraint – Right end simple support 0.246 0.169

1010
Figure 5. Relationship between C and β (Simple support at both ends).

Figure 6. Relationship between C and β (Left end warping restraint – Right end simple support).

(S is negative), then the strength is lower; if it is on the lower flange (S is positive), the strength
is higher. Although there are cases where the error from the approximation formula is signifi-
cant, as the evaluation was conducted on the relatively safe side, there is a trade-off in terms
of accuracy.
The approximate evaluation method for the elastic lateral buckling strength of H-shaped
beams under various boundary conditions can be summarized as follows.
First, determine the R of the beam to be studied from Eq. (4). Next, substitute ku and kβ and
the moment gradient correction factor C into Eq. (6) to obtain Kcr. Then, substitute the obtained
Kcr into Eq. (3) and evaluate the elastic lateral buckling strength.

3 ELASTIC LOCAL BUCKLING STRENGTH OF H-SHAPED BEAMS

3.1 Elastic local buckling strength estimation method of H-shaped beams


In this section, the calculation method for elastic local buckling strength by the theoretical
analysis based on energy method is described. The analysis model was assumed to be similar
to that in Chapter 2. Figure 7 shows the bending shear stress distribution of the web and
flange. Figure 8 shows the state of deformation of the beam during local buckling.

Table 8. Error of C between analysis value and approximate formula.


α=0 α = 0.5 α = 1.0 α = 1.5 α = 2.0

Boundary condition S=0 S = −1.0 S=0 S = 1.0 S = −1.0 S=0 S = 1.0 S = −1.0 S=0 S = 1.0 S = −1.0 S=0 S = 1.0

Simple support at both ends 1.9 7.0 7.5 12.0 5.8 7.8 37.0 6.2 11.3 21.8 7.6 13.6 24.9
Fixed support at both ends 1.7 10.6 11.1 17.7 7.8 17.4 40.0 10.4 12.4 22.8 11.9 16.4 27.6
Warping restraint at both ends 2.3 10.2 11.0 12.9 9.1 12.1 30.0 11.9 13.2 19.3 12.5 16.0 22.4
Left end fixed support – 4.1
10.4 19.3 33.0 12.5 13.8 22.1 27.7 24.3 33.2 27.7 33.7 42.7
Right end simple support
Left end fixed support – 6.5
11.0 12.0 21.4 6.8 24.0 41.8 10.8 14.9 25.0 12.3 16.0 25.4
Right end warping restrain
Left end warping restraint – 4.8
10.6 21.1 29.9 10.8 14.5 21.3 14.5 15.7 19.7 24.7 26.2 25.7
Right end simple support

Units [%]

1011
Figure 8. Deformation during
Figure 7. Stress distribution of plate elements. local buckling

The variables in Figure 7 and displacement functions are defined as follows.


  
4α 2 ð4α þ βÞ 2y
wðx; yÞ ¼  2 x þ x1 1 ð16Þ
l l bw
 
4α ð4α þ βÞ
f ðxÞ ¼  2 x2 þ x1 ð17Þ
l l
   
1 Af 8α 4α β
γðxÞ ¼ þ  xþ þ1 ð18Þ
6 Aw βl β λw
mπx
μm ¼ sin ðSimple support at both endsÞ ð19Þ
l
ðm  1Þπx ðm þ 1Þπx
μm ¼ cos  cos ðFixed support at both endsÞ ð20Þ
l l
mπx m ðm þ 1Þπx
μm ¼ sin  sin ðLeft end fixed support  Right end warping restrainÞ
l mþ1 l
ð21Þ
X X nπy
W¼ a μ sin
n mn m
ð22Þ
m bw
X X z
F1 ¼ a μ nπcosnπ sin
n mn m
ð23Þ
m bw
X X z
F2 ¼ a μ nπ sin
n mn m
ð24Þ
m bw

The strain energy increment ΔU1 in the web and the strain energy increment ΔU2, ΔU3 in the
flange are as follows.
ð l ð bw ( 2 2  2 2  2 2 )
1 ∂ W ∂ W ∂2 W ∂2 W ∂W
ΔU1 ¼ Dw þ þ 2 þ 2ð1   Þ dxdy ð25Þ
2 0 0 ∂x2 ∂y2 ∂x2 ∂y2 ∂x∂y

ð l ð bf ( 2  2 2 )
1 ∂2 F2 ∂ F2
ΔU2 ¼ Df þ 2ð1   Þ dxdz ð26Þ
2 0 bf ∂x2 ∂x∂z

ð l ð bf ( 2  2 2 )
1 ∂2 F1 ∂ F1
ΔU3 ¼ Df þ 2ð1   Þ dxdz ð27Þ
2 0 bf ∂x2 ∂x∂z

1012
Figure 9. Relationship between kσ and β.

The work of external force on the web ΔT1 and the work on the flange ΔT2 are as follows.

ð l ð b w (    )
1 4α þ β 4α 2y ∂W 2 ∂W ∂W
ΔT1 ¼ σcr tw 1 x þ 2 x2 1 þ 2γðxÞ dxdy ð28Þ
2 0 0 l l bw ∂x ∂x ∂y

ðl bðf ( ( 2   2 ))
1 4α þ β 4α ∂F1 ∂F2
ΔT2 ¼ σcr tf 1 x þ 2 x2  dxdz ð29Þ
2 l l ∂x ∂x
0 bf

The buckling conditional equation by the energy method is expressed by Eq. (30). This is
a simultaneous linear equation with respect to amn, and it is calculated as an eigenvalue problem.
Elastic buckling coefficient kσ can be calculated from the elastic buckling stress σcr using Eq. (31).

∂ðΔU1 þ ΔU2 þ ΔU3  ΔT1  ΔT2 Þ=∂amn ¼ 0 ðm ¼ 1; 2; . . . ; M; n ¼ 1; 2; . . . ; N Þ ð30Þ


  
12 1   2 bw 2
kσ ¼ σcr ð31Þ
π2 E tw

3.2 Analysis result


In this section, the elastic local buckling strength results calculated using the analysis method
from Section 1 are shown. Figure 9 shows an example of the results obtained from the ana-
lysis. The beam is the same as that used in Chapter 2, and there are three types of boundary
conditions. From this result, it is considered that the moment gradient β does not have
a considerable effect on the elastic buckling strength compared with elastic lateral buckling
strength. The increase in the strength caused by uniformly distributed load is also small com-
pared with the case of lateral buckling.
The elastic local buckling was investigated with one type of beam; however, detailed investi-
gation of other beams is given in other papers.

4 CONCLUSION

This study proposed methods to evaluate the elastic lateral buckling strength of H-shaped
beams subjected to bending moments and a uniformly distributed load. It also considered the
elastic local buckling strength. The following conclusions were drawn.

1013
1. Displacement functions under various boundary conditions were proposed, and calculation
methods for the elastic lateral buckling strength were obtained from a theoretical analysis
by the energy method.
2. Approximate equations for the buckling length coefficients and the moment gradient cor-
rection factor at each boundary condition were proposed. In order to evaluate the influence
of the cross-section on the buckling strength, an approximate expression considering the
cross-section index R was proposed.
3. The elastic lateral buckling strength was particularly affected by twisting owing to the pos-
ition at which the uniformly distributed load acted, and the more the point of action is on
the upper part of the beam, the more the strength tends to be disadvantageous.
4. A displacement function under various boundary conditions was proposed, and calculation
methods for the elastic local buckling strength were obtained from a theoretical analysis by
the energy method.
5. The influence of the moment gradient on the local buckling strength is small compared
with the case of lateral buckling.

NOTATION

α Coefficient indicating the magnitude of uniformly distributed load


β Moment gradient
l Length of H-shaped beam
h Distance from the shearing center of the beam to the point of application of
uniformly distributed load (negative on the upper flange side)
E Young’s modulus
Iω, J Bending torsion constant, Torsion constant
EIy Moment of inertia around weak axis
EIω, GJ Warping rigidity, Torsion constant
ku Buckling length coefficient for weak axial bending
kβ Buckling length coefficient for warping moment
Mmax Maximum moment in beam span
Ml Left end moment of beam
bw Width of web
bf Half the width of flange
tw Thickness of flange
tf Thickness of flange
ν Poisson's Ratio (=0.3)
Dw Bending Stiffness of web
Df Bending Stiffness of flange
Aw Cross sectional area of web
Af Cross sectional area of flange

REFERENCES

[1] Salvadori, M.G. 1956. Lateral Buckling of Eccentrically Loaded I-Columns. In Italian. Trans.
ASCE, Vol.121: pp.1163–1179.
[2] Ikarashi K., Tomo N., Wang T. 2011. Effects of Boundary Conditions and End Moment Ratio on
Elastic Lateral Buckling Strength of H-Shaped Beams. In Japanese. Journal of Structural and Con-
struction Engineering, Vol.670: pp.2173–2181.
[3] Nakamura T., Wakabayashi T. 1978. Approximate Solution to Correction Factor C of Elastic Lat-
eral Buckling Moment of H-Shaped Beams - Design Formula. In Japanese. Summaries of technical
papers of annual meeting, arch. Inst. of Japan: pp.1319–1320.
[4] Ikarashi K. 2003. Buckling Strength of Simply Supported Web Plate under The Action of Bending
Shear Stress. In Japanese. Journal of Structural and Construction Engineering, Vol.565: pp.135–141.
[5] Architectural Institute of Japan. 1980. Recommendation for Stability of Steel Structures.

1014
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

On the incorporation of cross-section restraints in Generalised


Beam Theory (GBT)

T.G. da Silva & C. Basaglia


School of Civil Engineering, Architecture and Urban Design, University of Campinas, Brazil

D. Camotim
CERIS, DECivil, Instituto Superior Técnico, University of Lisbon, Portugal

ABSTRACT: This paper reports the results of an investigation on the development, imple-
mentation and application of a Generalised Beam Theory (GBT) formulation for the local,
distortional and global buckling analysis of thin-walled restrained members, based on a cross-
section analysis procedure that incorporates elastic restraints and, therefore, requires less
deformation modes to obtain accurate analytical and numerical buckling solutions. Its cap-
abilities are illustrated by presenting and discussing numerical results concerning the buckling
behaviour of cold-formed steel studs braced by sheathing and purlins restrained by sheeting.
For validation and assessment purposes, some of these results are compared with values
yielded by the GBT-based code GBTUL2.0 and/or ANSYS shell finite element analyses.

1 INTRODUCTION

In Generalised Beam Theory (GBT), which incorporates genuine plate theory concepts
(thus accounting for cross-section in-plane and out-of-plane deformations), the cross-
section displacement field is expressed as a linear combination of deformation modes,
making is possible to write the equilibrium equations and boundary conditions in an
unique and very convenient format (e.g., Basaglia et al. 2015). The performance of a GBT
structural analysis involves (i) a cross-section analysis (determination of the deformation
modes and evaluation of the associated modal mechanical properties) and (ii) a member
analysis (modal solution of the structural problem under consideration).
In practical applications, thin-walled members are often continuously restrained along their
lengths – e.g., purlin-sheeting system, steel-concrete composite members or sheathed wall stud
systems. In order to incorporate such (elastic) restraints in a GBT analysis, two approaches
can be followed: (i) incorporate the restraints in the cross-section analysis, taking them into
account at the deformation mode determination stage (Schardt 1989 and Jiang & Davies 1997
adopted this approach for global and very specific distortional deformations) or (ii) include
the restraints only in the member analysis, as constraint equations, which means that the
deformation modes are calculated without considering the restraints (Camotim et al. 2008,
Basaglia et al. 2013 and Bebiano et al. 2018, authors of the code GBTUL2.0, adopted this
approach for arbitrary deformation patterns, which amounts to combining the conventional
deformation modes at the member analysis stage. This last approach requires larger deform-
ation mode sets and it may be argued that it somewhat “clouds” the structural interpretation
of the results.
The aim of this work to present and illustrate the implementation and application of
a novel GBT formulation for the local, distortional and global buckling analysis of thin-
walled restrained members, based on a cross-section analysis procedure that incorporates elas-
tic restraints and, therefore, is capable of providing accurate buckling results with only a few
deformation modes (often a single one). Moreover, it is also shown that the proposed

1015
formulation enables the development of analytical formulae to calculate critical buckling
loads of restrained members. Its capabilities are illustrated through the presentation and dis-
cussion of numerical results concerning the buckling behaviour of cold-formed steel (i) studs
braced by sheathing and (ii) purlins restrained by sheeting. For validation and assessment pur-
poses, some of these results are compared with values yielded by the code GBTUL2.0 (Bebiano
et al. 2018) and/or ANSYS (SAS 2013) shell finite element analyses.

2 GBT FORMULATION

The modelling of the restraint provided by the sheeting, wall or slab to the member involves
continuous translational and rotational elastic springs, which restrain the member transverse
displacements and mid-width rotations. Figure 1(a) illustrates the case of a purlin-sheeting
assembly: to simulate the restraint provided by the roof sheeting, continuous translational and
rotational elastic springs (stiffness KT and KR, respectively) are continuously attached to the
upper flange mid-points along the whole purlin length.
Consider the prismatic thin-walled open cross-section member, formed by n distinct plate/wall
elements, and restrained by continuous translational and rotational elastic springs, located,
respectively, at points PT and PR indicated in Figure 1(b), where x, s and z are local coordinates
along the member axis, cross-section mid-line and wall thickness, thus leading to member mid-
surface displacement components u(x,s), v(x,s) and w(x,s) expressed as

uðx; sÞ ¼ uk ðsÞ’k;x ðxÞ vðx; sÞ ¼ vk ðsÞ’k ðxÞ wðx; sÞ ¼ wk ðsÞ’k ðxÞ ð1Þ

where (.),x ≡ d(.)/dx, the summation convention applies to subscript k, functions uk(s), vk(s),
wk(s) are mid-line “displacement profiles” and φk(x) is a dimensionless amplitude function
defined along the member length – information on the derivation of these expressions can be
found in Silvestre & Camotim (2002).
The equations providing the member buckling behaviour, taking into account the presence
of elastic constraints (springs), are obtained by imposing the stationarity of the total potential
energy functional
ð ð
1 1
V¼ σij εij dO þ KD2r dx ð2Þ
2 2
O L

in the close vicinity of the member fundamental equilibrium path (adjacent equilibrium) – (i)
Ω is the member volume (n walls), (ii) σij and εij are the second Piola-Kirchhoff stress and
Green-Lagrange strain tensors, respectively, both comprising pre-buckling and bifurcated
components, (iii) L is the member length, (iv) K is the stiffness of a continuous (along
a longitudinal axis r) spring, (v) Δ are spring generalised displacements (translation or

Figure 1. Prismatic thin-walled member (a) with continuous elastic restraints/springs and (b) local
coordinate axes.

1016
rotation) and (vi) the summation convention applies to subscripts i and j. Then, the equilib-
rium equations defining the member buckling eigenvalue problem are obtained by (i) linearis-
ing the first variation (δ) of the total potential energy functional, at the fundamental
equilibrium path, and (ii) discarding the pre-buckling strains (e.g., Camotim et al. 2008), thus
yielding
ð

δV ¼ Cik ’k;xx δ’i;xx þ D1ik ’k;x δ’i;x þ D2ik ’k δ’i;xx
L ð3Þ

σx
þ D2ki ’k;xx δ’i þ Bik ’k δ’i þ λWj0 Xjik ’k;x δ’i;x dx ¼ 0

Wj0 ¼ Cjj ’0j;xx ð4Þ

with
ð ð
E
Cik ¼ E t ui uk ds þ t3 wi wk ds ð5Þ
12ð1  2 Þ
b b
ð
E
Bik ¼ t3 wi;ss wk;ss ds þ KT wi ðsPT Þwj ðsPT Þ þ KR wi;s ðsPT Þwj;s ðsPT Þ ð6Þ
12ð1  2 Þ
b
ð ð
G 3 E
D1ik ¼ t wi;s wk;s ds D2ik ¼ t3 wi wk;ss ds ð7Þ
3 12ð1  2 Þ
b b
ð
σx E
Xjik ¼ t uj ðvi vk þ wi wk Þds ð8Þ
Cjj
b

where (i) E, υ and G are the material Young’s modulus, Poisson’s ratio and shear modulus, (ii)
Wj0 are stress resultant profiles and (iii) λ is the load parameter.
The determination of the GBT deformation modes and evaluation of the corresponding
cross-section modal mechanical properties require the performance of a cross-section ana-
lysis – in open-section members it involves a sequential procedure comprising the following
major steps:
(i) Cross-section discretisation into n + 1 natural nodes (ends of the n walls forming the cross-
section) and m intermediate nodes (located within the walls). In open sections with branch-
ing natural nodes (nodes shared by more than two walls), the natural nodes must be still
subdivided into independent and dependent (several subdivisions are possible). Since one
must (i1) satisfy Vlasov’s null membrane shear strain assumption and (i2) ensure mem-
brane transverse displacement compatibility at all branching nodes, the dependent node
warping displacements cannot be chosen arbitrarily and must be appropriately selected
(Silvestre & Camotim 2002, Dinis et al. 2006).
(ii) Determination of the initial shape functions ui(s), vi(s) and wi(s), by sequentially imposing
unit warping displacements (u = 1) at each independent natural node and flexural displace-
ments (w = 1) at each intermediate node – concerning the imposition of unit displacements,
the cross-section end nodes are treated as both natural (independent or dependent) and
intermediate. Note that evaluating the flexural functions wi(s) involves solving a statically
indeterminate folded-plate problem, a task carried out here by means of the displacement
method – Figure 2 shows the geometry and a possible GBT discretisation of a lipped chan-
nel cross-section.

1017
Figure 2. Geometry and possible GBT discretisation of a lipped channel.

(iii) Calculation of tensors (5) to (8), on the basis of the initial shape functions and applied
loading. One obtains fully populated matrices (highly coupled equilibrium equations)
whose components exhibit no obvious structural meaning.
(iv) To uncouple the member equilibrium equation system as much as possible and, at the
same time, have stiffness matrix components with clear structural meanings, the simul-
taneous diagonalisation of the linear stiffness matrices Cik and Bik, given in (5)-(6), is per-
formed. This leads to the cross-section deformation modes (i.e., the final shape functions
uk(s), vk(s), wk(s)) and to the associated cross-section modal mechanical properties – the
new linear and geometrical stiffness matrix components, several of them with a clear/illu-
minating structural meaning. This procedure (the GBT “trademark”) makes it possible
to express the equilibrium equations in modal form, thus leading to a considerable
amount of interpretation and numerical implementation advantages.
Unlike the above conventional GBT cross-section analysis (for unrestrained cross-sections),
the determination of the GBT deformation modes (final base functions) is carried out through
a procedure involving one to three auxiliary eigenvalue problems, which are addressed separ-
ately next:
(i) Stage 1: Determination of the Distortional and Local Deformation Modes. Consider the aux-
iliary eigenvalue problem defined by
ð½Bik
 λk ½Cik
Þf
ak g ¼ 0; ð9Þ

which has n + m + 1 eigenvalues λk. Concerning this eigenvalue problem, it should be


pointed out that:
(i1) The null eigenvalues λk = 0 correspond to a subspace associated with cross-section rigid-
body motions. Since every vector in this subspace is an eigenvector, no rigid-body
deformation modes can be identified – this identification will be made in next stages.
(i2) The positive eigenvalues λk > 0 correspond to deformation modes involving either wall
transverse bending, warping displacements and fold-line motions (distortional deform-
ation modes) or only wall transverse bending (local deformation modes).
(i3) The base function change regarding the distortional and local deformation modes is
defined by matrix [A  1 ], formed by the (often normalised) eigenvectors with positive
eigenvalues.
(i4) If λ4 = 0, the diagonalisation of Stage 2 must be performed, to obtain the torsion
mode.
(i5) If λ4 > 0, the Stage 2 diagonalisation no longer has to be performed, since the restraint
precludes the occurrence of the torsion deformation mode – one then proceeds to Stage
3, in order to obtain the extension deformation modes, which are associated with λ1
and, if λ2 > 0 and/or λ3 > 0, also with the bending deformation modes.
(ii) Stage 2: Determination of the Torsion Deformation Mode. This stage involves sub-matrices
of [Cik] and [Dik], already expressed in terms of the eigenvector space base functions
obtained from the solution of (9) and denoted [C  1 ] and [D
 1 ] (similar procedures are car-
ik ik
 1 ] and [X
ried out for [B  1 ]). They are obtained as
ik 1ik

 1
¼ ½Y
½C 
T ½Cik
½ Y

 1
¼ ½Y
½D 
T ½Dik
½Y
 1
ð10Þ
ik 1 1 ik 1

1018
where [Y 1 ] is a sub-matrix formed by the null eigenvectors of (9). These two sub-matrices
are associated with the rigid-body (global) deformation modes, which means that [C  1 ] is
ik
fully populated (not diagonal). Then, a second auxiliary eigenvalue problem has to be con-
sidered, which is defined by
 
ð½D  1
Þ a1 ¼ 0
 1
 λk ½C ð11Þ
ik ik k

Concerning this eigenvalue problem, it should be noted that:


(ii1) The eigenvectors associated with the null eigenvalues defines a subspace associated
with cross-section rigid-body translations (no rotation) – axial extension mode and, if
λ2 > 0 and/or λ3 > 0, also the bending deformation modes.
(ii2) The eigenvalue λ4 > 0 corresponds to a cross-section in-plane rigid-body rotation (tor-
sion mode).
(ii3) The base function changes regarding the distortional and local deformation modes,
determined in Stage 1, and (iii2) the torsion mode, determined in Stage 2, are defined
by sub-matrix [A 2 ], obtained from [A
 1 ] by adding the torsion mode eigenvector, nor-
malised to exhibit a unit rotation.
(iii) Stage 3: Determination of the Axial Extension and Bending Deformation Modes. This
stage deals with sub-matrices of [Cik] and [X1ik], denoted [C  2 ] and [X
 2 ], if λ4 > 0 in
ik 1ik
 
Stage 1, or [Cik ] and [X1ik ], if λ4=0 in Stage 1. These sub-matrices are obtained from
1 1

 2
¼ ½Y
½C 
T ½Cik
½Y

 2
¼ ½Y
½X 
T ½X1ik
½Y

if λ4 40 in Stage 1 ð12Þ
ik 1 1 1ik 1 1

 2
¼ ½Y
½C  1
½ Y
 2
T ½C 
 2
¼ ½Y
½X  2
T ½X
 1
½ Y
2
if λ4 ¼ 0 in Stage 1 ð13Þ
ik ik 2 1ik 1ik

where [Y  2 ] is a sub-matrix formed by the null eigenvectors of (11). Since these sub-
matrices are associated with the rigid-body (global) translation deformation modes, [C 2]
ik
is fully populated (not diagonal). Then, a third auxiliary eigenvalue problem is required,
defined by
 2
ð½X  2
Þ a
 2
 λk ½C k ¼ 0 ð14Þ
1ik ik

Concerning this eigenvalue problem, it should be noted that:


(iii1) The null eigenvalue λ1 = 0 correspond to the axial extension deformation mode.
(iii2) The positive eigenvalues correspond to the bending deformation modes.
(iii3) The base function changes concerning the distortional, local and/or torsion deform-
ation modes, determined previously, and the axial extension and bending modes, just
determined, are defined by matrix [A  3 ], obtained from [A
 1 ] or [A
 2 ] by adding these
eigenvectors.
Finally, matrices [
uik ], [vik ] and [
wik ], containing the various cross-section normalised deform-
ation mode component functions, are given by

 3
T ½uik
½A
½uik
¼ ½A  3
½vik
¼ ½A
 3
T ½vik
½A
 3
½  3
T ½wik
½A
wik
¼ ½A  3
ð15Þ

After performing the cross-section analysis, one obtains the member GBT system of adjacent
equilibrium equations, which (i) is expressed in modal form as

 ik ’
C  ik ’
k;xxxx  D  ik ’
k;xx þ B k  λ X σx W 0’
 ¼0 ð16Þ
jik j k;x
;x

1019
and, together with the adequate end support conditions, (ii) defines the buckling eigenvalue
problem to be solved. In order to be able to carry out this task for members with arbitrary
support, it is necessary to use a GBT-based beam finite element (e.g., that developed by Silves-
tre & Camotim 2003).

3 NUMERICAL RESULTS

This section presents and discusses numerical results concerning the buckling behaviour of
cold-formed steel (E = 210GPa and υ = 0.3) purlins restrained by sheeting and studs braced by
sheathing. Figure 3(a) shows the dimensions of the two lipped channel cross-sections dealt
with in this work (Sections A and B), which are continuously restrained by different spring
arrangements (Constraintss I to IV), involving rotational and translational springs with stiff-
ness KR and KT, repectively (values given in Figure 3(a)) – four combinations of cross-section
dimensions and spring arrangement are considered. For each of them, Figure 3(b) displays the
8 most relevant deformation modes obtained by means of the proposed restrained cross-
section analysis procedure (mode 1 stands for axial extension) – for comparison purposes,

Figure 3. (a) Restrained lipped channel cross-sections and (b) 8 most relevant deformation modes for
each of them.

1020
Figure 4. Main features of the most relevant conventional lipped channel deformation modes yielded
by GBTUL2.0.

Figure 4 depicts the first 9 (conventional) deformation modes yielded by code GBTUL2.0 for
an unrestrained lipped channel cross-section. Table 1 show the components of the [C  ik ] (diag-
 
onal), [Bik ] (diagonal) and [Dik ] (almost diagonal) stiffness matrices concerning deformations
modes 2 to 9 of (i) Section A, unrestrained and with Constraint I, and (ii) Section B, unre-
strained and with Constraint III. The observation of the results presented in these two figures
and table prompts the following remarks:
(i) While Constraint I does not restrain major and minor axis bending (modes 2 and 3), Con-
straint III only does not restrain mode 2 – thus, C  22 and C
 33 are equal for the sections
unrestrained and with Constraint I, and C  22 are equal for the sections unrestrained and
with Constraint III.
(ii) Matrix D  ik tends to be “less diagonal” as the number of restraints increase – indeed,

matrix Dik is nearly fully populated for Constraint III and still “almost diagonal” for Con-
straint I.
Table 2 shows the critical buckling loads of uniformly compressed continuously restrained
simply supported steel studs exhibiting Section A – the restraint provided by the sheeting to
the studs is modelled through Constraint II. The results are obtained by means of GBT ana-
lyses with unrestrained (GBTUL2.0) and restrained deformation modes – for validation pur-
poses, some values yielded by ANSYS SFEA are also presented. Comparing the two sets of
buckling results leads to the following conclusions:
(i) First of all, the critical loads yielded by the ANSYS SFEA and the two GBT analyses
(including all deformation modes) are virtually coincident.
(ii) The GBT analyses with restrained deformation modes provide accurate buckling results
with only a single deformation mode – mode 7 for studs with lengths up to 250cm (local
buckling) and mode 2 for longer studs (minor-axis flexural buckling). Naturally, the con-
ventional GBTUL2.0 analyses must include more than one deformation mode to provide
accurate buckling results – considering only the dominant mode leads to errors that may
reach 19% (for short studs).
Attention is turned next to simply supported steel purlins restrained by steel sheeting and sub-
jected to uniform negative major-axis bending (bottom flange under compression) – the pur-
lins exhibit Section B and the sheeting restraints are modelled through Constraint IV with fully
prevented translation and rotation (springs with infinite stiffness). Figure 5 shows the signa-
ture curves of the unrestrained (KR = KT = 0) and fully restrained (KR = KT = ∞) purlins,
providing the variation of the critical buckling moment Mcr and mode shape with the length
L (logarithmic scale) – the buckling mode half-wave numbers are inside brackets. Besides the
results obtained through restrained-mode GBT analyses (solid and dashed curves, and buck-
ling mode shapes), the figure also display, for validation and comparison purposes, some crit-
ical buckling moment determined by means of GBTUL2.0 analyses (circles) – all analyses
adopt longitudinal discretisations into 10 finite elements and include all deformation modes.
The close observation of the buckling results displayed in this figure prompts the following
remarks:
(i) As before, the buckling moments provided by GBTUL2.0 and those obtained with the
proposed formulation virtually coincide (all differences are below 1.0%).

1021
Table 1. Components of C  kk , B
 kk and D
 ik stiffness matrices concerning modes 2-9 for unrestrained and restrained (Constraints I and III) lipped channel cross-
sections A and B – dimensions in cm, Young and shear moduli in kN/cm2.
Section A - Constraint I - C
 kk × 103 Section B - Constraint III - C
 kk × 104

852.4 216.9 9.514 0.309 0.334 0.010 0.014 0.012 2556 2.797 1.013 0.905 0.594 0.046 0.047 0.048
Unrestrained Section A (GBTUL2.0) - C
 kk × 103 Unrestrained Section B (GBTUL2.0) - C
 kk × 104

852.4 216.9 3947 0.301 0.325 0.010 0.014 0.012 2556 327.3 35163 1.038 1.135 0.043 0.047 0.048
Section A - Constraint I - B
 kk × 10−2 Section B - Constraint III - B
 kk × 10−2
0 0 0.092 6.863 11.21 22.76 142.0 172.4 0 0.004 2.162 3.625 374.4 34.16 278.9 856.9
Unrestrained Section A (GBTUL2.0) - B
 kk × 10−2 Unrestrained Section B (GBTUL2.0) - B
 kk × 10−2
0 0 0 2.967 6.474 22.68 141.8 172.0 0 0 0 1.789 5.103 31.62 278.8 858.7
Section A - Constraint I - D
 ik × 103 Section B - Constraint III - D
 ik × 104
0 0 0.016 0 0.245 0 1.141 0 0 0 0.106 -0.715 -0.261 -0.018 6.083 -0.019
0 0 0 0.205 0 0.25 0 -0.28 0 0.0764 -0.083 0.224 0.035 0 -0.222 0.587
0.016 0 0.109 0 -0.16 0 0.129 0 0.106 -0.083 5.114 -0.817 4.395 -3.612 -0.430 -2.657
0 0.205 0 2.305 0 0.779 0 0.231 -0.715 0.224 -0.817 3.738 1.262 0.315 2.480 -0.045
0.245 0 -0.16 0 2.3 0 0.134 0 -0.261 0.035 4.395 1.262 14.54 -14.51 1.815 3.911
0 0.25 0 0.779 0 2.537 0 -1.179 -0.018 0 -3.612 0.315 -14.51 15.70 0.127 -5.019
1.141 0 0.129 0 0.134 0 9.604 0 6.083 -0.222 -0.430 2.480 1.815 0.127 68.97 0
0 -0.28 0 0.231 0 -1.179 0 8.457 -0.019 0.587 -2.657 -0.045 3.911 -5.019 0 157.8
Unrestrained Section A (GBTUL2.0) - D ik Unrestrained Section A (GBTUL2.0) - D ik
0 0 0 0 -0.23 0 1.141 0 0 0 0 0 -0.919 0 -6.078 0
0 0 0 0.19 0 -0.25 0 0.28 0 0 0 -1.085 0 -1.431 0 -6.854
0 0 61.93 0 -6.91 0 -2.62 0 0 0 2355 0 53.01 0 32.01 0
0 0.19 0 2.23 0 -0.728 0 -0.159 0 -1.085 0 5.927 0 3.966 0 -2.714
-0.23 0 -6.91 0 2.24 0 -0.087 0 -0.919 0 53.01 0 5.962 0 -3.279 0
0 -0.25 0 -0.728 0 2.533 0 -1.169 0 -1.431 0 3.966 0 15.65 0 4.966
1.141 0 -2.62 0 -0.087 0 9.601 0 -6.078 0 32.01 0 -3.279 0 68.95 0
0 0.28 0 -0.159 0 -1.169 0 8.481 0 -6.854 0 -2.714 0 4.966 0 158.2
Table 2. Stud buckling results: ANSYS and GBT (GBTUL2.0 and restrained deformation modes) analyses.
Conventional Modes (GBTUL2.0) Restrained Modes

ANSYS All Modes Single Mode All Modes Single Mode

L Pcr Pcr Δ (%) Modal participations Pcr Mode Δ (%) Pcr Δ (%) Modal participations Pcr Mode Δ (%)

(cm) (kN) (kN) GBT/ (kN) Single (kN) GBT/ (kN) Single
ANSYS All ANSYS All

10 25.04 26.22 4.71 83%(7) + 16%(9) + 1% 31.32 (7) 19.5 25.46 1.68 91.10%(7) + 3.5%(6) + 3.4%(5) + 2% 25.79 (7) 1.3

50 24.32 24.41 0.37 85%(7) + 14%(9) + 1% 27.54 (7) 12.8 23.99 -1.36 92.5%(7) + 2.2%(6) + 2.1%(5) + 3.2% 24.19 (7) 0.8

100 24.30 24.37 0.29 86%(7) + 14%(9) 27.72 (7) 13.7 24.06 -0.99 92.2%(7) + 2.6%(6) + 2.5%(5) + 2.7% 24.22 (7) 0.7
150 24.31 24.39 0.33 85%(7) + 14%(9) + 1% 27.79 (7) 13.9 24.08 -0.95 93.1%(7) + 2%(6) + 1.95%(5) + 2.95% 24.25 (7) 0.7

200 - 25.57 - 84%(7) + 15%(9) + 1% 29.35 (7) 14.8 25.09 - 92.8%(7) + 2.5%(6) + 2.3%(5) + 2.4% 25.30 (7) 0.8

250 25.70 26.46 2.95 82%(7) + 18%(9) 30.43 (7) 15.0 25.96 1.00 90.2%(7) + 4.3%(6) + 4.2%(5) + 1.3% 26.17 (7) 0.8
300 - 23.78 - 99.9%(3) + 0.10% (9) 23.79 (3) 0.0 23.78 - 99.5%(2) + 0.5% 23.79 (2) 0.0

350 - 17.47 - 99.9%(3) + 0.10% (9) 17.48 (3) 0.1 17.47 - 99.7%(2) + 0.3% 17.48 (2) 0.1

400 13.35 13.38 0.23 100%(3) 13.38 (3) 0.0 13.38 0.23 100%(2) 13.38 (2) 0.0

450 - 10.57 - 100%(3) 10.57 (3) 0.0 10.57 - 100%(2) 10.57 (2) 0.0
500 - 8.56 - 100%(3) 8.56 (3) 0.0 8.56 - 100%(2) 8.56 (2) 0.0

Figure 5. Variation of the critical buckling moment and mode shape with the purlin length.

(ii) For L ≤ 252 cm, the buckling behaviour does not depend on the restraint conditions (the
solid and dashed curves coincide). This is because the critical buckling modes combine
local and/or distortional deformations that do not involve upper flange horizontal dis-
placements or rotations.
(iii) In the restrained purlins with L > 252 cm, the buckling mode shape is clearly lateral-
distortional, due to the full restraint of the upper flange horizontal displacements and
(mostly) rotations – this instability phenomenon has also been investigated by other
authors (e.g., Hancock et al. 2001).
(iv) In the unrestrained purlins with L > 252 cm, lateral-torsionalbuckling occurs and is associ-
ated with a very pronounced Mcr decrease. As shown by Basaglia et al. (2013), there is a
minute horizontal displacement of the point where the translational restraint is located – see
the buckling mode shape of the unrestrained purlin with L = 400 cm.
Finally, for a few purlins buckling in distortional or lateral-distortional modes, Table 3 pro-
vides the critical buckling moments obtained by means of GBT analyses including all deform-
ation modes and also either only the dominant (“restrained GBT”) or the two most relevant
(GBTUL2.0) ones. It is observed that, once again, the GBT analyses including a single
restrained deformation mode (3 or 4 – see Figure 3(b)) provide accurate buckling moments.
Conversely, the inclusion of the two most relevant deformation modes in the GBTUL2.0 ana-
lyses still leads to inaccurate buckling moments.

1023
Table 3. Purlin buckling results: GBT (GBTUL2.0 and restrained deformation modes) analyses.
Conventional Modes (GBTUL 2.0) Restrained Modes

All Modes Two Modes All Modes Single Mode

L Mcr Modal Participations Mcr Modes Δ (%) Mcr(kNcm) Modal Participations Mcr Mode Δ (%)

(cm) (kNcm) (kNcm) Two (kNcm) Single


All All

75 4933.78 52%(6) + 47%(5) + 1% 8608.13 (6) + (5) 74.5 4933.88 95%(4) + 1.1%(3) + 3.9% 5028.73 (4) 1.9

150 4934.20 52%(6) + 47%(5) + 1% 8235.79 (6) + (5) 66.9 4934.30 95%(4) + 1.1%(3) + 3.9% 5029.17 (4) 1.9
224 4936.10 52%(6) + 47%(5) + 1% 7870.98 (6) + (5) 59.5 4936.19 95%(4) + 1.1%(3) + 3.9% 5031.76 (4) 1.9

400 3551.75 52%(3) + 47(4) + 1% 71389.38 (3) + (4) 1910.0 3554.28 99.5%(3) + 0.5% 3566.53 (3) 0.3

500 3863.29 52%(3) + 47(4) + 1% 71510.10 (3) + (4) 1751.0 3867.27 99.4%(3) + 0.6% 3888.06 (3) 0.5

800 3552.10 52%(3) + 47(4) + 1% 72237.60 (3) + (4) 1933.7 3554.63 99.5%(3) + 0.5% 3566.88 (3) 0.3

4 CONCLUSION
The paper reported the results of an ongoing investigation on the development of a GBT formu-
lation for the buckling analysis of restrained thin-walled members, differing from the conventional
one in the cross-section analysis (it already incorporates the elastic restraints). Special attention
was paid to the procedures involved in the determination of the restrained deformation modes.
The application and capabilities of the above GBT formulation were illustrated through the pres-
entation and discussion of numerical results concerning cold-formed steel (ii) studs braced by
sheathing and (ii) purlins restrained by sheeting. For validation and assessment purposes, some
results were compared with values provided by the codes GBTUL2.0 (conventional GBT) and/or
ANSYS (SFEA). Besides the expected virtual coincidence of the results, it was found the proposed
GBT formulation is much more efficient than the conventional one. In particular, it becomes pos-
sible to obtain accurate buckling results with a single deformation mode, which enables the devel-
opment of analytical formulae to calculate critical buckling loadings of restrained members – this
feature, currently being exploited by the authors, will be addressed in future works.

ACKNOWLEDGMENTS
The second author gratefully acknowledges the financial support of Conselho Nacional de
Desenvolvimento Científico e Tecnológico (CNPq – Ministry of Science, Technology and
Innovation of Brazil), through project 308530/2016-0.

REFERENCES
Basaglia, C., Camotim, D. (2015). Buckling analysis of thin-walled steel structural systems using General-
ised Beam Theory (GBT), International Journal of Structural Stability and Dynamics, 15(1), 1540004.
Basaglia, C. Camotim, D. Gonçalves, R., Graça, A. (2013). GBT-based assessment of the buckling behav-
ior of cold-formed steel purlins restrained by sheeting, Thin-Walled Structures, 72(November), 217–229.
Bebiano, R., Camotim, D., Gonçalves R. (2018). GBTUL 2.0 – A second-generation code for the GBT-based
buckling and vibration analysis of thin-walled members, Thin-Walled Structures, 124(March), 235–257.
Camotim, D., Silvestre, N., Basaglia, C., Bebiano, R. (2008). GBT-based buckling analysis of thin-walled
members with non-standard support conditions, Thin-Walled Structures, 46(7–9), 800–815.
Dinis, P.B., Camotim, D., Silvestre, N. (2006). GBT formulation to analyse the buckling behaviour of thin-
walled members with arbitrarily “branched” open cross-sections, Thin-Walled Structures 44(1), 20–38.
Hancock, G.J., Murray, T.M., Ellifritt, D.S. (2001). Cold-Formed Steel Structures to the AISI Specifica-
tion, New York: Marcel Dekker Inc.
Jiang, C., Davies, J.M. (1997). Design of thin-walled purlins for distortional buckling, Thin-Walled Struc-
tures 29(1–4), 189–202.
SAS (Swanson Analysis Systems Inc.) (2013). ANSYS Reference Manual (version 15).
Schardt, R. (1989). Verallgemeinerte Technische Biegetheorie, Berlim: Springer Verlag. (German)
Silvestre, N., Camotim, D. (2002). First-order generalised beam theory for arbitrary orthotropic
materials, Thin-Walled Structures, 40(9), 755–789.
Silvestre, N., Camotim, D. (2003). GBT buckling analysis of pultruded FRP lipped channel members,
Thin-Walled Structures 81(18–19), 1889–1904.

1024
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Stability interaction effects in 3D steel frames—a case study

H.H. Snijder
Department of the Built Environment, Eindhoven University of Technology, Eindhoven, The Netherlands

L.H.J.D. Titulaer
Ballast-Nedam, Nieuwegein, The Netherlands

P.A. Teeuwen
Witteveen+Bos, Deventer, The Netherlands

H. Hofmeyer
Department of the Built Environment, Eindhoven University of Technology, Eindhoven, The Netherlands

ABSTRACT: 3D steel frames are usually assessed by checking their columns and beams sep-
arately using design rules, e.g. EN 1993-1-1 (Eurocode 3). For this, the force distribution in
the 3D steel frame is determined first, followed by cross-sectional resistance and member sta-
bility checks. When increasing the load on the frame, the first critical check defines the ultim-
ate load. This paper presents a case study in which a 3D steel frame is assessed both by code
checking and by more realistic numerical simulations by the finite element method (FEM). It
shows that the code checking approach overestimates the ultimate load of the FEM approach.
The FEM simulations show simultaneous failure of beams and columns and naturally take
into account the mutual stiffness interaction of unstable beams and columns. As the latter is
not the case for the design rules used in the code checking approach, the suggestion is made
that the current design rules predict too high values for this reason. As similar findings are
obtained for 2.5D and 2D frames, further research is needed, including full scale experiments.

1 INTRODUCTION

In practice, 3D steel frames are usually analysed by assessing the different columns and beams
separately using design rules, e.g. the ones of EN 1993-1-1 (Eurocode 3). The force distribution
in the 3D steel frame is determined first, followed by cross-sectional resistance and member sta-
bility checks for the beams and columns. The load can be proportionally increased until one of
the checks is violated, thus obtaining the ultimate load of the frame, which is expressed in the
ultimate load proportionality factor (LPF) αult,DR. Using the FEM, 3D steel frames can also be
analysed (A) as a whole, taking material (M) and geometric (G) non-linearities (N), and imper-
fections (I) into account, i.e. an assessment by GMNIA. In this way the ultimate load can also
be expressed in terms of an LPF: αult,FEM. The first approach, called the code checking
approach, is usually expected to be conservative in terms of ultimate load compared to the
more sophisticated second approach, called FEM approach. However, in the code checking
approach it is not obvious that all 3D instability effects are taken into account. Especially if
two or more members fail simultaneously by instability and plasticity, stability interaction
effects may cause the 3D frame to have a lower ultimate LPF αult,FEM than αult,DR. This would
mean that the code checking approach would not be sufficiently conservative. The design rules
for member stability used in the code checking approach have been derived for rather basic
cases with standard boundary conditions. However, in a 3D frame, the boundary conditions for
each member are complex and may change due to plasticity and stability effects with increasing
LPF. Nevertheless, these design rules are used in practice, also if the members form part of

1025
Figure 1. Geometry and loading of a 3D frame (left), a 2.5D frame (middle) and a 2D frame (right).

a 3D frame. It is expected that cases with columns and beams failing simultaneously may be
critical. Therefore, a research project was carried out to investigate these cases (Titulaer, 2018).
This paper presents the research for a case study where interaction effects are present, and com-
pares different analysis strategies, which resulted in different ultimate load predictions.

2 STEEL FRAMES CONSIDERED

The 3D steel frame as shown in Figure 1 (left) is used for the case study. Its dimensions equal
3m for all three directions. The columns are sections HEB140; the beams are sections IPE180.
The fillets of these sections have been neglected. The steel grade is S235. Figure 2 (left) shows
a connection between beams and column, which are welded and are assumed to be rigid and
full-strength. The weld dimensions have been neglected. At each column base, the displacements
in all three directions and the rotation about the longitudinal column axis are constrained.
A vertical concentrated load acts on each column equal to Fv = 96 kN. Two horizontal concen-
trated loads act at the beam level: Fh = 8.35 kN. Finally, uniformly distributed loads act on
each beam in vertical direction: qv = 13 kN/m.
For comparison two other frames are considered. Figure 1 (middle) shows a 2.5D frame
with similar properties as the 3D frame. At the top ends of the beam, the 2.5D frame is
restrained against out-of-plane displacements, leaving other out-of-plane displacements and
rotations possible. Figure 1 (right) shows a 2D frame with similar properties as the other two
frames, but now all out-of-plane displacements and rotations are restrained. For a better com-
parison with the 3D frame, the vertical concentrated loads on the columns of the 2.5D and 2D
frames have been increased to Fv = 115.5 kN to compensate for the uniformly distributed
loads on the removed beams. However, the out-of-plane bending moments at the top of the
columns of the 2.5D and 2D frame, associated with the removed loaded beams, have been
neglected to be able to compare the results of the 2.5D and 2D frame better.

Figure 2. Rigid full-strength welded connection (left) and section representation by shell
elements (right).

1026
3 FEM APPROACH

FEM is used to predict a load-displacement curve for the frames considered (Figure 1) by
GMNIA. Linearly increasing load control is applied, and for the solver the Riks (arc-length)
method is used to predict the behavior after the ultimate load. The applied load is expressed
as an LPF α with reference to the loads mentioned earlier. The ultimate load is characterized
by the ultimate LPF αult,FEM.
The finite element model has been developed in Abaqus and consists of S4R shell elements.
A mesh density study and validations have been carried out in Titulaer (2018), which includes
the selection of the elements. Figure 2 (right) indicates the shell element positions relative to
the cross-section, and Figure 3 (left) gives an impression of the mesh density and layout.
The material model for steel grade S235 is taken from EN 1993-1-5 (2006) and is shown in
Figure 3 (right), both in terms of engineering stress and strain and true stress and strain, the
latter used in the finite element model. Young’s modulus equals E = 2.1 x 105 N/mm2 and
Poisson’s ratio equals ν = 0.3. Plasticity is described by a Von Mises yield surface with associ-
ated plastic flow and isotropic hardening
Residual stresses are taken into account according to ECCS (1984), see Figure 4. For the
IPE180 beam h/b = 180/91 = 1.98 > 1.2 so Figure 4 (right) is applicable, while for the HEB140
column h/b = 140/140 = 1.0 < 1.2 so Figure 4 (left) applies.
Geometrical imperfections are taken into account using the Eigen buckling-mode method
(EBM) as presented in Liu et al. (2014). In this method, several Eigen modes determined by
a linear buckling analyses (LBA) are combined using a specific amplitude for each one. Using
the first three Eigen modes leads to a good trade-off between practical use and accuracy of the
EBM and the amplitudes of these Eigen modes (Aj) can be obtained as follows (Liu et al. 2014):

Aj ¼ Cj  F  H ð1Þ

where Cj is the contribution factor for buckling mode j (here C1 = 0.831, C2 = 0.132
and C3 = 0.037), F a factor depending on the number of Eigen modes considered (here

Figure 3. Mesh (left) and material model (right) used.

Figure 4. Residual stress patterns according to ECCS (1984) for rolled sections.

1027
Figure 5. First three Eigen modes as used in the EBM to define the geometrical imperfection.

F = 0.00163 for three Eigen modes) and H the height of the frame (here H = 3000 mm). Slightly
different numbers for F and Cj apply for 2D frames (Shayan et al. 2014). The first three Eigen
modes of the 3D frame are shown in Figure 5. They are combined according to Equation 1 into
the geometrical imperfection as used in the GMNIA.
As an alternative, the imperfections have been used as given in EN 1993-1-1 (2006). In this
code, so-called equivalent geometrical imperfections are given so residual stresses should not
be taken into account separately. This is called the method of initial geometric imperfections
(IGI) after Chan et al. (2005). Sway and bow imperfections should be combined such that
they have the most unfavourable effect, considering the different directions as indicated in
Figure 6 (translations in x and z-direction and rotation in the y-z-plane). If residual stresses
are nevertheless taken into account explicitly, the bow imperfection may be reduced to 1/1000
of the member length. In conclusion, six different imperfection types have been investigated in
the simulations:
– No imperfection (No);
– EBM and residual stresses (EBM-RS);
– IGI in y-direction without residual stress (IGI-y);
– IGI in z-direction without residual stress (IGI-z);
– IGI in y-z-plane rotational direction without residual stress (IGI-yz);
– IGI in y-direction with residual stress (IGI-y-RS).
Figure 7 shows the imperfect 3D frame for GMNIA calculations IGI-y, IGI-z an IGI-yz.
More detailed information on the imperfections is given in Titulaer (2018).
Figure 8 shows the load-displacement curves for the six different GMNIA calculations. The
horizontal axis shows the horizontal maximum displacement of the top of the front column
(Figure 9). Regardless the imperfection type used, the ultimate LPF is about αult,FEM = 0.375,
which is the value for EBM-RS. Since EBM-RS is a straightforward procedure to apply, this
is the selected method for all further GMNIA simulations. Imperfections according to EN
1993-1-1 (2006) give similar results as shown in Figure 8. The deformed 3D frame at ultimate
load is shown in Figure 9 (left), which also shows the deformed 2.5D and 2D frames at ultim-
ate load (middle and right). The deformed frames indicate simultaneous failure of a beam and
a column, further evidence of which is given in Titulaer (2018).

Figure 6. Directions of imperfections according to EN 1993-1-1 (2006).

1028
Figure 7. Imperfect 3D frames for GMNIA calculations IGI-y, IGI-z and IGI-yz (scaled
imperfections).

Figure 8. Load-displacement curves for the 3D frame with different imperfections.

Figure 9. Deformations (scaled) of frames at the ultimate load: 3D (left), 2.5D (middle) and 2D (right).

4 CODE CHECKING APPROACH

Two different approaches are considered for checking the frames by the design rules of Euro-
code EN 1993-1-1 (2006): the Sway-mode Buckling length Method (SBM) and the Amplified
Sway moment Method (ASM).
The SBM starts with a first-order linear elastic analysis of the sway frame, resulting in
internal force distributions as shown in Figure 10. Then cross-sectional resistance and member

1029
Figure 10. Internal force diagrams: normal force N (left), bending moment My (middle) and Mz (right).

stability are checked. Sway stability is included by the use of sway buckling lengths in the
column stability checks.
The principle of the ASM is illustrated in Figure 11. It uses a first order linear elastic ana-
lysis in combination with an amplification of the sway moments to calculate the internal
forces. The ASM is carried out by the following steps:
– add a horizontal support to obtain a non-sway frame;
– carry out a linear elastic analysis of the non-sway frame to determine the internal forces;
– determine the horizontal reaction force at the support;
– remove the horizontal support and position the reaction force on the sway frame;
– execute a linear elastic analysis of the sway frame to determine the internal forces;
– amplify the bending moments from the previous step by the amplification factor β of Equa-
tion 2 below;
– add these amplified bending moments to the internal forces in the non-sway frame (second
step above).
The amplification factor is:

β ¼ αcr =ðαcr  1Þ ð2Þ

where αcr is the elastic critical LPF at which the frame buckles elastically in its sway mode.
The ASM uses the non-sway buckling lengths for the columns since the effect of sway instabil-
ity is already included in the obtained internal forces, which are reported in Titulaer (2018).
Once the internal forces according to the SBM and the ASM are known, cross-sectional
resistance and member stability checks are carried out. As an example of a cross-sectional
resistance check, Equation 3 is given for the most general case of a normal force N, shear
forces Vy, Vz, and bending moments My, Mz simultaneously present in a cross-section. Equa-
tion 3 is taken from the Dutch National Annex (NA) to EN 1993-1-1 (2006).

Figure 11. Amplified sway-moment method.

1030
 α1  α2
My;Ed Mz;Ed
β0 þ β1  1:0 ð3Þ
My;N;V;Rd Mz;N;V;Rd

where α1, α2, β0 and β1 are coefficients, My.Ed and Mz,Ed are design bending moments about
the y- and z-axis of the cross-section caused by the applied load and My.N,V,Rd and Mz,N,V,Rd
are the design bending moment resistances about the y- and z-axis of the cross-section, taking
the effects of the normal force N and the shear force V into account; see also the Dutch NA to
EN 1993-1-1 (2006). As an example of a member stability check, the well-known Equation 4 is
given for a beam-column under simultaneous compressive normal force and bi-axial bending.
This design rule is obtained from EN 1993-1-1 (2006).

NEd My;Ed Mz;Ed


χY NRk
þ kyy My;Rk
þ kyz Mz;Rk
 1:0 ð4aÞ
γM1 χLT γM1 γM1

NEd My;Ed Mz;Ed


χz NRk
þ kzy M
þ kzz Mz;Rk
 1:0 ð4bÞ
γM1 χLT γ y;Rk γM1
M1

where for the meaning of the symbols, the reader is referred to EN 1993-1-1 (2006).
In the member stability checks, e.g. Equation 4, the elastic critical buckling force Ncr and
the elastic critical moment Mcr are needed to determine, amongst others, the reduction factors
χy and χz for flexural buckling and χLT for lateral torsional buckling. The elastic critical buck-
ling force can be determined by using the elastic critical buckling length in the well-known
Euler buckling formula. The elastic critical buckling force and moment can be determined by
either design rules (DR) as given in handbooks or via a FEM linear buckling analysis (LBA).
The DR approach is elaborated in Titulaer (2018), where the design rules given in the Dutch
NA to EN 1993-1-1 (2006) are used. The LBA approach is given by Equation 5.

Ncr ¼ NEd  αcr ð5aÞ

Mcr ¼ MEd  αcr;LTB ð5bÞ

where NEd and MEd are the design compressive normal force in the column and the design bend-
ing moment in the beam or column respectively, and αcr and αcr,LTB are the elastic critical LPF’s
for flexural buckling of the column and lateral torsional buckling of the column or beam,
respectively. As an example, Figure 12 shows the lateral torsional buckling modes of the beam
(left) and the column (right) for which the elastic critical LPF’s have been calculated by the LBA
method as given by Equation 5 to evaluate the elastic critical buckling force and moment.
The ultimate LPF based on a design rule αult,DR is defined as the load for which the
first time somewhere in the frame one of the design rules, e.g. Equations 3 or 4, yields
unity (1.0). Three different values of αult,DR are evaluated, namely for the:

Figure 12. Lateral-torsional-buckling modes for the beam (left) and the column (right).

1031
– SBM with design rules based elastic critical buckling forces and moments (SBM-DR);
– SBM with LBA based elastic critical buckling forces and moments (SBM-LBA);
– ASM with design rules based elastic critical buckling forces and moments (ASM-DR).

5 RESULTS, COMPARISONS, AND DISCUSSION

Figure 13 shows the results for the 3D frame. The load-displacement curve as obtained with the
FEM approach by a GMNIA with geometrical imperfections according to the EBM and with
residual stresses (EBM-RS) is compared to the ultimate LPF’s according to the three code
checking approaches. Figure 14 shows similar comparisons for the 2.5D and 2D frames.
For the 3D frame, Figure 13 shows that none of the code checking approaches is conservative
compared to the ultimate LPF αult,FEM = 0.375 of the FEM approach. Especially when the elas-
tic critical force and moment are based on design rules (SBM-DR and ASM-DR), αult,FEM is
significantly overestimated. However, the result is almost correct if the elastic critical force and
moment are based on an LBA (SBM-LBA). This suggests that coupled linear elastic stability,
as taken into account by an LBA but not by the design rules, is responsible for the differences.
For the 2.5D frame, Figure 14 (left) shows similar results as for the 3D frame. Again, the
result is better if the elastic critical force and moment are based on an LBA (SBM-LBA), but
the result is more unconservative than for the 3D frame. All code checking approaches overesti-
mate αult,FEM even more than for the 3D frame.

Figure 13. Load-displacement curve and ultimate LPF’s for the 3D frame.

Figure 14. Load-displacement curve and ultimate LPF’s for the 2.5D frame (left) and 2D frame (right).

1032
For the 2D frame, Figure 14 (right) shows that the SBM with the elastic critical force and
moment based on an LBA (SBM-LBA) is not the code checking approach closest to the ultim-
ate LPF αult,FEM. Note that this undermines the suggestion made for the 3D frame regarding
coupled linear elastic stability being responsible for the differences. Maybe in this case plasti-
city effects are responsible. All code checking approaches overestimate αult,FEM even more
than for the 3D and 2.5D frames.
For all frames considered, the FEM approach (EBM-RS) predicts a column and beam failing
simultaneously (Figure 9). The FEM approach is perfectly able to capture the associated effect
of the reduced support stiffness of the buckled column (due to flexural buckling and plasticity)
to the beam, and the reduced support stiffness of the buckled beam (by lateral torsional buck-
ling and plasticity) to the column. This unfavorable elastic-plastic stability interaction effect
may not be sufficiently covered by the code checking approaches.

6 CONCLUSIONS AND RECOMMENDATIONS

In this paper, a case study of a 3D steel frame was presented to explore stability interaction
effects. Also similar 2.5D and 2D frames were considered. Two code checking approaches, the
sway-mode buckling length method (SBM) and the amplified sway-moment method (ASM),
with their appropriate design rules for checking resistance and stability, were compared with
the finite element approach taking into account geometrical and material non-linearities with
imperfections (GMNIA). It can be concluded that the code checking approaches do not con-
servatively cover the ultimate load of the steel frames studied, in which the members fail sim-
ultaneously. However, using a linear buckling analysis (LBA) to determine the elastic critical
buckling force and moment of members yielded less unconservative results.
Extended research on stability interaction effects in steel frames is recommended since only
one specific 3D sway frame and its associated 2.5D and 2D sway frames have been studied in
this paper. Also research is needed on different geometries of non-sway frames. In general,
only limited research has been carried out on the stability interaction of members in steel
frames. Therefore, it is recommended to execute full-scale experiments, combined with finite
element analyses, to obtain further insights into the complex stability interaction effects in
steel frames.

REFERENCES

Chan, S. L., Huang, H. Y. & Fang, L. X. (2005). Advanced analyses of imperfect portal frames with
semi-rigid base connections. Journal of Engineering Mechanics 131(6): 633–640.
ECCS/TC8 ‘Stability’. 1984. Ultimate Limit State Calculation of Sway Frames with Rigid Joints. Brussels.
ECCS.
EN 1993-1-1. 2006. Eurocode 3: Design of steel structures- part 1-1: General rules and rules for buildings.
CEN. Brussels.
EN 1993-1-5. 2006. Eurocode 3: Design of steel structures- part 1-5: Plated structural elements. CEN.
Brussels.
Liu, W. & Rasmussen, K. J. R. & Zhang, H. (2014). On the modelling of geometric imperfections in 3D steel
unbraced frames. In R. Landolfo and F.M. Mazzolani (eds.), Eurosteel 2014 – 7th European Conference on
Steel and Composite Structures: 163-164 and 6 page paper on USB, Napoli, 10-12 September 2014. Brussels:
ECCS.
Shayan, S., Rasmussen, K. J. R., & Zhang, H. (2014). On the modelling of initial geometric imperfections
of steel frames in advanced analysis. Journal of Constructional Steel Research 98, 167–177.
Titulaer, L.H.J.D. 2018. Influence of stability interaction effects on the ultimate resistance of 3D steel
frames. MSc thesis. Report no. O2018.249. Eindhoven: Eindhoven University of Technology. Depart-
ment of the Built Environment.

1033
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Experimental investigation on the instability phenomenon


in stainless steel connections—plate curling

K. Sobrinho
PGECIV – Post-graduate Program in Civil Engineering, State University of Rio de Janeiro, Brazil

A. Tenchini, M. Cordeiro, P. Vellasco & L. Lima


Structural Engineering Department, State University of Rio de Janeiro, Brazil

J. Henriques
CERG – Construction Engineering Research Group, University Hasselt, Belgium

ABSTRACT: The overlap bolted steel connection with thin-plates can be subjected to the
occurrence of the phenomenon known as curling effect, which is able to influence the global
behavior and decrease its ultimate strength. For stainless steel joints, the mechanism is even
more important, because it is a material with high deformation capacity. Therefore, both
experimental tests and numerical analyses are presented in this paper in order to investigate
the influence of curling effect in overlap bolted connection. In addition, the lips were also
studied because they increase the stiffness for the out-plane displacement. The outcomes
shown that the curling effect reduced the ultimate bearing resistance. Comparing the actual
codes, the design load provided by they are lower than those are obtained in both experimen-
tal and numerical tests. In addition, it can be reported the efficiency from the lips for the fer-
ritic overlap connection.

1 INTRODUCTION

The connection plays a very important role in structure in order to transfer the forces among
the structural members and contribute in the global structural behavior. In nowadays, the
most commonly used types of connections are welded and bolted. The rivet connections also
were very common in the past. In details, the overlap bolted connections have been observed
in several civil constructions due to easy applicability. In structural design of these connec-
tions type, it is important to define the possible failure modes associated to applied load type,
geometric proprieties and material employed. It is recognized that these connection types can
present the following failure modes corresponding to net rupture, bearing, bolted shear and
yielding gross section. In addition, in thin-plates, there is another failure mode associated to
high compression stress in near hole known as curling. In fact, the curling effect can occur due
to compression deformation in the end-region of the connection, where the end-plate is fixed
by both nut and bolt, and at the other there is no restriction resulting in out-plane deform-
ation. This phenomenon has superior relevance in bolted connections with high distance
between the hole and edge plate, as well as, in materials with high deformation capacity, such
as, stainless steel (Henriques et al, 2018).
In fact, the stainless steel provides large deformation capacity in comparison with the
carbon one. Although there is still a limited examples of civil structures with stainless steel
members, the application of stainless steel for structural has been used more frequently in
recent years, mainly due to the increase of research on its use. This steel grade is recognized by
the presence of the chromium and nickel, and may contain molybdenum, iron and other elem-
ents. In special, stainless steel should contain at least 10.5% chromium. The interesting in this

1034
steel grade from engineer and researcher is associated to excellent proprieties in comparison
with traditional mild carbon steels, such as high corrosion resistance, durability, fire resistance
and high aesthetic value. On the other hand, the material cost of the stainless profiles is very
higher complicating the use in large scale in civil constructions. Recently, this idea was over-
come through of the study performed by Silva et al (2016) when the structural design is
addressed to assessment of the maintenance cost to be spent over the life of power transmis-
sion tower when there is comparison of mild carbon against stainless steel grades. Therefore,
stainless steel can provide several advantages related to carbon steel, which over time can
become a more economical solution, such as corrosion resistance, better behavior at high tem-
peratures when compared to carbon steel, higher reuse capacity, ductility and impact resist-
ance (Baddoo, 2008).
Considering the promising use of stainless steel in civil structures, this paper aims to
investigate the influence of this steel type in overlap shear connection using thin-plates
when it is possible to observe the curling effect. Several studies can be found in litera-
ture about the structural behavior of stainless bolted connections. In details, Kim et al
(2009) observed that in the bolted connections, the ultimate resistance increases in pro-
portion to distance between hole and edge-plate until the appearance of the curling
effect. The occurrence and magnitude of the influence of the curling effected the ultimate
resistance being associated to thickness plate and distance from hole and edge-plate. For
the described arrangements, this phenomenon reduced the connection ultimate resistance
by 11%, 16% and 14% for the plates with 1.5 mm, 3.0 mm and 6.0 mm, respectively. In
order to reduce the influence of the curling effect, Yancheng & Young (2014) investi-
gated through experimental tests the behavior of the stainless bolted connections using
lips. The idea was to assess the bolted connections capacity without the influence of the
curling effect since the lips increase the stiffness for the out-plane displacement. The
results shown that the nominal resistance obtained from actual design codes are gener-
ally conservative for both single and double shear bolted connections. On the other
hand, the failure modes observed in experimental tests are close to those predicted by
European code (EN 1993-1-4, 2006). Hence, the purpose of the present investigation is
to contribute with further experimental and numerical analyses in order to investigate
the curling effect in bolted stainless steel connections using lips. Hereafter, experimental
tests are presented where it was carried out two stainless steel grades: Austenitic and
Ferritic. In addition, a numerical model has been developed on basis of the experimental
tests and the results are compared in terms load-displacement curves and failure modes.

2 STRUCTURAL DESIGN OF OVERLAP BOLTED CONNECTIONS

2.1 Eurocode
According to EN 1993-1-4 (2006), the structural design of overlap bolted connections is
addressed to verification of the plastic resistance of the both gross and net cross-section, bear-
ing and limitation of the geometries on basis of the ultimate capacity. Considering the particu-
lar stainless behavior for gross cross-section, based on the yield stress, this criterion can
control the structural design and limit the connection resistance. The gross cross-section resist-
ance should be determined using the following equation:

A  fy
Npl;Rd ¼ ð1Þ
γM0

where, A corresponds to gross cross-section area, fy is the yield strength obtained from the
stress-strain curve. The partial safety factor γM0 is equal to 1.1. On the other hand, if this
expression is employed for the carbon steel connections with a distinct value equals to 1.0 is
adopted.

1035
The net cross-section resistance of a plate with holes is prescribed in EN 1993-1-4 (2006) on
basis of the following equation:
kr  Anet  fu
Nu;Rd ¼ ð2Þ
γM2

where, Anet is associated to net cross-section area, fu is the ultimate tensile strength of
the material and γM2 is a partial safety factor equal to 1.25. In this case, there is no difference
between the carbon or stainless steel connections for this partial safety factor. In contrast, the
value of the kr is equal to 0.9 for the carbon steel.
The bearing resistance is given in EN 1993-1-8 (2005) using a similar curve for carbon steel:

k1  αb  fu  d  t
Fb;Rd ¼ ð3Þ
γM2

where, d is the nominal bolt diameter, t corresponds to plate thickness, k1 depends on the
edge distance and inner bolts and αb is minimum value of the following based in geometric
relationships. For stainless steel connection, EN 1993-1-4 (2006) establishes that the ultimate
strength, fu, should be reduced due to a hole elongation limitation under serviceability loads.
This reduction is defined by combination of the yielding and ultimate strength of the material:

fu;Red ¼ 0; 5fy þ 0:6fu ð4Þ

Recently, it was published a new version of the Design Manual for Structural Stainless Steel
developed by The Steel Construction Institute (SCI) of for stainless steel structures in which
there is an important contribution in the bearing resistance of bolted connections. The SCI
P413 (2017) establishes that the use of the reduced ultimate strength to be applied in bearing
capacity is replaced by ultimate strength. In addition, this manual determinates that bolted
connection design needs to be verified based on two limit states: serviceability and ultimate.
This design strategy has been investigated in other study (Salih et al, 2011).

3 STUDY CASES

3.1 Experimental program


In this section, it is presented the experimental program carried out, which consists of eight
tests with and without lips being divided by stainless steel grades. For the first, two connections
with two shear planes were investigated where the external plates are controlling the structural
design. Here, the use of lips aim to mitigate the curling influence on the global behavior due to
stiffness increasing for out-plane displacements. For the second one, tests were carried out on
two shear planes with the central plate controlling the connection capacity. In these models, the
main objective is to analyze the behavior of the central plate, which is prevented from occurring
the curling effect. Therefore, with the tests carried out, it was possible to evaluate the influence
of the curling effect on austenitic and ferritic bolted connections steels.
A schematic drawing of the tests with and without lips is shown in Figure 1 where the
nominal value of the length of the stiffener L2 is equal to twice the value of e1, and its height,
h, equal to 10 mm. The plates have 3 mm thickness where the value of e1 is equal to 32 mm,
w corresponding to 50 mm, hole diameter equal to 13 mm and total length is 40 mm. In
addition, the bolted connections with lips are composed by mild carbon steel with 15 mm
thickness being designed in order to avoid a premature failure.
In order to identify the study cases, a nomenclature has been used being composed of
a code with two parameters. The first represents if the bolted connection is composed by outer
plates (OP) or inner plate (IN) to be considered as controlling member. The outer plates have
lips in order to increasing the stiffness against out-plane displacement. Thus, the aim is

1036
Figure 1. Geometries properties of the plates investigated.

Figure 2. Universal machine and LVDT position used in experimental tests.

avoiding the influence of the curling effect on bolted connection behavior, which may influ-
ence its maximum resistance (Sobrinho et al, 2018). The last parameter of the code represents
whether the bolted connections is with Austenitic (A) or Ferritic steel grades. The bolt used
for all tests was M12 class 12.9 type.
In order to perform the tests, the Losenhausen machine of 600kN was used, as shown in
Figure 2. The tests were instrumented with a linear differential transformer (LVDT) for meas-
uring the axial displacement of the tests and the curling displacement in the region between
the hole and the end of the plate. All the instruments were connected to the system Quantum
X-MX1615B universal data acquisition from HBM Test and Measurement.

3.2 Material characterization


The stress-strain curves were obtained for both ferritic and austenitic stainless steel grades using
the longitudinal tensile tests. The coupons tests were extracted considering the orientation of the
batch of stainless steel plate. In detail, it was fabricated twelve coupons tests being considered the
load axis for parallel, perpendicular and one direction corresponding to forty-five degrees of
batch of the plate. Figure 3 illustrates the three stress-strain curves found in tensile coupon tests.
In addition, the Table 1 reports the summary of main proprieties obtained from these
curves. As can observed, there is a notorious difference observed in structural behavior from
the both austenitic and ferritic stainless grades. The batch orientation has more influence for
the ferritic plates in comparison with the austenitic. Comparing the outcomes with the EN
1994-1-4 (2006), it can be observed a good correlation in terms of σ0.2%. On the other hand,
the ultimate strength from the Austenitic plate presented an higher difference. In general, the
values found in longitudinal tensile tests are superior to reported in EN 1993-1-4 (2006).

3.3 Numerical modelling


The finite element model used in this paper to investigate the tension capacity of overlap
bolted connections was developed by of software Abaqus 6.14 (2014). This element finite

1037
Figure 3. Stress-strain curves reported in tensile coupons tests.

Table 1. Main proprieties of coupon tests.


EN 1993-1.4 (2006)

E σ0,2% Ԑ0,2% σu Ԑu σ0,2% σu


CP [GPa] [MPa] [%] [MPa] [%] [MPa] [Mpa]

A00 202 275 0.336 860 55.2


A45 245 276 0.313 873 64.7 230 540
A90 258 279 0.308 879 62.0
F00 271 281 0.260 472 17.2
F45 254 322 0.286 484 15.6 260 450
F90 219 325 0.291 500 15.0

program is recognized by powerful tool that can incorporate material, geometric and bound-
ary non-linearity cause by nonlinear elasticity, plasticity, large displacement, contact problem,
etc. The numerical models were implemented using solid elements C3D8R defined by eight
nodes with three degrees of freedom per node: translations in the nodal x, y and z directions.
Concerning to adopted mesh, it was chosen a distribution which the proportions and size to
be adopted had the aim of avoid the numerical problems. In special, the mesh was refined
locally near the bolt hole for improved resolution of stress and strain due to be a region with
recognized high stress concentration. Contact surfaces were considered in bolt and plates to
better fit the adopted mesh distribution. The load was applied by means of axial load plate
displacements in reference node of the lateral face. In addition, all nodes of this face were con-
strained to reference node through of MCP-Tie.
The bolt material was idealized as linear elastic with Young modulus of 210 GPa and 0.3
Poisson coefficient. This strategy was also used in the mild carbon steel plate. On the other
hand, the stainless materials were modelled considering the stress-strain curves obtained in lon-
gitudinal tensile coupons tests for the same direction of the batch with load. Due to high strain
capacity from stainless steel plates, the stress-strain curves were converted to true stress versus
true strain where it is considered the large deformation observed in tensile coupons tests.
Therefore, a full nonlinear analysis was performed in all numerical models. The material
non-linearity was considered using a Von Mises yield criteria associated to a multi-linear
stress-strain relationship. The geometrical non-linearity was introduced in the model by using
an updated Lagrangean formulation.

4 RESULTS

Figure 4 illustrates the load-displacement curves for the four experimental tests together with
the numerical response. In addition, Table 2 reports the maximum load observed in both

1038
Figure 4. Load-displacement curves from both experimental and numerical tests.

Table 2. Maximum load observed in both experimental and numerical tests.


Test Code LoadEXP [kN] LoadNUM [kN] EXP/NUM

1 OP-A 71,33 68,31 1,04


2 IP-A 70,11 67,89 1,03
3 OP-F 53,93 53,5 1,01
4 IP-F 48,15 49,15 0,98
Média 1,02
C.O.V 0,03

experimental and numerical tests. As can be observed, it was possible to obtain a similar
behavior for the numerical analyses in comparison with the outcomes observed in experimen-
tal tests. Thus, it is possible to conclude that finite element analysis can consistently represent
the behavior of overlap bolted stainless steel connection submitted to shear.
Another important observation is related to maximum resistance observed for Austenitic
study cases. In fact, the ultimate strength observed in tensile coupon tests provided
a significant increasing of the bolted connection capacity. This issue resulted in bearing resist-
ance higher in compliance with the design code. In addition, the Austenitic bolted connection
reported an high axial deformation. In this case, a possible design on basis of the serviceability

1039
or ultimate limit state should be taking into account this important difference reaching a value
five times higher for the Austenitic study cases.
Analyzing the maximum load obtained for the connection with lips, it can be mentioned that
this system is more efficiency for case with ferritic stainless steel grade. There was a similar
behavior comparing the bolted connection with and without lips for the study cases with Aus-
tenitic steel grade. There is no difference for the resistance observed in both experimental and
numerical analyses. On the other hand, the case where the bolted connection with lips using
Ferritic steel grade presented an increasing of the maximum resistance equal to 10%.
Figures 5 and 6 show the deformed obtained in both analyses. As can be noted, there is an
out-plane displacement in study case with Ferritic steel grade. It is possible to note that there
is high stress concentration in near hole due to high deformation capacity from the Austenitic
one. In fact, the Ferritic study case provided a better stress distribution along the distance
between the end-plate to near hole. This fact is very important because the actual codes using
similar formulas in order to determine the bearing capacity of the overlap bolted stainless
steel connections.
Comparing the outcomes with the expression given by EN 1993-1-4 (2006) and SCI P413
(2017), Table 3 reports the ratio experimental and code provisions. It can be noted that there is
a better correlation in terms of maximum load capacity for the SCI P413 (2017). However, the
value provided by both design code are very restrictive to provide an adequate safe level with
economic aspects. In particular, EN 1993-1-4 (2006) does not distinguish for prediction of the
maximum load of connections with two planes in shear with the central or end plate control-
ling the structural design. The EN 1993-1-4 resulted in maximum load equal to 48.26 kN and
31.29 kN for austenitic and ferritic steel, respectively. The SCI P413 (2017), which has an equa-
tion for each model, provides a load of 49.54 kN for the OP-A type connection and 63.51 kN
for the IP-A. And for the ferritic study cases, values equal to 27.19 kN and 34.86 for OP-F and
IP-F, respectively.
In general, the results reported in Table 3 show that the use of the reduced ultimate
resistance provided a conservative safe level for the bearing resistance. This issue is not

Figure 5. Comparing the experimental and numerical responses of Austenitic study case with lips.

Figure 6. Comparing the experimental and numerical responses of Ferritic study case with lips.

Table 3. Comparison with code provisions.


EXP/EM 1993-1-4 EXP/SCI P413

OP-A IP-A OP-F IP-F OP-A IP-A OP-F IP-F

1,43 1,40 1,72 1,52 1,39 1,07 1,96 1,37

1040
addressed to SCI P413 (20017) as was mentioned in previous sections. In contrast, this
last code establishes a conservative value for the bolted connection when the outer plates
are responsible by controlling of the structural design. In fact, the code is very conserva-
tive for the cases where stainless thin-plates can be susceptible to curling effect. Thus, the
SCI P413 (2017) is not efficiency for the cases with Ferritic steel grade. Another issue is
addressed to difference found among the Ferritic and Austenitic steel grades. In particu-
lar, the relation between the ultimate strength observed in tensile coupons tests is nor
proportional to bearing resistance in experimental. This was confirmed in Table 3 where
it is not linear correlation for both structural codes because the expression are propor-
tional to ultimate strength from materials.

5 CONCLUSIONS

Experimental and numerical analyses have been studied in order to evaluate the influence of
the curling effect, as well as, the use of the lips in bolted connection to mitigate this effect.
Thus, a set of four study cases were selected modifying the structural scheme in order to inves-
tigate the bearing resistance comparing both Austenitic and Ferritic stainless steel grades.
Concerning to outcomes observed, it was possible to reported that the performance of the lips
was more evident for the cases where Ferritic steel grade is employed. This fact is related to
high deformation capacity from the Austenitic steel grade. In fact, the deformed observed in
both experimental and numerical analyses provided a reduced out-plane displacement for the
Ferritic study cases minimizing the curling effects.
Comparing the outcomes with the expression given by design codes, it can be noted that the
codes presented lower values for the determination of the bearing resistance. This is related to
use of reduced ultimate strength from EN 1993-1-4 (2006) and a value very restricted for the
overlap connection subjected to curling effect in SCI P413 (2017). Therefore, these issues
should be investigated in further analyses with a large number of the variables.

REFERENCES

Abaqus 6.14 2014. “Theory Manual and Users Manuals”, Dassault Systèmes Simulia Corp.
Baddoo, N.R. 2008. A review of research, applications, challenges and opportunities. Journal of Con-
structional Steel Research 64: 1199–1206.
Cai. Y. & Young, B. 2014. Structural behavior of cold-formed stainless steel bolted connections. Thin-
walled Structures 83: 147–156.
EN 1993-1-4 2006. Eurocode 3 – Design of steel structures: Part 1–4: General rules – Supplementary
rules for stainless steel. Brussels: Europen committee for standardization.
EN 1993-1-8 2005. Eurocode 3 – Design of steel structures: Part 1–8: Design of joints. Brussels: Europen
committee for standardization.
Henriques, J., Batista, G. Tenchini, A., Vellasco, P., Lima. 2018. Overlap shear connections in bearing in
thin-walled stainless steel structures. In Eighth International Conference on Thin-walled Structures;
Lisbon, 24–27 July 2018.
Kim, T.S., Kuwamura, H., Kim, S., Lee, Y., Cho, T. 2009. Investigation on ultimate strength of
thin-walled steel single shear bolted connections with two bolts using finite element analysis. Thin-
walled Structures 47: 1191–1202.
Salih, E.L., Gardner, L., Nethercot, D.A. 2011. Bearing failure in stainless steel bolted connections.
Engineering Structures 33: 549–562.
SCI P413 2017. Design manual for structural stainless steel. 4rd Ed. The Steel Construction Institute,
Building series.
Silva, G., Silva, A., Vellasco, P., Lima. L. 2016. Structural and economic assessment of stainless steel
power transmission towers. Metálica – Portuguese Steelwork Association Magazine 1: 12–18.
Sobrinho, K., Rodrigues, M., Tenchini, A., Vellasco, P., Lima, L., Henrique, J. 2018. Avaliação do efeito
curling em ligações aparafusadas em aço inoxidável submetida a esforço de tração uniaxial. Simpósio
de mecânica computacional; Vitória, ES novembro de 2018.

1041
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Experimental and analytical study of Cold-Formed Steel (CFS)


single-stud walls sheathed with FCB, CSB and MgO under
compression

Chanchal Sonkar
Academy of Scientific & Innovative Research (AcSIR), India
CSIR-Central Building Research Institute, Roorkee, India

Achal Kr. Mittal


CSIR-Central Building Research Institute, Roorkee, India

Sriman Kr. Bhattacharyya


Department of Civil Engineering, IIT Kharagpur, West Bengal, India

Sachin Kumar
CSIR-Central Building Research Institute, Roorkee, India

Abhinav Dewangan
National Institute of Technology, Raipur, India

ABSTRACT: Cold-formed steel members are widely used in residential, industrial and com-
mercial buildings as primary load bearing elements. It has been extensively used in lightweight
framing of low and mid-rise residential constructions. Limited studies have been carried out on
CFS wall panels sheathed with Fiber Cement Boards (FCB), Calcium Silicate boards (CSB), and
Magnesium Oxide (MgO) Boards under compression. In the present study, CFS single-stud walls
with FCB, Heavy Duty FCB (HDFCB), CSB, and MgO boards are experimentally tested under
axial loading applied at a constant rate. Analytical modelling of CFS single-stud wall panels with
the respective sheathing is carried out by calculating stiffness that the fastener-sheathing system
supplies to the stud as bracing. Elastic buckling analysis of the stud with sheathing based springs
is completed in CUFSM (Constrained and Unconstrained and Finite Strip Method) software ver-
sion 4.05, consecutively axial load carrying capacity has been calculated using Direct Strength
Method as per AISI S100. Analytical results seem to be in good agreement with the experimental
results and the ratio between the two is varying from 0.84 to 1.06. There is significant amount of
increment in the axial strength of CFS single-stud wall panels with the use of sheathing, which
acts as a restraint. The maximum increase in axial strength is due to HDFCB boards, which is
found to be 87% and 149% in one-sided and two-sided sheathed specimen respectively. The results
obtained are interesting and useful for the research, academic and industrial community working
in the area of CFS.

1 INTRODUCTION

Cold-formed steel (CFS) wall framing systems, which have the advantage of being environmen-
tally friendly, light-weight, aesthetically good and easy to construct. CFS are generally utilized as
load-bearing structural components in low- and mid-rise CFS structures and non-load bearing
structural components in other residential, commercial and industrial buildings. Previously,
numerous studies are conducted for assessing the behavioral pattern of the CFS wall-panel
(CFSWP) subjected to compression underneath axial loading, such as, Ye et al. 2016 investigated

1042
a complete of sixteen full-scale tests on CFS wall specimens with varied configurations of boards,
including fire-retardant GB, MgO board, OSB and CSB. The influence of different sheathing
boards, layer of sheathing boards, CFS stud section and spacing, and the joint detail of the
CFSWPs were tested. Vieira et al. 2012 evaluated the CFS studs with sheathings under the axial
load and found that that the load bearing CFS stud walls has positive impact on the stability and
strength because of the bracing supplied by sheathing. Schafer et al. 2010 performed the buckling
analysis of CFS steel members using the general boundary conditions in CUFSM. Vieira et al.
2012 gave the expression for the lateral stiffness of the sheathed of CFS wall-panels. Schafer et al.
2010, had investigated the rotational restraint and distortional buckling in the CFS framing
system. Tian et al. 2004 had analyzed different screw connection for the CFS structures sheathed
with the gypsum and cement-based boards.
The objective of the paper is to study the effect of different types of sheathing on the axial
strength of CFS single-stud wall panels as sheathing produces the confinement of the frame
which gives rise to strength of the CFSWP. Four types of sheathing boards i.e. Fiber Cement
Board (FCB), Heavy Duty Fiber Cement Board (HDFCB), Magnesium-Oxide Boards (MgO)
and Calcium Silicate boards (CSB) has been used for the present study. Screw spacing of
300mm has been kept constant for all the specimens. The study has been validated by using
analytical model utilized by AISI S-100 and Vieira & Schafer (2013).

2 EXPERIMENTAL STUDY

A test set-up has been fabricated and installed on 300-ton Instron UTM available in Struc-
tural Engineering Laboratory, CSIR-CBRI for testing of CFS single-stud wall panels under
compression. Tests set-up consists of 02 nos. of top and bottom plate girders designed using
IS 800: 2007 for a load of 200 Tons as shown in Figure 1.

2.1 Specimen configurations


Axial load tests are performed at CSIR-CBRI Roorkee. In this work nine CFS single-stud wall
specimen configurations as discussed in Table 1 were prepared in the laboratory and tested under
axial loading condition. CFS single-stud bare frame (W1) was fabricated using C-section (89 × 44
× 13 × 1 mm) and U-section (92 × 50 × 1 mm) as shown in Figure 2. One-sided and two-sided

Figure 1. Axial Load Test Set-up. (a) Schematic Diagram of set-up. (b) Experimental set-up.

1043
Table 1. Specimen details.
S. Sheathing Screw
No. Specimen Configuration Schematic Plan View Position
No. Layer Spacing

1 W1 CFS single-stud —– —– —–
2 W2 CFS single-stud with one-
1-0 one-side 300
sided FCB (8mm) sheathing
3 W3 CFS single-stud with one-
sided HDFCB (9mm) 1-0 one-side 300
sheathing
4 W4 CFS single-stud with one-
1-0 one-side 300
sided CSB (12mm) sheathing
5 W5 CFS single-stud with one-
1-0 one-side 300
sided MgO (10mm) sheathing
6 W6 CFS single-stud with two-
1-1 two-side 300
sided FCB (8mm) sheathing
7 W7 CFS single-stud with two-
sided HDFCB (9mm) 1-1 two-side 300
sheathing
8 W8 CFS single-stud with two-
1-1 two-side 300
sided CSB (12mm) sheathing
9 W9 CFS single-stud with two-
sided MgO (10mm) sheathing 1-1 two-side 300

Figure 2. CFS element cross-section. (a) Stud Cross-section. (b) Track Cross-section.

sheathed specimens (W2-W9) included four different types of sheathing FCB, HDFCB, CSB and
MgO boards as shown in Figure 3. The sheathing was kept 8.0 mm below the top track so that
when the load is applied on the panel, the load doesn’t come directly on the sheathing and whole
is restricted to the top track only.

2.2 CFSWP fastener configuration


Two types of screws were utilized for fabrication of specimens as shown in Figure 4. The self-
drilling dry wall screws of diameter 4.0 mm and length 24 mm; used to assemble the steel
studs with the top and bottom tracks. The self-drilling bugle head screws of diameter 4 mm
and length 32 mm; used for assembling the CFS frame and sheathing.

1044
Figure 3. Specimen configuration details. (a) Isometric view single CFS single-stud; (b) Isometric view
of one sheathed CFS single-stud frame; (c) Elevation & Plan of CFS single-stud frame.

Figure 4. Screw used for connecting CFS frame members; (a) Self-drilling dry wall screws (assemble the
steel studs with the top and bottom tracks). (b) Self-drilling bugle head screws (assembling the CFS
frame and sheathing).

2.3 CFSWP material configuration


The CFS frame members are fabricated using a galvanized steel sheet of nominal yield strength
of 350 MPa. The physical properties of the CFS material has been determined through tensile
coupon test as per IS 1608-2005 is observed to be an average of 403.56 MPa for tensile strength
and elastic modulus of 200.23 GPa . As per the literatures considered, assumed engineering
properties of the sheathing boards are {1} CSB (12 mm thick) U = 0.16, E = 6500 {2} FCB
(8 mm thick) U = 0.18, E = 5500 {3} HDFCB (9 mm thick) U = 0.18, E = 9000; {4} MgO
(10 mm thick) U = 0.20, E = 4500; Where, U = Poisson’s ratio, and E = Elasticity Modulus.

3 LOADING PROTOCOL

Loading was applied at a rate of 0.5 mm/min using 300 Ton Instron UTM available at Heavy
Testing Laboratory at CSIR-CBRI Roorkee. Test set-up was designed for a uniformly distrib-
uted loading on the top track. The specimen was fixed at the base.

4 INSTRUMENTATION

As shown in Figure 5, the compressive axial loading applied on the CFS single-stud walls with
and without sheathing by 300 Ton Instron UTM controlled. The application loads are

1045
Figure 5. Instrumentation for experiment. (a) Laser Displacement Sensors (L.D.S.) placed to record the
displacement. (b) Fixity of the CFS single-stud specimen at the bottom using MS plate and C-clamps.

measured by the load cells fixed with actuator. Laser displacement sensors were placed on the
specimen in-plane displacements as shown in Figure 5(a). The bottom track of the specimens
is fixed to the bottom plate girder with the help L-shape plates and C-clamps (Figure 5(b)).

5 ANALYTICAL STUDY OF CFSWP SUBJECTED TO COMPRESSION

Analytical modelling of CFS wall panels with sheathing is carried out by the method described
by calculating stiffness that the fastener-sheathing system supplies to the stud as bracing. The-
oretically, the fastener-sheathing system supplies three translational and three rotational
springs at every fastener location bracing the stud. Practically, a more limited set of springs in
the plane of the cross section, consisting of lateral translation (kx ), which is in the plane of the
sheathing, vertical translation (ky ), which is out of the plane of the sheathing, and rotational
stiffness (k ), which is in the plane of the cross section, are the most important. The springs
restrain against buckling modes associated with weak-axis flexure and torsion, while k
springs restrain flange (AISI S-100 and Vieira & Schafer 2013). The expressions for determin-
ing these restraint parameters are:

Lateral Translational Stiffness ðkx Þ


kx ¼ 1=ð1=kxl þ 1=kxd Þ ð1Þ
(Vieira & Schafer (2013))
Out-of-Plane Translational Stiffness (ky ) ðEIÞW π4 df
ky ¼ ð2Þ
(Vieira & Schafer (2013)) L4
Rotational Stiffness (k )
k ¼ 1=ð1=kc þ 1=kw Þ ð3Þ
(Vieira & Schafer (2013))
3Ed 4 t3 π
Where, kxl ðlocal translational stiffness) ¼ (Green et al. 1947; Winter
4t2board ð9d 4 π þ 16tboard t3 Þ
π Gtboard df wtf
2
1960); kxd ðdiaphragm translational stiffness) ¼ ; kc ¼ 0:00035Et2 þ 75;
L2
ðEIÞ
kw ¼ df w ; E ¼ Young’s modulus of the steel stud; d = fastener diameter; t ¼ flange thickness;
tboard ¼ board or sheathing thickness; G ¼ Shear modulus of the sheathing; wtf ¼ fastener tribu-
tary width; df ¼ distance between fasteners; L ¼ sheathing height; ðEI ÞW ¼ sheathing rigidity

1046
5.1 Elastic stability and direct strength method
The elastic stability of CFS stud is significantly altered by the presence of sheathing. Elastic
buckling analysis of the stud with sheathing based springs is completed in CUFSM (Con-
strained and Unconstrained and Finite Strip Method) software version 4.05. The basic method
proposed for strength determination is to correct the elastic buckling loads for the presence of
the sheathing and then to use existing design expressions; either the Direct Strength Method
(DSM) of AISI S-100, or the Effective Width Method (EWM) of the main specification of AISI
S-100 to find the strength. Consecutively axial load carrying capacity has been calculated using
Direct Strength Method as per AISI S100, according to which the elastic buckling loads; Local
Buckling (Pcrl ), Distortional Buckling (Pcrd ), and Global Buckling (Pcre ) will be obtained as per
the member strength determination procedure given by Vieira & Schafer (2013) using the prod-
uct of elastic buckling load factors and the squash load, Py obtained from CUFSM analysis to
get the elastic buckling loads. Then for predicting the nominal strength of the wall-panel by
DSM as per section A1.2.1 of AISI S-100. The nominal strength of the wall-panel will be deter-
mined by equation number (7) which is finally termed as predicted ultimate load in this study.
Global Strength,
 qffiffiffiffiffiffi
0:658λc Py
2

Pne ¼ for λc  1:5 and λc ¼


Py
ð4Þ
0:877Pcre for λc > 1:5 Pcre

Local-global interaction,
8
> Pne # for λc  0:776 sffiffiffiffiffiffiffiffi
<"  0:4  0:4 Pne
Pnl ¼ Pcrl Pcrl and λl ¼ ð5Þ
: 1  0:15 Pne
>
Pne
Pne for λc > 0:776 Pcrl

Distortional Buckling,
8
> Py # for λc  0:561 sffiffiffiffiffiffiffiffiffi
<"  0:4   Py
Pnd ¼ P crd Pcrd 0:6 and λd ¼ ð6Þ
>
: 1  0:25 Py for λc 40:561 Pcrd
Py Py

Nominal Strength,
Pn ¼ minðPne ; Pnl ; Pnd Þ ð7Þ

6 RESULTS AND DISCUSSIONS

The ultimate loading capacity of single-stud was 20.95kN. Flexural-torsional buckling was
observed at failure in the W1 specimen (Figure 6(a)). The increase in the axial strength of one-
sided CFS single-stud wall panels due to FCB, HDFCB, CSB and MgO were basically 77.08%,
87.11%, 66.25 and 82.71% respectively. The major mode of failure in all the one-sided CFS single-
stud wall panels was flexural-torsional buckling. In W2, W3, W4 and W5 location of buckling at
650 mm (Figure 6(c)), 950 mm (Figure 6(d)), 1100 mm (Figure 6(g, h)) and 800 mm (Figure 6
(j, k)) from the top track respectively. Local lip buckling was also observed at various locations in
both W2 and W3. Screw pull-through was observed in W2 and W5 (Figure 6(l)). Board separation
and cracking was observed at the peak load in W3 (Figure 6(e)) and W5 (Figure 6(l)) respectively.
Local web buckling was observed in W4 at the top stud-track junction (Figure 6(f)).
The increase in the axial strength of two-sided CFS single-stud wall panels due to FCB,
HDFCB, CSB and MgO were basically 110.16%, 149.02%, 141.19%and 108.40% respectively. In
W6 local flange buckling was observed at 600 mm (Figure 6(p)) from the top track. In W7, W8

1047
Figure 6. Failure pattern observed in different specimen.
(a) W1: FT-failure (b) W2: LB-failure (c) W2: FT-failure (d) W3: FT-failure
(e) W3: BS-failure (f) W4: WLB (g), (h) W4: FT-failure (i) W4: WLB
(j), (k) W5: FT-failure (l) W5: BC & SPT- failure (m), (n) W7: DB-failure (o) W7: LB in Stud
(p) W6: FB-failure (q), (r) W6: WLB (s) W6: BC-failure (t) W6: WLB
(u), (v) W8: DB-failure (w), (x) W8: LLB (y) W9: DB-failure
*note: FT: flexural-torsional; LB: local-buckling; BS: board-separation; WLB: web local-buckling; BC: board-
cracking; DB: distortional-buckling; LLB: lip local-buckling; SPT: screw pull-through; FB: flange local-buckling.

and W9 distortional buckling occurred at location of 300 mm (Figure 6(m, n)), 400 mm (Figure 6
(u, v)) and 2050 mm (Figure 6(y)) from the top track respectively. Local lip buckling observed at
several locations in W6 and W8 (Figure 6(w, x)). local buckling in stud (Figure 6(o)) at the
bottom stud-track junction W7. Local buckling waves were also generated in all the specimens
near the peak load. No screw pull-through and board cracking has been observed in W4 and W7.
Among the non-heavy-duty one-sided sheathed specimens, axial load carrying capacity W5
is highest whereas in two-sided specimens axial load carrying capacity of W7 is highest.

1048
Table 2. Test results.
Ultimate Load Predicted to
% increase in ultimate
Specimen Experimental Predicted Experimental test
load w.r.t. W1 specimen
Test (kN) (kN) ratio

W1 20.95 22.20 1.05 -


W2 37.10 39.45 1.06 77.08
W3 39.20 39.53 1.00 87.11
W4 34.83 33.83 0.97 66.25
W5 38.29 39.03 1.01 82.76
W6 44.03 43.78 0.99 110.16
W7 52.17 43.81 0.84 149.02
W8 50.53 43.36 0.86 141.19
W9 43.66 43.02 0.98 108.40

Therefore, MgO board performs well in one-sided configuration and CSB board is performs
better is two-sided configuration in terms of strength. The ultimate loading capacity increase
in W3 and W7 HDFCB sheathed specimen are 87% and 149% in comparison with CFS single-
stud. In this study the HDFCB sheathed specimen is found to be best performing sheathed
specimen among all the sheathing boards as it is as compressed boards of higher density and
elastic modulus. The major concern in this study behind fixing the length of all the specimen
to a single length is to observe the influence of type of sheathing and its elastic modulus on the
occurrences of failure pattern. The influence of asymmetrical and symmetrical boundary con-
dition imposed by application of sheathing boards on either side of CFSWP.

7 CONCLUSIONS

An axial compression test of 09 nos. of CFS single-stud walls was conducted to study the effects
of sheathing types on axial strength. The ultimate load of the wall studs obtained from experi-
ments was compared with that from analytical method discussed by Viera & Schafer 2013. It was
observed that there is significant amount of increment in the axial strength of CFS single-stud
wall panels with the use of sheathing. The following conclusions were drawn from this work:

1. The increase in the axial strength of one-sided CFS single-stud wall panels due to FCB,
HDFCB, CSB and MgO were basically 77.08%, 87.11%, 66.25 and 82.71% respectively.
2. The common buckling failure patterns observed in CFS single-stud bare frame and one-
sided sheathed specimens is flexural-torsional failure.
3. In one-sided sheathed specimens, constraints are only applied in one of the flanges of the
stud section, which gives origin for an eccentricity in the model, which leads to widening
of flanges resulting into flexural torsional failure.
4. The increase in the axial strength of two-sided CFS single-stud wall panels due to FCB,
HDFCB, CSB and MgO were basically 110.16%, 149.02%, 141.19%and 108.40%
respectively.
5. The type of failure pattern observed in the both side sheathed specimen is the distortional
failure because the stud has been constrained in both vertical direction and on both the
flanges which gave rise for occurring failure as distortional buckling.
6. Analytical results seem to be in good agreement with the experimental results and the ratio
between the two is varying from 0.84 to 1.06. It can be said that the discussed analytical
method by Vieira & Schafer 2013 can be used for prediction of axial strength of sheathed
CFSWP.
7. Among the non-heavy-duty boards, MgO and CSB boards carried maximum axial load
in one-sided and two-sided CFS sheathed single-stud wall panels respectively. Whereas in
the case of heavy duty FCB boards the axial strength approximately increased by 87%
and 149% for single side and both sides sheathing respectively.

1049
REFERENCES

AISI S100, 2016. North American Specification for Cold-Formed Steel Structural Members.
Green, G. G., Winter, G., and Cuykendall, T. R. 1947. Light gage steel columns in wall-braced panels.
Rep. 35, Pt. 2, Engineering Experiment Station, Cornell Univ., Ithaca, NY.
IS 800 (Indian Standard). 2007. General Construction in Steel- Code of Practice.
IS 1608 (Indian Standard).2005. MetallicMaterials-TensileTesting at Ambient Temperature.
Li, Z. and Schafer, B.W. 2010. Buckling analysis of cold-formed steel members with general boundary
conditions using CUFSM conventional and constrained finite strip methods.
Schafer, B.W., 2008. The direct strength method of cold-formed steel member design. Journal of con-
structional steel research, 64(7–8), pp.766–778.
Schafer, B.W., 2012. CUFSM 4.05–finite strip buckling analysis of thin-walled members. Baltimore,
USA: Department of Civil Engineering, Johns Hopkins University.
Schafer, B.W., 2013. Sheathing Braced Design of Wall Studs.
Schafer, B.W., Vieira Jr, L.C., Sangree, R.H. and Guan, Y. 2010. Rotational restraint and distortional
buckling in cold-formed steel framing systems. Revista Sul-americana de Engenharia Estrutural, 7 (1).
Vieira Jr, L.C.M. and Schafer, B.W. 2012. Lateral stiffness and strength of sheathing braced cold-formed
steel stud walls. Engineering Structures, 37, pp.205–213.
Vieira Jr, L.C.M. and Schafer, B.W. 2012. Behavior and design of sheathed cold-formed steel stud walls
under compression. Journal of Structural Engineering, 139 (5),pp.772–786.
Vieira, L. C. M., Jr. 2011. Behavior and design of sheathed cold-formed steel stud walls under compres-
sion. Ph.D. thesis, Johns Hopkins Univ., Baltimore.
Tian, Y.S., Wang, J., Lu, T.J. and Barlow, C.Y. 2004. An experimental study on the axial behaviour of
cold-formed steel wall studs and panels. Thin-walled structures, 42(4),pp.557–573.
Winter, G. 1960. Lateral bracing of beams and columns. J. Struct. Div. 102(1),77–92.
Ye, J., Feng, R., Chen, W. and Liu, W. 2016. Behavior of cold-formed steel wall stud with sheathing
subjected to compression. Journal of Constructional Steel Research, 116, pp.79–91.

1050
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Lateral-torsional buckling of stainless steel beams with slender


cross section

M. Šorf & M. Jandera


Czech Technical University in Prague, Prague, The Czech Republic

ABSTRACT: The contribution shows experimental and numerical research of welded slender
stainless steel I-section beams. Difference in behavior of stainless steel and common carbon steel
members is generally known, but design of stainless steel members has been established mainly
for hollow sections (CHS, SHS/RHS) as these are the typical stainless steel profiles. Currently,
open sections are also being used in structures and the design rules for both local buckling of
very slender sections as well as lateral torsional buckling reduction factors are based on very
limited experimental and numerical research. For this reason new research covering these phe-
nomena was started at the Czech Technical University in Prague. A numerical model was used
for design of test arrangement, the geometrically and materially nonlinear analysis with imper-
fection was made in software Abaqus. The experimental program consisted of six stainless steel
beam tests being used for a model validation. The tests employed two stainless steel materials
(austenitic and ferritic steel), one section slenderness (Class 4 for web and flange) and three
beam slenderness. A parametric study based on the validated numerical model will be used to
compare existing design procedures and their possible refinement for Class 4 sections.

1 INTRODUCTION

1.1 Stainless steel


Over the last several decades, popularity in use of stainless steel for structures increases. This
material is a specific type of steel which is highly alloyed containing more than 10.5% of chro-
mium. Austenitic, ferritic and duplex stainless steel are the most used stainless steel types in
structural applications. These groups are different in strength, ductility, weldability, toughness
and the ability to resist corrosive environment as result of using various alloying elements in
varying amount. The main reason for limited use is the initial cost of the material, which differs
significantly for each group of stainless steel and is much higher than for common carbon steel.
The important difference from common carbon steel is the material stress-strain diagram
behavior. Instead of typically linear behavior up to a visible yield strength for carbon steel,
the stress-strain curve for stainless steel has a more rounded response with no clearly defined
yield point. The material nonlinearity is the reason for the need of other design procedures for
stainless steel structures as the stiffness is reduced by yielding below the 0.2 % (yield) proof
strength and strain hardening is usually much higher.

1.2 Existing design procedure


Generally, open cross sections subjected to bending around the major axis with unrestrained
or partly restrained compressed flange or compressed web tend to fail with influence of lateral
torsional (global) buckling. Whereas slender cross section resistance may be governed by plate
(local) buckling. Both phenomena influence significantly design of steel beams.

1051
The resistance of beam subjected to lateral-torsional buckling should be determined according
to EN 1993-1-4, where supplementary rules are given for structural stainless steel, as follows:
Mb;Rd ¼ χLT Wy fy =γM1 ð1Þ

where Wy = Wpl, y for Class 1 or 2 section; Wy = Wel,y for Class 3; Wy = Weff,y for Class 4;
fy = yield strength; χLT = reduction factor accounting for lateral torsion buckling:

χLT ¼ 1
1 ð2Þ
LT þ ½LT 2  λLT
0;5
2

2
LT ¼ 0; 5ð1 þ αLT ðλLT  0; 4Þ þ λLT Þ ð3Þ
sffiffiffiffiffiffiffiffiffiffiffi
Wy fy
λLT ¼ ð4Þ
Mcr

where Mcr = critical moment; αLT = imperfection factor suggested as 0,34 for cold-formed and
hollow sections and 0,76 for open welded or other sections. As far as the authors are aware,
there were no experimental data for welded open sections when the procedure was codified.
Beams of comparatively stocky sections were tested and modelled by (Wang et al. 2014) and
(Yang et al. 2014). A wider test program is being prepared at KU Leuven, where preliminary
results were published (Fortan et al. 2016).

2 NUMERICAL MODELLING

A numerical model was primary used for the design of test arrangements and for the develop-
ment of an experimental program. The geometrically and materially non-linear analysis with
imperfection (GMNIA) was made in software Abaqus using shell element S4R (four-node
shell element with reduced integration).
The section resp. member imperfections were assumed by the lowest elastic buckling eigen-
mode (Figure 1) for local resp. global buckling. The geometric (local) and global imperfection
were determined as 0.8 of the fabrication tolerance. Amplitude for local imperfection was con-
sidered as 0.8b/100, where b is the web height. For global imperfection, there was considered
imperfection amplitude by value of 0.8L/750, where L is the distance between two cross-sections
where lateral torsional buckling is prevented. The amplitude for eigenmode with greater value
of αcr was reduced by 0.7 times as suggested in EN 1993-1-5.

Figure 1. Eigenmodes for buckling – local (left), global (right).

1052
3 EXPERIMENTS

3.1 Experimental program


The prepared experimental program consisted of 6 four point bending tests. There was
selected only one cross section (Class 4) made of two stainless steel grades and with three dif-
ferent lengths (2.4, 3.8 and 5.4 m span). Summary of beams is given in Figure 2, where the
section geometry, as well as the section and beam slenderness are given.
It can be seen that the test arrangement was identical for all beams and similar to the used
by (Prachař et al. 2016), with modification in the beam span and the distance between lateral
restraints only. The beam was supported at its ends under the lower flange. Lateral restraints
were used at ends and points of loading for both the upper and lower flange (Figure 3).

3.2 Material properties


Furthermore, several material tests were also carried out, measured material properties are
shown in Table 1. A two-stage Ramberg-Osgood material diagram (Gardner et al. 2004) was
used for description of the stress-strain diagram and used for validation of numerical models.

3.3 Measurement of imperfections


Before each test, the initial local and global imperfections were measured along three lines on
the web and along both edges of the upper flange. An example of measured amplitudes for the
longest beam is shown in Figure 4.

4 FE MODEL VALIDATION

The geometrically and materially nonlinear numerical model with imperfections (GMNIA) in
software Abaqus was used and validated on the tests. Section resp. member imperfections
were assumed by the lowest elastic buckling eigenmode for local resp. global buckling. The
amplitudes of imperfections were taken as measured for each specimen. The number of elem-
ents was selected as 10 per width of the flange and 33 per height of the web.

Figure 2. Experimental program overview.

1053
Figure 3. Test arrangement and beam failure modes.

Table 1. Material properties of tested specimens.

Type Modulus of R-O parameter


Yield strength Ultimate strength
of elasticity parameter n`0.2.1.0
steel [MPa] [MPa] [GPa] [-] [-]

1.4016 307.016 430.698 201.371 8.10 1.80


1.4016 - 306.83 - 429.36 - 202.86 - 7.90 - 1.79
1.4016 306.642 428.029 204.343 7.70 1.77
1.4301 297.787 637.413 195.693 5.60 2.30
1.4301 297.345 297.99 622.003 632.50 197.339 196.07 5.50 5.63 2.35 2.34
1.4301 298.834 638.079 195.188 5.80 2.37

In addition, the true stress-strain diagram was calculated from the measured material prop-
erties and implemented in the model. As no residual stresses were measured, the residual stress
pattern for welded stainless steel open I sections proposed by (Yuan et al. 2014) was used in
the model (Figure 5).
Specimens after tests can be seen in Figure 6, load-deflection diagrams for each beam are
shown in Figure 7.
Table 2 shows the comparison of the beam resistance obtained from experiment, FE model and
from the design procedure according to EN1993-1-4 with using measured material properties.
Failure modes for test 1 and test 3 are shown in Figure 8. The load-deflection curve comparison
of results from FE model and from the experiment for the shortest as well as for the longest beam
is in Figure 9, 10, where the influence of imperfections and residual stresses is also demonstrated.
Generally, the numerical model predicts the real behavior well as the predicted resistance is
in average 11,8% lower. Especially for longer beams, the prediction was very accurate. For
shorter beams, where the resistance is more influenced by local imperfections, the differences
in beam resistance prediction are higher. The difference is contributed to the simplification in
the local imperfection shape being considered by the first elastic buckling eigenmode. The real
shape of imperfection is more favourable for the beam resistance. More precise imperfection
modelling is planned in the future.
Furthermore, can be stated that existing design procedure from EN 1993-1-4 provides safe
but inaccurate results as can be seen in Table 2 and further improvement is possible.

1054
Figure 4. An overview of imperfection measurement including the scheme of measured points.

Figure 5. Implemented residual stresses.

1055
Figure 6. Beams after test.

Figure 7. Load-deflection diagrams for all tested beams.

Table 2. Beam resistance comparison.


Test Material Li LLTi Mb Rd[KNm]

[-] [-] [m] [m] Test FE model EN 1993-1-4

1 1.4301 5400 2500 76.1 77.0 1.2% 49.0 35.6%


2 1.4301 3800 1500 89.8 75.8 15.7% 62.2 30.8%
3 1.4301 2400 1000 90.8 77.8 14.4% 69.3 23.7%
4 1.4016 5400 2500 83.3 79.4 4.6% 50.6 39.3%
5 1.4016 3800 1500 98.7 79.9 19.0% 64.2 35.0%
6 1.4016 2400 1000 97.4 81.9 15.8% 71.5 26.6%

1056
Figure 8. Test and FE model failure mode for test 1 (right) and for test 3 (left).

Figure 9. FE model comparison for test 1.

5 CONCLUSIONS

Behavior of slender open section stainless steel beams loaded by major bending moment is
investigated in the paper. Numerical models created in software Abaqus using GMNIA were
validated based on experimental data. Furthermore, the comparison of results obtained
according to the existing design procedure was made.

1057
Figure 10. FE model comparison for test 3.

The following research will focus on more precise imperfection modelling for greater match
of results from experiment and FE model. Subsequently, the design procedure of stainless
steel welded open cross section resistance in bending as well as the lateral-torsional buckling
curves will be investigated in a numerical parametric study. Initial part of the study is expected
for publication on the conference.

ACKNOWLEDGEMENT

The support of the GAČR 17-24769S “Nonlinear stability and strength of slender structures
with nonlinear material properties” are gratefully acknowledged.

REFERENCES

Gardner, L., Nethercot, D. 2004. Experiments on stainless steel hollow sections - Part 1: Material and
cross-sectional behavior. Journal of Constructional Steel Research Vol. 60: 1291–1318.
Wang Y., Yang L., Gao B., Shi Y., Yuan H. 2014. Experimental Study of lateral-torsional buckling behav-
iour of stainless steel welded I-section beams. International Journal of Steel Structures 14, 411–420.
H.X. Yuan, Y.Q. Wang, Y.J. Shi, L. Gardner. 2014. Residual stress distributions in welded stainless steel
sections. Thin-Walled Structures: 79, 38–51.
Yang L., Wang Y. Q., Gao B., Shi Y. J., Yuan, H. X. 2014. Two calculation methods for buckling reduc-
tion factors of stainless steel welded I- section beams. Thin-Walled Structures: 83, 128–136.
Fortan M., Zhao O., Rossi B. 2016. Lateral torsional buckling of welded duplex stainless steel I section
beams. Sixth International Conference on Structural Engineering, Mechanics and Computation (SEMC
2016), Cape Town, South Africa.
Prachař, M., Hricák, J., Jandera, M., Wald, F., Zhao, B. 2016. Experiments of Class 4 open section
beams at elevated temperature. Thin-Walled Structures: 98(1), 2–18.

1058
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Improve load capacity calculations by considering realistic


imperfections induced by welding for plates and shells

C. Stapelfeld, B. Launert & H. Pasternak


Chair of Steel and Timber Structures, Brandenburg University of Technology, Germany

N. Doynov & V.G. Michailov


Department of Joining and Welding Technology, Brandenburg University of Technology, Germany

ABSTRACT: The topic of this paper is the application of an analytical numerical hybrid
model for a realistic prediction of weld imperfections. At the beginning, the analytical model,
its physical basis as well as the physical interrelationships are explained. This is followed by
the explanation of the coupling procedure between the analytical model and the numerical cal-
culation. The significance of this approach is proven by an application on a large welded
structure with several welds and a comparison of the calculated distortions with measure-
ments. Afterwards, the hybrid model is applied on the investigated stiffened structure for the
determination of the weld imperfections. An ultimate load analysis gives information about
the load carrying behavior under axial loading. The results are compared with the results of
ultimate load analysis from a literature example assuming different eigenvalues and scaling.
The comparison underlines the potential additional utilization of load bearing capacity by this
new approach.

1 INTRODUCTION

Welding distortions and residual stresses in terms of geometrical and structural imperfections
can significantly affect the quality and the stability behavior of welded structures such as stiff-
ened plates or shell. To simplify, in numerical strength calculations by the finite element
method (FEM) both types of imperfections, the geometrical and structural ones, are typically
combined being considered as equivalent geometrical imperfections. Values for standard cases
are included in EN 1993-1-5 in case of for example plated structures (Eurocode, 2017). Add-
itionally, global and local imperfections are to be distinguished. The global imperfections can
be defined as a bending of the whole structure with a maximum depending on its dimensions.
The local imperfections are mostly determined through a numerical eigenvalue analysis. The
challenge of a numerical ultimate load analysis is the detection of the lowest ultimate load by
the combination of different local and global imperfections. Numerous research works deal
with the numerical load capacity calculation of stiffened plates and shells (Beg 2010, Manco
2016, Degée 2008, Ghavamia 2006, Tran 2014). However, it remains unclear to some extend
how accurate these equivalent geometrical imperfections represent the actual deformations
and residual stresses caused by welds, especially for complex cases. The significance of numer-
ical load capacity calculations could be increased enormously if these imperfections were
known more exactly and could be considered directly during the computation.
Nowadays, the deformations and residual stresses can be determined by means of
a thermomechanical FE-simulation. This approach contains two steps: the calculation of the
transient temperature field followed by the calculation of the mechanics. Generally, such cal-
culations can provide very realistic and accurate values. However, the effort for the calibration
and validation of these simulations is huge and temperature-dependent thermophysical and
thermomechanical material properties are required. In addition, relevant structures and weld

1059
length are very large what leads to enormous calculation times and a huge demand of storage
capacity (Stapelfeld 2009). Simplified numerical approaches are available and able to remedy
this situation. However, the application of these models partly requires more expertise than
a conventional thermomechanical FE calculation [Duan 2007] or the simplifications are so
extensive that the weld imperfections calculated by the approach partially loses their validity
(Thikomirov 2008).
To answer the industrial needs, different modeling approaches based on the inherent strain
method exist. However, in the most works the leading hypothesis is that the inherent strain
depends only on the welding process and the material. The influence of the structure on the
inherent strain is seldom considered and limited to the structural strength (Ueda 1989, Ueda
1993, Mun 2011, Murakawa 2013).
The analytical-numerical approach (hybrid model) realizes a strong coupling between the
local and global mechanical effects (Michailov 2011). Thus, it can consider the interactions in
the structure, like the welding sequence or the variation of the structural stiffness as well as the
clamping conditions during the assembling process. Therefore, the model provides an uncompli-
cated and sufficient solution for calculating welding distortions. Currently the approach has
been successfully applied to ship and railway carriage structures, considering welding and ther-
mal straightening (Michailov 2011, Michailov 2014, Doynov 2018a, Doynov 2018b).

2 THE ANALYTICAL NUMERICAL HYBRID MODEL

2.1 General description of the approach


The basic idea of the coupled analytical numerical shrinkage force model is the linking of the
major advantages of both, analytical and numerical procedures. On the one hand, the match-
less marginal calculation time of the analytical shrinkage force model and its simple applica-
tion, and on the other hand, the possibility to conduct a FE simulation to calculate distortions
and residual stresses at any location of complex welded structures. All the determining factors
on quality and quantity of weld imperfections are initially passed to an analytical calculation
program, capturing the mathematical approach of the shrinkage force model. The program
calculates the shrinkage volume wpz, the inherent strain components εx, longitudinal and εy,
transversal to the weld as well as their centroids zc,x and zc,y for every single weld. The calcu-
lated entities are then transferred to the global finite element model in order to predict the
distortions of the structure after every welding operation (i.e., welds, assembly steps, changing
of fixtures, etc.). The calculated residual stresses in the domain of the next weld caused from
the previous welding operations serves as initial condition for the analytical calculation of the
shrinkage volume. It allows simultaneous variation of the geometry and clamping conditions
in the FE analysis. Thus, the welding sequence as well as the changing of geometry and clamp-
ing conditions can be taken into account. Finally, the results of the application of the hybrid
model are superposed with additional fabrication tolerances followed by a numerical load cap-
acity calculation (Figure 1).

2.2 The analytical model


The analytical solution for the calculation of the inherent strains in longitudinal and transver-
sal direction is derived from the shrinkage force model (Okerblom 1955, Kuzminov 1974,
Gatovskii 1980). This model contains a one-dimensional problem formulation for the longitu-
dinal strain and a two-dimensional plane stress problem formulation for the transverse one.
The following assumptions are made:
• The plane section hypothesis is valid;
• The width of the plastic zone is significantly smaller than the width of the plate;
• The welds are long (i.e., quasi-stationary temperature field);
• The welding speed is relatively high (i.e., fast moving heat source).

1060
Figure 1. Schedule of the load capacity calculation taking into consideration realistic weld imperfections.

The weld imperfections depend significantly on the maximum temperatures that every point verti-
cal to the weld direction is exposed to and the stiffness of the structure. Equations for the calcula-
tion of the maximum temperatures were derived by Rykalin (Rykalin 1957) constituting the basis
of the shrinkage force model. Two border cases are considered: a fast moving line source in a thin
plate with vertical isotherms penetrating the plate and a fast moving point source on a semi-
infinite body with circular isotherms around the weld. The proportion of each of the two border
cases to the resulting temperature field in the investigated weld structure depends on the heat
input per unit length, the geometrical properties as well as the heat exchange with the environ-
ment and is calculated iteratively. Considering further influencing factors, an axial force Fx is cal-
culated corresponding to the heat effect of welding (Kuzminov 1974):
rffiffiffiffiffi
2 α
Fx ¼ lnð2Þ qs EKχδ Kk Kσ ; ð1Þ
πe cρ

with the thermal expansion coefficient α, the specific heat capacity c, the density ρ and the
Young’s modulus E. Kχδ, Kk and Kσ are capturing the temperature field in medium thick
plates, the stiffness of the weld structure and the effect of existing stresses in the weld. Further-
more, the transversal shrinkage force Fy is calculated as follows:
"    2  2 ! !#
α lnð2Þ lnð2Þ 2 2  
Fy ¼ q s E þ 1 Kδ þ 1 Kav K 1 þ Kμ Kδ 
cρ e e 3π 3π ð2Þ
εF
 ½1 þ KC ð1 þ Kδ Þ
KW Kδ þ qs ð1  Kδ Þ;
θ

where Kδ captures the degree of heating through the thickness, Kav captures the influence of
stiffening cross-beams, Kμ determines the effect of longitudinal strains on the plastic transver-
sal strains, Kv considers the stretching of the temperature field, Kc is the degree of excessive
heat and Kw captures the effect of forced heat exchange. εF is the yield strain and θ is the pro-
portionality factor between the heat input per unit length and the cross section of the weld.
The axial shrinkage force Fx, eq. 1, is proportional to the width of the plastic zone:

Fx
bPZ ¼ : ð3Þ
εm Eδ

Here, εm is the averaged yield strain and δ is the plate thickness. The appropriate points of
action zc are equal to the centre of the zone of plastic deformations. They are significantly
influenced by the material and its properties as well as the heat input per unit length and the
plate thickness. Depending on the points of action, an equivalent linear strain distribution
or respectively stress distribution over the plates thickness can be calculated. Considering

1061
the point of origin in the centre of gravity of the plate’s cross section, the strain distribution
ε(z) follows as:

12εm zc
εðzÞ ¼ εm þ z: ð4Þ
δ2

2.3 The coupling procedure


For the coupling of the analytical shrinkage force model with the FE simulation, different
kinds of mechanical loads are available. The deformation state calculated by applying loads
and appropriate points of actions or alternatively eccentric pressures matches well with experi-
mental results [Stapelfeld 2016]. However, the calculated stress state in the structure is qualita-
tively wrong. Instead, specifying strains or stresses linearly distributed over the plates
thicknesses, (Figure 2), leads to correct stresses with tension in the weld and balancing com-
pressive stresses in the nearby regions. The procedure is already validated and verified with
Ansys®, LS Dyna®, Sysweld® and Abaqus®. When loading the FE model with longitudinal
and transversal stresses σ, the Poisson’s ratio ν must be considered:
εx;y E þ εy;x E
σx;y ¼ : ð5Þ
1  2

2.4 Application of the hybrid model on real structures


The hybrid model for calculating weld imperfections, mostly distortions, was already valid-
ated by means of numerous experimental welded butt and T-joints. The geometry, the mater-
ial as well as the welding technique and welding parameters were varied during the
experimental studies. The calculated distortions in the longitudinal and transversal direction
as well as the bending and the angular distortions always showed a good agreement with the
measured data (Stapelfeld 2016).
For the demonstration of the feasibility of the suggested model, the calculation procedure is
applied to a deck structure (20 m × 16 m) build from ship steel grade A (mild structural steel),
(Figure 3). The structure is welded with 109 welds in three manufacturing steps. For each step
measurements of the plane displacements at the plate edges were carried out. In the first
manufacturing step, 24 deck beams with thicknesses of 6 and 7 mm are welded to a 5 mm
thick base plate by Laser-MSG-Hybrid welding. The second manufacturing step involves the
joining of perforated stringers by double-sided manual GMAW welding. To simplify, the per-
foration is modelled by reduced thicknesses of the stringers being 6.5 mm and 5.6 mm. In the

Figure 2. Coupling procedure by means of linearly distributed strains (Stapelfeld 2016).

1062
Figure 3. Sketch of a large welded ship structure.

Figure 4. Scatter bands and averages of the experimental data compared with the calculations.

last manufacturing step, smaller parts and the walls are joined, again using manual GMAW
welding. The process parameters for the welding processes were assumed from the welding
procedure specifications provided.
The investigated structure is almost symmetrical and thus the welding of each single beam
as well as each single stringer has a similar effect on the measured welding distortions in longi-
tudinal and transversal direction. Therefore, the results of the measurements alongside the
seven paths in x direction and nine paths in y direction and the results of the hybrid calcula-
tion are summarized in scatter bands, (Figure 4). The real base plate is made up of several
small plates welded together. The resulting stresses and strains are not documented, thus the
calculated transversal distortions of manufacturing step 1 are too large. The results of the fur-
ther steps shows a good agreement.

3 CALCULATION OF WELD IMPERFECTIONS OF THE CURVED STIFFENED


STRUCTURE

3.1 The finite element model of the curved stiffened structure


For the application of the hybrid model and a subsequent load capacity calculation, a panel
similar to the panels of the Confluence bridge in France, introduced in [Tran 2014], was
chosen, (Figure 5). For defining the boundary conditions, a cylindrical coordinate system was

1063
Figure 5. Geometrical properties and boundary conditions of the curved panel.

created. The four edges of the panel are simple supported, uR ¼ 0. To restrain any movement
or rotation of the structure, two nodes in the middle of the curved edges are fixed alongside θ
and one node in the middle of the panel is fixed in the direction of the z coordinate. The steel
grade is S355. Correspondingly, the Young’s modulus is E ¼ 210 GPa, Poisson’s ratio is
 ¼ 0:3 and the yield strength is σF ¼ 355 MPa. A slope of E=100 is assumed for the harden-
ing and the ultimate stress is assumed to be reached at 470 MPa. The structure is discretized
with four node shell elements (S4R in Abaqus®) with three integration points over the elem-
ents thickness. A fine mesh with an approximated element edge length of 30 mm is used.
Considering the different thicknesses, the panel and the stiffeners are loaded with different
shell edge loads targeting a homogeneous axial pressure. For the verification of the loading as
well as the boundary conditions, the curvature of the plate was removed and a load capacity
calculation was carried out. The numerically calculated plastic normal force was about
27 MN and matches the analytical solution: F ¼ σF A ¼ 27:26 MN. For the determination of
the lowest ultimate load, several load deformation calculations were then executed under the
consideration of different buckle modes as well as combinations of them. Here, it turned out
that the imperfection according to the first buckle mode leads to the lowest load capacity.

3.2 Application of the hybrid model for the calculation of weld imperfections
The calculation of the mechanical loads with the analytical shrinkage force model requires
information about the type of joint and its dimensions, the material data as well as the welding
technique and the welding parameters. Here, the assumption is made that the stiffeners are only
welded one-sided. The material data correspond to the material data of the steel S355J2. The
welding technique is conventional MAG-welding with welding parameters targeting a fillet weld
with a design throat thickness of about 6.6 mm, which resulted in a heat input per unit length
of qs ¼ 2400 J/mm. In the example case, the linking between the analytical model and the sub-
sequent numerical calculation was done by linearly distributed initial stresses, Table 1.
For the proper loading of the FE model, two sections were created at each weld, representing
the idealized zone of plastic deformations. The width of the sections is specified by the width of
the analytically calculated plastic zone. When creating the single sections, the position of the
welds is considered by an offset of the center of the idealized plastic zone of half of the stiffeners

Table 1. Analytically calculated stresses for the numerical calculation of weld imperfections.
Curved panel Stiffener

Width of plastic pone bpz in mm 29.1 29.1


Location of the integration points Upper Middle Lower Upper Middle Lower
Longitudinal stresses σz in MPa 849.6 597.4 345.2 807.7 593.8 379.9
Transversal stresses σ in MPa 1313 801 290 1341 789 237

1064
Figure 6. Assembly of the FE-Model in the region of the welds (a) and out of plane deformations (b).

Figure 7. Asymmetrical deformation state of weld scenario 1(a) and symmetrical deformation state of
weld scenario 2 (b).

thickness, (Figure 6a). The initial stresses in longitudinal and transversal direction are defined
as “initial state” in Abaqus®. Geometric nonlinear behavior is considered within the elastic cal-
culation. The calculated deformation state indicates buckling of the curved panel in the positive
z-direction and a tilt over of the stiffeners towards the welds, (Figure 6b).
The calculation of the weld imperfections was done assuming two different positions of the
welds. In the first scenario, all the welds were done on same side of the stiffeners which leads
to an asymmetric deformation state, (Figure 7a). In the second scenario, all the left stiffeners
are welded on the left side and the right ones on the right side, (Figure 7b). This weld sequence
leads to a symmetrical deformation state.

4 RESULTS OF THE ULTIMATE LOAD CAPACITY CALCULATIONS

The ultimate load calculations were done assuming three kinds of imperfections. The first case
is a combination of the calculated geometrical and structural weld imperfections caused by weld
scenario 1 and manufacturing tolerances captured by a global buckle of b=1000 ¼ 4:8 mm. In
the second case, all imperfections in the structure are caused only by the eight welds being
welded according to weld scenario 1. The same calculation is again done with weld scenario 2
as a third case. The asymmetric and symmetric initial deformation state leads to asymmetric
and symmetric deformation states at ultimate load, (Figure 8). The comparison of the calcu-
lated ultimate loads with the results of a conventional calculation considering the scaled critical
first buckling mode shows an increase in load capacity of 16 % up to 57 %, (Figure 8).

1065
Figure 8. Out of plane deformation at ultimate load for weld scenario 1 (a) and weld scenario 2 (b).

Figure 9. Load displacements curves: three kinds of imperfections compared with the state of the art
calculation.

5 CONCLUSIONS

Numerous applications of the analytical numerical hybrid model for the calculation of weld
imperfections indicate the significance of the results. In the case of consideration, the calcu-
lated weld imperfections correlate well with known empirical values. The use of calculated
realistic weld imperfections instead of equivalent geometrical imperfections in the load cap-
acity calculation of the stiffened curved panel results in an increasing of the ultimate load in
the range of 16 % till 57 %. The approach gives the opportunity to analyze the influence of
different weld scenarios on the load capacity. A general statement on the effects of weld
imperfections in load capacity calculations cannot be given, because they are depending on
numerous influencing factors. A realistic consideration will be only possible by the application
of physically based approach of such type that has been presented in this paper. The future
potential of this method is thus very high. However, the approach in combination with load-
ing calculations has to be validated and calibrated by means of more experimentally deter-
mined load displacement curves. These works are ongoing at the moment.

ACKNOWLEDGEMENTS

These works are part of the IGF project No. 19173 BR of the German Research Association
for Steel Application (FOSTA). This project is kindly funded by the German Federal Ministry
of Economic Affairs and Energy (BMWi) by the AiF (German Federation of Industrial
Research Associations) as part of the program for support of the Industrial Cooperative
Research (IGF) on the basis of a decision by the German Bundestag.

1066
REFERENCES

Beg, D., Kuhlmann, U., Davaine, L. et al. 2010. Design of Plated Structures, Ernst W. & Sohn Verlag,
Berlin.
Degée, H., Kuhlmann, U., Detzel, A. et al. 2008. Der Einfluss von Imperfektionen in dünnwandigen
geschweißten Rechteckquerschnitten unter Druckbeanspruchung, Stahlbau 74 (4): 257–265.
Duan, Y.G., Vincent, Y., Boilot, F. et al. 2007. Prediction of welding residual distortions of large struc-
tures using a local/global approach, Journal of Mechanical Science and Technologie 21 (10): 1700–
1706.
Doynov, N., Michailov, V.G., 2018a. Distortion analysis of heat spot straightening thin-walled welded
structures: part 1: formation of the plastic deformation zone, Int J Adv Manuf Technol, 94: 667–676.
Doynov, N., Michailov, V.G., 2018b. Distortion analysis of heat spot straightening thin-walled welded
structures: part 2: analytical-numerical approach’, Int J Adv Manuf Technol, 95: 469–478.
Eurocode 3, 2017. Design and construction of steel structures (EN 1993-1-5).
Gatovskii, K.M., Karkhin, V.A, 1980, Theoria svarochnih deformaciy i napriajeniy, LKI, Leningrad (in
Russian).
Ghavamia, K., Khedmatib, M.R. 2006. Numerical and experimental investigations on the compression
behaviour of stiffened plates, Journal of Constructional Steel Research 62: 1087–1100.
Kuzminov, S.A. 1974, Svarochnie deformacii sudoviech korpusniech konstrukcii. Sudostroenie, Leningrad
(in Russian).
Manco, T., Martins, J.P., Rigueiro, C. et al. 2016. Numerical Analysis of Stiffened Curved Panels under
Compression, Eight International Conference on Steel and Aluminium Structures, Hong Kong.
Michailov, V.G., Doynov, N., Stapelfeld, C. et al, 2011. Hybrid model for prediction of welding distor-
tions in large structures’, Front. Mater. Sci., 5(2): 209–215.
Michailov, V.G., Stapelfeld, C., Doynov, N. 2014. Upgrade of an analytic-numerical hybrid model for the
distortion simulation of large structures, FOSTA VV mbH Düsseldorf (in German).
Mun, H.S., Jang, C.D. 2011. Prediction of welding deformation of hull panel blocks using an advanced
inherent strain analysis method considering the heat equivalent layer effect’, Met Mater Int 17(6): 993–
1000.
Murakawa, H., Okumoto, Y., Rashed, S. et al. 2013. A practical method for prediction of distortion pro-
duced on large thin plate structures during welding assembly’, Weld World 57: 793–802.
Okerblom, N.O 1955. Svarochnie napriajenia v metallokonstrukciiah, Moskau/Leningrad,Mashinostroenie
(in Russian).
Rykalin, N.N 1957. Berechnung von Wärmevorgängen beim Schweißen, VEB Verlag Technik, Berlin.
Stapelfeld, C. 2016. Vereinfachte Modelle zur Schweißverzugsberechnung, PhD Thesis, Berichte des
Lehrstuhls Füge- und Schweißtechnik der BTU Cottbus-Senftenberg, Band 10, Shaker Verlag,
Aachen.
Stapelfeld, C., Doynov, N. und Michailov, V. 2009. Hybride Berechnungsansätze zur Prognostizierung
und Minimierung des Verzugs komplexer Schweißkonstruktionen. Sysweld Forum 2009: 91–105.
Weimar.
Thikomirov, D., Rietman, B., Kose, K. et al. 2008. Computing Welding Distortion: Comparison of Dif-
ferent Industrially Applicable Methods” SHEMET 11: 195–202.
Tran, K.L., Douthe, C., Sab, K. et al. 2014. Buckling of Stiffened Curved Panels Under Uniform Axial
Compression, Journal of Constructional Steel Research 103: 140–147.
Ueda, Y, Kim, K., Yuan, M.G. 1989. A predicting method of welding residual stress using source of
residual stress (report I), Transactions of JWRI 18 (1): 135–141.
Ueda, Y, Yuan, M.G.1993. Prediction of residual-stresses in butt welded plates using inherent strains,
Journal of Engineering Materials and Technology – Transactions of the ASME115 (4): 417–423.

1067
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Plastic collapse loads of rectangular plate assemblies with constant


and linear load distribution

S. Stehr & N. Stranghöner


Institute for Metal and Lightweight Structures, University of Duisburg-Essen, Essen, Germany

ABSTRACT: The storage content of a rectangular silo causes loads normal to its vertical
walls which might lead to a plastic collapse of the structure in case of unfavourable design. The
corresponding design standards are EN 1993-1-7 and EN 1993-4-1. In the frame of the current
revision of EN 1993-1-7 it is envisaged to provide the designer with analytical expressions for
the determination of the plastic collapse load of rectangular plate assemblies under transverse
loads. For this reason, numerical investigations have been carried out at the Institute for Metal
and Lightweight Structures of the University of Duisburg-Essen regarding the determination of
plastic collapse loads of rectangular plate assemblies made of carbon steel. Herein, the plastic
collapse loads were evaluated using the methods of the modified Southwell plot and the Conver-
gence Indicator Plot. Within this contribution, first results are presented for a constant and
a linear load distribution as the extreme load cases of the Janssen load distribution.

1 INTRODUCTION

Silos can be built as with a circular or rectangular/square plan-form. Within this contribution,
rectangular silos made of carbon steel are covered which have a square plan-form. Rectangu-
lar plate assemblies provide an easy mounting and are of advantage in case only a limited area
is given. Silos may serve for storage of e.g. liquids or solids. However, depending on the stor-
age content, the load distribution normal to the vertical walls of the silo differs. Loads caused
by fluids may be represented by a distribution that corresponds to a linear slope over the fill-
ing height of the silo. In principle, the load resulting from a solid can be represented by the
Janssen distribution as described for silos in EN 1991-4. The Janssen distribution is limited on
the one side by a linear and on the other side by a constant load distribution, see Figure 1.
These two limiting load cases are considered in this contribution.
EN 1993-4-1 provides the principles for the structural design of steel silos of circular or rect-
angular plan-form. Besides this, EN 1993-1-7 should be used for the determination of the
resistance of a silo that is loaded by out of plane actions. Herein, the design rules for unstif-
fened or stiffened plates as part of a plate assembly such as a silo are given. In this contribu-
tion, the ultimate limit state of plastic collapse (LS 1) is considered. EN 1993-1-7 defines this
plastic collapse as the limit state where the plate assembly cannot longer resist loads without
a plastic mechanism arises and furthermore, excessive plastic deformations occur. Hence, the
maximum load that can be achieved is the plastic collapse load usually based on small deflec-
tion theory. In addition to a stress-based design by the Von Mises equivalent stress, supple-
mentary rules for the design by global analysis are given. Using a materially nonlinear
analysis, the load may be incrementally increased until the plastic collapse load is achieved.
Additionally, internal stresses and deflections based on small deflection theory can be calcu-
lated for plates loaded by a uniform pressure or a central patch.
In the frame of the current revision of EN 1993-1-7, numerical and analytical investigations
have been carried out at the Institute for Metal and Lightweight Structures of the University
of Duisburg-Essen regarding the determination of plastic collapse loads of rectangular plate
assemblies made of carbon steel. The final objective is to derive analytical expressions for the

1068
Figure 1. Linear, Janssen and constant load distribution in a rectangular silo made of a plate assembly.

determination of plastic collapse loads depending on the load distribution for implementation
in EN 1993-1-7. Within this contribution, first numerical and analytical results are presented
for constant and linear load distribution.

2 NUMERICAL MODEL

2.1 General
All finite element investigations were carried out by materially nonlinear analyses (MNA) using
four-node shell elements based on an ideal elastic-plastic stress-strain curve without any strain-
hardening defined by the following parameters: Young’s modulus E ¼ 210,000 N/mm², yield
strength fy ¼ 235 N/mm² and Poisson’s ratio ν ¼ 0.3.
From each FE simulation, firstly, (1) the maximum load factor was taken for the evaluation
of the plastic collapse loads. Secondly, due to difficulties in accurately determining the largest
load factor from the FE results, two additional methods have been applied: (2) modified
Southwell plot (Holst et al. 2005) and (3) Convergence Indicator Plot (Doerich & Rotter
2011) whereby the two latter ones rely on (1), see details in section 3.

2.2 Load cases


The simulated load cases result from possible storage loads on the vertical wall of a silo. As
already mentioned, the first load case is a linear load distribution which represents liquids and
the second load case is given by a constant load distribution. Thus, these two basic load cases
cover the upper and lower limit of the Janssen pressure distribution which is usually used to
determine storage loads caused by solids.
Regardless of any geometrical conditions, the initial load p0 was chosen to 1⋅10-3 N/mm2 for
each load case, see Figure 1. The self-weight of the plate assembly was considered in each
simulation, although the effect of the self-weight on the results is not decisive. The resulting
plastic collapse loads are given as a factor of the chosen load case.

2.3 Geometrical conditions


Each numerical simulation was carried out using a rectangular plate assembly model consist-
ing of four vertical rectangular plates each with the same thickness t1 ¼ t2 ¼ t3 ¼ t4 ¼ t, see
Figure 1. Neither a top nor a bottom plate were considered. All plate assembly models have
a square plan-form based on the horizontal dimensions d1 ¼ d2 and a height of d3.

1069
The first variable parameter is the height d3. Starting with a cubic plate assembly (d1 ¼ d2 ¼ d3),
five different d3/d1-ratios with 0.5, 1.0, 1.5, 2.0 and 3.0 were investigated. Within these investiga-
tions, the dimension d1 always remained unchanged. The thickness t of the plates was set as
the second variable parameter with d2/t-ratios from 50 to 500.

2.4 Boundary conditions


Three different sets of boundary conditions according to the definitions of EN 1993-1-6 were
investigated as illustrated in Figure 2. The first set of boundary conditions BC1f - BC1f
includes BC1f on the bottom as well as on the top edges of the plate assembly. BC1f is defined
by restrained displacements normal to the plate surface and in the plane of the plate surface
but with free rotation. The second set of boundary conditions BC1f - BC2f differs from the
first one in the chosen boundary condition on the top edges with BC2f (pinned), which is in
contrast to BC1f defined by free displacements in the plane of the plate. The third set of
boundary conditions BC1r - BC2f includes BC1r on the bottom edges and BC2f on the top
edges of the plate assembly. BC1r (clamped) is defined by restrained displacements normal to
the plate surface and in the plane of the plate surface and restrained rotation. In each case, the
four vertical edges of the plate assembly were fixed to each other so that all internal forces
could be transferred.

3 NUMERICAL DETERMINATION OF THE PLASTIC COLLAPSE


LOAD FACTOR ξpl,FE

In the frame of the numerical investigations, the load factor ξpl,FE has been determined in rela-
tion to the applied load p0 ¼ 1⋅10−3 N/mm2 describing the achievable load at which plastic
collapse occurs. For this, first, numerical load factor-displacement curves were obtained from
the finite element analyses at the point of extreme displacement at the height h and d2/2, see
Figure 3a. Herein, the highest load factor is ξpl,FE.
As not all load factor-displacement curves show a clear plateau, in some cases it is difficult
to determine the largest load factor ξa ¼ ξpl,FE accurately. Due to a possible deviation in the
accuracy of interpreting the results, the methods of the modified Southwell plot (MS) (Holst
et al. 2005) and of the Convergence Indicator Plot (CIP) (Doerich & Rotter 2011) have been
applied additionally.
For the modified Southwell plot (MS) the gradient at each point of the load factor-
displacement curve was plotted by ξa/δ2,x, see Figure 3b. This modification provides
a clearer view on the elastic range and the plastic collapse load factor. The elastic range is
presented by the vertical part on the right side followed by the elastic-plastic range in which
the gradient ξa/δ2,x decreases. Where the gradient ξa/δ2,x reaches its minimum, the plastic

Figure 2. Different sets of boundary conditions considered in the numerical investigations.

1070
Figure 3. Exemplary numerical load factor-displacement curve (a) and modified Southwell plot (b) for
h/d3 ¼ 0.5, d2/d1 ¼ 1.0, d3/d1 ¼ 1.0, d2/t ¼ 500, BC1f - BC2f and constant load distribution.

plateau is indicated. The interception point of the tangent at this point with the vertical axis
provides the value for the plastic collapse load factor ξpl,MS, see Figure 3b, which is more
precise than the value of ξpl,FE.
The Convergence Indicator Plot (CIP), illustrated in Figure 4a, is based on the previous
explained modified Southwell plot. Following the MS-procedure, the value of ξpl,MS is deter-
mined for each load factor increment ξa.
The CIP-curve indicates the plastic collapse load factor more accurate the closer ω, defined
in Equation 1, approaches a value of zero.

ξpl;MS  ξa
ω¼ ½
ð1Þ
ξa

The interception point of the tangent at the lowest value of ξpl,MS with the vertical axis pro-
vides ξpl,CIP, see Figure 4b, which is the most precise result for the plastic collapse load factor.
Values for ξpl,CIP were achieved for the constant and linear load distribution, for all three
boundary condition sets, illustrated in Figure 2, for d3/d1-ratios from 0.5 to 3.0 and for
d2/t-ratios from 50 to 500, always d2 ¼ d1, exemplary illustrated in Figure 5 for the constant
load distribution, d3/d1-ratios of 1.0 and 2.0 and BC1f - BC2f.

Figure 4. Exemplary Convergence Indicator Plot for h/d3 ¼ 0.5, d2/d1 ¼ 1.0, d3/d1 1.0, d2/t ¼ 500,
BC1f - BC2f and constant load distribution.

1071
Figure 5. Plastic collapse load factors ξpl,CIP for rectangular plate assemblies with BC1f - BC2f and
constant load distribution.

4 ANALYTICAL EXPRESSIONS

The final objective was to develop analytical expressions for the determination of the plastic
collapse load for all geometries, boundary conditions, and load cases in the scope of the inves-
tigations. For this purpose, the numerically derived load factors ξpl,CIP have been used as the
basis.
For the development of the analytical expressions, in a first step, each load case and boundary
condition have been treated independently. In a second step, a unified equation has been devel-
oped in which both load cases and three different boundary conditions are considered by simple
factors. Exemplary, in the following, the stepwise procedure of the development of the analytical
expression is presented for the constant load case and the boundary condition BC1f - BC2f.
Based on the values of ξpl,CIP, a first simple equation has been approximated by using the least
square method for each d3/d1-ratio to determine analytical load factors ξpl,calc, see Equation 2. In
principle, the values of ξpl,calc are calculated conservatively compared to the values of ξpl,CIP.
 2
t
ξpl;calc ¼ k  106  ½
ð2Þ
d2

where k is a factor depending on the d3/d1-ratio with k ¼ 3.9 for d3/d1 ¼ 0.5, k = 1.97 for
d3/d1 = 1.0, k = 1.56 for d3/d1 = 1.5, k ¼ 1.39 for d3/d1 ¼ 2.0 and k ¼ 1.25 for d3/d1 ¼ 3.0.
Based on these varying k-values, a more detailed equation has been approximated, again
using the least square method, which covers all investigated d3/d1-ratios, see Equation 3:

 2  2 !  2
d3 d3 d1 t
ξpl;calc ¼ k1 þ k2  þ k3  þ k4   10 
6
½
ð3Þ
d1 d1 d3 d2

where k1, k2, k3 and k4 are parameters depending on the boundary condition and load case,
see Table 1. Equation 2 and Equation 3 are valid for S235 as the numerical simulation was
based on fy ¼ 235 N/mm2.
All variants of Equation 3 provide load factors ξpl,calc that are conservative compared to the
values of ξpl,CIP obtained from the numerical analyses, see Figure 6. A statistical classification
has been carried out by determining the mean value correction factor b as well as the estimated
standard deviation sΔ according to EN 1990 on the basis of Equation 4 and Equation 5. Herein,
all boundary conditions, geometrical variations and both load cases were considered.

1072
Table 1. Parameters k1, k2, k3 and k4 for the effect of boundary condition and load case on the plastic
collapse load of a rectangular plate assembly.
Boundary condition at each end Parameters [-]

Load Case Bottom end Top end k1 k2 k3 k4

Constant BC1f BC1f 1.53 −0.215 0.0322 0.6161


BC1f BC2f 1.53 −0.215 0.0322 0.6161
BC1r BC2f 1.47 −0.165 0.0227 0.9
Linear BC1f BC2f d2/t 5 58.8 3.0 −0.546 0.0453 1.19
d2/t  58.8 3.02 −0.566 0.0638 1.19
BC1r BC2f d2/t 5 58.8 3.37 −0.782 0.0875 1.68
d2/t  58.8 3.03 −0.475 0.0418 1.89

Figure 6. Correlation of the load factors ξpl,CIP and ξpl,calc for rectangular plate assemblies with
BC1f - BC2f.

n 
P
ξpl;CIP;i  ξpl;calc;i
b ¼ i¼1 n  2 ð4Þ
P
ξpl;calc;i
i¼1
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 X n  2
sΔ ¼  Δi  Δ ð5Þ
n  1 i¼1

where
n is the number of ! numerical results,
ξpl;CIP;i
Δi ¼ ln is the logarithm of the error term and (6)
b  ξpl;calc;i
P
n
Δ ¼ 1n  Δi is the estimated mean value of Δi. (7)
i¼1

On this basis, for both load distributions, the mean value correction factor has been determined
to b = 1.01 and the estimated standard deviation to zero which are extraordinary good results.

1073
Exemplary curves of the analytically calculated plastic collapse load factors ξpl,calc based on
Equation 3 are illustrated in Figure 7 comparing both load distributions and Figure 8 compar-
ing different boundary conditions. The values for the plastic collapse load factors decrease with
increasing d2/t-ratios and increasing d3/d1-ratios. Furthermore, the plastic collapse load factors
for the linear pressure distribution are much higher than for the constant load distribution, see
Figure 7. The boundary conditions BC1f - BC1f and BC1f - BC2f yield to identical results for
the constant load case, see Figures 7 and 8. However, the plastic collapse load factors for BC1r -
BC2f are slightly higher than those for BC1f - BC1f or BC1f - BC2f as shown in Figure 8.

5 PROPOSED DESIGN CONCEPT

Based on Equation 3, which is approximated by the least square method, an analytical expression
for the determination of the plastic collapse load can be derived for design against the plastic fail-
ure limit state (LS1). As the design is based on pressure values, Equation 3 has to be generalized
taking into account the initial load p0, for which the load factors have been determined, to pro-
vide an analytical expression for the characteristic plastic reference resistance ppl,a,Rk for
a rectangular plate assembly with d1 = d2 and variable yield strengths. Equation 8 is the proposed
expression for the characteristic plastic reference resistance:

Figure 7. Plastic collapse load factors ξpl,calc for rectangular plate assemblies with BC1f - BC2f
comparing different load distributions.

Figure 8. Plastic collapse load factors ξpl,calc for rectangular plate assemblies with constant load distri-
bution comparing different boundary conditions.

1074
 2  2 !  2  
d3 d3 d1 t fy
ppl;a;Rk ¼ k1 þ k2  þ k3  þ k4   10 
3

d1 d1 d3 d2 235
 2  2 ! ð8Þ
d3 d3 d1 4:26  t2  fy
, ppl;a;Rk ¼ k1 þ k2  þ k3  þ k4  
d1 d1 d3 d22

with k1 to k4 according to Table 1. Herein, all investigated d3/d1-ratios, boundary conditions


and both constant and linear load distributions are covered.
Subsequently, the design value of the plastic reference resistance ppl,a,Rd of a rectangular plate
assembly of square plan-form with γM0 ¼ 1.0 for the plastic failure limit state (LS1) results to:

ppl;a;Rk
ppl;a;Rd ¼ ð9Þ
γM0

6 SIMPLIFIED ANALYTICAL MODEL

A simplification of Equation 8 is possible by relying the analytical expression on the plastic col-
lapse of an individual rectangular plate. Herein, the boundary conditions of this reference plate
are not of interest, but the correlation to the overall behaviour of the plate assembly. Exem-
plary, the simplified analytical model can be based on a reference plate with d2 ¼ d3 ¼ 1000 mm
and boundary conditions BC1f on all four edges. The characteristic value of the plastic reference
resistance of this individual rectangular reference plate is presented by Equation 10, which is
based on the expression for internal stresses of plates according to EN 1993-1-7:

fy  t2
ppl;p;Rk ¼ κ  ð10Þ
d2  d3

where κ is a value depending on the load case with κ 5.9 for constant and κ ¼ 11.325 for linear
load distribution.
Based on Equation 10 the characteristic value of the plastic reference resistance of
a rectangular plate assembly with d1 ¼ d2 can be determined by Equation 11:

ppl;a;Rk ¼ β  ppl;p;Rk ð11Þ

where β is the plastic reference resistance factor for a rectangular plate assembly, see Table 2.
Herein, the values of β result from simple correlations of the analytically determined plastic

Table 2. Plastic reference resistance factors β for a rectangular plate assembly of square plan-form.
Boundary condition at each end β [−]

Load Case Bottom end Top end d3/d1 = 0.5 1.0 1.5 2.0 3.0

Constant BC1f BC1f 2.81 1.42 1.12 1.0 0.9


BC1f BC2f 2.81 1.42 1.12 1.0 0.9
BC1r BC2f 3.6 1.61 1.21 1.05 0.92
Linear BC1f BC2f d2/t 5 58.8 2.82 1.39 1.06 0.9 0.71
d2/t  58.8 2.81 1.39 1.07 0.92 0.76
BC1r BC2f d2/t 5 58.8 3.65 1.64 1.22 1.02 0.82
d2/t  58.8 3.89 1.69 1.18 0.97 0.75

1075
collapse load factors ξpl,a,calc of a rectangular plate assembly with the analytically determined plas-
tic collapse load factors ξpl,p,calc of the individual rectangular reference plate, see Equation 12.

ξpl;a;calc
β¼ ð12Þ
ξpl;p;calc

7 CONCLUSIONS

The design of rectangular silos in the LS 1 can be simplified using analytical expressions for
the determination of plastic collapse loads. In this contribution, newly developed analytical
equations are presented which have been derived on the basis of numerical simulations for
constant and linear load distribution for rectangular steel silos of square plan-form. Further
investigations into the Janssen pressure distribution are currently carried out at the Institute
for Metal and Lightweight Structures at University of Duisburg-Essen in order to generate
general analytical expressions for implementation in the ongoing revision of EN 1993-1-7.

ACKNOWLEDGEMENTS

The research activities were supported by the work and comments of the CEN TC250 SC3
Project Team PT 05 “Shells and Related Structures”. Special thanks apply to Prof. J. Michael
Rotter (PT 05 Leader) and Chris J. Brown (PT 05 Member).

REFERENCES

Doerich, C. & Rotter, J.M. 2011. Accurate Determination of Plastic Collapse Loads From Finite Element
Analyses. Edinburgh: Institute for Infrastructure & Environment.
EN 1990:2002 + A1:2005 + A1:2005/AC:2010. Eurocode: Basis of structural design.
EN 1991-4:2006. Eurocode 1: Actions on structures – Part 4: Silos and tanks.
EN 1993-1-6:2007 + AC:2009 + A1:2017. Eurocode 3: Design of steel structures – Part: 1-6: Strength and
Stability of Shell Structures.
EN 1993-1-7:2007 + AC:2009. Eurocode 3: Design of steel structures – Part: 1-7: Plated structures subject
to out of plane loading.
EN 1993-4-1:2007 + AC:2009 + A1:2017. Eurocode 3: Design of steel structures – Part 4-1: Silos.
Holst, J.M.F.G., Doerich, C. & Rotter, J.M. July 2005. Accurate determination of the plastic collapse
loads of shells when using finite element analyses. Proc., Fourth International Conference on Advances
in Steel Structures, ICASS’05: pp. 1798-1794. Shanghai.

1076
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Stability of axially compressed cylindrical shells made of stainless


steel for different imperfection patterns

N. Stranghöner & E. Azizi


Institute for Metal and Lightweight Structures, University of Duisburg-Essen, Essen, Germany

ABSTRACT: The major difference between shells made of carbon steel and those made of
stainless steel is the nonlinear stress-strain behaviour of stainless steels below the 0.2 % proof
stress whereas carbon steel shows a typically linear elastic behaviour up to the yield stress. This
nonlinear stress-strain behaviour of stainless steel effects the buckling resistance of medium slender
shells. Furthermore, the shape of the imperfection has a significant influence on the buckling
resistance of shells. For these reasons, finite element investigations have been carried out validated
by experimental tests of axially loaded shells. Hereby, the parameters material nonlinearity and
imperfection pattern have been considered with the imperfection types (i) weld depression pattern
type A, (ii) sinusoidal (multiwave) imperfection and (iii) measured imperfections from test samples
for shell buckling tests. This paper presents the results of these investigations.

1 INTRODUCTION

Stainless steel shell structures are widely used in industrial applications with their key advan-
tages in corrosion resistance and durability. The load carrying behaviour of stainless steel
shell structures in pre- and post-buckling patch differs from those made of carbon steel due to
their different material characteristics (Hautala & Schmidt 1999; Gorbachov et al. 2017;
Stranghöner et al. 2019). While carbon steel can be characterised by a bilinear (elastic, per-
fectly plastic) stress-strain response, stainless steel is attributed by a rounded (nonlinear)
stress-strain behaviour below the 0.2 % proof stress with no sharply defined yield point. EN
1993-1-6 as the current European standard for the design of shell structures includes design
rules which were developed explicitly for carbon steel shell structures. Herewith, it does not
provide any specific design rules for the determination of the buckling resistance of cylindrical
shells made of stainless steel under meridional loading, but simply adopts the reduced tangent
modulus method which leads to an over-conservative design for slender and stocky shells.
Thin-walled shell structures most commonly fail by buckling under meridional compression
(Rotter 2004). Hautala & Schmidt (1998, 1999) already proposed slenderness and temperature
dependent buckling correction factors as a more efficient design approach specifically for medium
slender shells made of austenitic stainless steel under axial compression. Further experimental,
numerical and theoretical investigations into the shell buckling behaviour of meridionally loaded
duplex and ferritic stainless steel shells were carried out in the frame of the RFCS research project
BiogaSS (Thomas et al. 2017; Stranghöner & Gorbachov 2015; Gorbachov et al. 2017). The
objective was to develop buckling correction factors for austenitic, duplex and ferritic stainless
steel according to EN 1993-1-4 considering the different fabrication tolerance classes (FQCs) A to
C according to EN 1993-1-6. For the numerical simulation of the experimental tests of BiogaSS,
sinusoidal multiwave, single inward and outward imperfection patterns as well as measured imper-
fections of the test samples were considered. In these studies, it could be shown that multiwave
imperfection pattern reflects the experimental behaviour of the tested shells best. On this basis,
recent numerical investigations into the effect of the nonlinear material behaviour below the 0.2 %
proof stress have been performed which showed that the imperfections of the tested shells partly
superimposes the effect of the material nonlinearity on the buckling resistance so that the decrease

1077
due to the material nonlinearity became less with larger imperfections (Stranghöner et al. 2019).
Especially, this behaviour becomes obvious for FQC C with the largest imperfections.
However, the axisymmetric weld depression pattern type A as proposed by Rotter & Teng
(1989) was not included in this study. This imperfection pattern is widely used as a suitable
imperfection shape for exploring the imperfection sensitivity of cylindrical shells dominated
by meridional compression. Herewith, the focus of the presented study is on the comparison
of the effects of the weld depression pattern type A with the already investigated sinusoidal
(multiwave) imperfection pattern as well as measured imperfections on the buckling resistance
of meridionally loaded shells by numerical simulation of the experimental shell buckling tests
of BiogasSS (Thomas et al. 2017).

2 IMPERFECTION SENSITIVITY OF AXIALLY COMPRESSED SHELLS

Initial geometric imperfections are introduced into real shell structures (e.g. silos, tanks) during
the fabrication or the manufacturing process. Since the real failure behaviour of shell structures
under axial compression is strongly sensitive to imperfections, it is of exceptional importance to
incorporate the imperfections into the design process and herewith into numerical simulation. EN
1993-1-6 offers two types of nonlinear numerical analyses taking imperfections into account,
GNIA and GMNIA, either ignoring or including the influence of material nonlinearity. However,
the selection of a suitable imperfection pattern is one of the biggest challenges for designers,
because the imperfection sensitivity itself depends on the imperfection shape and this shape can be
changed independently of the boundary and load conditions as well as the shape of the shell (Jans-
seune 2016). A huge number of theoretical (e.g. Koiter 1945, Tennyson & Muggeridge 1969,
Hutchinson et al. 1971) and numerical (e.g. Rotter & Teng 1989; Schmidt & Winterstetter 2004)
investigations were carried in the past, aiming to study the effect of various imperfection patterns
on the buckling resistance of thin-walled shells. Since none of these studies were able to propose
or identify a universal applicable imperfection pattern (Jansseune 2016), three main philosophies
for classification and selection of an imperfection shape were proposed firstly by Schmidt (2000)
and subsequently in the Buckling of Steel Shells European Design Recommendations (Rotter &
Schmidt 2013) as follows:
• Most realistic imperfection shape: geometrical imperfections are considered in the FE
models as “realistic” as possible, whereby the deviation from the ideal-perfect shape of real
shell structure (Arbocz 1982; Song 2004; Coleman 1992; Teng et al. 2005) or of laboratory
shells should be measured for implementation into the numerical modeling. This approach
is not very common especially for practical use, as the design is usually carried out before
erection of a shell structure. Furthermore, it is highly time and cost consuming to extract
measured data and to apply them meaningfully in a numerical model.
• Worst possible imperfection shape: the imperfection shape can be adopted in simulations in
principle in such a way that shell buckling failures are reduced as much as possible to pro-
vide a safe lower bound for design. However, several analyses with different shapes must be
performed to compute and find the “worst possible” geometrical shape due to the fact that
the worst imperfection form depends on the shell geometry and loading, see Koiter (1963),
Greiner & Derler (1995). The application of this approach yields to an overconservative
buckling strength, which cannot be observed in practice and is not addressed as an econom-
ical solution for stability problems of thin-walled shells, see the historical background of
shell stability design rules in Schmidt (2018), Rotter (2017) and Rotter & Schmidt (2013).
• Simple equivalent imperfection shape: A simple “equivalent” shape can be implemented
into numerical shell models, which is perhaps not completely identical to the realistic imper-
fection pattern, but which is able to reduce the failure loads relatively in the same manner
as realistic imperfections. An equivalent imperfection pattern can either be extracted from
the failure modes of linear (e.g. Koiter 1945, Yamaki 1984) or nonlinear (e.g. Esslinger
et al. 1972, Schneider 2005) analyses considering the pre- and postbuckling region, or can
be assumed as a simple specific shape, e.g. sinusoidal (multiwave) or axisymmetric weld

1078
depression, which is representative for the surface deviations of real shell structures during
fabrication or execution. The last approach has been widely used as the most suitable
device to predict realistic failure loads by means of FE analyses.
In these studies, the focus was more on slender shells and their imperfection sensitivity under
meridional compression as they fail elastically and their stability is dominated by imperfections.
However, covering shells made of stainless steel, additional attention is necessary regarding
medium slender and slightly stocky shells, which fail by elastic-plastic buckling and where the con-
sideration of the material nonlinearity in the design process of stainless steel shell structures is
significant.

3 NUMERICAL MODELLING OF EXPERIMENTRAL STAINELSS STEEL SHELLS


OF BIOGASS

3.1 Introduction
In the frame of the presented study, the experimental shell buckling tests of BiogaSS were simu-
lated numerically using GMNIA analyses considering the three aforementioned imperfection
shapes: (i) weld depression pattern type A, (ii) the sinusoidal (multiwave) imperfection pattern and
(iii) measured imperfections. In total, 12 shell buckling tests were carried out under meridional
compression. The test samples were made of duplex (1.4462) and ferritic (1.4521) stainless steel
(Thomas et al. 2017, Gorbachov et al 2017, Stranghöner & Gorbachov 2015). The finite element
program ANSYS APDL v18.1 has been used for the numerical analyses. Assuming that
a bifurcation point would stay undetected by applying an arc-length method for nonlinear ana-
lyses (Riks 1979), it was required to ensure that the first reported negative eigenvalues of the
global stiffness matrix were detected in all geometrically nonlinear analyses. Thus, the nonlinear
equilibrium path near failure was automatically followed, aiming to avoid that no bifurcation
point with a negative global or local stiffness matrix stays undetected.

3.2 Geometry, material and boundary conditions


The geometries of the tested shells varied between diameters of 300 mm and 400 mm, wall thick-
nesses of 0.5/0.6 mm, 0.6 mm, 1 mm and 3 mm, a shell height of 350 mm and resulting r/t-ratios
between 50 and 400 (nominal dimensions), see Thomas et al. (2017), Stranghöner & Gorbachov
(2015). All numerical simulations were carried out considering the measured geometrical dimen-
sions and the measured true stress-train curves of the stainless steel plate materials which were
used for fabrication of the test shells (Thomas et al. 2017). Simply supported boundary conditions
of BC1f (all displacements restrained, but free to rotate about the circumferential axis) and BC2f
(free to displace meridionally and rotate about the circumferential axis) were assumed for the
bottom and top edge of the modelled cylindrical shells with BC1f and BC2f as defined in EN
1993-1-6.

3.3 Meshing details


For the determination of the optimal mesh size, preliminary studies were performed on perfect
cylindrical shells using a linear bifurcation analysis (LBA) as well as a geometrically nonlinear
elastic analysis (GNA). For the justification of the element density, a mesh refinement has been
progressively applied in the perfect length of the modeled shells, so that the LBAs or/and GNAs
provide approximately same results (max. difference 1 %). As a significant higher mesh resolution
is required to capture the local curvatures of a weld depression accurately, a mesh refinement (20-
time finer) was adopted in the vicinity and the local weld depression range. Four-node shell elem-
ents 181 with six degrees of freedom at each node were selected, whereby the element lengths in
the perfect range of the shell in longitudinal direction was connected to the classic buckling half-
wave and varied between 0.14 λcl to 0.4 λcl for r/t-ratios from 50 to 150 with

1079
pffiffiffiffi
π rt pffiffiffiffi
λcl ¼ 0:25
≈ 1:728 rt for  ¼ 0:3 ð1Þ
½12ð1  2 Þ

where λcl = classical axisymmetric buckle half-wavelength; υ = Poisson’s ratio, r = radius and
t = thickness of the shell.

3.4 Imperfection shape


EN 1993-1-6 gives only a general guide on the shape of imperfections. A sufficient number of
imperfection patterns (eigenmode affine, weld depression, collapse affine, post-buckling
affine, etc.) should be considered in a nonlinear buckling analysis of imperfect shells so that
the worst case can be identified. The definition of the imperfection shape in a unique and
repeatable manner is difficult, so that the judgement of the final calculation outcomes based
on nonlinear buckling analyses for an imperfect shell (e.g. GMNIA) can be considerably com-
plicated. Because of the above-mentioned reason, the eigenmode affine, collapse affine and
post-buckling affine patterns, which can be obtained from an LBA, GNA and GMNA ana-
lysis respectively, were not considered in this numerical study. Since all other mentioned
imperfection patterns seem to be suitable candidates for the nonlinear buckling analysis of
real imperfect shell structures, the already mentioned imperfection shapes (i) weld depression
pattern type A of (Rotter and Teng 1989), (ii) sinusoidal (multiwave) pattern and (iii) real
measured imperfections of the buckling test specimens were considered. The initial amplitude
and gauge length (wavelength) of the imperfections for the weld depression pattern type
A and sinusoidal imperfection pattern were assumed as defined in EN 1993-1-6.
EN 1993-1-6 distinguishes between three FQCs A to C with FQC A representing the least
and FQC C the worst imperfections. The imperfections should be considered depending on
the existing stress state which is in this case meridional compression. As the experimentally
investigated cylindrical shells were almost perfect in the meridional direction due to their
manufacturing process including subsequent rolling, FQC A was chosen for the numerical
investigations for defining the size of imperfection for weld depression type A and sinusoidal
imperfection. Besides this, additional numerical investigations have been carried out for FQCs
B and C for weld depression type A and sinusoidal imperfection.

4 COMPARISON OF EXPERIMENTAL AND NUMERICAL BUCKLING


STRENGTHS

Exemplary, experimental and numerical relative load-deformation diagrams of the test specimens
150-3-D1 and 150-1-D2 are given in Figure 1 considering all three kind of investigated imperfec-
tions. Both shells were made of duplex stainless steel 1.4462. Test sample 150-3-D1 is described
by a radius of r =151.04 mm, a wall thickness of t = 3.022 mm, a height of h = 349.31 mm, a r/
t-ratio of about 50, a relative slenderness of λx = 0.56 and a measured 0.2 % proof stress of
fy,exp,t-m = 789 N/mm2 in the transverse direction in the middle of the plate (measured values).
Test sample 150-1-D2 is described by a radius of r =150.23 mm, a wall thickness of t = 1.068 mm,
a height of h = 350.15 mm, a r/t-ratio of about 150, a relative slenderness of λx = 0.91 and
a measured 0.2 % proof stress of fy,exp,t-m = 759 N/mm2 in the transverse direction in the middle of
the plate (measured values). From Figure 1 it becomes obvious that the numerical simulations
based on all three kind of imperfection types simulate the ultimate experimental loads relatively
good. Nevertheless, for both shells, the simulations based on the weld depression type
A imperfection slightly overestimate the ultimate load and the sinusoidal imperfection leads to the
lowest ultimate loads comparable to the experimental ones. Shell 150-3-D1 has a relative slender-
ness of 0.56 where an influence of the material nonlinearity on the buckling strength is given
(Stranghöner et al. 2019). Even the simulation with measured imperfections yields to a higher
ultimate load for shell 150-3-D1. This can be explained by the fact that using measured geometric

1080
Figure 1. Experimental and numerical relative load-deformation curves for the two duplex test speci-
mens 150-3-D1 and 150-1-D2 with r/t = 50 and 150 respectively.

imperfections, material imperfections are not considered. Material imperfections are residual stres-
ses caused by rolling, pressing, welding, straightening etc., inhomogeneities and anisotropies etc.
EN 1993-1-6 specifies geometric imperfections for GMNIA and different FQCs which consider
both geometric and material imperfections. Shell 150-1-D2 with a relative slenderness of 0.91 is
more slender; in this case, the influence of the material nonlinearity is not significant, the shell
behaves elastic-plastically in the transition range to elastic failure. In a first step, from Figure 1 it
can be concluded that the sinusoidal imperfection type meets the ultimate loads best. Further-
more, the experimental load-deformation paths are not met by the numerical simulations as the
numerical deformations are always smaller than the experimental ones.
Figure 2 shows a diagram in which the relative experimental and numerical results are com-
pared for all 12 test samples considering FQC A and measured imperfections. Table 1 sum-
marizes the mean value correction factor b and estimated standard deviation sΔ which have
been determined according to EN 1990 with

n 
P 
Fx;exp  Fx;FEM
i¼1
b¼ n  2 ð2Þ
P
Fx;FEM
i¼1
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 X n  2
sD ¼  Di  D ð3Þ
n  1 i¼1

where
n is the number of experimental results,
 
Fx;exp
Di ¼ ln is the logarithm of the error term and ð4Þ
b  Fx;FEM

1 Xn
D¼  Di is the estimated mean value of Di : ð5Þ
n i¼1

1081
Figure 2. Experimental and numerical relative buckling strengths based on FQC A and measured
imperfections.

Table 1. Statistical evaluation acc. to EN 1990 for numerical prediction of ultimate relative buckling
strengths based on imperfections for FQC A and measured imperfections.
Mean value correction factor Estimated standard deviation
Imperfection pattern b [-] sΔ [-]

Weld depression type A 0.94 0.125


Sinusoidal (mulitwave) imperfection 1.01 0.132
Measured imperfection 0.96 0.145

Both parameters, the mean value correction factor b and estimated standard deviation sΔ, pro-
vide values of a comparable order of magnitude. The resulting mean value correction factors
show that a sufficient correlation can be achieved between the actual resistances of the tested
shells and the numerical predictions by adopting the considered imperfection patterns. On the
basis of these results, neither of the two investigated imperfection types can be claimed to be the
preferred one for the simulation of the experimental tests; on the contrary, both seem to be
equally well suited. Nevertheless, a mean value correction factor smaller 1 demonstrates that the
buckling strengths achieved by numerical simulation are not conservative. This means that using
a weld depression type A as an imperfection pattern for FE simulations might yield to slightly
unconservative values at least in the presented study for the experimentally tested shells.
Figure 3 presents the buckling correction factors ψ as defined in Hautala & Schmidt (1998,
1999) and Rotter & Schmidt (2013) as the ratio of two numerically determined buckling
strengths using the measured nonlinear stress-strain curve of the investigated stainless steel
shells and using an ideal elastic-plastic stress-strain response on the basis of the measured 0.2
% proof stress. The resulting buckling correction factors ψ show that the effect of the material
nonlinearity on the load capacity of duplex and ferritic stainless steel shells is reduced by
increasing the influence of imperfections from FQC A to C, see also Stranghöner et al. (2019).
Furthermore, it can be seen that there is nearly no influence of the imperfection type on the
buckling correction factor visible. Weld depression type A shows slightly smaller buckling
corrections factors compared to the sinusoidal imperfection but with deviations of about only

1082
Figure 3. Influence of the imperfection type on the buckling strength considering a bilinear or multi-
linear stress-strain relationship for FQC A to C according to EN 1993-1-6.

max. 0.3 %. Together with the fact that the sinusoidal imperfection yields to smaller ultimate
buckling strengths compared to the weld depression type A, the influence of the imperfections
pattern on the buckling correction factor has to be taken as fully negligible. Besides this, the
numerical ultimate loads of cylindrical shells with sinusoidal imperfection pattern are more
sensitive to wavelength and amplitude variations. Based on these results, it can be concluded
that the buckling resistance of cylindrical shells with weld depression type A is slightly more
sensitive to the material nonlinearity in comparison to that with sinusoidal imperfection pat-
tern although the effect is – for design purpose – negligible.
A comparison between the relative experimental and numerical failure loads taking into
account FQC A to C for all investigated shells is plotted in Figure 4. Both diagrams show that
FQC A meets mostly the experimental results best and FQC C yields to considerably lower
ultimate buckling strengths especially for those shells with higher relative buckling strengths. On
the basis of the presented study it can be concluded that in general both imperfection types,
weld depression type A and sinusoidal imperfection according to EN 1993-1-6, are suited to
simulate at least the experimental shell tests with a preference to the sinusoidal imperfection as
this type of imperfection leads to a slightly better fit. Nevertheless, this study does not give an
answer to the question which imperfection type reflects best real shell structures with their toler-
ances due to fabrication and erection. But it confirms the use of the weld depression type A as
a suitable imperfection in principle which is assumed to reflect best the imperfections in real
structures, see e.g. Teng & Rotter (1992).

5 CONCLUSIONS

The main objective of the presented study was to evaluate the effect of the imperfection types
weld depression type A and sinusoidal imperfection on the numerical determination of buck-
ling strengths of meridionally loaded shells made of stainless steel using GMNIA analyses.
For this reason, 12 experimental shell tests made of duplex and ferritic stainless steel have
been used as a basis for simulation taking into account their measured true stress-strain curves
and three different types of imperfections: measured imperfections, weld depression type
A and sinusoidal imperfection, the latter ones according to EN 1993-1-6 for the relevant FQC
A and furthermore, also for FQCs B and C.

1083
Figure 4. Experimental and numerical relative buckling strengths for different imperfection types.

Generally, both imperfection types according to EN 1993-1-6 are well suitable for simulat-
ing at least the experimental tests with a small preference of the sinusoidal imperfection as it
yields to a slightly better fit looking at the mean value deviation. As the weld depression type
A is commonly assumed to reflect best the imperfections in real structures, it can be concluded
on the basis of this study, that this imperfection pattern can be used for the numerical deter-
mination of the buckling strength of meridionally loaded shells made of stainless steel. How-
ever, it has to be kept in mind that weld depression type A imperfection might yield to higher
ultimate buckling strengths than the sinusoidal imperfection.

REFERENCES

ANSYS Mechanical APDL, Version 18.1.


Arbocz, J. 1982. The imperfection data bank, a mean to obtain realistic buckling loads. In Buckling of
Shells, Springer: 535–567.
Coleman, R., Ding, X., Rotter, J.M. 1992. Measurement of imperfections in full-scale steel silos, In 4th
International Conference on Bulk Materials 2 (92(7)): 467–472. Storage, handling and Transportation
Seventh International Symposium on Freight Pipelines. Institution of Engineers: Australia.
EN 1990:2002 + A1:2005 + A1:2005/AC:2010, Eurocode – Basis of structural design.
EN 1993-1-4:2006 + A1:2015, Eurocode 3: Design of steel structures – Part 1-4: General rules- Supplemen-
tary rules for stainless steels.
EN 1993-1-6:2007 + AC:2009 + A1:2017, Eurocode 3: Design of steel structures – Part 1-6: Strength and
stability of shell structures.
Esslinger, M., Geier, B. 1972. Gerechnete Nachbeullasten als untere Grenze der experimentellen axialen
Beullasten von Kreiszylindern. Der Stahlbau 41 (12): 353–360.
Gorbachov, A., Stranghöner, N., Rotter, J.M. 2017. Buckling behaviour of axially compressed cylin-
drical shells made of stainless steel, EUROSTEEL 2017, 13-15 September 2017: 828–837. Copenhagen:
Den-mark.
Greiner, R., Derler, P. 1995. Effect of imperfections on wind-loaded cylindrical shells. Thin-Walled Struc-
tures 23 (1-4): 271–281.
Hautala, K.T., Schmidt, H. 1998, Buckling Tests on Axially Compressed Cylindrical Shells Made of
Various Austenitic Stainless Steels at Ambient and Elevated Temperatures. Universität – GH Essen,
Forschungsbericht aus dem Fachbereich Bauwesen 76. Universität – GH Essen.

1084
Hautala, K.T., Schmidt, H. 1999, Buckling of axially compressed cylindrical shells made of austenitic
stainless. Light-Weight Steel and Aluminum Structures: 233–240. Amsterdam/New York: Elsevier Sci-
ence 1999.
Hutchinson, J.W., Tennyson, R.C. Muggerldge, D.B. 1971. Effect of a Local Axisymmetric Imperfection
on the Buckling Behavior of a Circular Cylindrical Shell under Axial Compression. In Reprinted from
American Institute of Aeronautics and Astronautics (AIAA) Journal 9 (1): 48–52.
Jansseune, A., De Corte, W., Belis, J. 2016. Imperfection sensitivity of locally supported cylindrical silos
subjected to uniform axial compression. International Journal of Solids and Structures 96: 92–109.
Koiter, W.T. 1945, On the Stability of Elastic Equilibrium (in Dutch). Ph. D. Thesis: Delft University.
Koiter, W.T. 1963. The effect of axisymmetric imperfections on the buckling of cylindrical shells under
axial compression. In Proc Koninklijke Nederlandse Akademie van Wetenschappen 66: 265–279.
Riks, E. 1979. An incremental approach to the solution of snapping and buckling problems. In Inter-
national Journal of Solids and Structures 15: 529–551.
Rotter J.M., Teng J.G., 1989, Elastic stability of cylindrical shells with weld depression. ASCE Journal of
Structural Engineering 115 (5): 1244–1263.
Rotter J.M. 2004, Cylindrical shells under axial compression. Buckling of Thin Metal Structures: 42–87,
In J.G. Teng and J.M. Rotter (ed.). Spon: London.
Rotter, J.M., Schmidt, H. 2013 (eds.). Buckling of steel shells, European design recommendations, 5th Edi-
tion, Revised 2nd impression. ECCS - European Convention for Constructional Steelwork.
Rotter, J.M. 2017. Shell buckling transformed: mechanics, design processes and their interrelation. Stahl-
bau 86 (4): 315–324.
Schmidt, H. 2000. Stability of steel shell structures: general report. Journal of Constructional Steel
Research 55 (1-3):159–181.
Schmidt, H., Winterstetter, T.A. 2004. Cylindrical shells under combined loading: axial compression,
external pressure and torsional shear. Buckling of Thin Metal Shells. IN J.G. Teng and J.M. Rotter
(ed.). Spon: London.
Schmidt, H. 2018. Two decades of research on the stability of steel shell structures at the University of
Essen (1985–2005): Experiments, evaluations, and impact on design standards, Advances in Structural
Engineering: Special Issue for Professor Rotter: 1–29.
Schneider, W., Timmel, I., Höhn, K., 2005. The conception of quasi-collapse-affine imperfections: A new
approach to unfavourable imperfections of thin-walled shell structures. Thin-Walled Structures 43:
1202–1224.
Song, C.Y., Teng, J.G., Rotter, J.M. 2004. Imperfection sensitivity of thin elastic cylindrical shells subject
to partial axial compression, International Journal of Solids and Structures 41: 7155–7180,
December 2004.
Stranghöner, N., Gorbachov, A. 2015. Experimentelles Tragverhalten axialgedrückter Kreiszylinderscha-
len aus nichtrostenden ferritischen und Duplex-Stählen (Experimental load bearing behaviour of
axially compressed cylinders made of ferritic and duplex stain-less steels), Stahlbau 84 (4): 275–284.
Stranghöner, N., Azizi, E., Gorbachov, A. 2019, Influence of material nonlinearity on the buckling resist-
ance of stainless steel shells. Journal of Constructional Steel Research, accepted for publication in Feb-
ruary 2019 (in press).
Teng, J.G., Rotter, J.M. 1992. Buckling of presurrized axisymmetrically imperfect cylinders under axial
loads, Journal of Engineering Mechanics 118 (2): 229–247, February 1992.
Teng, J.G., Lin, X., Rotter, J.M., Ding, X.L., 2005. Analysis of geometric imperfections in full-scale
welded steel silos, Engineering Structures 27 (6): 938–950.
Tennyson, R.C., Muggeridge, D.B. 1969. Buckling of Axisymmetric Imperfect Circular Cylindrical Shells
under Axial Compression. In American Institute of Aeronautics and Astronautics (AIAA) Journal 17,
(11):2127–2131.
Thomas, E., Soccol, D., Romero Barragan, M., Matres, V., Ohligschläger, T., Säynäjäkangas, J.,
Stranghöner, N., Gorbachov, A., Stehr, S., Brunstermann, R., Brinkmann, B., Widmann, R.,
Baddoo, N., Aggeloppoulos, E., Tholen, R. 2017. In: Innovative and competitive solutions using SS and
adhesive bonding in biogas (BiogaSS); Final Report, RFSR-CT-2012-00035, European Commission,
Research Fund for Coal and Steel: Luxembourg, Publications Office of the European Union.
Yamaki, N., 1984, Elastic Stability of Circular Cylindrical Shells. Elsevier Science. North-Holland.

1085
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Proposal for improving the consistency between Eurocode 3-1-8


and Eurocode 8-1

A. Stratan & D. Dubina


The Politehnica University of Timisoara, Romania

ABSTRACT: This paper makes on overview of provisions for design of joints in steel struc-
tures according to EN 1993-1-8 and EN 1998-1 and discusses the relationship between the two
codes. EN 1993-1-8 provides detailed design criteria and guidance on modelling of joints for
global structural analysis, being limited to joints subjected to predominantly static loading.
On the other hand, EN 1998-1 provides additional requirements for seismic design of joints in
steel structures, totally relying on design tools from EN 1993-1-8. Several inconsistencies
between the two codes concerning calculation of design resistance, modelling and strength
classification are discussed in this paper, and possible improvements are proposed.

1 INTRODUCTION

This paper discusses the application of two complementary Eurocodes for the design of moment-
resisting beam-to-column joints: EN 1993-1-8, addressing the design of joints in steel structures
and EN 1998-1, addressing the seismic design of structures. Reference is made to the current ver-
sion of the codes (EN 1993-1-8, 2005) and (EN 1998-1, 2004). A brief overview of the two codes is
given in the next sections and it is shown that seismic design of beam-to-column joints according
to EN 1998-1 heavily depends on the design tools in EN 1993-1-8. However, there are several
inconsistencies between the two codes which makes it difficult for the engineer to find a rational
way around. Possible improvements concerning the calculation of design resistance, modelling
and strength classification are proposed in this paper. Hopefully, this proposal may contribute to
improvements in the next generation of structural Eurocodes, which are currently in the process of
revision by CEN/TC 250 as part of the EC Mandate M/515 (NEN, n.d.).

2 OVERVIEW OF CODE PROVISIONS

2.1 EN 1993-1-8
EN 1993-1-8 provides rules for the design of joints in steel structures subjected to predominantly
static loading. It covers aspects related to the calculation of structural properties of joints (resist-
ance, stiffness and, to a certain extent, ductility), as well as rules for classification of joints by
stiffness and strength, and guidance on modelling of joints for global structural analysis. While
the application of basic principles would allow calculation of structural properties for joints sub-
jected to arbitrary loading (bending moments, axial and shear forces), detailed guidance is given
to moment-resisting joints only. This is especially true in what concerns classification by stiffness
and by strength, as well as guidance on modelling of joints for global structural analysis.
It is worth noting the meaning of terms “connection” and “joint” as used in EN 1993-1-8:
– The connection is defined as the location at which two or more elements meet. For design
purposes, it is the assembly of the basic components required to represent the behaviour
during the transfer of the relevant internal forces and moments at the connection.

1086
– The joint is the zone where two or more members are interconnected. A beam-to-column
joint consists of a web panel and either one connection (single sided joint) or two connec-
tions (double-sided joint).
In principle, EN 1993-1-8 provides two approaches for modelling of beam-to-column joints:
– A “general model”, in which the web panel in shear and the connections are modelled sep-
arately, Figure 1a.
– A “simplified alternative” in which the connection and the column web panel (CWP) are
considered being lumped in a single component, Figure 1b. Consequently, a double-sided
beam-to-column joint configuration has two moment-rotation characteristics, one for the
right-hand joint and another for the left-hand joint.
However, EN 1993-1-8 provides application rules for obtaining structural characteristics of
beam-to-column joints only for the “simplified alternative”. The code uses the component
method (Jaspart and Weynand, 2016) for deriving the moment resistance and stiffness of rota-
tional springs used to represent the joints in the global structural analysis.
Joints are classified by strength in EN 1993-1-8 as follows:
– Full-strength. A full-strength joint is defined as having a design moment resistance larger
than the one of the members it connects.
– Partial strength. A joint which does not meet the criteria for a full-strength joint or
a nominally pinned joint is classified as a partial-strength joint.
– Nominally pinned.
This classification is used to identify whether the effects of joint behaviour on the global struc-
tural analysis need be taken into account. According to Table 5.1 in EN 1993-1-8:
– If a joint is full-strength, it may be considered as continuous, i.e. the behaviour of the joint
need not be taken into account in case of global plastic analysis.
– If a joint is partial-strength, it may be considered as semi-continuous, i.e. the behaviour of
the joint shall be taken into account in the case of global plastic analysis.

2.2 EN 1998-1
EN 1998-1 applies to the design and construction of buildings in seismic regions. As stated by
the code, “attention must be paid to the fact that for the design of structures in seismic regions
the provisions of EN 1998 are to be applied in addition to the provisions of the other relevant
EN 1990 to EN 1997 and EN 1999”. EN 1998 contains only those provisions that, in addition
to the provisions of the other relevant Eurocodes, must be observed for the design of struc-
tures in seismic regions. It complements in this respect the other Eurocodes.

Figure 1. Modelling of beam-to-column joints in EN 1993-1-8 (2005): general model (a) and “simplified
alternative”(b).

1087
Design of earthquake-resistant steel buildings maybe accomplished following one of the fol-
lowing concepts: (1) low-dissipative structural behaviour; (2) dissipative structural behaviour.
Design using the concept of low-dissipative structural behaviour is the preferred approach in
low-seismicity regions and/or for structures with a small mass. Typical examples of such struc-
tures are industrial steel halls. According to EN 1998-1, the resistance of members and connec-
tions of steel structures designed using this approach should be performed in accordance with
(EN 1993-1-1, 2005) and (EN 1993-1-8, 2005), without any additional requirements.
On the other hand, design using the concept of dissipative structural behaviour is the
rational approach to the design of multistorey buildings. In this concept, the capability of
parts of the structures (dissipative zones) to dissipate seismic action through plastic deform-
ations is considered. The approach generally leads to more economical design, as it allows
structures to be designed for seismic forces considerably smaller than those applied in the case
of low-dissipative structural behaviour. The design may be accomplished based on a plastic
structural analysis (static or dynamic), which is the straightforward approach, as it models
explicitly the plastic response of dissipative zones and allows determining strength demands in
non-dissipative ones. However, since plastic structural analysis is time-consuming and requires
specialized knowledge, in practice, it is used only for the design of more important structures
or in research. In all other situations, the design is accomplished based on elastic structural
analysis, with the design seismic action reduced using the behaviour factor q. The overall duc-
tility of the structural system is enforced by observing a set of rules provided by EN 1998-1 at
the level of material, cross-section, members, connections and structural system.
EN 1998-1 allows that dissipative zones be located in the structural members or in the con-
nections. If dissipative zones are located in the structural members, the non-dissipative parts
and the connections of the dissipative parts to the rest of the structure shall have sufficient
overstrength to allow the development of cyclic yielding in the dissipative parts. Conversely, if
dissipative zones are located in the connections, the connected members shall have sufficient
overstrength to allow the development of cyclic yielding in the connections.
The adequacy of design (strength and ductility) of members and their connections in dissi-
pative zones should be supported by experimental evidence, in order to conform to the specific
requirements for each structural type and structural ductility class. This applies to partial and
full-strength connections in or adjacent to dissipative zones (EN 1998-1, 2004).
In beam-to-column joints of moment resisting frames, the column web panel should be
designed to resist the action effects corresponding to the development of plastic resistance in
the adjacent dissipative zones in beams or connections.

3 RELATIONSHIP BETWEEN EN 1993-1-8 AND EN 1998-1

As stated in EN 1993-1-8, it provides rules for the design of joints in steel structures subjected to
predominantly static loading. It covers aspects related to the determination of structural properties
of joints (resistance, stiffness and, to a certain extent, ductility), as well as rules for classification of
joints by stiffness and strength, and guidance on modelling of joints for global structural analysis.
In most practical situations, structures subjected to static loading, (persistent design situation in
(EN 1990, 2002)) are designed using elastic global analysis. Consequently, EN 1993-1-8 provides
only limited guidance on ductility (plastic rotation capacity) of joints. On the other hand, even if
guidance on elasto-plastic analysis exists, it is based on elastic-perfectly plastic behaviour of mem-
bers and joints, neglecting such effects as strain hardening.
EN 1998-1 applies to the design and construction of buildings in seismic regions, but it pro-
vides only additional rules to those of the other relevant Eurocodes. In relation to the design
of joints in steel structures, EN 1998-1 complements the rules in EN 1993-1-8. Consequently,
EN 1998-1 relies heavily on design rules in EN 1993-1-8.
Conceptually, seismic design of steel structures (seismic design situation in EN 1990) is different
from the non-seismic one (persistent design situation according to EN 1990) to the extent to which
the former exploits inelastic response of the structure. Irrespective of the type of global structural
analysis employed (elastic or plastic), a set of specific rules should be followed in order to

1088
guarantee a ductile response of the structure under seismic action. Conceptually, predetermined
dissipative zones should be designed to be ductile, while the non-dissipative ones should be
designed to withstand the maximum forces and moments generated by the yielded and fully
strain-hardened dissipative zones undergoing cyclic plastic deformations. The latter requirement is
known as “capacity design” (Landolfo et al., 2017).
It is to be underlined that the fundamental principles necessary to model a structure for
plastic analysis are the same in the case of seismic and persistent design situations. The dis-
tinctive features of seismic plastic analysis are the cyclic response of structural components
and the dynamic nature of the seismic action.
With respect to the design of connections in steel structures, there are several ways in which
EN 1998-1 interrelates with EN 1993-1-8 in accomplishing a ductile structural behaviour:
– EN 1998-1 references EN 1993-1-8 for determining the resistance and the stiffness of con-
nections. This is quite straightforward, as the cyclic response in the inelastic range would
rather affect the plastic deformation capacity and strain hardening, but not too much the
initial stiffness and resistance at yielding of connections.
– EN 1998-1 limits the categories of connections allowed for seismic applications, with
respect to those available in EN 1993-1-8. For example, bolted connections in shear of
bearing type (category A) and non-preloaded bolted connections in tension (category D)
are not allowed for seismic applications in dissipative zones.
– EN 1998-1 specifies ductility criteria under cyclic loading that should be fulfilled by connections
in dissipative zones. For example, beam-to-column connections in moment-resisting frames
should have a rotation capacity θp not less than 0,035 rad for structures of ductility class DCH
and 0,025 rad for structures of ductility class DCM designed for q > 2. Due to insufficiently
developed and unreliable analytical tools for assessing the cyclic response of connection in the
plastic range, EN 1998-1 requires that the rotation capacity is demonstrated experimentally.

4 INCONSISTENCIES AND PROPOSAL FOR IMPROVEMENT

4.1 Modelling of beam-to-column joints


Both dissipative and non-dissipative beam-to-column connections are allowed by EN 1998-1.
However, non-dissipative connections are preferred in practice, due to the difficulty of proving the
ductility of dissipative connections under cyclic loading, and requirements of more refined global
plastic analysis. To accomplish a ductile response of the joint region, a hierarchy of resistance of
the beam end, connection and column web panel is required by EN 1998-1, see Figure 2. In the
case of non-dissipative (brittle) connections, plastic deformations should occur at the beam end

Figure 2. Hierarchy of resistance in the beam, connection and column web panel according to EN 1998-1.

1089
(Figure 2a), while the connection should have sufficient overstrength (Figure 2b) to prevent its fail-
ure under the maximum moment that can develop at the plastic hinge location under cyclic load-
ing. Limited plastic deformations are also allowed to occur in the column web panel (Figure 2c).
The “simplified” joint model currently adopted by EN 1993-1-8 is not able to provide an
implementation of the requirements in EN 1998-1, as the three parts of the joint (connection,
connected member and column web panel) are merged into a single spring. Given the com-
plexity of the component method rules in EN 1993-1-8, engineers would usually rely on com-
mercial software for obtaining joint properties (resistance and stiffness). This makes it very
difficult, if impossible, to apply the requirements in EN 1998-1 using the tools in EN 1993-1-8.
Therefore, independent resistance checking of the connection and the column web panel is
necessary. In principle, this is possible to be accomplished using the “general approach” of model-
ling beam-to-column joints in EN 1993-1-8. However, the lack of application rules for the general
approach makes it mostly unusable in practice. Consequently, providing more detailed guidance
on the application of the “general approach” for modelling beam-to-column joints in EN 1993-
1-8, as well as the application of the component method, in this case, would be of great benefit in
achieving consistency between EN 1998-1 and EN 1993-1-8. The “general approach” of modelling
beam-to-column joints using distinct springs for the column web panel and the connection was
discussed by (Jaspart, 1991), (Charney and Downs, 2004) and (da Silva et al., 2016), among
others.
In addition to providing more accurate results, the refined modelling of joints using the
“general approach” brings the following benefits:
– Avoid the necessity of analysing distinct structural models for each load combination (da
Silva et al., 2016).
– Straightforward calculation of the design shear force in the column web panel by subtract-
ing shear force in the columns in equation 5.3 from EN 1993-1-8 (2005), which leads to
more economical design.
– Design verification efforts are reduced for double-sided beam-to-column joints. Firstly,
connections are designed to resist the applied bending moment. The same connection char-
acteristics may be appropriate for internal and external joints. Secondly, the column web
panel is checked for internal and external joints, under different load cases (gravity only or
gravity + wind, gravity + seismic loading, or any other relevant combination). It leads to
easy identification of the load combinations which would require strengthening of the
column web panel, a common design problem.
– The correct approach in modelling of varying moment distribution acting on the column
web panel in case of global plastic analysis with non-proportional loads (e.g. in case of non-
linear static analysis).
Though refined modelling of joints offers many benefits, it is strictly necessary only in case of
double-sided joints when the loading is non-proportional throughout the analysis. The simpli-
fied model is not appropriate in such cases, as the stiffness and strength of the springs in the
simplified model (Figure 1b) depend on the ratio of bending moments in the beams, and
would need to be updated (re-computed) at each step of the analysis. For all other cases, the
simplified model may be used, even if independent resistance checks are performed for the
connection and the column web panel.

4.2 Classification by strength

4.2.1 General
In EN 1993-1-8 a joint is considered full-strength if it fulfils the following requirement:

Mj;Rd  Mpl;Rd ð1Þ

which assumes zero post-yielding stiffness of the plastic hinge in the beam.

1090
The real response of members (Galambos, 1968) and joints (Weynand et al., 1996) is character-
ised by a significant amount of strain-hardening. Ignoring strain-hardening in a plastic global ana-
lysis significantly underestimates the forces and moments that cause brittle failure, which may lead
to unsafe design.
Variability of steel yield strength is another issue that might lead to unsafe design. Statistical
data obtained recently by (da Silva et al., 2017) and (Badalassi et al., 2017) shows that:
– mean yield strength of European steels is significantly larger than nominal values;
– mean yield strength is larger for lower steel grades than for higher-strength grades.
The last observation is particularly important for bolted beam-to-column connections in the
case of global plastic analysis, since:
– a beam fabricated from lower steel grades (S235/S275) may be characterised by a large
over-strength (ratio between mean and nominal strength), and consequently generate
a maximum moment in the plastic hinge significantly larger than the nominal one, while
– the bolted connection, and more specifically high-strength bolts have a significantly lower
over-strength.
Accounting for the effects of strain hardening and material variability, EN 1998-1 requires
that the resistance of non-dissipative (full-strength) joints Mj,Rd should be larger than the
resistance of the connected dissipative (ductile) member Mpl,Rd, amplified by material over-
strength factor (γov = 1.25) and a factor accounting for strain hardening (1,1):

Mj;Rd  1;1γov Mpl;Rd ð2Þ

A similar requirement exists in EN 1993-1-8:

Mj;Rd  1;2Mpl;Rd ð3Þ

which should be fulfilled if the “rotation capacity of the joint need not be checked”.
The condition (3) from EN 1993-1-8 is a simplified form of the requirement (2) from EN 1998-
1, with a single factor (1,2) accounting for strain hardening and material overstrength. Definition
(1) of a full-strength joint in EN 1993-1-8 allows a continuous model to be adopted (the behaviour
of the joint need not be taken into account in the global plastic analysis), and is formally correct
for elastic-perfectly plastic model. The continuous model implies that plastic deformations occur
exclusively in the connected member. Which is confusing, because according to the same code, this
is assured if the joint resistance is at least 1,2 times stronger than the connected member, according
to expression (3).

4.2.2 Full-strength macro-components


In the case of global plastic analysis, it should be shown that the structure has sufficient plastic
deformation capacity. This can be accomplished by designing the structure such that:
– plastic deformations occur in structural components which are ductile, and
– design resistance of brittle components is not exceeded under forces that cause full yielding
and strain hardening of ductile components.
To improve the consistency within EN 1993-1-8 and between the two codes, it is proposed
that in EN 1993-1-8 the definition of full strength macro-components (1) and the criteria for
disregarding the check of the rotation capacity of the joint (3) be merged into a single one,
and explicitly accounting for the effects of material over-strength and strain hardening. Thus,
for plastic analysis, non-ductile macro-components (connection or column web panel) should
be full-strength, by fulfilling the following criterion:

Rd;fs  γsh  γov  Rm ð4Þ

1091
where Rm is the plastic resistance of the connected member obtained without considering the
partial factor for material, evaluated in the section where plastic hinge is expected to occur;
Rd,fs is the design resistance of the non-ductile macro-component; γsh is the strain hardening
factor of the ductile macro-component (member); γov is the material overstrength factor of the
ductile macro-component (member) and represents the ratio between the mean and nominal
values of yield strength.
Expression (4) is defined in general terms, and generally refers to bending moment but may
apply to other relevant action effects (shear or axial force). In principle, it is appropriate for
the cases when the ductile and non-ductile macro-components are adjacent to each other.
When the ductile and non-ductile macro-components are located at a non-negligible distance
(e.g. in case of beam-to-column joints with haunches), a translation term would have to be
used. In the case of elastic-perfectly-plastic global analysis, it is convenient designing the full-
strength non-ductile macro-component for the maximum forces and moments obtained from
the analysis, multiplied by γshγov.
Values of γsh depend on the type of loading. The recommended value for members subjected to
static monotonic loading is γsh = 1,1. Evidence of larger values of the strain hardening γsh factor
exists, even under static monotonic loading (Faella et al., 2000), (Mazzolani and Piluso, 1996).
Even if potentially unconservative in some cases, the value of γsh = 1.1 is retained as
a simplification. Larger strain hardening would be required in the case of cyclic loading in EN
1998-1. Tentative values are γsh = 1,20 for bending moments, γsh = 1,50 for shear forces, γsh = 1,10
for axial forces.
Based on the SAFEBRICTILE project (da Silva et al., 2017), values for γov may be taken as
follows: γov = 1,45 for S235, γov = 1,35 for S275; γov = 1,25 for S355, γov = 1,2 for S460. These
values are representative of the modern European steel market.

4.2.3 Equal-strength macro-components


To cover the gap between full-strength and partial-strength macro-components, it is proposed
introducing a new strength classification: equal-strength. The design resistance of an equal-
strength macro-component (connection or column web panel) Rd,es, should be larger than or
equal to the plastic resistance of the member Rm:

Rd;es  Rm ð5Þ

Due to the variability of material characteristics and strain hardening in the macro-compo-
nents of a joint, plastic deformations may occur in both the member and in the “equal-
strength” macro-component. Therefore, both macro-components shall be ductile if the global
plastic analysis is used, and their resistance shall be modelled in the global analysis
(semi-continuous model).

4.2.4 Partial-strength macro-components


A macro-component (connection or column web panel) is partial-strength if it limits the
design resistance of the structural system. Partial-strength macro-components have a design
resistance Rd,ps which is smaller than the one of the connected member Rm:

Rd;ps 5Rm ð6Þ

A partial-strength macro-component shall be designed for the action effects in the relevant
design situation. If the plastic analysis is used, it should be shown that the partial-strength
macro-component have the necessary deformation capacity, and it shall be modelled as semi-
continuous.
If the structure is designed to develop plastic deformations exclusively in the partial-
strength macro-component, the adjoining member shall be designed for the relevant forces
that can be developed in the partial-strength macro-component, including the effects of mater-
ial overstrength and strain hardening:

1092
Rd  γsh  γov  Rps ð7Þ

where Rd is the design resistance of the connected member or other non-ductile macro-
component; Rps is the resistance of the partial-strength macro-component, obtained without
considering the partial factor for the material; γsh is the strain hardening factor of the partial-
strength macro-component; γov is the material overstrength factor of the partial-strength
macro-component.
Values of γsh depend on the type of loading. In lieu of more precise data, the same values as
for members may be used. However, it is known that certain types of connections (e.g. bolted
end-plate connections) may develop substantially larger hardening.

5 CONCLUSIONS

EN 1993-1-8 provides comprehensive procedures for the design of joints in steel structures sub-
jected to predominantly static loading. EN 1998-1 covers the design of joints in seismic-resistant
steel structures but relying heavily on the design tools from EN 1993-1-8 and providing only
additional requirements targeting a ductile response. Therefore, the consistency of the design
provisions in the two codes is highly important.
Areas in EN 1993-1-8 that may be improved include modelling of beam-to-column joints
and classification by strength. The authors suggest that three macro-components are expli-
citly considered in the design: the connection, the column web panel and the connected
member. For global analysis, the “general approach” which explicitly considers the three
macro-components, is more correct and avoids tedious iterations in the design process. Of
course, appropriate tools should be available in structural analysis software. For the design
process, independent modelling of the three macro-components allows establishing the hier-
archy of resistance necessary for obtaining a ductile response of the joint. Another possible
improvement concerns the classification of macro-components by strength, and it is sug-
gested that three categories be defined: full-strength, equal-strength and partial strength,
and accounting explicitly for material overstrength and strain hardening.

ACKNOWLEDGEMENTS

The research leading to these results has been funded from the European Community’s
Research Fund for Coal and Steel (RFCS) under grant agreement no 754048 RFCS-2016/
RFCS-2016 EQUALJOINTS - PLUS. This support is gratefully acknowledged.

REFERENCES

Badalassi, M., Braconi, A., Cajot, L.-G., Caprili, S., Degee, H., Gündel, M., Hjiaj, M., Hoffmeister, B.,
Karamanos, S.A., Salvatore, W., Somja, H., 2017. Influence of variability of material mechanical
properties on seismic performance of steel and steel–concrete composite structures. Bulletin of Earth-
quake Engineering 15, 1559–1607. https://doi.org/10.1007/s10518-016-0033-2.
Charney, F.A., Downs, W.M., 2004. Modeling procedures for panel zone deformations in moment resist-
ing frames, in: Proceedings of the Fifth International Workshop, Connections in Steel Structures
V. Amsterdam, The Netherlands, pp. 121–130.
da Silva, L.S., Simões, R., Gervásio, H., 2016. Design of steel structures: Eurocode 3: Design of steel
structures. Part 1-1, General rules and rules for buildings, 2nd edition. ed, ECCS Eurocode Design
Manuals. ECCS - European Convention for Constructional Steelwork.
da Silva, L.S., Tankova, T., Marques, L., Kuhlmann, U., Kleiner, A., Spiegler, J., Snijder, H. H.,
Dekker, R., Taras, A., Popa, N., 2017. Safety assessment across modes driven by plasticity, stability
and fracture. ce/papers 1, 3689–3698. https://doi.org/10.1002/cepa.425
EN 1990, 2002. Eurocode - Basis of structural design. European Committee for Standardization (CEN).

1093
EN 1993-1-1, 2005. Eurocode 3: Design of steel structures - Part 1-1: General rules and rules for build-
ings. European Committee for Standardization (CEN).
EN 1993-1-8, 2005. Eurocode 3: Design of steel structures - Part 1-8: Design of joints. European Commit-
tee for Standardization (CEN).
EN 1998-1, 2004. Eurocode 8: Design of structures for earthquake resistance - Part 1: General rules, seis-
mic actions and rules for buildings. European Committee for Standardization (CEN).
Faella, C., Piluso, V., Rizzano, G., 2000. Structural steel semirigid connections: theory, design, and soft-
ware. CRC Press, Boca Raton.
Galambos, T.V., 1968. Deformation and energy absorption capacity of steel structures in the inelastic
range. Steel Research and Construction Bulletin No. 8.
Jaspart, J.-P., 1991. Etude de la semi-rigidité des noeuds poutre-colonne et son influence sur la résistance
et la stabilité des ossatures en acier (PhD Thesis). Université de Liège, Belgium.
Jaspart, J.-P., Weynand, K., 2016. Design of joints in steel and composite structures, ECCS Eurocode
design manuals. ECCS/Ernst & Sohn.
Landolfo, R., Mazzolani, F., Dubina, D., Silva, L.S.D., D’Aniello, M., 2017. Design of Steel Structures
for Building in Seismic Areas. ECCS, Ernst & Sohn, Berlin.
Mazzolani, F., Piluso, V., 1996. Theory and Design of Seismic Resistant Steel Frames, 1 edition. ed.
CRC Press, London.
NEN, n.d. Call for Tender for the development of the second generation of Structural Eurocodes [WWW
Document]. URL https://www.nen.nl/Normontwikkeling/Eurocodes-2020.htm (accessed 12.14.17).
Weynand, K., Jaspart, J.-P., Steenhuis, M., 1996. The stiffness model of revised Annex J of Eurocode 3,
in: Connections in Steel Structures III. Elsevier, pp. 441–452. https://doi.org/10.1016/B978-008042821-
5/50100-0.

1094
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Assumption of imperfections for the LTB-design of members based


on EN 1993-1-1

Richard Stroetmann
Institute of Steel and Timber Structures, Technical University of Dresden, Germany

Sergei Fominow
Institute of Numerical Methods and Structural Analysis, TH Mittelhessen University of Applied Sciences,
Giessen, Germany

ABSTRACT: The stability design of members and structures by a geometrically non-linear


analysis with equivalent geometric imperfections (GNIA) is a commonly used method. In the case
of lateral torsional buckling (LTB) of members, the rules in EN 1993-1-1 are unsatisfactory. With
the current code but also with the new draft prEN1993-1-1 results are achieved, some of which
are significantly on the safe side but some also on the unsafe side. The reasons for this are the
limitation to bow imperfections e0 out of plane, an inappropriate differentiation related to the
cross-section shape and yield strength, as well as neglecting the influence of the moment distribu-
tion over the member lengths. Parameter studies have shown that a differentiation is required due
to the different structural behavior of members loaded with pure bending, pure compression, or
a combination of bending and compression. In this article current research results are presented,
which consider the case of pure bending My of I and H sections. Dependencies on the shape of
sections, cross-sectional resistance models and the influence of steel grades are analyzed with
respect to the assumption of equivalent geometrical imperfections.

1 INTRODUCTION

The design process by a geometrical nonlinear analysis using equivalent initial imperfections
(GNIA) will be carried out by applying equivalent initial geometric imperfections, determination
of internal forces using geometrical non-linear analysis and verification of the cross-section resist-
ance at the most unfavorable position. The LTB resistance according to this design method
depends on the size and shape of the geometrical imperfection, the cross-sectional shape, the
moment distribution and possible additional compression forces, the resistance model and the
steel grade. In EN 1993-1-1 imperfections for the LTB-design of members are defined in chapter
5.4.3. The shape is specified as an initial bow imperfection out of plane. The amplitude is given as
k · e0, where k = 0,5 is the recommended value and e0 is defined in table 5.1 and 6.2 of the code.
For rolled I-sections the relevant buckling curve for flexural buckling depends on the h/b ratio,
the plate thickness and steel grade. For slender cross-sections with h/b > 1.2 this results in smaller
imperfection amplitudes than for compact cross-sections with h/b ≤ 1.2. Kindmann & Beier-Tertel
2010 have shown that this assumption is not suitable for lateral torsional buckling. Cross-Sections
with h/b ≤ 2.0 are more favorable than those with h/b > 2. This is also considered in the selection
of the LTB curves according Table 6.5 of EN 1993-1-1. In the German National Annex of
EN1993-1-1 other amplitudes of imperfection are specified which are derived from the assignment
to the LTB-buckling-curves. The k-factor is given as 1.0 in the medium range of slenderness (0,7 ≤
λLT ≤ 1,3). This rules are adopted pand modified
ffiffiffiffiffiffiffiffiffiffiffiffiffi
ffi in the latest version of prEN1993-1-1 by consider-
ing the material parameter ε ¼ fy =235 as an amplifier. The shape of the equivalent geometrical
imperfection is still specified as a bow-imperfection out of plane (Table 1).

1095
Table 1. Initial bow-imperfections e0 for rolled I-sections for the LTB-design of members by GNIA.
current EN 1993-1-1 prEN 1993-1-1
Elastic cross- Plastic cross-
elastic analysis plastic analysis section verifi- section verifi-
Limits e0 =L e0 =L Limits cation ε  e0 =L cation ε  e0 =L

h=b  1:2 1/400 1/400 h=b  2:0 1/250 1/200

h=b41:2 1/500 1/300 h=b42:0 1/200 1/150

Figure 1. Shapes of imperfections – bow- and mode-imperfections.

Investigations by Ebel 2014 on different structural systems and loads have shown, that the
structural behavior may be better reflected by considering the shape of the first LTB buckling
mode for the equivalent geometrical imperfection. Snijder, van der Aa, Hofmeyer & van Hove
2018 also considered imperfection shapes based on LTB modes and developed a proposal for
geometric and material nonlinear analysis. In Hajdú, Papp & Rubert 2017 a proposal is pre-
sented, which considers the combination of compression and bending. The imperfection shape
also based on the relevant elastic critical LTB mode.
In the following parameter studies, both imperfection shapes, the sinusoidal bow imperfec-
tion out of plane (IMP 1) and the first eigenmode for LTB (mode-imperfection, IMP 2) are
investigated (Figure 1). In this context, a comparison of the results is made, which also enables
a comparison with the rules in EN1993-1-1. For both imperfection shapes, the amplitude e0
corresponds to the maximum amplitude.
The aim of the present investigations is to derive imperfection approaches for the LTB design
by GNIA, which take into account the essential influences of cross-sectional shape, yield
strength, moment distribution and model for the cross section resistance. For reasons of simpli-
fication, only linear interaction formulas are used to determine the cross-section resistance.
– Interaction formula 1 (I-1)
Elastic verification of the cross-sectional resistance. The limit criterion is fulfilled, if the von
Mises stress σv reaches the yield strength fy at the most critical point of the cross-section.
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
σV ¼ σ x 2 þ 3  τ 2  fy ð1Þ

– Interaction formula 2 (I-2)


Linear plastic interaction. The influence of shear may lead to a reduction of the full plastic
cross section resistances.

1096
N My Mz B
þ þ þ  1:0 ð2Þ
A Mpl;y Mpl;z Bpl

– Interaction formula 3 (I-3)


Linear plastic interaction with limitation. The full plastic cross section resistances are limited to
1.25 times of the elastic values. Background of the limitation is that in the case of LTB the
cross-section is usually not completely plasticized when the structural resistance is reached. Usu-
ally the plastic zones reach only over a part of the flanges but not the entire cross section. The
new design rules for flexural buckling in prEN1993-1-1 (see Lindner, Kuhlmann & Jörg 2017)
consider also this limitation. For LTB of I- and H-sections the limit is relevant for Mz and B.

N My Mz B
þ þ þ  1:0 ð3Þ
A Mpl;y 1:25  Mel;z 1:25  Bel

2 PARAMETER STUDIES FOR PURE BENDING

2.1 Calibration of imperfections by comparison of results of GMNIA and GNIA calculations


In a comprehensive parametric study, numerous LTB resistances of hot-rolled I-sections consider-
ing a wide range of member lengths, section shapes and load types were calculated by GMNIA
with ANSYS. For this purpose, a hybrid shell-beam model was created modeling the I-section as
a combination of shell and beam elements. The shell elements were located in mid-plane of the
flanges and the web. The beam elements were arranged to consider the rolled radii at the intersec-
tion of the flange and the web. Initial geometric imperfections were implemented into the FE
models considering the first buckling mode from an elastic buckling analysis. The amplitude was
scaled to L/1000. For the residual stress patterns, the recommendations of the ECCS were con-
sidered. The magnitudes of the stress patterns depending on the height-to-width (h/b) ratio of the
cross-section and are independent of the yield stress. The values of 0.5 · 235 MPa for h/b ≤ 1.2
and 0.3 · 235 MPa for h/b > 1.2 were assumed with a triangulated distribution over the width of
the flanges and the web. Further information are given in Stroetmann & Fominow 2018.
Based on LTB resistances according to GMNIA, equivalent geometric imperfections were
derived using geometric nonlinear analysis (GNIA). A single span beam with fork end condi-
tions consisting of open I- and H-sections was considered as the structural system (Figure 2).
The calculation of the required amplitude of imperfection e0 was an iterative process with the
aim of achieving the same LTB resistance as with GMNIA. They are defined in relation to the
member length L as a non-dimensional j-value (Eq. (4)).

L
j¼ ð4Þ
e0

In the parameter study, the maximum beam length was limited to exclude irrelevant lengths in the
calibration of imperfections. In general the limitation was set to L/h = 50 and for very slender
cross-sections to L/h = 40. In the parameter studies presented here, the limitation can be recog-
nized by the fact that the j-values end within the slenderness range shown. This is the case if the
beams are only slightly prone to lateral torsional buckling, e.g. those with HE200B sections.

2.2 Influence of the cross-sectional shape and interaction formulas


The required imperfection amplitudes e0 are significantly influenced by the cross-sectional
shape and the model for determining the resistance. In the parameter study, the dependencies

1097
Figure 2. Structural system and loads.

Table 2. Ratios h/b and Iy/IT of the investigated I- and H-sections.


Section h/b Iy/IT Section h/b Iy/IT Section h/b Iy/IT

IPE80 1.74 114.8 HE200B 1.00 96.1 HE200M 1.07 41.1


IPE200 2.00 277.9 HE360B 1.20 147.9 HE320M 1.16 45.4
IPE500 2.50 539.8 HE400B 1.33 162.0 HE340M 1.22 50.6
IPE600 2.73 558.1 HE800B 2.67 379.6 HE400M 1.41 68.9
HE200A 0.95 175.7 HE1000B 3.33 515.8 HE650M 2.19 178.3
HE360A 1.17 222.1 HE360AA 1.13 324.5 HE800M 2.69 268.2
HE400A 1.30 238.5 HE400AA 1.26 369.0 HE1000M 3.34 424.9
HE550A 1.80 317.9 HE600AA 1.90 613.3
HE800A 2.63 508.2 HE800AA 2.57 813.4
HE1000A 3.30 649.4 HE1000AA 3.23 1007.5

of the h/b- and Iy/IT-ratio were examined. For this purpose, 27 rolled I- and H-sections were
considered, which are listed in Table 2.
Figure 3 shows the minimum j-values for the described range of beam lengths. From the evalu-
ation for the h/b-ratio it can be seen that at h/b = 1.2 there is a strong increase in the minimum
j-values due to the different assumptions for the residual stresses in the GMNIA. However, the h/
b-ratio is generally not significant as the minimum j-values for cross-sections are similar for h/b ≈
1.2 and h/b > 3. The dependence on the Iy/IT ratio is pronounced for the assumption of mode
affine imperfections IMP-2. This applies in particular to beams subjected to transverse loads.
If the cross-sectional resistance is determined elastically, significantly lower amplitudes are
required with respect to the linear plastic interaction formula. However, the reduction of the
full plastic moment resistance causes only a slight reduction of the required amplitudes of the
imperfections. Concerning the imperfection shape, pure bow-imperfections require signifi-
cantly higher amplitudes than mode-imperfections.

2.3 Influence of the steel grade


Figure 4 shows the course of the required j-values for a compact (HE200B), a medium
(HE400A) and a slender cross-section (HE1000A) in the steel grades S 235, S 355 and S 460
considering the elastic and the plastic cross-sectional resistance. In most cases, the beams of
the higher steel grades require slightly larger amplitudes e0. There are exceptions for the slen-
der cross-sections in the medium slenderness range, where larger amplitudes are required. The
reason for this is that the occurring shear stresses reduce the load-bearing capacity calculated
by GMNIA. For slendernesses around 1.0, beams with cross-sections HE1000A have rela-
tively short lengths. For higher steel grades the beam length decreases for the same slenderness

1098
Figure 3. Minimum j-values for sections in the steel grade S355 according to Table 2, depending on the
interaction formulae, imperfection shapes as well as ratios h/b and Iy/IT.

and the influence of shear stress increases. The superposition of the influences of shear, bend-
ing and torsion leads to a lower load-bearing resistance at GMNIA.

2.4 Influence of the bending moment distribution


In a further parametric study different moment distributions for bending about the major axis
have been analysed, see Table 3. They were generated by end moments and uniform distrib-
uted or concentrated loads, acting on the top chord of the cross-sections, see Figure 2.
Figures 5 and 6 show the j-values for the cross-sections HE200B and HE1000A considering
the elastic and plastic cross-sectional resistances and imperfections IMP-1 and IMP-2. The sig-
nificant dependence on the moment distribution is recognizable. The approach of mode
imperfections leads to larger j-values and thus to smaller amplitudes e0. While the imperfec-
tion form IMP-2 is significantly influenced by the moment distribution My, the bow imperfec-
tion IMP-1 is independent. If the bending moment My at mid span of the beam is small, the
design point relevant for the cross-sectional check is asymmetrical. To achieve the same design
result with IMP-1, the amplitude e0 must be increased.
As illustrated in Figure 6, the load cases LC 1 and LC 2 require the largest imperfection
amplitudes when the mode-imperfection shape IMP-2 is applied. The moment distributions
caused by uniform loads (LC 3, LC 6 and LC 7) require slightly less imperfections. Compara-
tively small values resulting for concentrated loads (LC 8, LC 9 and LC 10).

3 CONCLUSION AND OUTLOOK

The stability design of steel structures using equivalent geometric imperfections is frequently
used in the case of flexural buckling, but is rather used for lateral and flexural torsional

1099
Figure 4. j-values for sections HE200B, HE400A and HE1000A in the steel grades S 235, S 355 and
S 460.

Table 3. Overview of the investigated load cases for bending about the major axis.

1100
Figure 5. Required j-values for bow-imperfections IMP-1 for sections HE200B and HE1000A in S 235,
considering interaction formulae I-1, I-2 and various moment distributions.

Figure 6. Required j-values for mode-imperfections IMP-2 for sections HE200B and HE1000A in
S 235, considering interaction formulae I-1, I-2 and various moment distributions.

1101
buckling. This is partly due to the demand of more complex calculation software, but partly
also due to missing or unclear specifications for imperfection assumptions and cross section
resistances to be used for the calculations.
As imperfection assumptions, sinusoidal bow-imperfections out of plane of the members
and the scaled mode-imperfections are common. The amplitudes of the equivalent geometric
imperfections, which have to consider, depending on many influences. These include the
cross-sectional shape, the loads and its distributions, the structural system and the boundary
conditions, but also the type of structural mechanic calculations and the model for the cross-
section resistance. Thus, for the elastic stress analysis and the linear plastic sectional inter-
actions, different assumptions of imperfections are necessary in order to achieve equal load
bearing capacities from more accurate calculations.
In this article, parameter studies are presented in which imperfection measures e0 for the LTB-
design of members with I- and H-sections under pure bending My were derived. The values are
based on the results of more detailed structural analysis according to the GMNIA. The imperfec-
tion amplitudes e0 were derived by a comparison of the results of a linear elastic calculation of
internal forces and the assessment of the sectional resistance with three different interaction
formulas.
The parameter studies show, that the mode-imperfection approach leads to significant smaller
imperfection amplitudes e0 than the approach of sinusoidal bow-imperfections in y-direction.
The higher the cross-section can be used by selecting the interaction formula for the cross-
sectional resistance, the greater the amplitude e0 of the imperfection must be applied.
The parameter studies on the influence of the cross-sectional shape of hot rolled sections reveal
that the ratio Iy/IT better describes the tendency for the imperfection size e0 than the ratio of h/b.
Discontinuities arise in that, from h/b > 1.2 the assumed residual stress decreases from 0.5 · fy,S235
to 0.3 · fy,S235. The studies of the effect of the yield strength related to the value e0 show that this
is only slight and is overestimated with the provisions in the current draft prEN1993-1-1.
The shape of the moment distribution has a very strong influence on the imperfection assump-
tion e0. For strongly asymmetric moment diagrams with alternating signs, the approach of sinus-
oidal bow imperfections e0 is unsuitable. With the assumption of mode imperfections it is possible
to react to the special bearing conditions of a structural system and its moment distribution.
The results of the parameter studies show the way forward for calibrating imperfections for
a codification approach based on EN1993-1-1. The evaluation of the data, in combination
with the necessary simplifications for the design practice, leads to corresponding definitions of
imperfection values e0 and the necessary differentiations.

REFERENCES

ANSYS, Finite Element Software Package, Version 18, ANSYS Inc., Canonsburg, PA, USA.
EN 1993- 1-1. 2005. Design of steel structures – Part 1-1: General rules and rules for buildings, 2005-05.
prEN1993-1-1. 2018. Design of steel structures – Part 1-1: General rules and rules for buildings, final
draft, 2018-07.
Ebel R. 2014. Systemabhängiges Tragverhalten und Tragfähigkeiten stabilitätsgefährdeter Stahlträger
unter einachsiger Biegebeanspruchung. In: Dissertation, Ruhr-Universität Bochum, Schriftenreihe des
Instituts für Konstruktiven Ingenieurbau, Heft 2014-03.
Hajdú G., Papp H. & Rubert A. 2017 Vollständige äquivalente Imperfektionsmethode für biege- und
druckbeanspruchte Stahlträger. In Stahlbau 86 (2017), P. 483–496. Berlin: Verlag Ernst & Sohn.
Kindmann R. & Beier-Tertel J. 2010. Geometrische Ersatzimperfektionen für das Biegedrillknicken von Trä-
gern aus Walzprofilen – Grundsätzliches. In Stahlbau 79 (2010). P. 689–697. Berlin: Verlag Ernst & Sohn.
Lindner, J., Kuhlmann, U. & Jörg, F. 2017: Initial bow imperfections e0/L of Table 5.1 of EN 1993- 1-1.
Doc. CEN-TC 250-SC3-WG1 N0155 (2017).
Snijder, H.H., van der Aa, R.P., Hofmeyer, H. & van Hove B. 2018. Lateral torsional buckling design
imperfections for use in non-linear FEA. In: Steel Construction 11 (2018), No. 1.
Stroetmann R. & Fominow S. 2018. Imperfections for the LTB-design of members by geometrical non-
linear analysis. Eighth International Conference on THIN-WALLED STRUCTURES - ICTWS 2018,
Lisbon, 24–27 July, 2018.

1102
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Welds on high-strength steels—influence of the welding process and


the number of layers

R. Stroetmann & T. Kästner


Technische Universität Dresden, Dresden, Germany

ABSTRACT: In addition to the development of a new design approach for welded joints of
high-strength steels, the influence of the welding process and the number of weld layers was exam-
ined within the ongoing AiF-FOSTA research project P1020. Hereby, the load-bearing character-
istics of fully mechanized welds was compared to manual produced welds. Furthermore, the
influence of different number of weld layers was examined considering two filler-metals. These
examinations include tensile tests and hardness measurements in two different levels of the weld.

1 INTRODUCTION

The mechanical properties of welds are influenced by a variety of factors. As part of the AiF-
FOSTA research project P1020 the influence of the manufacturing process of the high-
strength steels and the welding process on the mechanical properties of welds was examined.
For this purpose, experimental tests were carried out to study the influence of the automation
level of the welding process and number of layers. Within this examination, under- and over-
matching material combinations were considered. The aim was to quantify these influences on
the strength and ductility of the welds.

2 CHARACTERIZATION OF THE BASE AND FILLER METALS

For the examinations two thermomechanical rolled fine grain steels S500ML (EN 10025-4)
and S700MC (EN 10149-2) as well as two quenched fine grain steels S690QL and S960QL
(EN 10025-6) were selected. All base materials had a plate thickness of 20 mm.
The thermomechanical rolled fine grain steels (TM-steels) have a lower carbon equivalent
than the quenched fine grain steels (QT-steels). The difference results mainly from the signifi-
cantly lower carbon contents. Lower carbon equivalent results in a lower preheat temperature
for the same process parameters. Thus, the TM steels can be welded at lower preheat temper-
atures than the QT steels. For TM-steels, however, care must be taken to minimize the energy
input during the welding process, as the mechanical properties of these steel grades are largely
achieved by hot rolling in the not re-crystalized austenite and the transition temperature from
austenite to ferrite. Higher heating than the transformation temperature Ac1 leads to second-
ary crystallization processes that affect the mechanical properties.
In (Stroetmann et. al. 2018a), the authors examined the influence of the peak temperature
and cooling time t8/5, representative for the areas of the heat-affected zone and the used pro-
cess parameters, on the tensile strength of the base metals. It could be shown that the tensile
strength of the S700MC steel for each peak temperature and cooling time was below the
strength of the unaffected base metal. A reduction in the strength in the area of the heat-
affected zone took place through the welding process. For the other base metals, in some
cases significantly higher strengths were achieved in the area of the heat-affected zone than for
the unaffected base material.

1103
Table 1. List of filler metals.
Filler metal Designation Standard

Union K 40 G 35 A M23 Z2Si1 EN ISO 14341-A


Union K 52 G 42 4 M21 3Si1 EN ISO 14341-A
Union MoNi G 62 5 M21 Mn3Ni1Mo EN ISO 16834-A
Union X 85 G 79 5 M21 Mn4Ni1,5CrMo EN ISO 16834-A
Union X 90 G 89 6 M21 Mn4Ni2CrMo EN ISO 16834-A

The choice of welding consumables was made taking into account the mechanical prop-
erties of the base metals. Only solid wire electrodes for metal active gas welding accord-
ing to EN ISO 14341 and EN ISO 16834 were used for the investigations carried out.
The welding consumables used in the research project and their complete designation are
given in Table 1.

3 EXAMINATIONS OF THE MECHANICAL PROPERTIES OF WELDS

Within the ongoing AiF-FOSTA research project P1020 a new test specimen was devel-
oped to determine the mechanical properties of welds. A flat tensile test specimen based
on EN ISO 6892-1 and EN ISO 4136 with a centric hole in the weld root area was
used. In conjunction with the contour of the test specimen, a clear definition of the weld
area and the notch effect at the root of the weld as well as the transition to the base
metal takes place. Therefore, test results are comparable and reproducible. Furthermore,
the failure of the test specimen occurs in the weld itself. This is also the case for over-
matching welds where the strength of the welds is higher than that of the base metal.
This failure is necessary for a quantitative determination of the mechanical properties of
the weld. Figure 1 shows the geometry of the test specimen and the weld area before
and after the test procedure.
The specimen is part of a new test procedure for welded joints, designed as a comparative
test. Using the new design approach, the mechanical properties of the welds can be transferred
to the component level taking into account the type of joint and the dimensional ratios
(Stroetmann et. al. 2018a, 2018b).
To determine the mean value of the weld stress the force (F) of the testing machine will be
divided by the net section area of the test specimen (S0) as shown in equation 1. The yield
and tensile strengths determined this way are test specific mechanical properties of the weld.
The mean strain will be derived from the ratio of the elongation (ΔL) to the initial gauge
length (L0) (see equation 2). This is not a local but a mean value over the measuring range.
It serves as a comparative value for assessing the ductility of the welds.

Figure 1. Geometry of the test specimen and range of initial gauge length before and after the test.

1104
σ w ¼ F = S0 ð1Þ

ε ¼ ΔL = L0 ð2Þ

3.1 Test program


The mechanical properties of welds are influenced in different ways. From interest are the
combination of base and filler metal, the execution parameters and the number of layers of
the welds. These influences are partly dependent on each other. Therefore, several parameter
studies were performed to examine the influences on the mechanical properties within the
framework of the AiF-FOSTA research project P1020.
In addition to the determination of the mechanical properties, metallographic examin-
ations were carried out. These include the preparation of macro-sections of each weld,
microstructure analyses and hardness measurements. Furthermore, miniature tensile spe-
cimens were taken from the area of the base material, the heat-affected zone and the
weld metal.

3.2 Manufacturing of the test specimens


The welds of the samples were produced under laboratory conditions fully mechanized at the
Professorship for Welding Technology of the TU Chemnitz as well as manually by industrial
partners. The fully mechanized specimens were welded at a specially designed test setup
including a six-axis welding robot and a synchronized rotatable table, where they were stored
at certain points. For reasons of comparability, the electrical power of the welding process
was chosen on an almost constant basis for all examinations. During the complete welding
process, the execution parameters and the distortion forces were measured and recorded. The
desired cooling times t8/5 were achieved by varying the welding speed, preheating and interpass
temperatures. By applying thermo-elements in the liquid weld metal, the cooling time was
measured for each welding bead. In addition, macro-sections of each weld were prepared and
metallographic examinations were carried out. The microstructure of selected test specimens
was compared with the results of the quenching tests to evaluate the measuring systems and
the configuration of the test setup.
To produce the flat tensile specimens from the welded plates, thin segments were cut out
and the contour of the sample were milled. Then, the hole was made in the area of the weld,
taking into account the structure of the macro-sections. During production, the temperatures
of the specimens were limited so that no microstructural changes occurred.
The opening angle of the welds was for all specimens 90°. Two different geometries were
used. A reduction of the specimen width (b0) for the plates with the two-layer welds was neces-
sary due to the smaller weld volume compared to the three-layer welds. The dimensions of the
two-layer specimen, including the diameter of the hole (d), were scaled accordingly and the
initial gauge length was reduced based on the change in the net cross-sectional area. The speci-
men dimensions are summarized in Table 2. Thereby the terms and formula symbols of EN
ISO 6892-1 were used. The test of the specimens was performed strain-controlled with
a displacement rate of 0.5 mm/min.

Table 2. Specimen dimensions and initial gauge lengths in mm (see Figure 1).
Specimen geometry b0 t d L0

Geometry A 15.0 5.0 5.0 25.0


Geometry B 13.5 4.5 4.5 20.0

1105
3.3 Influence of the welding process
For the assessment of the influence of the welding process, the base metals S500ML, S700MC
and S690QL were welded with filler metals G42, G62 and G79. The process parameters were
selected in a way that a cooling time t8/5 of almost 12 seconds was achieved for each material.
For the assessment of the influence of the welding process, the base metals S500ML, S700MC
and S690QL were welded with filler metals G42, G62 and G79. The process parameters were
selected in a way that a cooling time t8/5 of almost 12 seconds was achieved for each material
combination. The test were carried out with the specimen geometry A. Figure 2 shows the
macro-sections of a fully mechanized (left) and a manually (right) produced weld.
The fully mechanized produced weld has an approximately symmetric structure in which
the penetration of the roots of the individual weld beads is low. In contrast to this, the weld
structure of the manually produced weld is more asymmetrical and there is a clear penetration
of the individual weld beads. In the visible areas of the heat-affected zones, there are clear
differences in the geometric characteristics and overlapping areas. The volume of the manually
produced weld is larger than that of the fully mechanized process. Representative for the test
with the filler metal G79, the stress-strain curves of the combination with the steel S690QL are
shown in Figure 3.
An influence of the manufacturing process on the mechanical strength properties of the
weld can be recognized only to a small extent. The average tensile strength of manually pro-
duced welds is 3% lower than for fully mechanized produced welds. For the yield strength,
however, the difference is 2% and therefore negligible. There are slight differences in the duc-
tility of the welds and the associated elongation at break. One reason for this may be the
described differences in the geometry of the welds. For this purpose, numerical investigations
are currently carried out to assess the influence of local strengths and strains within the weld
cross-section and the heat-affected zone.
For the undermatching material combination S690QL-G42 the stress-strain curves are
shown in Figure 4. The behaviour with regard to the strength properties is similar to that of
the overmatching combination S690QL-G79. The mean tensile strength of the manual pro-
duced welds is about 1% below the fully mechanized produced specimens. For the yield
strength the difference is 3%. It is slightly larger than for the overmatching material combin-
ation. Further differences lie in the ductility of the manually produced welds, which on aver-
age exceeds the values of the fully mechanized produced welds.
Table 3 shows the averaged tensile strengths of all tests depending on the filler metal. It can
be seen that the tensile strengths of the manually produced welds are always slightly lower than
those of the fully mechanized produced welds. However, the differences are within 3.8% of the
standard deviation for the tensile strength. The influence of the degree of automation on the

Figure 2. Macro sections of welds produced fully mechanized (left) and manually (right).

1106
Figure 3. Stress-strain curves of the material combination S690Q – G79 for fully mechanized and
manually produced welds.

Figure 4. Stress-strain curves of the material combination S690Q-G42 for fully mechanized and manu-
ally produced welds.

Table 3. Average tensile strength in N/mm² and tensile


strength ration of the manually and fully mechanized produced
welds.
Filler metal G42 G62 G79

fwu - manual 638 777 857


fwu - full mechanized 646 806 884
Tensile strength ratio 0,99 0,96 0,97

strength characteristics of the welds can be considered as low for the same cooling time t8/5. The
influence of the weld geometry on the ductility will be examined in more detail.

1107
3.4 Influence of the number of weld layers
The base materials S700MC and S690QL were welded with the filler metals G42 and G79 to
examine the influence of the number of weld layers on the mechanical properties of welds. In
addition, the material combination S960QL–G42 was examined. The specimens were welded
fully mechanized with a cooling time t8/5 of 12 seconds.
Figure 5 shows the macro sections of welds with three (left) and two layers (right). For the
3-layer welds a total of 8 weld beads were introduced, for the 2-layer welds 4 weld beads. With
the large opening angle and the chosen cooling time, two weld beads were needed to get
a complete flange connection in the 3rd layer of the weld.
The stress-strain curves of the material combination S700MC-G42 shown in Figure 6 are
exemplary for the behaviour of under-matching welds. The tensile strength of the weld with
three layers is 4 %, the yield strength 7 % lower than for the weld with two layers. In addition,
the ductility of the welds with two layers is slightly higher. This behaviour can also be
observed for overmatching welds. In contrast to the under-matching weld, the characteristic
strength values are at a similar level.
Table 4 summarizes the average values of the yield and tensile strength as well as the corres-
ponding ratio values of the welds with three and two layers. For under-matching welds, there is
a slight influence of the number of layers on the strength. However, in the case of overmatching

Figure 5. Macro section of a weld with three (left) and two layers (right).

Figure 6. Stress-strain curves of the material combination S700M–G42 for two- and three-layer welds.

1108
Table 4. Average strengths in N/mm² for welds with three and two layers and their ratio values.
Filler metal G42 G79 Filler metal G42 G79

fwu – 3 layer 658 884 fwy – 3 layer 541 752


fwu – 2 layer 687 885 fwy – 2 layer 584 749
Tensile strength ratio 0.96 1.00 Yield strength ratio 0.93 1.01

welds this was not observed. One reason for this behaviour are the differences in the alloying con-
cepts of the filler metals. The G79 filler metal used is highly alloyed and has, among other things,
a fin-grain-forming alloying element in titanium. When the weld metal will be heated above the
Ac3 temperature, such as in the case of multilayer welds, a finer-grained microstructure will be
produced, which leads to higher strengths. The filler metal G42 does not contain any fine-grain-
forming alloying elements, multilayer welds result in locally coarser microstructures and a drop in
strength.
The mixing of liquid filler metal and molten base material during the welding process can also
have an influence on the results. With a lower number of weld layers, the proportion of melted
flank material is higher than for more layers. The distribution of these proportions in the weld
metals depends on the dominant melt flow. For undermatching welds, the influence of the molten
flank material can be higher as for overmatching welds were the strength differences between base
and filler metal are small. However, the test results do not provide any unambiguous information
for this assumption. In addition to the tensile tests carried out, hardness measurements were per-
formed on welds with three layers in two planes. The tests were carried out in accordance to EN
ISO 9015-1 using the Vickers hardness test method with a test load of 9.81 N. The results and the
position of the measuring points for the hardness measurement on the weld S700MC–G42 with
three weld layers and a cooling time of 20 seconds is shown in Figure 7.
An analysis of the hardness values in the heat-affected zone shows that the values are below the
hardness of the base material. This can be seen in the upper (layer 1) and lower measuring series
(layer 2). It can be stated that the maximum hardness values of the heat-affected zone at the lower
measurements are lower than those of the upper measurements due to the repeated thermal influ-
ence. In addition, the expected drop in hardness between the heat-affected zone and the weld
metal is clearly shown in both measurement series. The hardness values of the upper measure-
ments are higher in the edges than in the middle area. This results from welding of the cover layer,
which was carried out with two welding beads. From this, multiple microstructural

Figure 7. Hardness curves of the material combination S700M-G42 for a cooling time of 20 seconds.

1109
transformations and tempering effects were happening in this area. This is also indicated by the
increasing tendency of the hardness from left (first cover layer) to the right (second cover layer) in
the middle of the weld metal. Quantitatively, differences in hardness of up to 35 HV1 can be
observed in this area. The hardness values in the lower measuring range are above those of the
centre of the weld metal in the upper measuring range, as these have been subjected to less thermal
influence. Taking into account a correlation between tensile strength and hardness, the series of
hardness measurements confirm qualitatively the results of the tensile tests.

4 SUMMARY

In this paper, experimental investigations on the influence of the welding process and the
number of layers on the strength and ductility of the welds are presented. The investigations,
which include undermatching and overmatching welds, have been carried out with a new test
procedure, which is part of a new design approach for welded joints in the frame of the
ongoing AiF-FOSTA research project P1020. It could be determined with the scope of the
examinations carried out, that the influence of the base material on the mechanical properties
of the welds is negligible. The filler metal is decisive for the load-bearing capacity of the weld.
This behaviour was observed for welds produced fully mechanized under laboratory condi-
tions and in welds produced manually by industrial partners. A comparison of the degree of
automation of the welding process has shown that with manually produced welds approxi-
mately the same strength values can be achieved as with welds produced fully mechanized.
Studies on the influence of the number of layers on the strength and ductility of welds have
shown that a differentiation must be made between the individual welding consumables.
While no influence of the number of layers on the strength characteristics was found for the
high-alloyed filler G79, the tensile and yield strength decreased for the low alloyed filler G42
with higher number of layers. One reason for this lies in the different alloying concepts of the
welding consumables and the associated chemical composition of the weld metal. Fine grain
alloying elements have a positive influence on the strength of multi-layer welds. In addition,
there are slight differences in the ductility of the welds depending on the number of layers.
Among other things, these can be attributed to the structure of the weld metal. The assump-
tion, which takes into account a mixing of the molten base material and liquid filler metal
during the welding process, could neither be confirmed nor denied by the experimental tests.
For this reason, further investigations are necessary taking into account a lager selection of
mismatching ratios and filler metals with different alloying concepts.
The results of the presented investigations will be considered in the development of the new
design approach for welds. Detailed knowledge about various influencing parameters on the
strength and ductility of welds allow an economical design of welded joints.

ACKNOWLEDGMENT

The IGF-project 19043 BR/P1020 “Economic welding of high-strength steels” of the FOSTA –
Forschungsvereinigung Stahlanwendung e. V. Düsseldorf, is encouraged within the program for
the promotion of Industrielle Gemeinschaftsforschung (iGF), funded by the Federal Ministry of
Economics and Energy on the basis of a resolution of the Deutscher Bundestag. The authors of
this paper would like to thank for this funding and the participation of the industry partners.

REFERENCES

Stroetmann, R. & Kästner, Th. & Hälsig, A. & Mayr, P. 2018a. Influence of the cooling time on the mech-
anical properties of welded HSS-joints. Steel Construction 11 (2018), Nr. 4, S. 264–271.
Stroetmann, R. & Kästner, Th. & Werner, L. 2018b, Welds for high-strength steels – Development of
new design rules, 40th IABSE Symposium – Tomorrow´s Megastructures 2018, Nantes.

1110
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Validation of the Overall Stability Design Methods (OSDM) for


tapered members

J. Szalai
ConSteel Solutions Ltd., Budapest, Hungary

F. Papp & G. Hajdú


Faculty of Architectural, Civil and Transportation Engineering, Széchenyi István University, Győr,
Hungary

ABSTRACT: Two new stability design methods are demonstrated and validated: the Overall
Strength Reduction Method (OSRM) and the Overall Imperfection Method (OIM). Both
methods are based on the linear buckling analysis (LBA) of global structural models and use the
standard reduction curves. The OSRM is formulated in the classic way using generalized slender-
ness and reduction factors while the OIM uses equivalent amplitude for the buckling mode based
geometrical imperfection. These new design methods cover all types of buckling modes, which
can be calculated by LBA of structural models composed of tapered members with arbitrary sup-
port conditions and subjected to any complex loading. This paper clarifies the mechanical inter-
pretation and proper calculation of all the components of the two methods in case of tapered
members with arbitrary support conditions. The validation is performed on GMNIA results for
several different buckling situations of tapered members proving the accuracy of the OSDM.

1 INTRODUCTION

In Szalai & Papp (2019) a new design methodology is presented which utilizes the overall Linear
Buckling Analysis (LBA) results of any structural model. The Overall Stability Design Methods
(OSDM) have two equivalent alternatives based on the same underlying mechanics: the Overall
Strength Reduction Method (OSRM) which uses the traditional reduction factor for the calcula-
tion of the design buckling resistance and the Overall Imperfection Method (OIM) which calcu-
lates the equivalent amplitude for the buckling mode based geometrical imperfection. Both
methods are based on an essential underlying assumption which states that any complex global
buckling mode calculated by the LBA can be classified into finite number of fundamental buck-
ling mode types which are significantly different in terms of various mechanical characteristics
(loading, mode shape displacement components etc.). It is also assumed that the well-known, cali-
brated standard buckling curves are solutions for some of these fundamental buckling modes (for
flexural buckling or LTB of doubly symmetric cross-sections etc.) which can be used within the
OSDM to ensure their reliability level. Accordingly the methodology consists of two basic steps:
(1) a universal transformation method which converts the real structural model with a certain
complex buckling problem into a properly defined equivalent reference member which is
a prototype model of the corresponding fundamental buckling mode type – this is the ultimate
generalization of the effective length (or equivalent member) method to any buckling problem
(2) a closed-form analytical solution for the reference member which is based on the stand-
ard buckling curves corresponding to the equivalent fundamental buckling mode type – this is
the ultimate generalization of the beam-column buckling strength interaction equations.
The calculation scheme of the OSDM with this two basic steps is illustrated in Figure 1.
In this paper the concrete calculation steps for both methods are presented with a benchmark
example on a tapered member, then the results of the numerical validation is shown.

1111
Figure 1. Calculation scheme of the OSDM.

2 STEPS OF THE OSDM

In this section the calculation steps of the OSDM are presented based on Szalai & Papp
(2019) where a bit more detailed background is given for the formulas. In this paper for the
sake of simplicity and brevity the following assumptions are made:
1. The pre-buckling deformations from initial loading are negligible so the second order
effects are due entirely to the imperfections. The handling of second order effects due to
loading is described in Szalai (2017) and Szalai (2018).
2. The real structural model is a single web tapered member with doubly symmetric I cross-
section and arbitrary load and support condition. Application of the OSDM to plane
frames is demonstrated in Szalai & Papp (2019).

2.1 Numerical calculations on the real structural model

2.1.1 Step 1: Structural analysis


Two numerical structural analysis results are required for the OSDM performed on the perfect
real structural model: linear elastic analysis (LA) and linear buckling analysis (LBA). The fol-
lowing results are used further (x denotes the longitudinal coordinate along the reference line
of the tapered member):


– first order internal forces and moments: S I ðxÞ ¼ N I ðxÞ; M I v ðxÞ; M I z ðxÞ; BI ðxÞ
– elastic critical buckling load factor: αcr
– buckling mode shape: ηcr ðxÞ ¼ ½0; wcr ðxÞ; vcr ðxÞ; ’cr ðxÞ

2.2 Forward model transformation


In this phase an equivalent reference member (ERM) is defined by the previous results
obtained on the real structural model. The ERM in general is a straight, prismatic, simply sup-
ported member subjected to uniform compression force and/or bending moments. For the
proper definition of the ERM the following data should be determined:
– geometry: cross-section and member length
– member loads (causing uniform compression force and/or bending moments)
– buckling mode type (one of the fundamental buckling modes)
These set of data is defined through a suitably selected equivalent point (ep) along the real
structural model where the second order internal stress utilization effect from the buckling

1112
mode shape is the highest (or the second order flexural curvature of the compressed flange
from the buckling mode is the highest).

2.2.1 Step 2: Determination of the equivalent point (ep)


The recovery of the equivalent point is practically done by calculating the internal force and
moments due to the deformation of the buckling mode shape along the longitudinal member
axes of the real structural model:


S cr ðxÞ ¼ 0; M cr y ðxÞ; M cr z ðxÞ; Bcr ðxÞ ð1Þ

and calculating the corresponding cross-section resistances (which is varying along the tapered
member axis) using the appropriate classes:

1 1 1 1
RðxÞ ¼ ð2Þ
Nsec ðxÞ My;sec ðxÞ Mz;sec ðxÞ Bsec ðxÞ

then the linear utilization function form these internal force and moments can be finally
determined:

1
Usec;cr ðxÞ ¼ RT ðxÞS cr ðxÞ ¼ ð3Þ
αsec;cr ðxÞ

where αsec;cr ðepÞ is the corresponding linear load multiplication factor (LMF). The equivalent
point x = ep is where the utilization function of Eq. (3) takes the highest value:

Usec;cr ðepÞ ¼ max Usec;cr ðxÞ ! min αsec;cr ðxÞ ! x ¼ ep ð4Þ

From the equivalent point the following data can be received directly:
– the cross-section of the ERM and its class and resistances
– the uniform compression force and/or bending moments acting on the ERM

2.2.2 Step 3: Buckling mode classification through the equivalent point


For the calculation of the length of the ERM the proper buckling mode type should be defined,
since the obtained cross-section and loads can generally lead to various buckling mode types. The
classification into a fundamental buckling mode class (BMC) is done by the buckling mode shape
components and the loading at the equivalent point of the tapered member according to Table 1.

2.2.3 Step 4: Equivalent length of the reference member


The length of the ERM (Leq) can be calculated from the well-known analytical formulae for
the proper BMC where the only unknown parameter is the length (Table 2.).

2.3 Analytical solution of the ERM


Once the ERM is fully defined it is solved based on the generalized Ayrton-Perry formulation
defined by Szalai (2017).

2.3.1 Step 5: Equivalent imperfection factor


For the different fundamental BMCs the equivalent imperfection factors are shown in Table 3.
For the pure modes these are the calibrated standard imperfection factors of the Eurocode 3,
and for the coupled mode it is based on the interpolated formulation derived in Szalai (2017).

1113
Table 1. The fundamental Buckling Mode Classes (BMC).
Cross-section at Active load Buckling mode shape
BMC the ep components at the ep component(s) at the ep Buckling mode type

BMC_01 NI fwcr g strong axis flexural buckling


BMC_02 NI fvcr g weak axis flexural buckling
BMC_03 doubly NI f’cr g torsional buckling
BMC_04 symmetric MI y fvcr ; ’cr g lateral-torsional buckling
BMC_05 N I ; MyI fvcr ; ’cr g coupled lateral-torsional
buckling

Table 2. Equations including the Leq equivalent length


which can be expressed for the ERM.
BMC_01 αcr N I ¼ Ncr;y ¼ π2 EIy =L2
BMC_02 αcr N I ¼ Ncr;z ¼ π2 EIz =L2 
BMC_03 αcr N I ¼ Ncr;x ¼ 1=r2 GIT þ π2 EIw =L2
p ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
BMC_04 αcr MyI ¼ r Ncr;z Ncr;x
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 ffi
BMC_05 αcr MyI ¼ r Ncr;z  αcr N I Ncr;x  αcr N I

Table 3. The equivalent imperfection factors.


 
BMC_01 ηERMNy ¼ αy λy  0; 2
BMC_02 ηERM
Nz ¼ αz ðλz  0; 2Þ
BMC_03 ηERM
Nx ¼ αz ðλT  0; 2Þ
BMC_04 2
ηERM
My ¼ ðλLT =λz Þ αLT ðλz  0; 2Þ
MyI
NI
Nsec My;sec αsec;a ERM αsec;a ERM
BMC_05 ηERM ¼ ηERM þ μ I ηERM
My ¼ η þμ η
coupled MI
þ My;sec
N MyI αsec;N N αsec;My My
Nsec þ My;sec
NI y N
Nsec

2.3.2 Step 6 in OIM: Second order effect on the reference member


From this point there are different calculation steps for the OIM and OSRM. In case of the
OIM the basic result of the ERM which is used for the backward transformation is the proper
value of the second order effect which can be calculated from the equivalent imperfection
factor. It is done in the form of linear load multiplication factor as defined in Szalai & Papp
(2019) using the general form of the imperfection factor as derived in Szalai (2017) as follows:
 
αsec;a 1
αII;ERM ¼ ERM 1  ð5Þ
sec;cr
η αcr

2.3.3 Step 6 in OSRM: Equivalent reduction factor


The base for the backward result transformation in the case of OSRM is the reduction factor
of the ERM. It is a straightforward calculation using the generalized slenderness
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
(λERM ¼ αsec;a =αcr ) and the equivalent imperfection factor in the well-known Ayrton-Perry
based reduction formula:

1 1 2

χERM ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi where  ¼ 1 þ ηERM þ λERM ð6Þ
2 2
 þ 2  λERM

1114
2.4 Backward result transformation
When the ERM has the proper solution its results should be transformed back to the real
structural model to reach the final buckling solution. The steps are now going again on two
paths for the OIM and the OSRM.

2.4.1 Step 7 in OIM: Equivalent geometrical imperfection


In the OIM the correct design amplitude is to be determined for the applied equivalent geomet-
rical imperfection with the shape of the calculated buckling mode of the tapered member.
Important to note that the buckling mode shape (ηcr ðxÞ) can have arbitrary amplitude when cal-
culating S cr ðxÞ of Equation 1 and it is linearly dependent on the actual amplitude. Accordingly
the proper amplitude scale factor (δeq ) can be calculated from the equality of the second order
II
effect of the real structural model at the equivalent point (Usec;cr ðepÞ) and the ERM (Usec;cr
II;ERM
):

1 1 ηERM 1
II
Usec;cr ðepÞ ¼ Usec;cr
II;ERM
! δeq ¼ ð7Þ
αsec;cr ðepÞ αcr  1 αsec;a 1  α1
cr

The correct amplitude of the equivalent geometrical imperfection of the real structural model
which after some arrangement takes the following form:
αcr
ηcr;eq ðxÞ ¼ δeq ηcr ðxÞ ¼ ηERM αsec;cr ðepÞηcr ðxÞ ð8Þ
αsec;a
The final step of the OIM is running a second order analysis and a cross-section check on the
real structural model with the equivalent geometrical imperfection.

2.4.2 Step 7 in OSDM: Non-uniform equivalent reduction factor


Since by definition the reduction factors includes all the second order effects due to the proper
equivalent geometrical imperfection in the OSRM the equivalency relationship means the
equality of the reduction factors of the ERM (χERM ) and the real model at the equivalent
point (χea ðepÞ). Important to note that the equivalent reduction factor of the real structural
model is not constant in general but its distribution follows the distribution of the second
order effects defined by Equation 3. Accordingly after some manipulation – described in
Szalai (2018) – the final from of the equivalent non-uniform reduction factor along the real
structural model can be written as follows:
1
χeq ðxÞ ¼ χERM mðxÞ ¼ χERM ð9Þ
αsec;cr ðepÞ αsec;a ðxÞ
χERM þ ð1  χERM Þ
αsec;cr ðxÞ αsec;a ðepÞ

The final check of the buckling resistance along the real tapered member takes the following form:

1
1 ð10Þ
χeq ðxÞαsec;a ðxÞ=γM1

3 ILLUSTRATIVE EXAMPLE

The calculation steps of the OSDM is demonstrated on a tapered member (Figure 1) with
S235 material, symmetric I section of flanges 300 mm/14 mm and web 290-435 mm/8 mm,
span of 8629 mm, simple supports at the ends and a lateral restraint at the top flange at
a distance of 4314 mm from the deeper section end. The member is subjected to uniform com-
pression (N = 379,64 kN) and uniform bending moment (My = 233,36 kNm). The dominant
LT buckling mode of the tapered member is shown in Figure 2. All the necessary calculations
are performed in ConSteel 12 (2018) software using 7 DOF beam-column element accounting
for the tapering effect, the steps are shown in Table 4.

1115
Figure 2. Geometry, support and load condition of the tapered member example.

Figure 3. LTB mode shape of the tapered member example.

Table 4. General steps of OIM and the OSRM for tapered member.
Designations of parameters Notation Dimension Value

Step 1.1 Linear elastic analysis (LA)


– Axial force NI kN 379.64
– Bending moment MIy kNm 233.36
Step 1.3 Linear buckling analysis (LBA)
– Elastic critical load factor αcr 4.45
– Amplitude of buckling shape on the vcr.max mm 9.698
beam centroid
Step 2.1 Internal forces and moment due to
buckling mode shape
– Bending moment in the equivalent Mcrz(ep) kNm 21.26
point
– Bimoment in the equivalent point Bcr(ep) kNm2 2.86
Step 2.2 Cross-section resistance
– Axial compression Nsec(ep) kN 2530,2
– Bending around major axis My,sec(ep) kNm 346,9
– Bending around minor axis Mz,sec(ep) kNm 14,91
– Bimoment Bsec(ep) kNm2 2,293
Step 2.3 Linear load multiplication factor in αsec,cr(ep) 1.085
the equivalent point
Step 2.4 Location of the equivalent point ep mm 6615
Step 3.1 Buckling-active internal forces and
moments
(Continued )

1116
Table 4. (Continued )
– Axial force Na(ep) kN 379.64
– Bending moment around the major axis May(ep) kNm 233.36
Step 3.2 Classification of the buckling mode BMC_05
Step 4 Equivalent length of the reference Leq mm 4170
member
Step 5 Equivalent imperfection factor
– Imperfection factor for FB αz 0.490
– Reduced slenderness for FB λz 0.584
– Imperfection factor for FB ηN 0.188
– Imperfection factor for LTB αLT 0.378
– Reduced slenderness for LTB λLT 0.516
– AP imperfection factor for LTB ηMy 0.113
– Modifying factor μ 0.979
– Equivalent imperfection factor ηeq 0.125
Step 6.1 Buckling-active linear multiplication αsec,a(ep) 1.226
factor (in equivalent point)
Step 6.2 Second order effects linear multipli- αIIeq,cr 7.608
cation factor
Step 7 Equivalent scale factor δeq 0.492
Final check
– Second order internal forces and
moments due to equivalent geometric imper-
fection (in critical point)
● Axial compression NII,imp kN 379.6
● Bending around major axis MII,impy kNm 243.8
● Bending around minor axis MII,impz kNm 10.06
● Bimoment BII,imp kNm2 1.320
Step 8 Maximum cross-section utilization (in Umax 0.985
critical cross-section)
Ultimate load factor αu 1.008
Step 6.1 Buckling-active linear multiplication αsec,a(ep) 1.226
factor (in equivalent point)
Step 6.2 Equivalent reduced slenderness λERM 0.678
Step 6.3 Equivalent reduction factor χERM 0.832
Step 7.1 Non-uniform load effect modification mcr 1.023
factor (in critical cross-section)
Step 7.2 Non-uniform reduction factor (in χeq,cr 0.850
critical cross-section)
Step 8 Maximum buckling utilization (in crit- Umax 0.990
ical cross-section)
Ultimate load factor αu 1.002

4 VALIDATION

In order to validate the accuracy of the proposed OSDM in the buckling design of tapered
members numerical GMNIA performed on three different support and load configurations:

Case 1: simple support, uniform compression and bending moment


Case 2: simple support, uniform compression, linear bending moment
Case 3: simple support with an intermediate compressed flange restraint, uniform compres-
sion and bending moment
In each cases several different cross-section types, tapering ratio and slenderness values are
considered. The detailed description of the numerical model can be found in Hajdú (2019).

1117
The normalized buckling resistance results from the GMNIA calculations (re) and from the
proposed OSDM (rt) is plotted in Figures 4–6.

Figure 4. Validation plot for Case 1.

Figure 5. Validaqtion plot for Case 2.

1118
Figure 6. Validation plot for Case 3.

5 CONCLUSIONS

Two new buckling design methods are presented for the buckling verification of tapered mem-
bers. The Overall Imperfection Method (OIM) uses directly a properly scaled buckling mode
based equivalent imperfection and the Overall Strength Reduction Method (OSRM) general-
ized the classical buckling design based on reduction factors. The calculation steps of the
methods are presented in general and illustrated on a benchmark example with a tapered
member with intermediate restraint. Finally the methods are validated against GMNIA results
performed on tapered members with different cross-sections, slenderness, support and load
conditions. The methods are suitable for software implementation providing a fully automatic
and economic way of buckling design for any tapered members.

REFERENCES

ConSteel 12 2018. Structural analysis and design software, www.consteelsoftware.com.


Hajdú, G. 2019. The Validity of the Universal Transformation Method in Global Buckling Design. The
14th Nordic Steel Construction Conference, September 18–20, 2019, Copenhagen, Denmark.
Papp, F. & Szalai, J. & Movahedi, R. M. 2019. Out-of-Plane Buckling Assessment of Frames through
Overall Stability Design Method. The 14th Nordic Steel Construction Conference, September 18–20,
2019, Copenhagen, Denmark.
Szalai, J. & Papp, F. 2019. New stability design methodology through overall linear buckling analysis.
The 14th Nordic Steel Construction Conference, September 18–20, 2019, Copenhagen, Denmark.
Szalai, J. 2017. Complete generalization of the Ayrton-Perry formula for beam-column buckling
problems. Engineering Structures 153:205–223.
Szalai, J. 2018. Direct buckling analysis based stability design method of steel structures. Ninth International
Conference on Advances in Steel Structures (ICASS2018) 5–7 December, 2018, Hong Kong, China.

1119
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Stability design of cable-stayed columns

T. Tankova, L. Simões da Silva & J.P. Martins


Institute for Sustainability and Innovation in Structural Engineering, University of Coimbra, Portugal

ABSTRACT: Pre-stressed stayed columns offer several advantages among traditional ones:
they have enhanced buckling resistance in comparison to conventional columns provided by
the pre-stressed cables and cross-arms; they are also aesthetically appealing, combining
a strong architectural effect with a top notch of engineering solution. In this paper, design rule
for stability design of pre-stressed cable-stayed columns is presented. It is based on the well-
known Ayrton-Perry format, i.e., on a combination of first and second order effects. It is con-
sistent with the design rules for uniform columns in Eurocode 3.

1 INTRODUCTION

The design of long span structures implies using slender columns whereby buckling dominates the
load-carrying capacity. In such cases, using high strength steel has little, or no effect, on the load
carrying capacity and so the common solution is to increase the cross-section dimensions. How-
ever, improvement of the stability resistance can be achieved by the adoption of structural solu-
tions or configurations which increase the buckling resistance, such as cable-stayed columns. In
fact, adding stays and cross-arms to slender columns can improve both the elastic buckling load
and the load carrying capacity as they provide translational and rotational restraint along the
length.
Cable-stayed columns have been studied in the past by several researchers. First attempts
are found by Chu & Berge (1963) and Mauch & Felton (1967) in pursuit of the buckling load
of cable-stayed columns. Many studies were developed in the 70s, starting with the work by
Smith et al. (1975) whereby expressions for the determination of the maximum critical load in
symmetric and asymmetric buckling shapes were proposed. Numerical solutions of the prob-
lem can be found by Khosla (1975) and Hathout (1977). A significant advance in the know-
ledge of the behaviour of cable stayed columns is the paper by Hafez et al. (1979) who
proposed expressions for the optimum pre-stressing force. Furthermore, the effect of initial
imperfections on the buckling load was investigated based on experimental tests by Wong &
Temple (1982) and later extended using FE simulations, Temple et al. (1984). Smith (1985)
provided analytical studies to distinguish between buckling modes.
Later on, the second-order instability of stayed columns was derived including initial imper-
fections by Chan et al. (2002). The advantage of using split up cross-arms was studied numer-
ically and compared with other type of cross-arms by Van Steirteghem et al. (2005).
A series of full-scale tests were performed in Araujo et al. (2008, 2009), which were used for FE
simulations and an extensive parametric study was performed to assess the effect of different
parameters on the buckling behaviour. The post buckling of the stayed column was assessed by
the application of the Rayleigh-Ritz method and the results were compared with numerical simu-
lations in Saito & Wadee (2008, 2009a, 2009b). Osofero et al. (2012a, 2012b) summarized the
assessment of the effects of imperfection shape, orientation and magnitude on the buckling behav-
iour of columns. The sensitivity of the load-carrying capacity to the geometry of the stayed
column, the initially applied pre-stress level within the stays and the initial global imperfection
was investigated numerically. A general design procedure for pre-stressed stayed columns with
a single cross-arm system was proposed (Wadee et al., 2013). Finally, Serra et al. (2015) and

1120
Martins et al. (2016) presented the results of an experimental study on 12 and 18 m pre-stressed
stayed column with single and double cross-arms.
In this paper, a design approach in an Ayrton-Perry format is presented and explained. Firstly,
the essential background for the buckling resistance of columns is briefly summarized. Further-
more, the necessary adjustments to account for the enhanced structural behaviour are explained.
The proposed expressions are calibrated against the experimental results obtained in Serra et al.
(2015).

2 BUCKLING RESISTANCE: BACKGROUND

2.1 Introduction
In this section the theoretical background behind the adopted approach is briefly presented. It
is based on the current design format of Eurocode 3 as a reduction factor to the plastic resist-
ance of the column. For that, the background of the design rules for flexural buckling of pris-
matic columns is presented first. Then a short discussion on the possible buckling modes of
stayed columns is given. Finally, the assumptions behind the proposed model are presented.

2.2 Flexural buckling of simply supported columns


The behaviour of a perfect simply supported compressed column with length L is described by
the following differential equation and boundary conditions

EIv00 þ Nv ¼ 0 with vð0Þ ¼ vðLÞ ¼ 0 and v00 ð0Þ ¼ v00 ðLÞ ¼ 0 ð1Þ

The solution of Equation (1) yields the elastic critical load and mode shape:

π2 EI πx
Ncr ¼ with vðxÞ ¼ v sin ð2Þ
L2 L

In a similar way, it is also possible to describe the imperfect simply supported columns
(Figure 1):

EIv00 þ Nv þ Nv0 ¼ 0 with vð0Þ ¼ vðLÞ ¼ 0 and v00 ð0Þ ¼ v00 ðLÞ ¼ 0 ð3Þ

which leads to relationship between the amplitude of the lateral displacement and the ampli-
tude of the initial imperfection:
N
v ¼ v0 ð4Þ
ðNcr  NÞ

The design equation is established on the basis of linear yield criterion for the imperfect
column. It accounts for the axial load and second order bending moment that arises due to
initial imperfections:

N M II N N ðv þ v0 Þ
þ ¼ 1:0 þ ¼ 1:0 ð5Þ
Afy Wfy Afy Wfy

Figure 1. Column buckling: deformed configuration.

1121
where N is the applied axial force, MII is the bending moment due to second order effects, Afy is
the column’s plastic axial capacity and Wfy is the column’s elastic bending. Inserting Equation (4)
into Equation (5) it is possible to express the equation as a function of the initial imperfection:

N N v0
þ ¼ 1:0 ð6Þ
Afy Wfy ð1  N=Ncr Þ

Defining
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi N
λ¼ Afy =Ncr and χ ¼ ð7Þ
Afy

it is possible to conclude that


v0 A χ χ   χ
χþ 2
¼ 1:0 ! χ þ η 2
¼ 1:0 ! χ þ α λ  0:2 2
¼ 1:0 ð8Þ
Wz ð1  χλ Þ ð1  χλ Þ ð1  χλ Þ

where χ is the reduction factor of the column’s axial capacity, η is the generalized imperfection
factor as defined in EC3, which was calibrated to account for initial out-of-straightness and
residual stresses.

2.3 Flexural buckling of stayed simply supported columns


Stayed columns have the advantage of increased buckling load due to the restraint provided
from the cross-arms and the prestressed cables. The elastic buckling load for these columns
was derived by Smith (1975) and was further developed by Hafez et al. (1979) as mentioned
earlier. Generally, two buckling modes are possible: i) symmetric, when the restraint provided
by the cross-arms and stays is not completely sufficient to prevent the displacement at mid-
span (Figure 2, left); and ii) asymmetric, when the restraint at midspan is sufficient and the
buckling is also accompanied by buckling of the cross-arms (Figure 2, right). The earlier stud-
ies on stayed columns report that the predominance of one mode or another depends only on
the geometrical characteristics of the stayed column such as ratio between lengths of the cross-
arms and column and diameter of stays.
The application of the linear yield criterion, as it was shown in the previous section, requires
the consideration of the different buckling modes, where the resistance is given by the lowest.
If Mode 1 (Figure 2, left) is considered, the linear yield criterion can be expressed by Equation
(9), where the additional term accounting for the restraint introduced by the stays is accounted
by an extra bending moment and the critical buckling force calculated using linear buckling
analysis in the presence of the prestressing force (Ncr,a).

Figure 2. Stayed columns buckling modes (Smith et al., 1975).

1122
N M II Q1 L
þ  ¼ 1:0 ð9Þ
Afy Wfy 2Wfy

This moment is calculated using the Equation (42) from Smith et al. (1975) using the reaction
Q1 given by:

Q1 ¼ 2Ks sin2 ðαÞδc ¼ 2Ks sin2 ðαÞv ð10Þ

Hence, the retraining bending moment also depends on the lateral displacement at midspan,
so that Equation (9) becomes

N N v0 Ks sin2 ðαÞvL


þ   ¼ 1:0 ð11Þ
Afy Wfy ð1  N Ncr;f Þ Wfy

After straight-forward transformations, Equation (11) yields:


!
χ Ks sin2 ðαÞL
χþη 1 ¼ 1:0 ð12Þ
ð1  χλ Þ
2 Ncr;f

whereby the slenderness and the reduction factor are defined as:
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi N
λ¼ Afy =Ncr;f and χ¼ ð13Þ
Afy

with
Na
Nf ¼ Na þ 4Tf sin α ¼ þ 4Tini cos α ð14Þ
C2
Ncr;a
Ncr;f ¼ Ncr;a þ 4Tf sin α ¼ þ 4Tini cos α ð15Þ
C2

where Na is the applied force, Tini is the initial prestress in the cables, Ncr,a is numerically
obtained from a LBA (linear buckling analysis) and C2 is given by (Hafez et al. (1979)):

2cos2 α
C2 ¼ 1 þ ! ð16Þ
1 2sin2 α
Kc þ
Ks Kca

1 
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi with  ¼ 0:5 1 þ ηtot þ λ
2
χ¼ ð17Þ
 þ 2  λ
2

The last equation constitutes the proposed method to compute the buckling resistance of pre-
stressed stayed columns corresponding to Mode 1. Since buckling Mode 2 is characterized by
a complete restraint at midspan, therefore, the application of Equation (12) should be cor-
rected with the bending moment arising from the reaction Q2.

1123
3 VALIDATION

3.1 Introduction
The proposed procedure is based on the Ayrton-Perry format currently used for the stability
design of columns according to EN 1993-1-1. It makes use of the critical buckling force calculated
using linear buckling analysis in the presence of the prestressing force (Ncr,a). Furthermore, the
imperfection factors to be used were calibrated based on an experimental investigation (Serra
et al. (2015) and Martins et al. (2016)).
In this section, the proposed approach is validated against experimental results by Serra
et al. (2015). Firstly, the parameters of the experimental programme are presented. The add-
itional calibration of imperfection factors is performed and finally, the results obtained using
the proposed method and method by Wadee et al. (2013) were compared.

3.2 Experimental programme


The experimental programme covered 10 columns, using three different cross-section geometries,
two different cable diameters were tested for each column. The experimental programme covers
steel grades S355 and S690; two lengths 12 and 18m; and also, for each column 5 levels of initial
prestress were tested. The summary of these parameters is given in Table 1 and more details can
be found in Serra et al. (2015). In this study only the specimens with single cross-arms were used.

3.3 Calibration
Based on the analytical proposal from Section 2.3, the imperfection ηtot was calibrated against the
numerical results using Equation (18). It was split into three terms: i) ηEC3 which is the generalized
imperfection directly taken from Eurocode 3 curve a0 (αEC3 = 0.13) for S690 steels and from curve
a (αEC3 = 0.21) for S355 steels; ii) Cpr, which accounts for the cable stiffness; and iii) ηpr, additional
imperfection. These parameters are aggregated in one global parameter following Equation (12).

Table 1. Experimental programme.


Column Column Number of
length cross-Section cross-arms Steel Cables

Code 12m 18m CS1 CS2 CS3 4 8 S355 S690 10mm 13mm

C01-C1 X X X X X
C01-C2 X X X X X
C02-C1 X X X X X
C02-C2 X X X X X
C03-C1 X X X X X
C03-C2 X X X X X
C04-C1 X X X X X
C04-C2 X X X X X
C05-C1 X X X X X
C05-C2 X X X X X
C06-C1 X X X X X
C06-C2 X X X X X
C07-C1 X X X X X
C07-C2 X X X X X
C09-C1 X X X X X
C09-C2 X X X X X
C10-C1 X X X X X
C10-C2 X X X X X

where: CS1 – CHS 101.6 x 8.0; CS2 – CHS 139.7 x 6.3; CS3 – CHS 177.8 x 6.3.

1124
χexp
χexp þ ηtot ¼ 1:0 ð18Þ
1  χ λnum
2
exp

ηtot ¼ ηpr  ηEC3  Cpr ð19Þ

where
!
Ks sin2 ðαÞL
Cpr ¼ 1 ð20Þ
Ncr;f
 
ηEC3 ¼ α λ  0:2 ð21Þ

In Equations (18) and (19) the only unknown is ηpr:


 
1  χexpλnum
2
1:0  χexp
ηpr ¼ ð22Þ
χexp  Cpr  ηEC3

The parameter ηpr was chosen according to Equation (20). It is set to unity once the prestress
is equal to 0 and thus allowing for direct correspondence with the Eurocode 3 expression for
columns without stays. Once the prestress is non-zero, it depends on two coefficients pi and qi,
which in this case were chosen according to the steel grade. Nonetheless, global coefficients
could be also adopted which would lead to slightly different accuracy.

 1 Ti ¼ 0

 p1 Ti =As þ p2
ηpr ðTi Þ ¼  2 ð23Þ
 i s Þ þ q2 Ti =As þ q2
ðT =A
 Ti =As in MPa

The application of the proposed design procedure is summarized in Figure 3. It starts with the
value of Ncr,a. Next Ncr,f is calculated in step 2 (during this step the values of axial stiffness of
the column, cables and cross-arms are also necessary to calculate the value of C2). In step 3

Figure 3. Application of the method.

1125
Figure 4. Comparison of different methods against experimental results.

Table 2. Comparison of results.


Pu,exp/Nu,Wadde (L/1000) Pu,exp/Nu,Wadde (L/400) Pu,exp/Nu,Wadde (L/200) Pu,exp/Nu,proposed

Average 0.99 1.23 1.57 1.02


St. Dev. 0.26 0.40 0.33 0.19
CoV 26% 32% 21% 19%
R2 0.68 0.60 0.83 0.87

the slenderness is evaluated and in step 4 the equivalent imperfection is calculated according
to the level of initial prestress in the cables. With this step concluded, it is possible to calculate
the reduction factor χ in step 5 and the maximum applied load in step 6.

3.4 Comparison
Using the method that results from the calibration explained above, it is possible to calculate
the values of the ultimate loads. Figure 4 summarizes the scatterplots of experiments and
numerical analysis and compares them with the now calibrated and with the method proposed
by Wadee et al. (2013) for an amplitude of the geometrical imperfection of L/1000.
The scatter of results is presented in Figure 4 and Table 2 show a better correlation (higher
value of r-square and lower value of CoV) for the proposed method.

4 CONCLUSIONS

In this paper, a methodology for the design of cable stayed columns was presented. It is based
on the Ayrton-Perry format and thus consistent with the current Eurocode 3 prescriptions for
flexural buckling. The method can distinguish between buckling modes and it is easily applied
using a step-by-step procedure as summarized by the flowchart (Figure 3).
The calibration of the method against experimental results was also demonstrated. It was
done only for Mode 1, because of the available experimental results. When the method was
compared to other design proposal from the literature, it was shown to provide better results
in terms of mean value and standard deviation. Nevertheless, the results presented in this
paper are based on limited number of experimental results. Therefore, the concept can be fur-
ther extended on the basis of advanced numerical simulations and the investigation may be
deepened in order to provide expressions which are applicable to a larger range of cases.

ACKNOWLEDGEMENTS

This work was partly financed by:

1126
– the Research Fund for Coal and Steel under grant agreement RFSR-CT-2012-00028
(HILONG)
– FEDER funds through the Competitively Factors Operational Programme - COMPETE
and by national funds through FCT – Foundation for Science and Technology within the
scope of the project POCI-01-0145-FEDER-007633.

REFERENCES

Araujo, R.R., Andrade, S.A.L., Vellasco, P.C.G.d.S., da Silva, J.G.S., Lima, L.R.O. 2008. Experimental and
numerical assessment of stayed steel columns, Journal of Constructional Steel Research, 64, 1020–1029.
Araujo, R.R. 2009. Comportamento Estrutural de Colunas de Aço Estaiada e Protendida, Civil Engineer-
ing Department, Pontifical Catholic University of Rio de Janeiro, Brazil.
Chan, S.-L., Shu, G.-P. Lü, Z.-T. 2002. Stability analysis and parametric study of pre-stressed stayed
columns, Engineering Structures, 24, 115–124.
Chu, K.H., Berge, S.S. 1963. Analysis and design of struts with tension ties, Journal of Structural Div-
ision, 89, 127–163.
Gkantou, M; Tran, A; Martins, J. P; Ellen, M; Koltsakis, E; Afshan, S; McCormick, F; Veljkovic, M;
Manoleas, P; Baniotopoulos, C; Remde, C; Baddoo, Nancy; Simões da Silva, Luís; Chen, A;
Theofanous, M; Herion, S; Gardner, L; Aggelopoulos, Eleftherios; Fleischer, O; Cederfeldt, L. (2017).
High Strength Long Span Structures (HILONG), Grant agreement RFSR CT 2012-00028, final report,
ISBN 978-92-79-65601-9.
Hafez, H.H., Ellis, J.S., Temple, M.C. 1979. Pre-tensioning of single cross-arm stayed columns, Journal
of the Structural Division, 105, 359–375.
Hathout, I.A.-S 1977. Stability analysis of space stayed columns by the finite element method, University of
Windsor.
Khosla, C.M. 1975. Buckling loads of stayed columns using the finite element method, University of Windsor.
Martins, J.P., Shahbazian, A., Simões da Silva, L., Rebelo, C. R., Simões, 2016. Structural behaviour of
prestressed stayed columns with single and double cross-arms using normal and high strength steel.
Archives of Civil and Mechanical Engineering, 16, 618–633.
Mauch, H.R., Felton, L.P. 1967. Optimum design of columns supported by tension ties, Journal of Struc-
tural Division, 93, 210–220.
Osofero, A.I., Wadee, M.A., Gardner, L. (2012a). Experimental study of critical and post-buckling
behaviour of prestressed stayed columns, Journal of Constructional Steel Research, 79, 226–241.
Osofero, A.I., Wadee, M.A., Gardner, L. (2012b). Numerical Studies on the Buckling Resistance of Pre-
stressed Stayed Columns, Advances in Structural Engineering, 16 487–498.
Saito, D., Wadee, M.A. 2008. Post-buckling behaviour of prestressed steel stayed columns, Engineering
Structures, 30, 1224–1239.
Saito, D., Wadee, M.A. 2009a. Buckling behaviour of prestressed steel stayed columns with imperfections
and stress limitation, Engineering Structures, 31, 1–15.
Saito, D., Wadee, M.A. 2009b. Numerical studies of interactive buckling in prestressed steel stayed
columns, Engineering Structures, 31, 432–443.
Savin, I.V. 1977. Prestressed Load -Bearing Metal Structures, MIR Publishers, Moscow.
Serra, M., Shahbazian, A., Silva, L. S., Marques, L., Rebelo, C., da Vellasco, P. C. G. S. 2015. “A full
scale experimental study of prestressed satyed columns”, Engineering Structures, 1000, 490–510.
Simulia (2014) ABAQUS FEA (version 6.14).
Smith R., Mc Caffrey G.T., Ellis J.S. 1975. Buckling of a single cross-arm Stayed Column. Journal of the
Structural Division, 101, 249–268.
Smith, E.A. 1985. Behaviour of Columns with Pretensioned Stays, Journal of Structural Engineering, 111,
961–972.
Temple, M., Prakash, M., Ellis, J. 1984 Failure Criteria for Stayed Columns, Journal of Structural Engin-
eering, 110, 2677–2689.
Van Steirteghem, J., De Wilde, W.P., Samyn, P., Verbeeck, B.P., Wattel, F. 2005. Optimum design of
stayed columns with split-up cross arm, Advances in Engineering Software, 36, 614–625.
Wadee, M.A., Gardner, L., Osofero, A.I. 2013. Design of prestressed stayed columns, Journal of Con-
structional Steel Research, 80, 287–298.
Wong, K.C., Temple, M.C. 1982. Stayed columns with initial imperfection, Journal of the Structural Div-
ision, 108, 1623–1641.

1127
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Influence of geometrical imperfection of rib stiffeners


on beam-to-column joint behaviour

R. Tartaglia, M. D’Aniello, G.M.Di Lorenzo & R. Landolfo


Department of Structures for Engineering and Architecture, University of Naples “Federico II”, Naples, Italy

ABSTRACT: The introduction of rib stiffeners in the extended end-plate joints can guaran-
tee a beneficial increase of both the strength and ductility of the connection. Indeed, the pres-
ence of the rib influences the yield line distribution in the tensile zone of the connection, the
depth of internal lever arm as well as the resistance of compression components of the joint.
However, unexpected phenomena can develop if the stiffeners are not properly aligned with
the beam web. Therefore, in this work, a parametric finite element (FE) analyses was con-
ducted to investigate the influence of the rib constructional imperfection on both the local and
global joint response. In particular, the variation of the compression center position and the
development of additional internal actions in the bolt rows was monitored for five different
rib misalignments. The results show that some geometrical imperfections could have
a beneficial effect on the joint capacity limiting the beam out-of-plane mechanism when the
plastic hinge develops at the beam extremity.

1 INTRODUCTION

Steel rib plates are commonly used to increase the strength and the stiffness of steel beam–
to-column joints in moment resisting frames (Giordano 2017, Montuori 2015, 2016, 2017a,
and b, Nastri 2018). Among the wide variety of stiffened bean-to-column connections,
extended end-plate stiffened by rib plates on both the tension and compression side, are
commonly adopted for seismic applications because they are less expensive than haunched
end-plate connections and characterized by symmetric hogging and sagging behaviour.
Notwithstanding the key role of the stiffeners, their design, verification and fabrication is
considered troublesome by European engineers and constructors due to the limited guidance
given by the current Eurocodes and the increase of constructional costs as respect to unstif-
fened connections. The nonlinear monotonic and cyclic behaviour of stiffened connections
can be predicted using experimental tests (D’Aniello et al. 2018), analytical approaches and
sophisticated finite element models, methods which can be impractical and inconvenient in
current design practice. With this regard, Kurejková and Wald (2017) recently presented
a promising method to predict the response of steel joints with different types of stiffeners (i.e.
rib plates and haunches) based on the “research finite element model” (RFEM). However, as
highlighted by the same Authors, the RFEM is still less applicable in current practice owing
to the difficulties in setting the geometrical and mechanical imperfections in advanced numer-
ical models as well as the time-consuming calculations.
Recently, a European seismic pre-qualification procedure of ESEP joints has been devel-
oped within the RFCS EQUALJOINTS project (Landolfo et al 2018) on the basis of both
experimental tests and finite element simulations. In the framework of this project, the present
paper describes and discusses the influence of the rib stiffeners focusing on their construc-
tional imperfections for different beam-to-column assemblies. Indeed, although the imperfec-
tions are generally considered as highly affecting the response of slender systems as light
weight structures (Fiorino et al. 2016, 2017, 2018, Pali et al. 2018), they can also affect the
local response of heavy steel systems. Following this consideration, in this study the yield line

1128
distribution and the relevant variation of internal forces into the rib and the bolt rows are
monitored. These results allow to characterize the evolution of the internal lever arm, the out-
of-plane bending and torsional moment developing when the connected beam experiences
large plastic rotations as well as the forces acting on lateral torsional restraints.
The paper is organized in two main parts, as follows: the investigated parameters and the
design assumptions of the joints are briefly presented in the first part; the results of the FE
parametric study are presented and critically discussed in the second part.

2 FRAMEWORK OF THE STUDY

2.1 The investigated joint assemblies


The beam-to-column assemblies constituting the examined joints have been extracted from
a set of reference buildings, designed according to EN1993:1-1 (2005) and EN1998-1 (2005).
The selected beam-to-column assemblies are the following:
• beam IPE360 – column HEB280 (hereinafter corresponding to the joints labelled as “ES1”);
• beam IPE450 – column HEB340 (hereinafter corresponding to the joints labelled as “ES2”);
• beam IPE600 – column HEB500 (hereinafter corresponding to the joints labelled as “ES3”).
All the joints have been designed to guarantee the activation of the plastic hinge at the beam
extremity, preserving the connection and the column web panel that should remain in elastic
range. The joints were designed according to the procedure described within EQUALJOINTS
project; to fulfil the capacity design requirements the Eq. (1) is satisfied:

Mwp;Rd  Mc;Rd  Mc;Ed ¼ α  ðMB;Rd þ VB;Ed  sh Þ ð1Þ

Mwp;Rd ¼ Vwp;Rd  z ð2Þ

In Equation 1, Mwp,Rd is the flexural strength corresponding to the capacity of column web-
panel (see Eq. (2), Vwp,Rd is the column web shear resistance, z is the internal level arm, Mc,Rd
is the flexural strength of the connection zone, Mc,Ed is the design bending moment at the
column face, MB,Rd is the design bending strength of the beam, VB,Ed is shear force corres-
ponding to the occurrence of the plastic hinge in the connected beam, sh is the distance
between the applied shear and the column face. α depends on the design performance level
and it is given by γsh × γov.
The geometrical features of all the investigated joints are summarized in Table 1 and Figure 1.

2.2 Investigated parameters


The construction imperfections of the stiffeners, namely misalignment of the rib with respect
to the beam and the connection (Figure 2), was investigated. These constructional faults can
easily occur in ordinary production conditions, and when recognized during the quality

Table 1. Joints geometrical dimensions.


Continuity Supplemen-
End-Plate Rib Bolts
plates tary web plate

Joint hEP bEP tEP hRib LRib d e w p1 p2 bCP tCP Side tSWP
ID mm mm mm mm mm mm mm mm mm mm mm mm - mm

ES1 760 260 25 200 235 30 50 150 75 160 222 14 2 8


ES2 870 280 25 210 250 30 50 150 75 180 234 15 2 10
ES3 1100 280 30 250 295 36 55 160 95 210 232 20 2 15

1129
Figure 1. Joint geometrical features.

Figure 2. Position of the rib stiffeners for the investigated joint configurations.

control of the execution they have to be rectified process resulting in an increased unitary
costs. Therefore, in order to understand how this type of defect can compromise both the
global and the local joint behaviour, four configurations of constructional faults were investi-
gated. As shown in Figure 2, the stiffeners (both on tension and compression side) were shifted
from the vertical symmetry axis of a distance equal to ± their thickness (that is constant and
equal to 20mm for all the investigated beam-to-column assembly).

3 FE MODELLING DESCRIPTION

The finite element models (FEMs) were developed using ABAQUS 6.14 (2014). All modelling
assumptions and their validation against experimental tests carried out within EQUALJOINTS
project are as described by the Authors in previous publications (Tartaglia et al., 2018a and
2018c). Therefore, for the sake of brevity only the main features of the models are summarized
hereinafter. The beam-to-column joints were modelled considering a sub-assemblage obtained
by extracting the beam and the column at the inflection points of the bending moment diagram
induced by shear type lateral loads on the reference MRFs (Nastri 2017, 2018). The boundary
conditions and assumed length of the members are reported in Figure 3a.

1130
Figure 3. Boundary condition and element mesh dimension.

All elements were discretized using C3D8I solid element type (i.e. 8-node linear brick, incom-
patible mode), while the mesh dimension change in function of the model parts. According to
a sensitivity analysis performed by the Authors and described in previous research (Tartaglia
et al. 2018d), the adopted mesh dimensions are 12.5, 10 and 15mm respectively for bolt, end-plate
and steel profiles (beam and column) with at least two elements through the element thickness.
The beam, the column and all the plates were considered as made out of European S355
steel with the average yield strength equal to γov × fy. The von Mises yield criterion was
adopted to model steel yielding; plastic hardening was simulated using both nonlinear kine-
matic and isotropic hardening law on the basis of the data provided by Dutta et al. (2010).
The material of the welds was modelled with an elastic perfectly plastic constitutive law, with
yield stress set equal to 460MPa.
The non-linear response of the adopted high-strength pre-loadable bolts was modelled
according to D’Aniello et al. (2017). The pre-tensioning was modelled using the “Bolt load”
option available in ABAQUS and the clamping force as recommended by EN1993:1-8 (2005).
The geometrical imperfections of beam profiles due to mill tolerances given by EN 10034
(1993) were accounted for. Contacts were modelled considering both the normal and the tan-
gential behavior. Hence, “surface-to-surface” interactions were used to model the contacts
between (i) end-plate and column flange, (ii) bolt head and end-plate and nut and column
flange, (iii) shank and the corresponding surface of the holes. Both fillet and full-penetration
welds were connected to the corresponding parts by means of “Tie” constraints.

4 RESULTS

The influence of constructional imperfections on the global and local response of the joints
was investigated by monotonic analyses with respect to a references structure called “C0”
where the rib both on the tension and compression side are perfectly aligned with the beam
axis. The influence of the imperfections on the joint behaviour is not function of the assembly
dimensions reason why in the following only the results of the intermediate joint (ES2-Full
strength) are described.

1131
Figure 4a depicts the monotonic moment-rotation curves that are almost overlapped up to
3.5% of chord rotation. For larger imposed rotation, the results are apparently surprising.
Indeed, the constructional imperfections C3 and C4 exhibit larger capping rotation than the
reference “perfect” configuration C0. The local behaviour of the connection is more affected
by the imperfections, which induce non-symmetrical distribution of internal forces and
deformations.
Indeed, except for the configuration C1, all the other cases with imperfection have the com-
pression center closer to the beam flange than the reference assembly C0 (see Figure 4b), with
a consequent decrease of the internal level arm.
The different types of behaviour mostly depend on the position of the stiffener on the com-
pression side. Indeed, only the C1 configuration has the rib in compression perfectly aligned
with the beam web, thus being effective to transfer the compression forces. In addition, the
eccentricity of the rib on the tension side provides a beneficial effect because it restrains the
beam flange of the plastic hinge.
In all the other cases, the eccentricity of the rib in the compression side impairs the effectiveness
of the transfer mechanism of compression forces, thus the compression center moves from the rib
web to the beam flange. However, some differences can be recognized between the C2, C3 and
C4 configurations. The latter cases show a similar variation of the compression with respect to
the reference joint C0, while the C2 configuration is the closest joint C0. These differences are due
to the position of the stiffener on the tension side. In the C3 and C4 configurations both the ribs
in the tension and compression side are not aligned with the beam web; contrariwise in C2 config-
uration the rib in tension is perfectly centered with the beam. The misaligned stiffeners restrain
the buckling of the beam flange more effectively than the cases with aligned ribs, thus providing
larger bending strength and rotation capacity to the plastic hinge of the beam of C3 and C4 con-
figurations, despite the smaller internal lever arm of the connection (see Figure 4).
This latter consideration is better clarified analysing the distribution of contact forces
(CPREES) between the end-plate and the column (see Figure 5). For C1, the CPREES distri-
bution shows that the rib web can transfer compression forces increasing the imposed rotation
up to 6% of rotation, where the internal distribution of contact forces changes due to the
buckling of the beam flange (see Figure 5a). In the case of C2 (where the rib on the compres-
sion is closer to a bolt alignment) the contact forces are almost equally distributed in the beam
and in the rib for small rotations (e.g. 2%), but increasing the imposed rotation the compres-
sion forces mostly concentrate in the beam (see Figure 5b) with asymmetric distribution into
the beam flange due to its plastic buckling, whose deterioration effects are magnified by the
eccentricity of the stiffener on compression side. A similar trend can be also observed for the
C3 and C4 configurations (see Figures 5c and d, respectively).
Even on the tension side, the constructional imperfections induce non-symmetric distribu-
tions of bolt forces, especially in the outer bolt rows adjacent to the rib stiffener. Figure 6
shows the evolution of bolt forces in the four upper bolts of the ES2-F assemblies. As it can

Figure 4. Results ES2-F joints under monotonic loads.

1132
Figure 5. CPREES distribution for all the investigated joint configurations.

Figure 6. Bolt internal forces.

be recognized, the larger differences occur in the outer bolt row (i.e. bolts 1 and 2), while the
bolt row close to the beam flange shows negligible differences. Indeed, the first bolt row is
generally subjected to forces smaller than those acting in the second row: The latter is directly
subjected to the larger tensile force transferred by the beam flange which additionally restrains

1133
the yield line pattern. Hence, the additional effects due to the imperfections are more evident
in the bolts 1 and 2 (compare Figure 6a,b to Figure 6c,d).
The variation of the internal forces in the bolts differs with the type of imperfection. If the
rib on tension is aligned to the beam web (e.g. C2 configuration), the bolt forces are almost
symmetrically distributed regardless of the misalignment of the stiffener in the compression
side. In the cases with misaligned rib on the tension side, the equivalent T-Stub per bolt row is
non-symmetric and the bolt forces increase in the bolt closer to the stiffener. For instance, in
the C1 configuration the rib is closer to the bolts on the right side of the connection and, com-
paring the results of Figures 6a and b, it can be observed that the force in the relevant bolt 2 is
larger owith about 30kN than the paired bolt 1. The same trend can be also observed for the
C3 and C4 configurations where the ribs are closer to bolt 1 that experiences forces 7% larger
than its paired bolt 2 in both cases.

5 CONCLUSIONS

The constructional imperfections of the rib stiffener may unexpectedly have beneficial influ-
ence on the global response curves of full strength joints. Indeed, the eccentricity of the stiff-
ener allows restraining the buckling of the beam flange and the torsional deformations of the
plastic hinge, thus increasing the capping rotation and the overall ductility. However, the bolt
forces increase due to the additional secondary effects induced by the rib eccentricity. Indeed,
the internal forces could overcome the bolt’s resistance and reduce the joint ductility; this
could be very dangerous in the case of partial strength connections where an accurate estima-
tion of the bolt’s internal forces is crucial to avoid a brittle failure and dissipate the energy in
the end-plate.

REFERENCES

D’Aniello, M., Cassiano, D., Landolfo, R. 2017. Simplified criteria for finite element modelling of Euro-
pean preloadable bolts. Steel and Composite Structures, 24(6),643–658.
D’Aniello, M., Tartaglia, M., Costanzo, S., Campanella, G., Landolfo, R., De Martino, A. 2018. Experi-
mental Tests on Extended Stiffened End-Plate Joints within Equal Joints Project. Key Engineering
Materials, 763, 406–413.
Dassault (2014), Abaqus 6.14 - Abaqus Analysis User’s Manual, Dassault Systèmes Simulia Corp.
Dutta, A., Dhar, S., Acharyya, S.K. 2010. Material characterization of SS 316 in low cycle fatigue
loading. Journal of Materials Science, 45, 1782–1789.
EN 10034, 1993, Structural Steel I and H Sections: Tolerances on Shape and Dimensions. European
Committee for Standardization; Brussels, Belgium.
EN 1993 1-8 2005, Eurocode 3: Design of Steel Structures. Part 1–8: Design of Joints. European Commit-
tee for Standardization; Brussels, Belgium.
EN 1998-1 (2005), Design of Structures for Earthquake Resistance - Part 1: General Rules, Seismic
Actions and Rules for Buildings. CEN.
EN 1993:1–1 (2005), Design of Steel Structures - Part 1–1: General rules and rules for buildings. CEN.
EQUALJOINTS – European pre-QUALified steel JOINTS: RFSR-CT-2013-00021. Research Fund for
Coal and Steel (RFCS) research programme.
Fiorino, L., Terracciano, M.T., Landolfo, R. 2016. Experimental investigation of seismic behaviour of
low dissipative CFS strap-braced stud walls. Journal of Constructional Steel Research, 127, 92–107.
Fiorino, L., Macillo, V., Landolfo, R. 2017. Shake table tests of a full-scale two-story sheathing-braced
cold-formed steel building. Engineering Structures, 151, 633–647.
Fiorino, L., Pali, T., Landolfo, R. 2018. Out-of-plane seismic design by testing of non-structural light-
weight steel drywall partition walls. Thin-Walled Structures, 130, 213–230.
Giordano, V., Chisari, C., Rizzano, G., Latour, M. 2017. Prediction of seismic response of moment
resistant steel frames using different hysteretic models for dissipative zones. Ingegneria Sismica - Inter-
national Journal of Earthquake Engineering, 34(4),42–56.
Kurejková M., Wald, F. 2017. Design of haunches in structural steel joints. Journal of civil engineering
and management, 23(6): 765–772.

1134
Landolfo et al. 2018. European pre-QUALified steel JOINTS – EQUALJOINTS: final report. Euro-
pean Commission Research Programme of the Research Fund for Coal and Steel, Technical
Group: TG S8.
Montuori, R., Nastri, E., Piluso, V. 2015. Advances in theory of plastic mechanism control: Closed form
solution for MR-Frames. Earthquake Engineering and Structural Dynamics, 44(7),1035–1054.
Montuori, R., Nastri, E., Piluso, V., Troisi, M. 2016. Influence of the cyclic behaviour of beam-to-
column connection on the seismic response of regular steel frames. Ingegneria Sismica - International
Journal of Earthquake Engineering, 33(1),91–105.
Montuori, R., Nastri, E., Piluso, V. 2017. Influence of the bracing scheme on seismic performances of
MRF-EBF dual systems. Journal of Constructional Steel Research, 132, 179–190.
Montuori, R., Nastri, E., Piluso, V., Troisi, M. 2017. Influence of connection typology on seismic
response of MR-Frames with and without ‘set-backs’. Earthquake Engineering and Structural Dynam-
ics, 46(1),5–25.
Nastri, E. 2018. Design and assessment of steel structures in seismic areas: Outcomes of the last Italian con-
ference of steel structures. Ingegneria sismica - International Journal of Earthquake Engineering, 35(2),1–4.
Pali, T., Macillo, V., Terracciano, M.T., Bucciero, B., Fiorino, L., Landolfo, R. 2018. In-plane
quasi-static cyclic tests of non-structural lightweight steel drywall partitions for seismic performance
evaluation. Earthquake Engineering & Structural Dynamics, 47(6),1566–1588.
Tartaglia, R., D’Aniello, M. De Martino, A. 2018a. Ultimate performance of external end-plate bolted
joints under column loss scenario accounting for the influence of the transverse beam. Open Construc-
tion and Building Technology Journal. 12, 132–139.
Tartaglia, R., D’Aniello, M., Zimbru, M., Landolfo, R. 2018b. Finite element simulations on the ultim-
ate response of extended stiffened end-plate joints. Steel and Composite Structures, 27(6),727–745.
Tartaglia, R., D’Aniello, M., Landolfo, R., 2018c. The influence of rib stiffeners on the response of
extended end-plate joints. Journal of Constructional Steel Research, 148, 669–690.

1135
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

The fire behaviour of extended stiffened joints designed for seismic


actions

R. Tartaglia, M. Zimbru, A. Linguiti, M. D’Aniello & R. Landolfo


Department of Structures for Engineering and Architecture, University of Naples “Federico II”, Naples, Italy

F. Wald
Department Steel and Timber Structures, Czech Technical University in Prague, Prague, Czech Republic

ABSTRACT: The structural behaviour of steel moment resisting frames (MRFs) is strongly
dependent on the beam-to-column joint behaviour. The role of the joints is crucial especially
under accidental natural and human induced actions, as in the cases of earthquake and fire
scenarios, which can occur subsequently after severe seismic events in urban areas. The study
summarized in this paper aims at investigating the fire behaviour of seismically designed
extended stiffened end-plate joint by means of finite element analyses (FEAs). The joint per-
formance was investigated considering two scenarios: (i) in the first case the assemblies were
subjected only to the fire action; (ii) in the second scenario the fire actions were applied to
seismically damaged joints. The numerical results show that the fire action changes the
restraining capacity of the joint, and local failure can also occur, especially when fire occurs
after severe seismic damage.

1 INTRODUCTION

During the second part of the 20th century, the necessity of integrating Fire Engineering in
the design of buildings heightened because of the countless losses of human lives and destruc-
tion of material goods.
In this field, important results come from the Cardington experimental campaign (Newman
et al., 2004). Steel beams subjected to the fire action develop large deflections as a consequence
of progressive loss of stiffness ultimately behaving like cables hanging from the joints. As
observed during the Cardington fire test (Newman et al., 2004), the beams can reach high tem-
peratures with large deformations if the connection can withstand the significant catenary
actions that develop. The significant axial forces developed must be resisted by the joint simul-
taneously with design levels of shear force and bending moment while the local ductility
demand is increased as well (Wald et al. 2009a).
The current European fire design regulations were put to test in experimental campaigns on
structures (Wald et al., 2006) and several studies investigated the behaviour of traditional
bolted steel connections under fire actions using both experimental and numerical approaches
(Liu, 1998, Al-Jarbri et al. 2008, Qian et al. 2008, Yu et al. 2009b, c, Dai et al 2009, Strejček
et al. 2010, Garlock & Selamet 2010, Wang et al., 2011, Wang & Wang 2013).
Studies show (Wald et al 2009b) that, if not properly designed, there is a clear possibility of
the fracture in the joint, which could lead to both the fire spreading and the building’s pro-
gressive collapse. Several of the shear and partial-strength connection typologies investigated
proved to have poor behaviour at elevated temperatures (Yu et al. 2009b, c). The traditional
end-plate connections (flush, extended, flexible and spaced) were thoroughly investigated by
Wang & Wang (2013) who proposed the use of reduced beam section or channel extended
end-plate to improve their fire performance. The Authors highlighted the limited capacity of
these configurations and pointed out that by upgrading the joint, its response under fire is

1136
significantly improved. In European framework, Wald et al. (2009b) demonstrated, based on
the Cardington tests that, with regard to bolted joints, the Eurocode design prescriptions are
conservative.
The recently developed extended stiffened end-plate (ESEP) joints were not investigated
under fire action. The joint configurations designed for seismic applications proved to have
superior strength and stiffness, when compared with gravity load designed end-plate connec-
tions, while the ductility is improved as well.
An important aspect regarding the structural robustness is related to the fire resistance of
damaged joints. Indeed, differently from light weight cold formed structures (Fiorino et al.
2016, 2017a,b, Macillo et al. 2014), the steel frames under fire may guarantee sufficient robust-
ness provided that the joints do not fail prematurely.
Nowadays, the possibility of fire occurring after an earthquake has been investigated by
research ventures like PEER in California (USA). In Europe, experimental and numerical
investigations carried out by Pucinotti et al. (2011) on welded beam to column composite
joints subjected to fire after seismically induced damage, showed the limited impact of the
plastic deformations on the fire performance of the joint.
The main objective of the current work is to examine the robustness of ESEP joints
designed according to newly proposed design strategy, under accidental fire action. An
important aspect analysed is the impact the seismic damage has on the joint capacity when the
fire occurs. The joints designed within the framework of the EQUALJOINT research project
(Landolfo et al. 2018) focused on three design criteria: full, equal and partial strength joints,
out of which the first two were considered for the current investigation. In the framework of
the research carried out in the past years (Yu et al. 2009a, Garlock & Selamet, 2010, Wang &
Wang, 2013, Shakil et al., 2018), the numerical models were often used and continually
improved providing significant help both in the interpretation of the experimental results and
in performing additional investigations with good accuracy of results.
In this paper the fire performance of extended end-plate joints was investigated considering
two scenarios. In the first case the fire action was applied on the undamaged joints while in
the second case, the joints were pre-damaged by subjecting them to a cyclic loading protocol.
The second scenario is representative of the fire occurring after a seismic event and the
selected damage levels correspond to varying levels of imposed chord rotation. The beam-to-
column assemblies were investigated assuming both the case of protected (P) and non-
protected (NP) joints.

2 FRAMEWORK

The joint assemblies have been extracted from reference buildings conforming with the current
state-of-practice in Europe which are detailed with moment resisting frames (MRF) in one direc-
tion and concentric braced frames (CBF) transversally. The structure plan is square (3 x 3 bays)
with uniform span lengths in both principal directions and the storey height was considered
4,5 m for the ground storey and 3,5m for all upper storeys. The structures were designed assuming
medium and high seismicity level (PGA = 0.25g and 0.35g, respectively) and soil Type
C according to the definition of EN 1998-1 (2005). The vertical loads have been selected according
to the type of use (residential/office) and have values of 5 and 3 kN/m2 respectively for permanent
and live loads. Steel S355 has been used for all frame elements. Depending on the geometry of the
connected members (i.e. beams and columns), the analysed set of beam-to-column assemblies
includes three types of bolted extended stiffened end-plate joints: EXS1 (with a IPE 360 beam
and a HE 280 B column), EXS2 with a IPE 450 beam and HE 340 B column) and EXS3 with
a IPE 600 beam and HE 500 B column). The geometrical configuration of the joints and the com-
ponents are detailed in Figure 1 and Table 1.
All the assemblies have been designed as both full and equal strength joints. As explained in
more detail by Tartaglia et al. 2018a, the full-strength joints allow the formation of the plastic
hinge at the beam extremity, leaving the joint in elastic range. On the other hand, the equal

1137
Figure 1. Plan layout of structure and joint characteristics.

strength joints are designed to dissipate the seismic energy in both the beam and the connec-
tion. The fire response of these two joint typologies is comparatively analysed in this work.
Ideally, in case of a seismic event, the moment resisting frame (MRF) is expected to develop
plastic hinges at the beam extremities and achieve a global plastic mechanism. The damage
accumulated in the plastic hinge can compromise the joint behaviour if the fire is onset after
a strong earthquake, and the fire resistance is affected in two ways: (i) reduced capacity of the
joint and (ii) loss of effectiveness of fire protection.
To have a comprehensive understanding of the joint fire behaviour, two scenarios have
been considered for the fire action onset moment and both protected (P) and non-protected
(NP) joints were analysed for the second scenario. The investigation key aspects are herein-
after described.

3 FINITE ELEMENT SIMULATIONS

Finite element analyses (FEAs) were carried out using the finite element software package
ABAQUS 6.14 (Dassault Systemés, 2015); in the following paragraph the main model features
were described, while the modelling procedure and the hypothesis made are presented in
a previous study of the joint seismic behaviour (Tartaglia et al. 2017a, and 2018c). On the
other hand, all the modelling assumptions regarding the fire action and the variation of the
material properties in function of the temperature are in line with the EN1993-1-2 (2005)
assumptions and the procedure described by Yu et al. (2009b) and Qiang et al. 2014a and b.
The geometric properties of the investigated EXS joints are summarized in Table 1. The
beams and the columns have the same length of 3.4m and were extracted from the MRF using
the sub-structuring methodology. The beam local imperfections, due to the mill tolerance
allowed by EN 10034 (1993) were accounted for by an initial buckling analysis, hence the
most severe out-of-square buckling modes were selected as proposed by Tartaglia et al
(2018b).
All assemblies are made by the European S355 steel with the exception of Grade 10.9 for
the pre-loadable bolts. The materials’ elastic and plastic properties, at ambient temperature
(20°C), were obtained from a set of coupon tests performed within the EQUALJOINTs
research project. In the FE model, the Von Misses criteria and the combined (i.e. isotropic
and kinematic) plastic hardening was introduced (in line with Dutta et al. 2010). The bolts’
behavior was modelled by multilinear force-displacement curve described by D’Aniello et al.
(2017). To model the shank necking and the fracture in the threaded area the ductile damage
was introduced in the model as proposed by Pavlovic et al (2015). The material of the welds
was modelled by an elastic perfectly plastic constitutive law, with the yield stress equal to 460
MPa, which corresponds to an electrode grade A46 (as given by EN ISO 2560, 2009).
For increasing temperature, the material properties were evaluated according to EN1993-1-2
(2005) and are reported in Figure 3a. Hence, the reduction coefficient k was used for both the
structural steel and bolts materials, for the material strength and Young modulus reduction. An
example of the stress strain relationship varying the temperature was reported in Figure 3b.

1138
Table 1. Features of the designed joints.
Supplementary
End-Plate Rib Bolts Continuity plates
web plate
hEP bEP tEP bRib aRib n° d e w p1 p2 bCP tCP Side tSWP
Performance
Joint ID level mm mm mm mm mm - mm mm mm mm mm mm mm - mm
EXS1-F Full St. 760 260 25 200 235 12 30 50 150 75 160 222 14 2 8
EXS1-E Equal Str. 600 280 18 120 140 8 27 50 160 160 180 222 14 1 8
EXS2-F Full Str. 870 280 25 210 250 12 30 50 150 75 180 234 15 2 10
EXS2-E Equal Str. 770 300 20 160 190 8 30 55 160 200 260 234 15 1 8
EXS3-F Full Str. 1100 280 30 250 295 12 36 55 160 95 210 232 20 2 15
EXS3-E Equal Str. 1100 300 22 250 295 8 36 55 160 95 210 232 20 1 15
The interaction of the joint assembly parts in contact was modelled considering “Hard Con-
tact” for the normal behaviour and “Penalty”, with a friction coefficient equal to 0.3, for the
tangential behaviour. “Tie” constraints were assumed in place of the parts welded together.
The boundary conditions (BC) are modelled to mimic as closely as possible the inter-
action with the structure the assembly is extracted from. At the beam end, the conditions
are dependent on the step and the actions applied; hence for the step where the fire
action and gravitational loads are applied. a double pendulum constraint is placed to
simulate the beam continuity, while only a vertical restraint is considered for the applica-
tion of the cyclic loading. The torsional restraints out of the length of plastic hinge are
placed to simulate the restraining conditions imposed by the slab. The spacing of lateral
torsional restraints was taken according to the lateral-torsional stable length segment pro-
posed by EN 1993-1, clause 6.3.5.3. The columns are pinned at the bottom (all DOF
blocked except in plane rotation) and have a roller at the top (free in plane rotation and
displacement along the column axis).
In order to assess the joint behaviour under varying levels of internal forces, four alterna-
tives were investigated. In the first case, the external loads were evaluated compliant to the
structural scheme and the fire loading combination (FLC) and for the other three cases, the
vertical loads were assumed equal to 25, 50 and 75% of the beam plastic shear resistance (iden-
tified hereinafter as 0.25Vpl,B ed, 0.5Vpl,B and 0.75Vpl,B respectively).
The clamping force simulating the tightening of bolts was evaluated depending on the bolt
diameter according to EN 1993-1-8 (2005) and applied in the middle face of the bolt shanks
using the “Bolt Load” command. The cyclic loading protocol given by AISC 341 (2016) for
the qualification of beam-to-column connections was applied. The uniform distributed load
q [kN/m2], evaluated as specified in the previous chapter, was applied on the surface of the
upper beam flange.
Finally, the application of the fire load consisted in varying the temperature of the struc-
ture’s elements not protected against fire. The temperature distribution in time for beams and
columns are those evaluated following the prescription of Chapter 4.2.5.1 of EN 1993-1-2
(2005), which is based on the nominal fire curve ISO834. According to Ding and Wang
(2009), for the investigated joint configuration, the temperature can be considered uniform in
all the elements of the connection and the magnitude is a factored value of the bottom beam
flange temperature (the reduction was assumed as equal to 80%). Three sides were considered
exposed to fire for the non-protected beams, the slab protecting the top side, and for the non-
protected columns, all sides were considered exposed to fire. By way of example, the tempera-
ture-time fire curves are shown in Figure 5b and c respectively for an IPE360 and an HE280B
member.
The models were discretized using C3D8I solid element type (i.e. 8-node linear brick, incom-
patible mode). Based on the mesh sensitivity study performed by Tartaglia et al. (2018b) the
maximum elements size used in the end-plate, bolts and beam elements is respectively equal to
12.5, 10 and 15mm.

4 NUMERICAL RESULTS

In this chapter the results of the two investigated scenarios were reported in terms of tempera-
ture-rotation curves, equivalent plastic strain distribution (PEEQ) and the time–temperature
curves to define the joint fire resistance.

4.1 The behaviour of undamaged seismic joints


The first scenario concerns the study of the beam–to-column joints subjected directly to the
fire action. Figures 2 a and b show for EXS2-F and E the results in terms of temperature-
rotation curves. Independent from the design criteria (full or equal) and from the assembly
size, the joint performance is strongly influenced by the shear demand, showing a large

1140
Figure 2. Temperature rotation curve and PEEQ for EXS2 full (a) and equal strength joint (b).

decrease of the maximum temperature achievable for increasing values of shear actions.
Indeed, the increase of vertical loads in the EXS1-F joint for instance, leads to a decrease of
the joint maximum temperature equal to 24% (from 774 to 591°C).
Once the maximum temperature is reached, the joints behave as a hinge, showing
a constant increase of rotation (the horizontal plateau). As it can be noted in Figure 2 for
EXS2-F at 600°C, a large amount of shear plastic deformation concentrates at the beam
extremity on the joint side and at the beam tip, where the symmetry constraint prevents the
free rotation, in line with the sub-structuring hypothesis. The connection parts (bolts, end
plate, rib stiffeners) do not develop plastic deformations nor in the case of full or equal-
strength joints, the damage remaining concentrated in the beam.
The analyses of equal and full-strength joints show no appreciable differences (the tempera-
ture-rotation curve, the PEEQ). Indeed, the largest differences in terms of maximum resisting
temperature is of about 3%. This result is explained by the identical shear resistance of the two
joint configurations, which despite having different connections (i.e. number and dimensions
of bolts) have a capacity governed by the connected beam.
An important and well-known difference between the full and equal strength joint is the
different connection masses associated with a larger steel consumption for the former. As
reported in Table 1, the full-strength joints are heavier compared to the equal strength, espe-
cially for the EXS1 and EXS2 cases where also the number of bolts changes. The larger mas-
siveness of full-strength joints should influence the temperature propagation in the
connection, and therefore, the joint behaviour. Figure 3 summarizes the shear and tensile
forces in the first bolt line for EXS2 F and E.
The shear force in the bolts linearly increases during the loading step, while in the second
step, when the external load is maintained constant, the shear action increases slowly, reaching
a value 20% larger compared to the one from the previous step. Therefore, the fire action
implies a larger shear demand on the bolts. However, in all the examined cases, the maximum
increase for the full-strength joint is equal to 5% and it never leads to the bolts’ shear failure.
The resilience of the joint bolts is a consequence of the local resistance checks introduced in
the seismic design requirements which are aimed to avoid brittle failures in the joint.
The normal force distribution in the bolts for the full strength and equal strength joints ana-
lysed in the four cases of external force applied show the same initial equal clamping forces
and the degradation in the force when the thermal action is applied in the second step. The
bolts practically lose their clamping due to relaxation and start working as non-preloaded
bolts, the tensile forces in the bolts decreasing and reaching a limit value function of the verti-
cal loads applied. Owing to the lower temperatures in the connection area, the material

1141
Figure 3. Shear and tensile forces in the first bolt row.

preserves sufficient resistance to carry the demand from the external forces applied and as
already shown, no plastic deformations were observed. The results presented for the EXS2
joints are representative of what was observed in the EXS1 and EXS3 assemblies hence, the
latter will not be presented.

4.2 Influence of the cyclic damage on the joint fire behaviour


The second investigated scenario concerned the study of the joint fire behaviour subsequent to
the application of a cyclic loading, simulating thus fire after a damaging earthquake. From
the point of view of joint fire protection, two cases were investigated: (i) non-protected joints
(NP), for which the fire load was applied on all joint components and, (ii) protected joints (P)
for which all the joint components are protected against the fire, with the exception of the
zones where the plastic deformations occur during the seismic events. Figure 4 shows com-
paratively the results of the protected (P) and non-protected (NP) configurations of both full
and equal-strength joints. Similar to the non-protected joints previously presented, assuming
fire protection is used, the joint maximum capacity is not influenced by the cyclic plastic

Figure 4. Temperature-rotation curves for NP and P joints.

1142
damage imposed. Independently from the seismic design criteria, the influence of the fire pro-
tection is strongly depended from the joint dimensions. Hence the effectiveness of the protec-
tion is more evident for the EXS1 joint where the ultimate temperature increases from 800°C
for the NP model to 1000°C for the P model, respectively. Increasing the assembly dimensions,
the differences between the protected and non-protected joint is reduced becoming negligible
for EXS3.
This difference is mainly due to the expected type of failure. Indeed, as shown for the
undamaged models, all the joints show a shear failure concentrated in the beam at the rib
extremity. Therefore, considering that all joints are subjected to the same external action, the
stress ratio of the smaller beam is higher, hence, the degradation of the steel properties during
the fire leads to a poorer joint performance.

5 CONCLUSIONS

Based on the results hereby presented the following conclusions can be drawn:
• The joint resistance degradation and the maximum resisting temperature are function of
the shear action on the beam. Indeed, an increase of the vertical loads leads to a strong
degradation of the maximum resisting temperature. This resistance degradation is inde-
pendent from the joint dimensions and from the seismic design performance investigated
(full and equal strength).
• Differently from what was observed in literature, all the investigated joints show a shear
failure mode mainly concentrated in the beam, leaving the bolts almost elastic. The investi-
gated joints were designed to resist the seismic action, and particular attention was given to
the local hierarchy between the end-plate and bolts. This local capacity design leads to
slightly oversized bolts, perfectly able to resist to the catenary action actin developing
under the fire action.
• The joint pre-damage does not influence the joint ultimate fire capacity; however, plastic
cumulated damage localized in the connection or at the beam extremity (in case of equal
and full-strength joints, respectively) imply a reduction of the structure stiffness, with
a consequent increase of beam deflection. This phenomenon is more evident when a large
shear is applied on the frame.
• The effectiveness of the fire protection is function of the joint stress ratio. In the investi-
gated cases, keeping constant the vertical loads, the fire protections give beneficial effects
only for the smaller assemblies, while the heavier joints did not show any significant differ-
ences between the protected and non-protected configuration.

REFERENCES

Al-Jabri, K.S., Davison, J.B. Burgess, I.W. 2008. Performance of beam-to-column joints in fire-A review.
Fire Safety Journal 43: 50–62.
Dai, X.H. Wang, Y.C. Bailey, C.G. 2009. Effects of partial fire protection on temperature developments
in steel joints protected by intumescent coating. Fire Safety Journal 44: 376–386.
D’Aniello, M., Cassiano, D., Landolfo, R. 2017. Simplified criteria for finite element modelling of Euro-
pean preloadable bolts. Steel and Composite Structures 24(6): 643–658.
Dassault Systèmes - Simulia Inc. 2015. Abaqus analysis 6.14 user’s manual.
Ding, J. Wang, Y.C. 2009. Temperatures in unprotected joints between steel beams and concrete-filled
tubular columns in fire. Fire Safety Journal 44: 16–32.
Dutta, A. Dhar, S. Acharyya, S.K. 2010. Material characterization of SS 316 in low-cycle fatigue
loading. Journal of Material Science 4: 1782–1789.
Fiorino, L., Terracciano M.T., Landolfo, R. 2016. Experimental investigation of seismic behaviour of
low dissipative CFS strap-braced stud walls. Journal of Constructional Steel Research 127: 92–107.
Fiorino, L. Macillo, V., Landolfo, R. 2017a. Shake table tests of a full-scale two-story sheathing-braced
cold-formed steel building. Engineering Structures 151: 633–647.

1143
Fiorino, L., Shakeel S., Macillo, V., Landolfo, R. 2017b. Behaviour factor (q) evaluation the CFS braced
structures according to FEMA P695. Journal of Constructional Steel Research 138: 324–339.
Garlock, M. E. and Selamet, S. 2010. Modeling and Behavior of Steel Plate Connections Subject to Vari-
ous Fire Scenarios. Journal of Structural Engineering 136(7): 897-906.
Landolfo, R. et al. 2018. European pre-QUALified steel JOINTS – EQUALJOINTS: final report. European
Commission Research Programme of the Research Fund for Coal and Steel, Technical Group: TG S8.
Liu, T.C.H. 1998. Effect of connection flexibility on fire resistance of steel beams. Journal of Construc-
tional Steel Research 45: 99–118.
Macillo, V., Iuorio, O., Terracciano, M.T., Fiorino, L., Landolfo, R. 2014. Seismic response of Cfs
strap-braced stud walls: Theoretical study. Thin-Walled Structures 85: 301–312.
Newman G.M., Robinson J.T., Bailey C.G. 2004. Fire safety design: a new approach to multistorey
steel-framed buildings. The Steel Construction Institute, UK.
Pavlovic, M., Heistermann, C., Veljkovic, M., Pak, D., Feldmann, M., Rebelo, C., Da Silva, L.S. 2015.
Connections in towers for wind converters. Part I: Evaluation of down-scaled experiments. Journal of
Constructional Steel Research 115: 445–457.
Pucinotti, R., Bursi, O.S., Demonceau, J.F. 2011. Post-earthquake fire and seismic performance of
welded steel–concrete composite beam-to-column joints. Journal of Constructional Steel Research 67:
1358–1375.
Qian, Z.H. Tan, K.H. Burgess, I.W. 2008. Behavior of Steel Beam-to-Column Joints at Elevated Tem-
perature: Experimental Investigation. Journal of Structural Engineering 134: 713–726.
Qiang, X. Bijlaard, F.S.K. Kolstein, H. Jiang, X. 2014a. Behaviour of beam-to-column high strength
steel endplate connections under fire conditions - Part 1: Experimental study. Engineering Structures
64: 23–38.
Qiang, X. Bijlaard, F.S.K. Kolstein, H. Jiang, X. 2014b. Behaviour of beam-to-column high strength steel
endplate connections under fire conditions - Part 2: Numerical study. Engineering Structures 64: 39–51.
Shakil S., Lu, W., Puttonen, J. 2018. Response of high-strength steel beam and single-storey frame in
fire: Numerical simulation. Journal of Constructional Steel Research 148: 551–561.
Strejček, M. Wald, F. Sokol, Z. 2010. Column web panel at elevated temperature. Fire Technology 46:
37–47.
Tartaglia, R., D’Aniello, M. 2017a. Nonlinear performance of extended stiffened end plate bolted
beam-to-column joints subjected to column removal. The Open Civil Engineering Journal 11: 369–383.
Tartaglia, R., D’Aniello, M., Landolfo R. 2018a. The influence of rib stiffeners on the response of
extended end-plate joints. Journal of Constructional Steel Research, 148: 669–690.
Tartaglia, R., D’Aniello, M., Zimbru, M., Landolfo, R. 2018b. Finite element simulations on the ultim-
ate response of extended stiffened end-plate joints. Steel and Composite Structures 27(6): 727–745.
Tartaglia, R., D’Aniello, M., Rassati, G.A., Swanson, J., Landolfo, R. 2018c. Influence of composite
slab on the nonlinear response of extended end-plate beam-to-column joints. Key Engineering Mater-
ials, 763, 818-825.
Wald, F., Simoes da SIlva, L., Moore, D.B., Lennon T., Chladna, M., Santiago A., Benes, M., Borges, L.,
2006. Experimental behavior of a steel structure under natural fire. Fire Safety Journal 41: 509–522.
Wald, F. Sokol, Z. Moore, D. 2009a. Horizontal forces in steel structures tested in fire. Journal of Con-
structional Steel Research 65: 1896–1903.
Wald, F., Chlouba, J., Uhlir, A., Kallerova, P., Stujberova, M. 2009b. Temperatures during fire tests on
structure and its prediction according to Eurocodes. Fire Safety Journal 44: 135–146.
Wang, Y.C. Dai, X.H. Bailey, C.G. 2011. An experimental study of relative structural fire behaviour and
robustness of different types of steel joint in restrained steel frames. Journal of Constructional Steel
Research 67: 1149–1163.
Wang, M., Wang, P. 2013. Strategies to increase the robustness of endplate beam-column connections in
fire. Journal of Constructional Steel Research 80: 109–120.
Yu, H. Burgess, I.W. Davison, J.B. Plank, R.J. 2009a. Tying capacity of web cleat connections in fire,
Part 1: Test and finite element simulation. Engineering Structures 31: 651–663.
Yu, H. Burgess, I.W. Davison, J.B. Plank, R.J. 2009c. Experimental investigation of the behaviour of fin
plate connections in fire. Journal of Constructional Steel Research 65: 723–736.

1144
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Buckling length assessment with finite element approach

T. Tiainen, K. Mela & M. Heinisuo


Tampere University, Finland

ABSTRACT: In the design of steel frames, the consideration of stability and buckling is an
important issue. It can be done in multiple ways. If the concept of buckling length is used,
widely used procedure is to calculate the eigenmodes and corresponding eigenvalues for the
frame and by using them define buckling length of the members with the well-known Euler’s
equation. However, it maybe difficult to tell, which eigenmode should be used for the defin-
ition of the buckling length of a specific member. Conservatively, the lowest positive eigen-
value can be used for all members. In this contribution, two methods to define the buckling
length of a specific member are considered. The first one uses geometric stiffness matrix
locally and the other one uses strain energy measures to identify members taking part in
a buckling mode. Compared to simplified approaches presented in literature the approaches
based on the finite element discretization have certain advantages. First, the method is applic-
able to any kind of distributed loading. Secondly, also tapered members can be handled with
the technique. Moreover, the out-of-plane buckling behavior and with suitable element the
lateral buckling loads can be also be assessed. The applicability and features of the methods
are shown in a numerical 3D example. Both methods can be relatively easily implemented into
automated frame design procedure. This is essential when optimization of frames is
considered.

1 INTRODUCTION

In design of skeletal steel structures, stability is in important role. According to European


design standards EN 1993-1-1 (2006), either effective length approach or approach based on
geometrically non-linear analysis can be used. As the analysis software and sufficient compu-
tational power have become available, the latter option has become more and more tempting.
However, if automated design procedure such as optimization is used, usually the analysis will
have to be performed dozens, hundreds or even thousands of times and to keep the needed
computational effort at an acceptable level. Thus effective length approach may be more
efficient.
In certain structures, such as tubular trusses, buckling length factors are given by EN 1993-
1-8 (2006). For frames, multiple methods have been proposed for finding the effective length
of a single member. Widely used simplified approach has been presented by Dumonteil
(1992). In his contribution, the transcendental stability equation is solved approximately with
simplified formulas. Multiple extensions for this work have been carried out by several other
authors. For example, semi-rigid joints have been considered by Maigerou et al. (2006).
Webber et al. (2015) have proposed an extension to cover the effect of axial force in columns
adjoining the considered member as well as the effect of axial force in other columns in the
same floor. In the examples, it is demonstrated that this approach gives very accurate values
in comparison to results given by a finite element software.
Even if the presented simplified methods can be considered accurate enough to be applic-
able with design codes, they do not necessarily fit well in integrated design systems. For
example, in approach proposed by Webber et al., the user needs to identify other columns in
the floor which is not always straightforward task in a complicated structure.

1145
In this contribution, the approach based on finite element discretization (Tiainen & Heini-
suo, 2018) is adopted. The reference describes application and performance of two different
methods in planar structures. In this contribution, the two methods are extended to cover 3D
structures. The performance of the proposed extension is illustrated in a simple 3D frame
structure calculation example.

2 FINITE ELEMENT BASED METHODS

The basis of both methods are presented in the reference Tiainen & Heinisuo (2018). Both
methods rely on the well-known finite element approach where basis is the eigenvalue problem
 
K þ λKg q ¼ 0 ð1Þ

where K is the stiffness matrix, Kg is the geometric stiffness matrix, λ is the eigenvalue
and q is the eigenvector representing the buckling mode. When using the finite element
model to assess the buckling length of a certain member, the designer needs to manually
scroll the eigenmodes until the mode containing buckling of the particular member.
When the correct mode is found, the buckling length around local axis y can be calcu-
lated as
sffiffiffiffiffiffiffiffiffiffiffiffi
EIy
Lcr;y ¼π ð2Þ
λj jNi j

where E is the Young’s modulus for the material, Iy is the second moment of area of the cross-
section, and Ni is the axial force in the member i. Similarly, if the correct eigenpair is found
for buckling around the local z axis the buckling length can be written as
sffiffiffiffiffiffiffiffiffiffiffiffi
EIz
Lcr;z ¼π ð3Þ
λj jNi j

2.1 Local geometric stiffness approach


In the local geometric stiffness approach, the geometric stiffness matrix is only applied to elem-
ents belonging to the member whose buckling length is being assessed. This can be expressed as

K þ λKig q ¼ 0 ð4Þ

where
X
Kig ¼ Keg ð5Þ

where the sum is taken over the element belonging to member i.


The extension to 3D can be done by applying the geometric stiffness matrix partly only
including those elements of the matrix being part of buckling behavior in the plane in
question. This means that for each member, two limited eigenvalue problems need to be
solved.

1146
2.2 Strain energy based method
The strain energy based method supposes that in the relevant buckling mode, the compressed
member will have substantial share of total strain energy. By identifying in-plane and out-of-plane
elements in the geometric stiffness matrix, for each member in-plane and out-of-plane buckling
loads can be obtained in similar manner.
In the well-known linear finite element framework, the element strain energy is calcu-
lated as
1
E e ¼ qT ke q ð6Þ
2

where ke is the element stiffness matrix and q is the vector of displacements. The member
strain energy can be calculated as

1 X e
E m ¼ qT kq ð7Þ
2

where sum is taken such that elements belonging to the member in question are taken into
account. Respectively, for the whole structure, the total strain energy can be calculated when
the global stiffness matrix K is used:

1
E ¼ qT Kq ð8Þ
2

Figure 1. The example 3D frame topology, dimensions, loads and column numbering.

Table 1. Buckling length factors both proposed approaches and manual assessment.
Local kg Energy based Manual

Columns y z y z y z

1 0.66 0.61 0.64 1.39 0.64 1.39


2 0.66 0.61 0.64 1.39 0.64 1.39
3 0.65 0.70 0.97 0.98 0.97 0.98
4 0.65 0.70 0.97 0.98 0.97 0.98
5 0.66 0.61 0.64 1.39 0.64 1.39
6 0.66 0.61 0.64 1.39 0.64 1.39

1147
Figure 2. First eight buckling modes for the example frame.

1148
The ratio for each member in a deformed shape can be thus calculated as

Em
Rm ¼ ð9Þ
E

In 3D case, the member stiffness matrix has to be replaced by a partial stiffness matrix including
only elements needed for flexural behavior in the plane in question.
The criterion for a member to take part in a buckling mode is proposedly proportional to
it’s share of the total energy. The original proposal by Tiainen & Heinisuo (2018) was

1
Rm  ð10Þ
n

where n is the number of members in the structure.

3 NUMERICAL EXAMPLE

Consider a 3D frame in Figure 1. All the members are of cold-formed rectangular hollow section
with outer dimension 100 mm and wall thickness 5 mm. All joints are supposed ideally rigid but
the ends of diagonal braces are supposed hinged. The task is to find buckling length for each
column in both planes. The material Young’s modulus is 210 GPa. The well-known
Euler  Bernoulli beam theory assumptions and corresponding finite elements are used. Each
member is modeled with five elements. The local axes of the columns are defined such that buck-
ling in the plane where the frame is braced is considered the local z axis and the other main axis is
the z axis.
Both of the methods were applied to the problem. The resulting buckling length factors can be
seen in Table 1. Moreover, in Table 1 there are the results for each column obtained by manual
evaluation inserting λ values to Eqs. 2 and 3. For the manual evaluation, consider the 8 lowest
modes shown in Figure 2. Clearly, in the first mode all the columns buckle simultaneously in
a sway mode around their z axis. In the second mode, the middle columns buckle around their y
axis. The corner columns’ lowest buckling mode around their y axis is the 8th mode. As both the
structure and loading are symmetric, these modes cover all the relevant modes for the columns.
Both methods seem to capture the non-sway mode (mode 8 where corner columns buckle
around their local y axis). However, the local geometric stiffness matrix based method seems
to fail in modes exhibiting sway of the whole frame (modes 1 and 3 in Figure 2). The strain
energy-based method seems to perform well in this example finding the correct eigenmodes
for every column. The criterion of a member to take part in a buckling mode is the same pro-
posed by Tiainen & Heinisuo (2018). However, as found in Tiainen & Heinisuo (2018), the
criterion is not general. The strain energy shares are seen in Tables 2 and 3.

Table 2. Shares of strain energy in y axis buckling for eight first modes.
Mode

Column 1 2 3 4 5 6 7 8

1 0 0 0 0 0 0 0 21
2 0 0 0 0 0 0 0 21
3 0 23 0 42 0 42 0 1
4 0 23 0 42 0 42 0 1
5 0 0 0 0 0 0 0 21
6 0 0 0 0 0 0 0 21

1149
Table 3. Shares of strain energy in y axis buckling for eight first modes.
Mode

Column 1 2 3 4 5 6 7 8

1 10 0 8 0 1 0 1 0
2 10 0 8 0 1 0 1 0
3 17 0 13 0 39 0 42 0
4 18 0 13 0 41 0 39 0
5 10 0 8 0 1 0 1 0
6 10 0 8 0 1 0 1 0

4 CONCLUSIONS

In this paper, the extensions from 2D to 3D for two finite element based buckling length
assessment methods are presented. Based on the example 3D frame it can be said that the
results obtained in 2D apply also in 3D. The approach based on local geometric stiffness
matrix may fail if the relevant buckling mode is a sway mode. In non-sway modes, the
approach is rather accurate. The energy-based method seems to correctly predict the buckling
lengths in the example. However, the question of a general criterion for the energy share pre-
sent in 2D remains unanswered.
Both of the methods contain several benefits in comparison to many approaches found in
the literature. Both can be used with non-prismatic members or distributed axial loads. Also,
different beam behaviour assumptions may be used such as the Timoshenko beam theory.
Moreover, in future research on the topic, the torsional modes in axially loaded members as
well as critical bending moment resulting in lateral torsional buckling could be considered.

REFERENCES

Dumonteil, P. 1992. Simple equations fo effective length factors. Eng J AISC: 29(3):111–115.
EN 10219-2, Cold formed welded structural hollow sections of non-alloy and fine grain steels. Part 2:
Tolerances, dimensions and sectional properties. CEN, 2006.
EN-1993-1-1. Eurocode 3: Design of steel structures. Part 1-1: General rules and rules for buildings.
CEN, 2006.
EN-1993-1-8. Eurocode 3: Design of steel structures. Part 1-8: Design of joints. CEN, 2006.
Mageirou, G.E. and Gantes, C.J. 2006. Buckling strength of multi-story sway, non-sway and
partially-sway frames with semi-rigid connections. Journal of Constructional Steel Research: 62
(9):893–905.
Tiainen, T. and Heinisuo, M. 2018. Buckling length of a frame member. Journal of Structural Mechanics:
51(2):49–61.
Webber, A., Orr, J., Shepherd, P., and Crothers, K. 2015. The effective length of columns in multi-storey
frames. Engineering Structures: 102:132–143.

1150
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Experimental and numerical analysis of the local and interactive


buckling behaviour of hollow sections

A. Toffolon, A. Müller & A. Taras


Institute of Structural Engineering, Bundeswehr University Munich, Germany

I. Niko
Department of Steel and Timber Structures, Slovak University of Technology, Slovakia

ABSTRACT: Inadequate knowledge regarding specific local and interactive behavior of slender
high-strength steel (HSS) hollow sections presents an obstacle in implementing these sections in the
construction practice. The current approach in the standard is simplified and offers overly conserva-
tive results. Dealing with non-standard cross-sections only expands on these difficulties. This paper
contributes to the on-going RFCS project “HOLLOSSTAB”, which has been dealing with the
aforementioned issues. The project design proposal is based on the “Overall interaction concept”
(OIC), and a new set of design rules for hollow sections is currently being developed. In this concept,
linear buckling analysis (LBA) is used to obtain the slenderness of the member, and Geometrically
and Materially Non-linear Imperfection Analysis (GMNIA) is used to determine an “overall” buck-
ling reduction factor. Extensive experimental tests are used to validate the method, by correlating
the experimental results with numerical test results (GMNIA-real) and statistically analyzing them.
The focus of the tests in this paper is the use of high-strength steel, ranging from S500 to S890, used
for cylindrical hollow-sections, rectangular hollow-sections and hexagonal hollow-sections. The
state-of-the-art measuring tools available offer the possibility of precise reverse-engineering process,
creating a numerical model of experimental test. It is possible to reach an accurate prediction of
ultimate load capacity of the specimen. The buckling shape can also be checked, using digital image
correlation (DIC), comparing numerical models with real shape of tested specimen. The paper aims
to validate the numerical models, as well as validity of assumptions on imperfection amplitude.

1 INTRODUCTION

Manufacturers of racking systems for large logistics centers face the challenge of producing very
light-weight, economical structures that nevertheless meet very high standards with respect to stiff-
ness and strength. In order to meet this challenge, in the market an increasingly number of high-
strength steel profiles and stiffened cold formed cross-sections are found. However, current design
standards, such as the Eurocode 3, do not sufficiently address the resistance of such sections – as
well as more common square and rectangular hollow sections - against local (L) and lip-stiffener
= “distortional” (D) buckling modes, especially for general combinations of loading (compression
and mono- or biaxial bending). This paper describes an experimental and numerical study of the
local buckling behavior of a wide variety of hollow sections, with the aim of calibrating
a numerical campaign within the framework of the on-going RFCS project “HOLLOSSTAB”.
Cold-formed sections with and without lip stiffeners, circular hollow sections, hexagonal
hollow sections both mild-steel and high strength steel are tested in pure compression, bending
and bending + compression, with various slenderness ranges for local buckling. Finally, the
results of the parametric study will be used for the development of appropriate design rules for
these sections, with a continuous representation of strength throughout slenderness ranges, using
the “Overall Interaction Concept” (Boissonnade et al. 2017) as a conceptual basis for the develop-
ment of the design rules. This concept – which is similar to the Direct Strength Method (DSM)

1151
(Schafer 2008) used in North America for the design of cold-formed steel open cross-sections -
makes use of the results of (numerical) linear buckling analyses (LBA) for the whole member to
determine the slenderness and consequently an “overall” buckling reduction factor.

2 SCOPE AND METHODOLOGY

As stated above, the strength and stability of hollow sections of various shapes and steel grades
are being studied within the scope of the European research project HOLLOSSTAB by means of
an extensive experimental and numerical test campaign and parametric study. Among the various
section types considered, unstiffened and stiffened, double- and mono-symmetrical cold-formed
sections are being tested at the Chair of Steel Structures of Bundeswehr University Munich.
A review of international tests on high-strength cold-formed hollow sections was recently
given in (Jia-Lin Ma et al. 2015), (Jia-Lin Ma et al. 2017), (Gardner et al. 2017), (Wang et al.
2010), (Kim DK et al. 2014), (Schillo & Feldmann 2015) and (Toffolon & Taras 2017).
This section of the paper describes the scope and methodology of this series of tests and
numerical analyses.

2.1 Types of studied cross-section and materials


Several distinct types of cold-formed and hot rolled cross-section are the subject of study
during the HOLLOSSTAB project and are illustrated in Figure 1:
i. rectangular (SHS, RHS) hollow sections produced in accordance with the European fabri-
cation standard EN 10219, which applies to cold-formed sections, with steel grades from
(normal strength) S355 to (high strength) S700.
ii. cold-formed and welded sections of bespoke, stiffened shape, produced primarily for stor-
age racking applications by voestalpine Finaltechnik Krems GmbH in Krems, Austria.
The two shapes studied within the scope of this paper are designated VHPS (“S” for “stiff-
ened”) and VHPT (“T” for T-shaped) and are shown in Figure 1a and b.

Figure 1. Stiffened section types studied at Bundeswehr University Munich within the scope of the
RFCS-HOLLOSSTAB project: a) “VHPS”; b) “VHPT”; c) Hexagonal hollow section; d) Rectangular
hollow section, e) Square hollow section, f) Circular hollow section.

1152
iii. circular hollow sections produced in accordance with the European fabrication standard EN
10219 and EN 10210, with steel grades from (normal strength) S355 to (high strength) S890.
iv. hexagonal hollow sections produced in accordance with the EN10210 with S355 steel grade.

2.2 Experimental methodology and test campaign


A total of 68 full-scale tests presented on cold-formed members and hot rolled members of
short length (L = 800 mm) and longer tests (L = 2000 mm), with varying wall thicknesses and
load eccentricities. These tests were carried out in the 10MN 4-column test rig in the laboratory
of Bundeswehr University Munich. Stub column tests were carried out with a centric load appli-
cation (in the centroid axis, determined analytically in the case of the VHPT section). The bend-
ing moment was introduced by adding eccentricity to the test specimen by employing a stiff
lever arm plate. Figure 2b and c illustrate the employed experimental loading scheme.
The overarching aim of the experimental test campaign is to obtain a reliable basis for the
calibration of numerical (FEM-based) simulations of an even wider set of load parameters
and cross-sectional configurations. The simulation of the experiments thus employs a process
of reverse engineering which is based on the real geometry of the specimen, as well as the
measured stress-strain curve of the material, in the FEM modelling of the experimental tests.
To facilitate this, each specimen’s geometric shape and the shape deviations from the ideal
geometry have been measured with a 3D scanning system made by Zeiss ©. 3D spline curves were
laid over the point cloud obtained from the 3D scan and imported into the finite element simula-
tion. Furthermore, a DIC (Digital Image Correlation) measurement system was employed, see
Figure 2a. At each test time step (consistent with the experimental test duration) two pictures were
taken with GOM Aramis high-resolution cameras, as the basis for the derivation of the deform-
ations and local strains in a randomly applied speckle field on the specimens, see Figure 3b.

2.3 Numerical methodology and campaign


The mentioned reverse engineering process consists in replicating the experimental test in
a “numerical test”, i.e. in a geometrically and materially non-linear analysis on the imperfect
geometry (GMNIA), with the highest possible accuracy. A GMNIA analysis with the meas-
ured geometrical shape of the sections and material law obtained from tensile coupon tests
can lead to minimum (<5%) deviations to the ultimate load of the buckling tests if the meshing

Figure 2. a) point cloud of the 3D scan data; b) spline curves approximating the real geometry c) experi-
mental test in the the 10 MN test rig; d) 3d schematic representation of the test setup.

1153
Figure 3. a) deformed shape of a VHPS stub-column test of a GMNIA-MEAS numerical analysis with
Abaqus; b) specimen deformed shape of the corresponding experimental test measured by the Aramis system.

and modelling of boundary conditions is accurate and representative. This type of GMNIA
analysis is denoted by “GMNIA-MEAS” in the present paper.
The comparison between the “GMNIA-MEAS” and the experimental test results allows for
a fine-tuned calibration of generalized, somewhat simplified FEM-models, which may then be
used in a broader parametric study. DIC data was used in order to compare both the global short-
ening and the local deformation and buckling phenomena in the numerical and experimental tests,
see Figure 3.
In this calibration, the first step consists in the determination of the mesh density, element type
and boundary conditions that best describe the experimental test, provided that the geometry and
material law are precisely modelled. In the main step, a simpler, more generalized model is devel-
oped and calibrated, with a simplified definition of the material law and the geometrical imperfec-
tions, yet the same FEM mesh size and element types as the ones validated through the
calibration to the experimental tests. The main simplification thus consists in the determination of
an equivalent imperfection shape for the GMNIA calculation. In the study presented in this
paper, the imperfect geometry was derived from the first buckling mode of a Linear Buckling
Analysis (LBA), with the amplitude for the buckling waves calibrated as described in Section 4.
The proprietary software Simulia ABAQUS was used for all numerical simulations, with
linear isoparametric shell elements with reduced integration (element type S4R). As a result of
the model calibration, it was found that a mesh density with a minimum of 60 elements in
circumferential and 200 elements in longitudinal direction was sufficient to obtain results with
an average difference between the “GMNIA-MEAS” and the experimental test result of less
than 5% in terms of ultimate buckling load.

3 EXPERIMENTAL TEST RESULTS AND NUMERICAL MODEL CALIBRATION

3.1 Overview of test results


A synthetic overview of the experimental test results, on cold-formed sections which were car-
ried out as described in Section 2 of this paper, is given in Table 1. Thereby, the eccentricity

1154
Table 1. Experimental test results for the bespoke cross-sections, RHS and SHS, CHS and HEX.
Cross-section B/D H T L e FExp,max

No. Type Steel Grade [mm] [mm] [mm] [mm] [mm] [kN]

1 CHS S355 357 - 9 800 0 1901.6


2 CHS S355 244 - 6.3 800 0 2094.3
3 CHS S355 193.7 - 8.8 800 0 1389.9
4 CHS S355 219.1 - 6.3 800 0 1804.7
5 CHS S355 195 - 6 800 0 3402.6
6 SHS S355 140 140 4 800 0 861,54
7 SHS S355 140 140 4 800 14.6 868
8 VHPS S350GD 140 140 2,5 800 0 623,9
9 VHPS S350GD 140 140 2,5 800 15 614,8
10 VHPS S350GD 140 140 3,5 800 0 1002,4
11 VHPS S350GD 140 140 3,5 800 15 995,9
12 VHPT HX460 140 140 3,5 800 10 1106,9
13 VHPT HX460 140 140 3,5 800 0 1099,2
14 VHPT HX460 140 140 2,5 800 0 584,5
15 VHPT HX460 140 140 2,5 800 10 594,3
16 VHPS S350GD 140 140 2,5 800 137 208,2
17 VHPS S350GD 140 140 2,5 800 314 99,0
18 VHPS S350GD 140 140 3,5 800 91 328,3
19 VHPS S350GD 140 140 3,5 800 308 156,0
20 VHPT HX460 140 140 2,5 800 195 121,1
21 VHPT HX460 140 140 2,5 800 372 63,3
22 VHPT HX460 140 140 3,5 800 195 224,9
23 VHPT HX460 140 140 3,5 800 372 118,3
24 RHS S355 300 150 6 800 0 1583
25 RHS S355 300 150 6 800 17.9 1508
26 RHS S355 300 150 8 800 0 2808
27 RHS S355 300 150 8 800 17.35 2532
28 SHS S355 200 200 5 800 0 1228
29 SHS S355 200 200 5 800 20.92 1224
30 SHS S355 200 200 8 800 0 2917
31 SHS S355 200 200 8 800 20.9 2414
32 CHS S355 323.9 - 8 800 0 3711
33 CHS S355 323.9 - 6.3 800 0 2414
34 CHS S355 323.9 - 5 800 20.9 1668
35 RHS S355 300 150 6 800 57 1310.3
36 RHS S355 300 150 6 800 297 585
37 CHS S355 323.9 - 5 800 57.05 1356
38 CHS S355 323.9 - 5 800 407.05 429
39 CHS S355 323.9 - 6.3 800 57.05 1824
40 SHS S355 140 140 4 2000 137 262
41 VHPS S350GD 140 140 2.5 2000 138.5 203.2
43 RHS S355 300 150 6 2000 57 1290
44 CHS S355 323.9 6.3 2000 57.05 1290
45 RHS S355 300 150 8 800 57 2122
46 RHS S355 300 150 8 800 232 1081
47 SHS S355 200 200 5 800 107 616
48 SHS S355 200 200 5 800 457 194
49 SHS S355 200 200 8 800 107 1442
50 SHS S355 200 200 8 800 457 454
51 HEX S355 250 - 8 2000 75.7 1485.27
52 CHS S770 193.7 - 8.8 800 127.2 1640.64
53 HEX S355 250 - 8 800 75.7 1638.23

(Continued )

1155
Table 1. (Continued )
Cross-section B/D H T L e FExp,max

No. Type Steel Grade [mm] [mm] [mm] [mm] [mm] [kN]

54 HEX S355 250 - 8 800 425.7 457.31


55 HEX S355 250 - 8 800 0 2934.9
56 HEX S355 250 - 8 800 62.8 2118.14
57 CHS S770 193.7 - 8.8 800 0 4511.4
58 CHS S890 219.1 - 10 800 0 7001.68
59 SHS S500 200 200 4 800 0 1112
60 SHS S500 200 200 4 800 63.05 865.072
61 SHS S500 200 200 4 800 107 622.591
62 SHS S500 200 200 4 800 457 212.163
63 SHS S500 200 200 4 2000 107 579.673
64 SHS S500 200 200 5 800 0 1736.293
65 SHS S500 200 200 5 800 63.9 1334.598
66 SHS S500 200 200 5 800 107 895.787
67 SHS S500 200 200 5 800 457 294.552
68 SHS S500 200 200 5 2000 107 852.664

“e” determines the amount of bending. It was applied in direction of the “web” of the VHPT
and RHS sections, and so as to cause additional compression on the wider stiffener in the
VHPS sections. The numbering indicated in the table followed the overall progress of the pro-
ject’s test campaign.

3.2 Calibration of the numerical model


With the chosen FEM modelling technique and discretization, described in Section 2, the
GMNIA-MEAS model is able to approximate the resulting maximum force of the experimen-
tal test with an average error of less than 5% in terms of ultimate load.
The overall evaluation of the calibration work on the various tested cross-section shapes
load eccentricity values led to the conclusion that an equivalent imperfection based on the
LBA first buckling mode with an amplitude b/400 is the most representative “even-
numbered” value for the test series, and was thus taken to be generally more representative
for the local buckling behavior of the these bespoke (and other flat-surfaced) sections than
the value of e0 = b/200 contained in Eurocode design provisions (EN1993-1-5, Annex C).
This is also in line with the findings by Rusch and Lindner 0: they also determined that amp-
litudes of b/400 are most suitable to represent the so-called Winter curves for local buckling
in numerical analyses.
Figure 4 shows a justification for the employed value of b/400 as an equivalent geometric
imperfection for local buckling for the subsequent parametric studies. The same procedures
as above were applied to a number of different cross-sections belonging to the experimental
tests shown in Table 1. In this case, the results obtained in the GMNIA-MEAS calculation
on the basis of the measured geometry and tensile coupon tests from the flat sides of the
section lead to an over-prediction of the ultimate strength, even though the shape of the
load-deformation curve matches the experimental result quite accurately. Figure 4b on the
other hand shows that the amplitude of b/400 once again leads to very accurate predictions
of the buckling strength (albeit not the precise deformation path) of the test. The higher
results for the calculation with the coupon test results can in this case by explained with the
omission of residual stresses in the model inputs. In the subsequent parametric studies, it
was thus deemed to be appropriate to apply the amplitude of b/400 without residual stresses
to all studied sections.

1156
Figure 4. a) load deformation diagram of the experimental test No.30 (SHS 200 x 5,0) measured with
DIC and traditional methods, plotted against the GMNIA-MEAS calibration model results for; b) vary-
ing imperfections GMNIA analyses plotted against the GMNIA with measured values.

4 SUMMARY AND CONCLUSIONS

This paper discussed work carried within the European research project HOLLOSSTAB,
during which new design rules for hollow sections with innovative shapes and/or steel grades
are developed on the basis of the “Overall Interaction Concept” (OIC).
The focus of the paper was put on the calibration of the GMNIA model for the numerical
model for a further extensive numerical campaign.
The results presented in this paper showed the great opportunity for an increase in accuracy
in the understanding of the experimental mechanical model due to the choice of state-of-the-
art measuring techniques in the reverse engineering process. Further work is being focused on
the combination of these numerical models with extensive numerical campaigns, with the pro-
posal of different design curves in the overall design approach.

ACKNOWLEDGMENTS

The authors would like to acknowledge the funding received by the European Community’s
Research Fund for Coal and Steel (RFCS) under grant agreement No. 709892 - HOLLOS-
STAB. Additionally, the authors express their gratitude to the engineering department of
voestalpine Krems Finaltechnik GmbH, in particular to Mr. P. Häckel, for the contributions
and technical support provided during the study on VHPT and VHPS sections.

1157
REFERENCES

Boissonnade N., Hayeck M., Saloumy E. & Nseir J., 2017. An Overall Interaction Concept for an alter-
native approach to steel members design, Journal of Constructional Steel Research, Vol. 135, pp.199–
212, Elsevier, London/Amsterdam.
EN 10210, 2006. Hot finished structural hollow sections of non-alloy and fine grain steels, CEN – Euro-
pean Committee for Standardization, Brussels.
EN 10219, 2006. Cold formed welded structural hollow sections of non-alloy and fine grain steels, CEN –
European Committee for Standardization, Brussels.
EN 1993-1-1, 2005. Eurocode 3. Design of steel structures – Part 1-1: General rules and rules for build-
ings, CEN - European Committee for Standardization, Brussels.
EN 1993-1-5 Eurocode 3. 2006. Design of steel structures - Part 1-5: General rules - Plated structural
elements, CEN - European Committee for Standardization, Brussels.
Gardner L., Saari N. & Wang F., 2010. Comparative experimental study of hot-rolled and cold-formed
rectangular hollow sections.
Jia-Lin Ma, Tak-Ming Chan & Ben Young, 2015. Experimental Investigation on Stub-Column Behavior
of Cold-Formed High-Strength Steel Tubular Sections.
Jia-Lin Ma, Tak-Ming Chan & Ben Young, 2017. Tests on high-strength steel hollow sections: a review,
Structures & Buildings, 170 (9), Pages: 621–630.
Kim DK, Lee CH & Han KH et al. 2014. Strength and residual stress evaluation of stub columns fabri-
cated from 800 MPa high-strength steel, Journal of Constructional Steel Research 102: 111–120.
Rusch A. & Lindner J., 2001. Überprüfung der grenz (b/t)-Werte für das Verfahren Elastisch-Plastisch,
Stahlbau, 70, pp. 857–868, Wiley/Ernst&Sohn Berlin.
Schafer, B.W., 2008. Review: The Direct Strength Method of cold-formed steel member design, Journal
of Constructional Steel Research, Vol. 64, Issue 7–8, pp. 766–778, Elsevier.
Schillo N. & Feldmann M., 2015. Local buckling behaviour of welded box sections made of
high-strength steel, Steel Construction, 8 (3), Pages: 179–186.
Toffolon A. & Taras A. 2017. Numerical investigation of the local buckling behaviour of high strength steel
circular hollow sections, Proc. of Eurosteel 2017, Sept. 15–17, Copenhagen, pp. 3603–3612, Ernst & Sohn,
Berlin.
Wang J, Afshan S & Gardner L, 2017, Axial behaviour of prestressed high strength steel tubular
members, Journal of Constructional Steel Research, Vol: 133, Pages: 547–563, ISSN: 0143-974X.

1158
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Proposal of a design curve for the overall resistance of cold-formed


rectangular and square hollow sections

A. Toffolon & A. Taras


Institute of Structural Engineering, Bundeswehr University Munich, Germany

ABSTRACT: The design of slender square (SHS) and rectangular (RHS) hollow-sections made
of high-strength steel is often governed by the combined effect of local (L) and global (G) instabil-
ity phenomena. Common design rules, e.g. those found in the Eurocodes, penalize these slender
sections through a conservative omission of various mechanical effects and a categorization of
cross-sections in distinct cross-sectional class, from “Class 1” for stocky sections to “Class 4” for
very slender ones, that intrinsically leads to discontinuities in the strength representation. This
paper discusses the results of an extensive numerical campaign and a subsequent definition of
a continuous design curve for the treatment of local buckling. The study represents one of the
steps of the RFCS research project “HOLLOSSTAB”, during which new design rules for High
Strength Steel (HSS) hollow sections are developed on the basis of an “Overall Interaction Con-
cept” (OIC). This concept – similarly to the Direct Strength Method (DSM) used in North Amer-
ica for the design of cold-formed steel open cross-sections – makes use of the results of
(numerical) linear buckling analyses (LBA) for the cross-section and member to determine the
slenderness and consequently an overall buckling reduction factor. The paper discusses the effect-
iveness of this approach in the general framework of buckling design checks for hollow sections,
reviews existing rules and their implications, shows the numerical campaign results and introduces
initial design formulas based on the overall buckling resistance approach.

1 INTRODUCTION

Manufacturers of steel buildings and racking systems for large logistics centers face the chal-
lenge of producing very light-weight, economical structures that nevertheless meet very high
standards with respect to stiffness and strength. The use of High-Strength Steel (HSS) is an
effective strategy to meet these challenges. However, current design standards, such as the
Eurocode (EN 1993-1-1, 2005), do not consistently assess the resistance of such sections
against local (L) buckling modes with sufficient accury, especially for general combinations of
loading (compression and mono- or biaxial bending). This paper describes a numerical study
of the local buckling of rectangular and square hollow sections, with an aim of developing
first proposals for the local buckling design of these sections. The study is carried out within
the framework of the on-going RFCS project “HOLLOSSTAB”, during which new design
rules for innovative hollow sections of various shapes and steel grades are being developed.
The Geometrically and Materially Nonlinear Analyses with Imperfections (GMNIA) are per-
formed in a larger parametric study calibrated in previous steps (Toffolon & Taras, 2018), and
the results of the parametric study will be used for the development of appropriate design
rules, with a continuous representation of strength throughout slenderness ranges, using the
“Overall Interaction Concept” (Boissonnade, Hayeck, Saloumy, & Nseir, 2017). This concept,
which is similar to the Direct Strength Method (DSM) (Schafer, 2008) used in North America
for the design of cold-formed steel open cross-sections, makes use of the results of (numerical)
linear buckling analyses (LBA) for the whole member to determine the slenderness and conse-
quently an “overall” buckling reduction factor.

1159
2 SCOPE AND METHODOLOGY

As stated above, the strength and stability of hollow sections of various shapes and steel
grades are the object of study of the European research project HOLLOSSTAB by means of
an extensive experimental and numerical test campaign and parametric study. Among the
various section types considered, unstiffened and stiffened, double- and mono-symmetrical
cold-formed sections are being tested at the Chair of Steel Structures of Bundeswehr Univer-
sity Munich. The exclusive focus on cold-formed cross-sections in this paper is justified by
comparative studies on RHS and SHS cold-formed, hot-formed and hot-finished cross-
sections, which confirmed a marked difference in the mechanical properties and need for
a revision of design rules, see (Ma, Chan &Young, 2016) and (Ma, Chan &Young, 2017).
Strain hardening has been shown to be the main influence parameter, among other features
such as geometric imperfections, see e.g. (Gardner, Saari & Wang, 2010). The next sections of
the paper describe the scope and methodology of this series of tests and numerical analyses.

2.1 Numerical methodology and campaign


Two types of cold-formed cross-sections are the subject of the present study and are part of
the HOLLOSSTAB project, see Figure 1.
A significant number of tests on high-strength cold-formed hollow sections were recently
published internationally by Wang, Afshan & Gardner (2017), Kim, Lee & Han (2015),
Schillo & Feldmann (2015) and Toffolon & Taras (2017). The latter were used, employing
a reverse engineering process, as the basis for the calibration of an equivalent imperfection
shape that could representatively be used in an extensive GMNIA parametric study. Thereby,
the imperfect geometry for the GMNIA calculations was derived from the first buckling mode
of a Linear Buckling Analysis (LBA), with the amplitude for the buckling waves calibrated as
e0;loc ¼ b=400, see the authors’ dedicated paper (Toffolon et al.) at this conference. The general
procedure adopted in this parametric study is schematically represented in Figure 2.

3 THE OIC FRAMEWORK

In the European research project HOLLOSSTAB, the Overall Interaction Concept (OIC) is
used as the main reference framework and representation tool for the analysis, representation
and design method calibration for the resistance of both members and cross-sections. As stated
in the introduction, this concept makes use of a generalized, “overall” definition of slenderness
and of the buckling reduction factor in order to define the ultimate buckling capacity of cross-
sections and members failing in local (L), distortional (D), global (G) or interactive (I) buckling.
In this paper, only local buckling is considered. Therefore, only the overall slenderness λL for
local buckling, as well as the corresponding buckling knock-down factor χL, will be presented
and discussed in the following. The OIC adopts a series of steps, illustrated in Figure 3.

Figure 1. Cold-formed rectangular (SHS, RHS) acc. To EN 10219 (2006), with steel grades from
(normal strength) S355 to (high strength) S700.

1160
Figure 2. a) imperfection and mode shape from the LBA analysis; b) Example of LBA results for the
analyzed cross-sections; c) stress-relationship used in the numerical simulations.

Figure 3. Illustration of the OIC-steps; a) shows the 5 steps of the OIC method, where a design curve
provides an overall knock-down factor given the overall slenderness; b) shows the necessary parameters
for the OIC-step calculations, with alternative versions of CS resistance definition.

Alternative definitions of the slenderness values and achievable load amplification factors
based on the OIC approach have been considered within the study presented in this paper. In
alternative to the representations based on a knock-down factor applied to the full ideal-
plastic reference resistance Rpl as shown in Figure 3, representations using a linear-plastic
resistance RLIN-PL and the elastic resistance REL were also considered, see section 5.

4 NUMERICAL STUDY AND COMPARISON WITH CODE PROVISIONS

4.1 General remarks


A general numerical test campaign was conducted with a comprehensive parameter set
for geometries and steel grades, as well as a representative range of combinations of
applied compression forces and (bi-axial) bending moments (see Table 1 for reference).
The choices for the model geometry and material are coherent with the scope of the
paper, the analysis of the cold formed cross-sections according to EN10219. Differently
from the previous analysis, the input length is given as a fraction of the critical global
(Euler) buckling length for Ncr=Npl.

1161
Table 1. Parameters for the SHS and RHS numerical campaign.
Thickness L/Lcr Steel grade Imperfection Φy Φz
[mm] [-] amplitude [°] [°]

2.0 0.1 S355 b/400 0 0


2.5 0.15 S460 15 15
3.0 0.2 S550 30 30
3.0 S700 45 45
4.0 60 60
5.0 75 75
6.3 90 90
8.0
10.0
12.0

4.2 Local buckling of SHS and RHS sections for compression and biaxial bending
The numerical study considered for this paper consisted of around 9000 calculations, which them-
selves form only a smaller part of the total parametric study carried out in the context of the EU-
RFCS project HOLLOSSTAB. The procedure adopted to cover the overall field of possible.
N + My + Mz combinations is shown in Figure 4 for a single cross-section, an RHS section
with an H/B ratio of 2. The continuous surface is the chosen, arbitrary “reference load”, which
was given in the n/my/mz space by the equation of a sphere: n² + my² + mz² = 1. The dots show
the points corresponding to the full plastic resistance of the section, which is asymmetrical with
respect to the origin and thus has a markedly varying distance from the reference load surface.
In Figure 5, all results obtained during the parametric study conducted for this paper are
plotted using the three definitions for the slenderness and resistances adopted before, where
the reference cross-sectional resistance varies between the non-linear plastic, linear plastic and
elastic resistances.
As can be seen in the figures, the results for pure compression, bending about the y-axis and
bending about the z-axis generally scatter fairly homogeneously around the traditional Winter-
curve formulation, as long as plastic resistances are used for reference. If the elastic resistance
is used, see Figure 5c, the bending resistances are raised significantly, as should be the case due
the difference between the elastic and plastic reference bending resistance. However, for all
other results, for which the mix of bending moments and compression is made to vary, the

Figure 4. The cross-section resistance (“R’pl”) in the normalized 3D plot with n, my and mz in two dif-
ferent views for a RHS with H/B=2.

1162
Figure 5. Parametric study on the cross-sectional resistance of SHS and RHS sections in the OIC
format: a) results based on the non-linear plastic resistance Rpl; b) results bases on the linear-plastic
resistance; c) results based on the elastic resistance; d) results based on the non-linear plastic resistance,
plotted over a definition of the slenderness that used the linear-plastic resistance.

position and width of the scatter band is strongly influenced by the chosen representation.
While the majority of the results lie somewhere between the Winter and the European column
buckling curve “b” (chosen here purely for reference) if χL is based on the non-linear cross-
sectional capacity, these same results come to lie above the Winter curve for a representation
of χL that is based on the linear-plastic and (even more so) the elastic resistance.

5 DEVELOPMENT OF A DESIGN REDUCTION FACTOR

5.1 General remarks and methodology


In the following, the results from the extensive numerical campaign are collected and evaluated in
order to calibrate design curve proposals. Since the worked data have shown a strong correlation
with the classic Winter curve, the Winter format is the only consistent choice for a new

1163
Figure 6. a) Collected results of the entire SHS and RHS numerical campaign: b) selected step for the
evaluation and 5% lower bound of the χL results is chosen.

design curve. This approach finds support in similar design frameworks such as the Eurocode
(EN 1993-1-5), and the direct strength method (Schafer, 2018). In Taras (2016) the same design
approach was systematically applied for the global buckling case of steel beam-columns.
The parameters to be tuned from the scattered data comprise three coefficients, n1, n2 and c,
used as follows in a Winter-type formula for the local buckling knock-down factor χL:

n1 n2
χ L ¼ λL  cλL ð1Þ

which is a more general formulation of the Winter curve:

1 2
χL ¼ λL  0:22λL ð2Þ

The calibration of the three coefficients was carried out by a numerical optimization proced-
ure. The λL range of the scattered data was divided into n-steps of 0.01. In Figure 6, an over-
view of the collated data is shown in the OIC-format, while Figure 6b shows an interval of
0.01 at around λL =0.75. The values of χL that belong to the 5% lower population are con-
sidered further, while the rest of them is discarded. The procedure is repeated for each interval,
and for each 0.01 interval a mean value of the lower bound range of the buckling reduction
factor is obtained.

5.2 First proposals and discussion


In the following, two initial proposals for the local buckling resistance of SHS and RHS are pre-
sented, which make reference to the non-linear plastic and the linear-plastic cross-sectional resist-
ance, respectively. A proposal for both definitions is then provided, using the methodology
described in 5.1.
As is expected from the previous considerations in section 4, the non-linear plastic definition
of the cross-sectional resistance leads to fairly low values of the corresponding buckling reduc-
tion factor, and the 5% lower bound confirms this trend, with the additional feature of having
even more scatter in this range than in the overall data pool. This is shown in Figure 7a, where
the points fall even below the (otherwise unrelated, in its origins) EC3 column buckling curve

1164
Figure 7. a) Design curve based on RPL,L and corresponding fitting data from the 5% lower bound dataset
of the RHS and SHS extensive numerical campaign; b) same result for the different definition of RLIN-PL,L.

“b”. It shall be also noted that the data points follow locally a slightly different pattern than the
design curve, which can be described as follows:

0:75 1:15
χ L ¼ λL  0:37λL ð3Þ

Figure 7b represents the lower bound dataset from the linear plastic definition for the cross-
section base resistance. There is much lower scatter in the lower 5% fractile data for this repre-
sentation. The design curve for this case is:

0:87 1:31
χ L ¼ λL  0:28λL ð4Þ

As a direct result of the definition of RLIN-PL,L the curve describes higher values of the knock-
down factor, and is found to come to lie much closer to the Winter domain.
The advantage of choosing the non-linear plastic definition for the section resistance lies in
a mechanically more consistent inclusion of all material non-linear effects, such as the material
strain-hardening. However, when it comes to the design curve definition, a mechanically more
straightforward definition of the reference resistance – such as RLIN-PL,L – leads to a better descrip-
tion of the behavior of the “critical” dataset, which is the dataset for the design curve calibration.
Another positive aspect of the curve based on a linear-plastic reference resistance is the con-
venient position close to the traditional definition of the Winter curve, approaching the corres-
ponding border of the Winter-curve domain of EN 1993-1-5. This increases the consistency of
any design proposal with, whereas the one based on the non-linear plastic resistance tends to
be reduced to a mere dataset curve fitting.

1165
6 SUMMARY AND CONCLUSIONS

This paper discussed work carried within the European research project “HOLLOSSTAB”,
during which new design rules for hollow sections with innovative shapes and/or steel grades
are developed on the basis of the “Overall Interaction Concept” (OIC).
The focus was laid on the development of initial proposals for the design against local buck-
ling of unstiffened SHS and RHS sections. The results presented in this paper showed the
great opportunity for an increase in accuracy, as well as some challenges, presented by the use
of OIC-type design formulae. This paper contains first proposals for local buckling strength
curves that apply for cold-formed sections. Further work will focus on the finalization of
these proposals, their combination with global buckling cases, on the underlying level of reli-
ability, and on the gain in economy given by the proposed design methodology in practical
applications.

ACKNOWLEDGEMENT

The authors would like to acknowledge the funding received by the European Union’s Research
Fund for Coal and Steel (RFCS) under grant agreement No. 709892 - HOLLOSSTAB.

REFERENCES

Boissonnade, N., Hayeck, M. & Saloumy, E. & Nseir, J., 2017. An Overall Interaction Concept for an
alternative approach to steel members design, Journal of Constructional Steel Research, Vol. 135, pp.
199–212, Elsevier, London/Amsterdam.
EN 10219, 2006. Cold formed welded structural hollow sections of non-alloy and fine grain steels, CEN –
European Committee for Standardization. Brussels.
EN 1993-1-1, 2005. Eurocode 3. Design of steel structures – Part 1-1: General rules and rules for build-
ings, CEN - European Committee for Standardization, Brussels.
EN 1993-1-5, 2006. Eurocode 3. Design of steel structures - Part 1-5: General rules - Plated structural
elements, CEN - European Committee for Standardization. Brussels.
Gardner, L., Saari, N. & Wang, F., 2010. Comparative experimental study of hot-rolled and cold-formed
rectangular hollow sections, Thin-Walled Structures.
Kim D.K. Lee C.H. & Han K.H., 2014. Strength and residual stress evaluation of stub columns fabri-
cated from 800 MPa high-strength steel, Journal of Constructional Steel Research. 102: 111–120.
Ma, J.L. & Chan, T.M. & Young, B. 2016. Experimental Investigation on Stub-Column Behavior of
Cold-Formed High-Strength Steel Tubular Sections. Journal of Structural Engineering.
Ma, J.L. & Chan, T.M. & Young, B., September 2017. Tests on high-strength steel hollow sections: a
review, Structures & Buildings, 170 (9), Pages: 621–630.
Schafer, B.W., 2008. Review: The Direct Strength Method of cold-formed steel member design. Journal
of Constructional Steel Research. Vol. 64, Issue 7–8, pp. 766–778, Elsevier.
Schafer, B. W. 2018. Advances in the direct strength method of thin-walled steel design, Eighth Inter-
national Conference on Thin-Walled Structures (ICTWS).
Schillo, N. & Feldmann, M. 2015. Local buckling behaviour of welded box sections made of
high-strength steel, Steel Construction, 8 (3), Pages: 179–186.
Taras, A. 2016. Derivation of DSM-type resistance functions for in-plane global buckling of steel
beam-columns, J. of Constructional Steel Research, 105, pp. 95–113, Elsevier London/Amsterdam.
Toffolon, A. & Taras, A. 2017. Numerical investigation of the local buckling behaviour of high strength
steel circular hollow sections, Proc. of Eurosteel 2017, Sept. 15-17, Copenhagen, pp. 3603–3612, Ernst
& Sohn, Berlin.
Toffolon A., Müller A., Niko I., Taras A., 2019. Experimental and numerical analysis of the local and
interactive buckling behaviour of hollow sections, Proceedings of SDSS 2019, Prague.
WangJ, AfshanS & GardnerL, 2017. Axial behaviour of prestressed high strength steel tubular members,
Journal of Constructional Steel Research, Vol: 133, Pages:547–563.

1166
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Behavior of extended end-plate connections under cyclic


alternate loading

I.C. Tomăscu & R.M. Bâlc


Technical University of Cluj-Napoca, Cluj-Napoca, România

ABSTRACT: Behavior of steel frame structures, subjected to static and dynamic loadings, exten-
sively studied over the past decades, can be controlled and influenced by the connections configur-
ation. Thus, special attention has been paid over the years to the evaluation of the three basic
characteristics of steel connections, strength, stiffness and ductility, and their impact on the global
structure behavior. Starting from this, the present paper introduces a study of a beam-to-column
bolted stiffened double extended end-plate connection. Based on the experimental and numerical
models and through a parametric study, analytical models capable of describing its behavior have
been established. Following the nonlinear analysis carried out in ABAQUS, the cyclic behavior of
the connection was evaluated in terms of stress distribution, its equivalent plastic strain, local phe-
nomena evolution and moment-rotation curves. The results of these studies are analyzed and com-
pared, and finally some conclusions can be drawn regarding the behavior of this typology of joints.

1 INTRODUCTION

Joints of the steel moment resisting frames (MRFs) have a fundamental role in defining the behav-
ior of this category of structures. Stiffness, strength and rotation capacity of MRFs can be influ-
enced and even controlled by a proper design of the connections between vertical and horizontal
members. This fact has been proven over time by unannounced brittle failure of joints, especially
welded, caused by natural disasters (Northridge 1994, Kobe 1995, Tohoku 2011); thus, design of
MRFs joints being the subject of numerous and intensive research studies (Nogueiro et al. 2006b,
Kartal et al. 2010, Razavi et al. 2012, Augusto et al. 2013 & 2017, Sofias et al. 2014, El-Khoriby
2016, Haghollahi & Jannesar 2018).
Special attention has been given to the extended end-plate connections which, although
having an increased moment capacity, being very close to the rigid behavior, can provide suffi-
cient flexibility to the structure and exhibit a ductile behavior.
Experimental research, time-consuming and resource-intensive, is used more and more
often to validate numerical and analytical models, which are more accessible in current design
practice. In this respect special attention is paid to the material models suitable for the simula-
tion of the cyclic behavior of structures.
This paper presents the analysis of the behavior of steel bolted extended end-plate moment con-
nections under the cyclic alternate loading. Nonlinear analysis with FEM was conducted in
ABAQUS (Hibbit et al. 2011) and was developed in accordance with the experimental test per-
formed by the authors (Mureșan 2017). The numerical results were compared to those obtained
from the experimental test and were used to create an analytical model derived from the Richard-
Abbott model.
The paper reveals aspects regarding stress state and deformations in the structure, failure
mode, moment-rotation curves evolution. The most sensitive areas of the connection are
detected as well as the components that provide the greatest deformability of the structure
without producing unexpected failure modes.

1167
2 EXPERIMENTAL TEST

The experimental research project consists of a steel beam-to-column bolted double extended end-
plate connection, subjected to cyclic alternate loading, Figure 1. The connection members were
welded H-shape (column) and I-shape (beam) profiles and were connected by high strength bolts
M22/10.9. The loading was applied by means of two different hydraulic actuators which were
placed to the end of beam. These actuators were used to apply, under displacement control, the
desired displacement history at the beam end, as a function of monotonic tests, according to the
procedure recommended by ECCS (ECCS 1986).
Monotonic tensile tests were performed for establish the mechanical properties of specimens,
which are illustrated in Table 1.
Starting from the measured entities values, force and displacement, the connection bending
moment was calculated. Beam rotation over the column axis consists of column web panel rotation
and connection rotation. Experimental tests revealed that the column web panel rotation is very
small, its contribution being insignificant. Thus, the relative rotation of the connection was calcu-
lated according to the Equation 1,

D
¼ ð1Þ
lb

Figure 1. Details of the studied structure.

1168
Table 1. Mechanical properties of the specimen.
Element part fy (MPa) fu (MPa) εu

Beam flange 423 545 0.07


Beam web 230 397 0.16
End-plate 350 535 0.10
Column flange 360 552 0.11
Column web 355 543 0.10
Stiffeners 423 545 0.07

Figure 2. Member contribution to the overall rotation: column, beam and joint contribution.

where ϕ = connection relative rotation; Δ = beam end displacement; lb = beam length.


The measured beam end displacement have been corrected by substracting the elastic contri-
bution due to the column and beam flexural deformability, Figure 2, according to the Equa-
tions 2 and 3,

D ¼ Dmeasured  Delastic ð2Þ

lc  lb2  P P  lb3
Delastic ¼ þ ð3Þ
12EIc 3EIb

where lc = column length; P = force applied to the beam end; E = elastic modulus; I = moment
of inertia.

3 FEM MODEL

The numerical model was developed by means of ABAQUS v.6.11 (Hibbit et al. 2011), Figure 3.
The finite element type C3D8I, 8-node brick element, incompatible mode, was mainly adopted to
discretize all components and revealed quite fairly the deformation of the connection components
subjected to bending (e.g. end-plate, column flange). In addition, to avoid possible distortion of
the connection elements, 2 or 3 layers were considered in their thickness. The finite element type
C3D6, 6-node wedge element, was used to mesh the bolts and the circular areas around the holes.
For the rigid plates employed for constraints definition and for loading application the finite
element type R3D4, 4-node rigid element, was adopted. Welds have not been modeled, account-
ing to their good execution proven by experimental test.
The beam-to-column load transmission was achieved by the interaction between the end-
plate and column flange by tightening the preloaded bolts. This interaction is essential in the

1169
Figure 3. Finite element model of typical four-bolt stiffened double extended end-plate connection

connection behavior simulation and was achieved by defining a small sliding surface-to-
surface contact. “Hard contact” using augmented Lagrange formulation was considered for
the normal direction to resist penetration and for tangential contact was considered
a frictional contact using penalty stiffness formulation, friction coefficient μ = 0.3. The same
type of interaction was applied to the bolt head/end-plate and nut/column flange interfaces.
A tangential frictionless contact was considered between the bolt hole and the bolt shank.
Experimental tests performed on steel specimens subjected to monotonic and cyclic alter-
nate loadings revealed that material behaves differently depending on loading type, so that the
simple elasto-plastic isotropic or kinematic hardening models do not accurately describe the
seismic behavior of the material (Shi et al. 2011). For this reason, material was defined using
the combined isotropic/kinematic hardening model available in ABAQUS (Chaboche 1989,

1170
Figure 4. Cyclic loading protocol.

Shi et al. 2011, Augusto et al. 2013), which uses Von Mises yield criterion and a flow rule and
an evolution law of backstresses (α) for a half-cycle, Equation 4,

Ck  γ εpl 
αk ¼ 1e k ð4Þ
γk

where Ck and γk are material constants; Ck/γk = maximum change in backstress.


Material plastic behavior was defined by pairs of stress and strain values. These values were
achieved from monotonic coupon tensile tests: 397 MPa - ultimate stress and 0.16 - ultimate
strain. For bolts a bi-linear elasto-plastic material model was considered, with ultimate stress
of 1000 MPa and ultimate strain of 0.0048.
Load was applied in two steps. The first one consists in the bolts preloading, using the “bolt
load” option available in ABAQUS. The preloading force introduced in bolts was 216 kN. In
the second step the cyclic alternate loading was applied by imposing a boundary condition
with amplitude. The cyclic loading history is plotted in the Figure 4.

4 ANALYTICAL MODEL

4.1 Introduction
In the last decades many simplified models have been developed to predict the connections
behavior, in order to be implemented in the global structural design of steel frames (Rui &
Simoes da Silva 2001, Nogueiro et al. 2006a). Starting from the analysis of these models, this
study proposes two simplified models for predicting the behavior of a four-bolt stiffened
double extended end-plate connection, related to both monotonic and cyclic loadings, applic-
able in the current design of structures. These models are based on the experimental, numer-
ical and component method moment-rotation curves and represent adaptations of the
Richard-Abbott model (Richard & Abbott 1975). This model was chosen according to the
parameters defining the moment-rotation curve and taking into account the characteristics
provided by the component method (ASRO 2006), the shape of the curve and the possibility
to fit the corresponding curve to that obtained on the numerical model validated by the
experimental test (Mureșan 2017, Tomăscu & Bâlc 2018).

4.2 Monotonic loading


The Richard-Abbott model (Richard & Abbott 1975) comprises four parameters: initial stiff-
ness, Rki, plastic moment capacity, M0, hardening stiffness, Rkp and connection shape param-
eter, n. This model proposes that the initial stiffness and moment capacity to be determined
using the component method (ASRO 2006), and this was done for a series of specimens

1171
considered in a parametric study previously performed by the authors (Mureșan 2017); the
other two parameters, hardening stiffness and shape parameter, are calibrated to the following
values: Rkp = Rki/100, n = 1.9 ÷ 2.1 (Mureșan 2017).
Thus, determining the moment-rotation curve of the bolted extended end-plate connec-
tion, based on the proposed analytical model, assumes calculating the initial stiffness
and the moment capacity using the component method and then replacing them in the
Equation 4,

0:99  Rki  θr
M¼ þ 0:01  Rki  θr ð4Þ
 2 12
ki θr
1 þ 0:99R
M0

where M = bending moment of the connection; θr = relative rotation of the connection.

4.3 Cyclic loading


For cyclic behavior prediction we used the model proposed by Rui and Simoes da Silva for
composite joints (Rui & Simoes da Silva 2001, Nogueiro et al. 2006a), derived from the same
Richard-Abbott model (Richard & Abbott 1975). Parameters in the equations describing the
loading and unloading branches consist in the same four parameters to be found in the model
for monotonic behavior prediction, for which the same notations were kept. Thus, the harden-
ing stiffness was considered 1 percent of initial stiffness, Rkp = 1/100 Rki and the shape param-
eter was considered within the same value range, n = 1.9 ÷2.1.
The loading and unloading branches are described by the Equation 5, respectively the
Equation 6. The next loading branch uses the same expression as the unloading branch, but
the Mp and θp coordinates are replaced by Mn and θn.
 
Rki  Rkp  θr
M¼   þ Rkp  θr ð5Þ
ðRki Rkp Þθr n 1=n
1þ 
M0 
   
Rki  Rkp  θp  θr  
M ¼ Mp   n 1n  Rkp  θp  θr ð6Þ
ðRki Rkp Þðθp θr Þ
1 þ  2M0



Based on these expressions and using the parameters set in the previously proposed analytical
model, the moment-rotation curve of the connection under cyclic alternate loading was obtained,
Figure 5.

5 RESULTS

Analyzing the von Mises stress state in the connection elements close to the failure, stress con-
centrations can be observed in both tension and compression areas, particularly in the nut-
column flange interaction area, Figure 6.
Also, local buckling appears in the compressed beam flange and high stresses appear in
the bolts. The gap between the end-plate and the column flange extends almost up to the
beam flange (the theoretical center of compression) and it is more pronounced in the case of
experimentally tested specimen, Figure 7, while the bending deformation of the end-plate is
very low.
Connection moment-rotation hysteretic curves exhibit a good agreement in terms of plastic
moment capacity, Figure 8. The modified Richard-Abbott model reveals an energy dissipation

1172
Figure 5. Moment-rotation curve of the connection using the proposed analytical model.

Figure 6. Von Mises stress distribution in end-plate, bolts and column flange.

capacity close to that in the experimental model. The cyclic curves resulting from the experi-
mental test reveal stiffness degradation in the last cycles due to the detachment of plates and
progressive deterioration of bolts and nuts. These phenomena do not develop in the numerical
model due to the contact between the bolt head and the end-plate, respectively between the
nut and the column flange. The numerical model evolves with tight loops, smaller rotation
and exhibits in the last cycles stiffness degradations explained by stress and strain concentra-
tions that develop in the compressed areas, in the end-plate and in bolts, close to yielding
phase.

1173
Figure 7. Connection failure: experimental vs. numerical model.

Figure 8. Comparison of the moment rotation curve of the connection.

6 CONCLUSIONS

This paper investigates the cyclic behavior of a beam-to-column extended end-plate connec-
tion using experimental, numerical and analytical models.

1174
The analysis of the cyclic moment-rotation curves on the three models reveals good approxima-
tions of the stiffness and last moment values in the first cycles.
The numerical model was developed using ABAQUS finite element analysis. The material mod-
eling for the cyclic behavior uses a combined isotropic/kinematic hardening model available in
ABAQUS.
The analytical model based on the Richard-Abbott model was developed starting from the
moment-rotation curves determined according component method together with the same curves
plotted from the monotonic experimental test.
The numerical moment-rotation curves in the first cycles are close enough of those determined
with the modified Richard-Abbott model, but evolve with stiffness degradations, accentuated in
the last cycles, caused by the intense local phenomena that develop in the connection area.
The connection failure occurs with greater deformations in the numerical model and plates
detachment is more pronounced in the experiment.

REFERENCES

ASRO. SR-EN 1993-1-8. 2006. Eurocod 3: Proiectarea structurilor de oțel Partea 1-8: Proiectarea îmbinărilor.
Augusto, H. et al. 2013. Numerical simulation of partial-strength steel beam-to-column connections
under monotonic and cyclic loading. Proceedings of the Congress on Numerical Methods in Engineer-
ing. Volume 1: 121–140. Bilbao. 25-28 June 2013. Spain.
Augusto, H. et al. 2017. Cyclic behaviour characterization of web panel components in bolted end-plate
steel joints. Journal of Constructional Steel Research. Volume 133: 310:333.
Chaboche, J.L. 1989. Constitutive equations for cyclic plasticity and cyclic viscoplasticity. International
Journal of Plasticity. Volume 5: 247–302.
European Convention for Constructional Steelwork (ECCS). Recommended Testing Procedure for Assess-
ing the Behaviour of Structural Steel Elements under Cyclic Loads. 1986.
El-Khoriby, S. et al. 2016. Simplified modelling of steel frame connections under cyclic loading. Proceed-
ings of The 2016 Structures Congress (Structures 16). Jeju Island, Korea, August 28-September 1, 2016.
Haghollahi, A. & Jannesar, R. 2018. Cyclic Behavior of bolted extended end-plate moment connections with
different sizes of end plate and bolt stiffened by a rib plate. Civil Engineering Journal. Vol. 4, No. 1: 200–211.
Hibbit, Karlsson & Sorenson, Inc. 2011. Abaqus User’s Manual.
Kartal, M.E. et al. 2010. Effects of semi-rigid connections on structural responses. Electronic Journal of
the Structural Engineering. Volume 10: 22 –35
Mureșan, I.C. 2017. Analiza neliniară a îmbinărilor grindă-stâlp cu placă de capăt extinsă și șuruburi.
Teză de doctorat. Universitatea Tehnică din Cluj-Napoca.
Nogueiro, P. et al. 2006a. Numerical implementation and calibration of a hysteretic model for cyclic
response of end-plate beam-to-column steel joints under arbitrary cyclic loading. Computational
Methods in Engineering and Science. EPMESC X, Aug. 21-23, 2006, Sanya, Hainan, China.
Nogueiro, P. et al. 2006b. Experimental behaviour of standardised european end-plate beam-to-column steel
joints under arbitrary cyclic loading. Stability and Ductility of Steel Structures. Lisbon, Portugal, Septem-
ber 6-8, 2006.
Razavi, M. et al. 2012. Computational benchmarks in simulation of cyclic performance of steel connec-
tions using three dimensional nonlinear finite element method. i-manager’s Journal on Structural
Engineering. Volume 1. No. 3. Issue Sep-Nov 2012: 15 –25.
Richard, R.M. & Abbott, B.J. 1975. Versatile elasto-plastic stress-strain formula. J. the Engineering
Mechanics Division. ASCE, 101 (EM4): 511 –515.
Rui, S. & Simoes da Silva, L. 2001. Cyclic behaviour of end-plate beam-to-column composite joints.
Steel and Composite Structures. Vol. 1, No. 3: 355 –376.
Shi, Y. et al. 2011. Experimental and constitutive model study of structural steel under cyclic loading.
Journal of Constructional Steel Research. Volume 67: 1185 –1197.
Sofias, C.E. et al. 2014. Experimental and FEM analysis of reduced beam section moment endplate con-
nection under cyclic loading. Engineering Structures. Volume 59: 320 –329.
Tartaglia, R. et al. 2015. Numerical investigation on the seismic response of bolted extended stiffened
end-plate joints. 8th International Conference on Behavior of Steel Structures in Seismic Areas. Shang-
hai, China, July 1-3, 2015.
Tomăscu, I.C. & Bâlc, R.M. 2018. Analytical approach for the moment-rotation prediction of four-bolt stiff-
ened double extended end-plate connections. Proceedings of the C65 International Conference: Tradition
and Innovation. 65 Years of Higher Education in Civil Engineering in Transilvania. 15–17 November 2018.
Cluj-Napoca, Romania.

1175
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Thermorheological testing and modelling of seismic bearing


elastomers

C. Treib, M.A. Kraus & A. Taras


Institute of Structural Engineering, University of German Armed Forces, Munich, Germany

ABSTRACT: This paper investigates the thermoviscoelastic behavior of two commonly


used bridge bearing elastomers: natural rubber (NR) and chloroprene rubber (CR). Aiming at
an improved and thus more economic modelling of bridge bearings, especially at seismic load-
ing, the energy dissipated to heat is assessed to predict the actual temperature inside the bear-
ings and the corresponding material behavior.
Carrying out thermomechanical experiments such as ‘Dynamic Mechanical Thermal Analysis’
(DMTA) allows a strain rate dependent, that is viscoelastic, material characterization. Conse-
quently, such experiments were carried out to provide data to calibrate material parameters for
a constitutive model. Viscoelastic characteristics can mechanically be described via a Generalized
Maxwell Model. A mathematical approach to represent the model is the so called ‘Prony-Series’,
whose parameters are fitted to experimental data. Assuming thermorheologically simple material
behavior, the parameter estimation in the frequency domain replaces time consuming relaxation
and retardation experiments by applying the so-called Time-Temperature-Superposition Principle
(TTSP). Investigations on the linearity-bounds in the stress-strain-time-relations as well as experi-
ments on the dissipated energy and the resulting temperature increase in the specimens were also
conducted. Finally, the parameters found within this paper are compared to former experimental
results from DMTA investigations with other batches of NR and CR with different test specimen
geometries. The outlook provides ideas and hints on future research needs with respect to the dis-
sipative self-heating of elastomeric seismic bearings.

1 INTRODUCTION AND STATE OF THE ART

Elastomeric bearings are a widely spread component to insulate structures from vibrations,
either induced by traffic running over it or seismic hazards from below. Their damping cap-
abilities, aside from mass inertia mainly driven by the dissipated energy, protect structural
components from excessive excitation and even collapse. More and more material mixtures
are being used to cope with rising demands, e.g. higher damping requests (Okui et al. 2019).
Nevertheless, not even for the mainly used elastomeric bearing materials (natural rubber NR,
chloroprene rubber CR), the coupling of stress and strain induced mechanical response to the
temperature inside the bearing is sufficiently captured in the design process (Dippel et al.
2015). Common models to simulate the material response of elastomer bearings consider stiff-
ness values measured at 23°C and isothermal conditions (DIN EN 1337-3; DIN EN 1998-2,
DIN EN 15129). However, during service time seismic or cyclic loading introduces dissipative
self-heating inside the bearing. In order to prevent damage due to miscalibration caused by
the omission of temperature changes in the constitutive model of NR or CR rubber, it is
necessary to further investigate the time and temperature dependence of these materials.
In recent research, many examples can be found, that state a strong correlation between
temperature and mechanical deformation. According to Okui et al. (2019), the importance of
self-heating of High Damping Rubber Bearings (HDRB) under low ambient temperatures is
highlighted. Strong dependence of mechanical behavior on the material temperature is as well
stated in several more publications e.g. Dippel et al. (2015), Guo et al. (2018), Lion (1997),

1176
Johlitz et al. 2016, Köppel et al. 2018). Parameters influencing the dissipation of heat are
strain rate, frequency, maximum strain amplitude, geometry of test specimens (especially the
surface-volume-ratio), filler content, heat flux over boundaries and heat transfer to the envir-
onment, cf. Dippel et al. (2015), Johlitz et al. (2016) and Ovalle Rodas et al. (2014).
The focus of this paper is on the investigation of NR and CR with respect to their time-
temperature dependence, assuming them as thermorheologically simple materials with viscoelastic
behavior and dissipative heating potential. By using the DMTA, stiffness-damping-temperature
relations can be established from the isothermal measurements at different frequencies or respect-
ively times. From the obtained data, ‘Prony-Series’ with associated TTSPs are calibrated. Further
evaluations concentrate on the hysteretic response, i.e. the amount of energy transformed into
heat. The results of this paper on NR and CR are checked against recent findings of Köppel et al.
(2018) in order to gain insight on the variations due to different production batches of the
samples.

2 THEORETICAL BACKGROUND

The theory of viscoelasticity and thermorheology is well documented in literature, see e.g.
Marques & Creus (2012), Wrana (2014), Tschoegl (1989) for a more detailed introduction to the
time and temperature dependence of polymeric materials. Elastomer-specific consideration on
thermomechanics can be found in Findley & Davis (2013) and Schwarzl (1990). However, within
this paper a short introduction into the main concepts of thermomechanics and viscoelasticity is
given.

2.1 Thermorheology of polymers


Every polymer shows two characteristic stiffness areas: a brittle, glassy state at low temperat-
ures and a rubber-like behavior at high temperatures (cf. Figure 1). The area in between is
called glass-transition region, in which the stiffness decreases significantly (Schwarzl 1990),
and which is commonly characterized by the glass-transition temperature Tg Below that glass-
transition temperature Tg the material behavior is energy elastic (brittle) whereas above it is
entropy elastic (rubbery).
Elastomers are a specific form of polymer with a widely meshed structure. The low increase
in stiffness observable in the entropy-elastic range depends on the degree of chemical cross-
linking of molecules. Compared to the other types of polymers, thermoplastics and duroplas-
tics, classified according to polymeric structure and mechanical behavior, elastomers do not
melt but decompose at a certain temperature due to the cross-linking.
Within this paper, two elastomeric materials are investigated: NR and CR, each of Shore
Hardness 60. Block (2010) estimated the Tg of NR and CR to be Tg,NR ≈ - 60°C and Tg,CR ≈ -
37°C. Determining the Tg either by the phase shift, tanðδÞ, or with the turning point of the
storage modulus, Köppl et al. (2018) found ranges of Tg,NR = [- 57°C;-50°C] and Tg,CR =
[-35°C; 30°C].

Figure 1. Temperature dependent behavior of polymeric interlayers based on Schneider et al. (2016).

1177
Figure 2. Generalized Maxwell-Model.

2.2 Theory of linear viscoelasticity and the generalized Maxwell-Model


Linear viscoelasticity describes a time- and temperature-dependent material behavior in case
of small strains that is linear in time and stress-strain. Evaluation of the experimental data will
reveal if this assumption is justified.
To mechanically describe the correlation between stiffness and time, a Generalized Max-
well-Model (Figure 2) is used within this paper. The Generalized Maxwell Modell can be
described mathematically in the time and frequency domain by adding time dependent, decay-
ing parts to an equilibrium stiffness. The Prony Series associated to the Generalized Maxwell
Model in the time (Eq. 1) or frequency domain (Eq. 2) reads:
XK t
E ðtÞ ¼ E þ Eb e ~τk
k¼1 k
ð1Þ

XK ω2 τ2k XK ωτk
E  ðωÞ ¼ E þ Eb
k¼1 k
þi Ebk ð2Þ
1 þ ω τk
2 2 k¼1 1 þ ω2 τ2k

with: ~τk ¼ ηk =Ek

In the frequency domain the real part of the complex modulus is called Storage modulus and
the imaginary part is named Loss modulus. The nomenclature mirrors the relation of the
moduli to the energy stored and dissipated respectively within cyclic loading.

2.3 Time-Temperature-Superposition-Principle
The Prony-Series representing the Generalized-Maxwell Model as given in Eq. (2) until now
covers solely a time-dependency, but as discussed before, elastomeric material behavior is as
well strongly influenced by temperature. It was mentioned above, that the ‘Time-Temperature-
Superposition-Principle’ (TTSP) is based on the similarity of time and temperature dependence
of the stiffness modulus. In literature, different mathematical forms for a TTSP are documented
but omitted for reasons of brevity within this paper (Schwarzl 1990, Kraus & Niederwald
2017). The TTSP with one functional form can be used for thermorheologically simple materials
(Schwarzl 1990, Tschoegl 1989), which NR and CR will be considered as in the following.
Though measuring the stiffness modulus at different isothermal temperatures in a chosen
time or frequency range, a continuous Master Curve for a reference temperature can be
constructed from that short time measurements applying a shift factor to the frequency axis
(cf. Figure 3).
Introducing the TTSP into the Prony Series in the frequency domain mathematically reads:

XK ξ 2  τ2k XK ξ  τk
E  ðω; T Þ ¼ E∞ þ Eb
k¼1 k
þ i Ebk ð3Þ
1 þ ξ  τk
2 2 k¼1 1 þ ξ 2  τ2k

where ξ ¼ aT ðTjTref Þ  ω is defined as the reduced frequency.

1178
Figure 3. Structure of stiffness modulus in frequency domain.

The Prony parameters ek (or Ebk ), ~τk and E0 (or E) as well as the parameters of the TTSP
need to be determined from experimental results. In earlier times, the experimental data basis
for that was provided by carrying out isothermal relaxation or creep experiments over decades
of time, which were then repeated at different temperatures, cf. Sec. 3.1. Due to the enormous
time effort, the significantly less time-consuming approach using DMTA result data is favored
now (Schneider et al. 2016, Kraus & Niederwald 2017). Within this paper the raw data of
DMTA experiments are used to calibrate the parameters of TTSP and Prony-series according
to the identification method ‘GUSTL’ (Kraus & Niederwald 2017). This method calibrates
a Prony-Series with an associated TTSP using a fast, non-negative least squares algorithm.

2.4 Energy dissipated during cyclic loading


Every material that shows a hysteresis loop in its stress-strain curves is dissipating energy,
where the greatest part of this energy is transformed into heat.
Considering the uniaxial tension loading and a sinusoidal strain history within the con-
ducted DMTA tests, Bergström (2015) suggests Eq. (4) to compute the energy loss per unit
reference volume and load cycle by integrating the stress over one load cycle with respect to
strain. Solving that integral of Eq. (4) delivers Eq. (5):
ð 2π=ω
uloss ¼ σðtÞ_εdt ð4Þ
0

uloss ¼ π  σa  εa  sinðδÞ ¼ π  ε2a  E 00 ð5Þ

where σa ; εa = amplitude of oscillation; δ = angle of phase shift; E 00 = Loss modulus


Eq. (5) is used to assess the dissipated energy during DMTA experiments and the resulting
increase in temperature of the test specimen, cf. Sec 4.3.

3 EXPERIMENTAL SETUP

Within this section, at first the general experiment setup is outlined followed by a summary of
conducted tests and the obtained results.

3.1 Operational principle of the DMTA


Figure 4 shows the dynamic mechanical thermal analysis (DMTA) testing equipment
‘EPLEXOR 2000N’ from the manufacturer GABO available at the site of the University of
German Armed Forces Munich.
With this DMTA, the following principle operation modes can be conducted amongst
others:
• application of forced sinusoidal vibrations
• simultaneous measurement of force und displacement

1179
Figure 4. Fundamental assembly of a DMTA and entire test bay according to Halm.

• determination of the phase shift tanðδÞ between force and displacement


• Direct calculation of complex modulus split in real and imaginary part
• repeated measurements for different temperatures and frequencies

3.2 Conducted experiments


The DMTA tests are conducted at different temperatures, where at each temperature
a sinusoidal steady state oscillation is run for multiple frequencies. In that way, frequency and
temperature dependence of the polymer stiffness are measured simultaneously.
The samples are clamped in the DMTA with an initial clamp distance of 20 mm. Clamping the
elastomer specimen is tricky, as for the case of high strains slippage might occur during the testing.
Hence prior calibration and validation experiments have to be carried out, starting from small
strain amplitudes to large ones in order to possibly detect problems due to slip of the specimen.
According to Block (2010), the temperature range of interest to civil engineering applications
for elastomer bearings is -50°C to 70°C whereas DIN EN 1337-3 sets the temperature interval
for elastomeric bearings from -25°C to 50°C. Within this experimental study, the temperature
program for the DMTA tests is set to start at the highest test temperature with subsequent
decreasing of the temperature for the consecutive tests. A temperature step of 2°C is chosen to
ensure a sufficiently smooth Master Curve as well as a large overlap of the stiffness moduli,
which are obtained from frequency sweeps at different temperatures. A sufficiently fine sam-
pling of the temperature and frequency axis is necessary to determine the time-shift and thus
TTSP accurately. Testing polymers from warm (T > Tg) to cold circumvents physical ageing
effects. At each new temperature step the temperature equilibrating time is 3 min to avoid
a temperature gradient in the specimen. Table 1 summarizes the settings of the experiments:

Table 1. Test Parameters of the DMTA – Temperature-Frequency Sweeps.


Sample Bearing elastomer NR Bearing elastomer CR

Rectangle: Rectangle:
L × W × T: L × W × T:
Sample geometry [mm] 30 × 10 × 2 30 × 10 × 2
Test mode Temperature- Frequency-Sweeps in Tension
Temperature program [°C] +50: -80 +50: -30 +80: -50 +50: -30
Temperature steps [°C] 2 5 2 5
Frequency program [Hz] 1 to 20 1 to 20 1 to 20 1 to 20
Logarithmically equally spaced
6 10 6 10
frequencies [-]
Contact Force [N] 0,15 0,15
Static strain [%] 1.00 10.00 1.00 10.00
Dynamic strain [%] 0.01 0.10 7.00 0.01 0.10 7.00
Setting Number (1) (2) (3) (4) (5) (6)

1180
4 RESULTS AND DISCUSSION

4.1 Master Curves according to GUSTL


Having obtained the stiffness-temperature-frequency raw data with the settings as described in
Sec. 3 now Master Curves can be calibrated by using the method GUSTL, cf. Sec. 2.3.

4.2 Dissipated heat


Using the formulas by Bergström (2015) as presented in Sec. 2.4 together with the experimen-
tally obtained Loss Moduli for NR and CR for different strain amplitudes, temperatures and
frequencies, the potential energy for dissipation can be computed. For reasons of brevity,
within this paper only some graphs can be shown, Figures 6 and 7 show the loss energy for
the different experimental settings
Both materials show a peak energy loss, as can be expected, close to their Tg but at the
lowest tested frequency. At higher temperatures frequency dependent behavior changes to the
opposite, though, the amount of energy loss decreased significantly. Comparing the results of
setting (2) and (3) respectively (5) and (6), a remarkable shift of the maximum loss from below
30°C to about 15°C is recognized.
Considering the same settings (2), (3) and (5), (6) both an amplitude and temperature
dependent increase in the area enclosed by the hysteresis graph can be recognized, cf.
Figure 8. Whereas temperature rise lowers the energy loss, an increase in strain amplitude
supports energy dissipation. Another material property, the storage modulus, reacts in
a similar manner with respect to temperature and amplitude or rather strain rate. Lower
temperatures as well as higher amplitudes steepen the inclination of the ellipsis as more
Maxwell elements in the Prony-series are active, which is more pronounced for the case of
CR compared to NR for the small strain experiments. In general CR reacts more sensitive
with respect to temperature changes, which may be caused by the proximity of the investi-
gated temperature range to the Tg of CR. While for the small strain experiments (2) and
(5) ellipses can be recognized in Figure 8, for the cases (3) and (6) mild nonlinearities can
be seen in the hysteresis plots. This gives rise to investigations to a greater degree in that
direction with even higher strain amplitudes in order to assess the further evolution of the
nonlinearities in that region.

4.3 Approximate computation of increase in temperature


In order to determine the scale of the increase in temperature in the center of an elastomeric
bearing due to dissipation, the temperature rise of the specimen is approximately computed by
Eq. (8) under neglection of further heat fluxes to the outside. For any material the inner
energy dU and the volume work dQ relate according to:

Figure 5. Master Curve of the Storage Modulus regarding setting (1), left and (4), right.

1181
Figure 6. Calculated energy loss (uloss) setting (1), (2) and (3) for natural rubber.

Figure 7. Calculated energy loss (uloss) for the settings (4), (5) and (6) CR.

Figure 8. Exemplary hysteretic behavior for settings (2), (3), (5) and (6).

1182
Table 2. Specific heat capacity and mass of specimens.
Sample Bearing elastomer NR Bearing elastomer CR
Setting Number (2) (3) (5) (6)

Mass [g] 0.845 0.833 0.913 0.916


Specific heat capacity [J/gK] 1.91-2.08 2.20

dQp ¼ dU þ dQ ¼ dH ð6Þ

This enthalpy change requires an amount of heat being supplied:


ð
Qp ¼ m  cp ðT ÞdT ð7Þ

Without heat transport over the surface and neglecting the volume work as well as the tem-
perature dependence of the specific heat capacity cp the increase in temperature of a specimen
is calculated accordingly to:

uloss
ΔT ¼ with Qp ¼ cp  m  ΔT ¼ uloss ð8Þ
cp  m

Within this paper, only exerpts of all conducted computations can be shown for reasons of
brevity. For a specimen of setting (3) with 7% dynamic strain, the energy loss within one load
cycle theoretically generates a rise of about 0.015°C for ambient temperatures below -15°C
(Figure 9) and even less for higher temperatures. Since the energy loss calculates per volume,
it is possible to refer the temperature rise to e.g. liter (Figure 9) which is easier to compare to
the size of a real bearing.

5 CONCLUSION AND OUTLOOK

In the present paper, Prony Series’ parameters and associated TTSP for small and large strains
are determined and discussed. Considering the low temperature rise due to dissipative heating
in the specimen used in DMTA analysis, there is no dissipation induced error in these param-
eters. They will be incorporated into Finite-Element-Software to numerically investigate the
experimental findings or else serve as validation base to other authors. Further analyzing the
hysteretic behavior, the dissipated energy maximum regarding temperature scale is dependent

Figure 9. Approximated temperature rise of specimen setting (3) and the temperature rise per liter cal-
culated from the setting (3).

1183
on frequency and strain amplitude. Dissipative heating should not be neglected in the design
process regarding the increase in damping at higher frequencies as well as the temperature rise
per volume. As could be shown, the temperature rise inside the bearing can be approximated
through
Prony Series’ parameters based on DMTA experiment data making use of the TTSP and
the assumption that no energy is lost at the surface.
In order to be able to include dissipative heating in simulations of elastomeric bearings with
high accuracy, several further investigation steps are recommended. First it is suggested to
assess dissipative heating by a thermocouple inside the specimen. Then a possible increase in
nonlinearity at higher strain rates compared to the experiments shown but within the limit of
applications, needs to be checked. In addition, the influence of the chemical composition of
the elastomers should be further investigated. Finally, in order to verify the simulated dissipa-
tive heating in an elastomeric bearing, tests on an instrumented real product are planned.

REFERENCES

Bergström, J. 2015. Mechanics of Solid Polymers- Theory and Computational Modeling. PDL Handbook
Series. USA: FluoroConsultants.
Block, T. 2010. Verdrehwiderstände bewehrter Elastomerlager. Dissertation. Ruhr-Universität Bochum.
Dippel, B. & Johlitz, M. & Lion, A. 2015. Thermo-mechanical couplings in elastomers - experiments and
modelling. ZAMM Zeitschrift für Angewandte Mathematik und Mechanik 95 (11): 1117–1128.
Findley, W.N. & Davis, F.A. (revised ed.) 2013. Creep and Relaxation of Nonlinear Viscoelastic Mater-
ials. New York: Courier Corporation.
Guo, Q. & Zaïri, F. & Ovalle Rodas, C. & Guo, X. 2018. Constitutive modeling of the cyclic dissipation
in thin and thick rubber specimens. ZAMM - Journal of Applied Mathematics and Mechanics/Zeits-
chrift für Angewandte Mathematik und Mechanik (series number if necessary) 98 (10): 1878–1899.
Halm, H. & Deckmann, H. Dynamic-Mechanical Thermal Analysis of Polymers and Solids.
Johlitz, M. & Dippel, B. & Lion, A. 2016. Dissipative heating of elastomers: a new modelling approach based
on finite and coupled thermomechanics. Continuum Mechanics and Thermodynamics 28 (4): 1111–1125.
Köppl, J. & Kraus, M.A. & Mangerig, I. 2018 Thermorheological Testing and Modeling of a Bridge
Slide-Bearing Elastomer. Proceedings of the 12th Japanese German Bridge Building Symposium.
Kraus, M.A. & Niederwald, M. 2017. Generalized collocation method using Stiffness matrices in the con-
text of the Theory of Linear viscoelasticity (GUSTL). Tech. Mech. (37) 1: 82–106.
Lion, A. 1997. A physically based method to represent the thermo-mechanical behaviour of elastomers.
Acta Mechanica 123 (1–4): 1–25.
Marques, S. & Creus, G. 2012. Computational Viscoelasticity. Berlin: Springer.
Okui, Y. & Nakamura, K. & Sato, T. & Imai, T. 2019. Seismic response of isolated bridge with high
damping rubber bearings: self-heating effect under subzero temperatures. Steel Construction 12: 2–9.
Ovalle Rodas, C. & Zaïri, F. & Naït-Abdelaziz, M. 2014. A finite strain thermo-viscoelastic constitutive
model to describe the self-heating in elastomeric materials during low-cycle fatigue. Journal of the
Mechanics and Physics of Solids 64 (1): 396–410.
Schneider, J. & Kuntsche, J. & Schula, S. & Schneider, F. & Wörner, J.-D. (vol. 2) 2016. Glasbau – Grun-
dlagen, Berechnung, Konstruktion. Springer.
Schwarzl, P.D.F.R. 1990. Polymermechanik. Berlin, Heidelberg, New York: Springer.
Tschoegl, N. 1989. The Phenomenological Theory of Linear Viscoelastic Behavior - an Introduction. Berlin:
Springer.
Wrana, C. 2014. Polymerphysik. Springer Spektrum.

1184
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Built-up cold-formed steel beams with web openings

V. Ungureanu
Politehnica University of Timisoara, Timisoara, Romania
Timişoara Branch, Romanian Academy, Romania

I. Both, C. Neagu & M. Burca


Politehnica University of Timisoara, Timisoara, Romania

D. Dubina
Politehnica University of Timisoara, Timisoara, Romania
Timişoara Branch, Romanian Academy, Romania

A.A. Cristian
Technical University of Civil Engineering of Bucharest, Bucharest, Romania

ABSTRACT: Cold-formed steel profiles are lightweight elements which can be assembled in
a numerous variety of shapes considering either truss structures or corrugated web beams. Espe-
cially in residential or office buildings, adjustments are required for the service installations. The
web openings represent a weak point in a beam and special attention must be considered to main-
tain the initial capacity. Previous experimental tests were performed on built-up beams with lipped
channel sections as flanges and trapezoidal corrugated steel sheets as web. The connection of the
beam components was performed by two methods, i.e. resistance spot welding and MIG brazing.
The paper presents the experimental investigations on two full-scale beams with different strength-
ening solutions for the web openings, in function of the welding technique, i.e. a reinforcing steel
plate was spot welded to the corrugation of the web and a border type frame was MIG brazed on
the opening perimeter. A lesser influence of the web opening was observed for the beam connected
by MIG brazing, a superior bearing capacity being obtained.

1 INTRODUCTION

The results of a previous experimental investigations on built-up corrugated web beams


(CWB), entirely made of cold-formed elements connected by resistance spot welding or by
MIG brazing (Ungureanu et al., 2018a, 2018b), suggested the possibility of using these solu-
tion at a larger scale having a high bearing capacity and a ductile response.
Due to their strength-to-weight ratio, the corrugated web beams represent an attractive
solution for residential/office buildings, but the building services sometimes interfere with the
structural elements requiring adjustments of the elements. Web openings are a common solu-
tion for multi-storey steel structures. The castellated beams are known for their high material
saving and the possibility for large spans as well as trusses. The solution of using cold-formed
steel elements for trusses has the disadvantage of part joining which represents a time con-
sumer. An automated process may be used in the case of welded elements, but light gauge
steel elements are difficult to be welded. Still, the advances in the automotive industry allow
the connection of thin steel plates either by resistance spot weld (SW) or by MIG brazing.
The sensitivity of thin-walled cold-formed steel elements and structures to imperfections is very
well known (Blum & Rasmussen, 2018; Cardoso et al., 2019; Dubina, 2008) and the web opening
in a CW beam certainly reduces, even more, its bearing capacity.

1185
Two solutions for reinforcing the web opening of the corrugated web beam are investigated in
the paper, in function of the connecting technique. For the case of the resistance spot welding,
a plane steel sheet positioned in the web plane can be conveniently welded while for the case of
MIG brazing, a plate perpendicular to the web plane is more suitable to be used. Based on the
experimental results, the behaviour of these beams is presented by their bearing capacity and
deformation, as well as by the evolution of the components instabilities.

2 EXPERIMENTAL PROGRAM

In a previous experimental program, 5 built-up corrugated web beams (2 using resistance spot
welding and 3 using MIG brazing) were tested with various distributions of the web panels
function of the sheet thickness (Ungureanu et al., 2018a, 2018b). The current investigations
highlight the effect of a web opening in two CWB built-up beams, one using the spot welding
(CWB-SW) and the second one using MIG brazing (CWB-CMT).

2.1 The test setup


The built-up beams were tested in a planar rigid frame with both ends fixed to the frame.
A 500 kN actuator loaded the beam through a leverage system that uniformly distributed the
load in 4 points in order to simulate a uniform distributed load. Restrictions for the out-of-
plane displacements were applied by a separate structure in two locations. Figure 1 presents
the setup of the specimens’ 6 point bending test.
In order to simulate the quasi-static loading regime, a rate of 2 mm/s was applied by the
actuator. The force was recorded by the actuator load cell, while the vertical displacements
were monitored at each quarter of the span by wire linear transducers (see Figure 2a). The
relative displacement between the flanges and the end-plates and, the deformation of the end-
plate was recorded by linear displacement transducers, as shown in Figure 2b.

2.2 The specimens


Compared to the previous built-up CWB without web openings, for which the web consisted of
individual corrugated panels of approximately 1 m, the web of the current specimens was made of
a single piece, i.e. a continuous web. Thus, the built-up of the beams consisted in three stages: 1)
connecting the shear panels to the corrugated web, 2) connecting the flanges to the web and 3)
connecting the supporting device to the beam. The final operation consisted of machining the web
opening. A gauge nibbler was used to cut the perimeter of the web opening after performing 4
holes in the corners of the web opening. Although the connection between the parts of the beam
was made by two different welding techniques, the manufacturing process is presented in Figure 3
for the spot-welded specimen only. Having the same base material as for the built-up beams

Figure 1. Test setup of the built-up beams.

1186
Figure 2. Monitoring displacement and deformations: a) vertical displacement, b) beam ends.

Figure 3. Stages of the built-up process.

without web opening, the mechanical properties are consistent with the values determined in
(Ungureanu et al., 2018a) for the corresponding steel sheet thicknesses. The 1.0 mm sheet for the
corrugated web is classified as S250GD+Z while the 1.2 mm and the 2.0 mm thicknesses for the
shear panel and for the flanges respectively have a yielding characterised by S350GD+Z steel.
The beams consisted of the following components: (1) corrugated steel sheet for the web -
1.0 mm; (2) additional shear panels - flat plates of 1.2 mm; (3) two back-to-back lipped channel
sections for flanges - 2 × C120/2.0; (4) reinforcing profiles U150/2.0 used under the load applica-
tion points; (5) bolts M12 grade 8.8 for flange to end plates connections, as presented in Figure 4.
It must be mentioned that the corrugation height of the SW beam was 60 mm, while the
corrugation height in the case of the MIG brazed beam was 45 mm.
The web opening dimension considered a reasonable height for the service installations while
the length was limited by the distance between the corrugations, such that an optimised position
of the spot welding to be applied, i.e. a minimum distance to the edge of the corrugation to be
assured. In the case of MIG brazed specimen, the length of the opening is not limited by
a minimum distance to the edge of the corrugation, but it was chosen for the similarity to the SW
beam specimen. The position of the web opening was chosen to avoid the maximum bending
moment and shear force, close to the supports or the middle of the beam, respectively.
Similar welding techniques used for the built-up beam were considered also for the reinforcing
solution of the web opening. In order to apply the welding, different configurations were selected.
For the beam built-up by spot welding, a 2.0 mm thick flat steel plate was welded on the contour
of the opening, with the dimensions given in Figure 5. Only one reinforcing plate was used due to
the limited possibility of welding another plate on the opposite side of the corrugation. The
reinforcing plate was also bent to 90° on both sides parallel to the flanges. These lips were also

Figure 4. Components of CWB specimens (all dimensions in mm).

1187
Figure 5. Reinforcing the web opening of the CWB-SW-WO specimen.

Figure 6. CWB-SW with web opening specimen.

spot welded to the flanges. The undeformed shape of the built-up beam in the experimental stand
is presented in Figure 6.
A convenient configuration for the reinforcing of the web opening was performed for the CWB
specimen built-up using MIG brazing. A 1.2 mm steel plate was bent in order to take the shape of
the web opening. Since the corrugation height of the web was 45 mm, a wider plate was necessary
to allow the MIG brazing. Also, to facilitate the insertion, the reinforcing plate was conceived by
two U shaped pieces of 80 mm width. For the side parallel to the flanges, the brazing was per-
formed alternatively on each corrugation, while for the vertical sides the brazing was applied as
intermittent segments Figure 7 presents the solution for the web opening reinforcing. The full-
scale beam specimen of the built-up beam connected by MIG brazing is presented in Figure 8.

Figure 7. CWB-CMT with web opening specimen.

Figure 8. CWB-CMT with web opening specimen.

1188
2.3 Results
The response of the beam is assessed not only by the bearing capacity but also the failure
mechanism that conducted to the collapse. Starting with the general view, Figure 9 presents
the deformed shape of the beam specimen built up by spot welding. A significant deformation
is noticed on the left side of the web opening.
On single parts, the deformation and failure observation were recorded in the following
order: (1) shear buckling of the shear panels, (2) deformations of the corrugated web in the
corner of the opening, (3) distortions of the web corrugations close to the end of the beam, (4)
shear buckling of the corrugations, (5) shear buckling of the corrugated web sheet (connecting
the shear buckling of the corrugations), 6) failure of the spot welds after increase of the previ-
ous deformations and (7) buckling of the flanges under the load application point. The first
four instabilities occurred in the first part of the capacity degradation, as shown in Figure 10,
while (5), (6) and (7) were noticed after large deflection of the beam with plastic deformations
(see Figure 11).
The force-displacement curve is compared in Figure 12 with the curves obtained from the CWB
without web opening (Ungureanu et al., 2018a). It must be mentioned that, although the general
configuration is similar, in the previous tests the web was composed of corrugated web plates with
approximately 1 m, with thicknesses varying along the beam. Therefore, a direct specification of
the force reduction is incorrect. Qualitatively, it can be observed the tested beam has a stiffness of
12106 N/mm, of the same order as the previously tested one. On the other hand, a smaller capacity
was expected compared to the previously tested beams, where the thickness of the corrugated web
panels near support was higher, i.e. 1.2 mm. The second beam specimen, built-up by using MIG
brazing, presented also an increased deflection in the web opening side but only in the last stage of
the testing, see Figure 13.

Figure 9. Deformed shape of the CWB-SW beam with web opening.

Figure 10. Instabilities of the CWB-SW beam during plastic response initiation.

Figure 11. Deformations of the CWB-SW beam with plastic deformations.

1189
Figure 12. Force-displacement curve for the specimens using spot welding.

Figure 13. Deformed shape of CWB-CMT beam with web opening.

However, during testing, the instabilities and local failures were very symmetric as presented in
a sequence of intermediate deformed shapes of the beam depicted in Figure 14. The reinforcing of
the web opening together with the MIG brazing lead to rigid components of the corrugated beam,
above and below the web opening.
The degradation of the elastic response was initiated by a limited shear buckling of the shear
panels, less obvious than the shear buckling of the corrugation of the web. Ultimately, the increase
of displacement led to linking the individual buckling into the shear buckling of the corrugated
web (see Figure 15). During the expansions of the shear buckling of the web to the web opening,
the buckling of the shear plate and buckling of the flanges were observed, as shown in Figure 16.

Figure 14. Evolution of the deformed shapes of the CWB-CMT specimen.

1190
Figure 15. Instabilities of the CWB-CMT beam during plastic response initiation.

Figure 16. Deformations of the CWB-CMT beam in the plastic stage.

Compared to the results obtained for the built-up CWB beams without web openings
(Ungureanu et al., 2018b), the force-displacement curve of the current test suggests
a minimum effect of the web opening on the bearing capacity. Both the initial stiffness, 27746
N/mm, and the ductility of the beam is comparable to the previously tested beams, as pre-
sented in Figure 17. Similar to the case of the spot welded beam, the two cases (specimens
with and without web opening) were not identical, but the bearing capacity of the CWB beam
with web opening is situated between the specimen CMT 1 having both the corrugated web
and shear panels thicknesses of 1.2 mm, and specimen CMT 2 with the steel sheets thickness
of the corrugated web and shear panel of 1.0 mm.
Since the currently tested specimens were identical from the configuration and components’
thicknesses point of view, except the connecting techniques and the reinforcing solution, the
responses of the two beams can be compared. A higher rigidity can be seen in the MIG brazed
specimen, along with an increased ductility (see Figure 18). The increased rigidity of the MIG
brazed specimen is due to the restraint of the corrugations against distortions. The spot-
welded specimen allows distortion of the corrugation, the initial shape of the corrugation
being deformed in early stages of the loading, thus reducing the initial rigidity.

Figure 17. Force-displacement curve for the MIG brazed specimens.

1191
Figure 18. Comparison of CWB with web openings.

3 CONCLUSION

The use of the web openings in corrugated web beams made of cold-formed steel elements is pos-
sible only by strengthening the affected area. Using a similar joining technique for the strengthen-
ing as the ones used in the fabrication of the beam, an adequate bearing capacity was obtained.
In the case of spot-welding technology, a steel plate parallel to the web plane can be used as
reinforcement, surrounding the web opening. Combined with small rigidity of the corruga-
tions due to the discrete connection of the spot weld, only 61% of the bearing capacity of the
MIG brazed specimen was reached. The less stable web of the SW beam, in the opening area,
leads to a weak point which constitutes the main deflection source of the beam.
Connecting the corrugated web to the flanges by MIG brazing increases the rigidity of the
beam as well as the rigidity of the corrugations. Together with the border type reinforcement
of the web opening, the remaining of the web in the opening area can transmit the shear force
without large deformations leading to a bearing capacity of similar magnitude as the previ-
ously tested specimen without web opening.
In terms of ductility, the MIG brazed specimen allows a higher deflection of the beam, while the
SW specimen failure modes concentrate the damage in the web opening area, reaching the collapse
at smaller deflection.

ACKNOWLEDGEMENT

This work was supported by a grant of the Romanian Ministry of Research and Innovation, pro-
ject number 10PFE/16.10.2018, PERFORM-TECH-UPT - The increasing of the institutional per-
formance of the Politechnica University of Timișoara by strengthening the research, development
and technological transfer capacity in the field of “Energy, Environment and Climate Change”,
within Program 1, Subprogram 1.2.

REFERENCES

Blum, H.B. & Rasmussen, K.J.R. 2018. Experimental investigation of long-span cold-formed steel
double channel portal frames. Journal of Constructional Steel Research 155: 316–330.
Cardoso, F.S., Zhang, H., Rasmussen, K.J.R. & Yan, S. 2019. Reliability calibrations for the design of
cold-formed steel portal frames by advanced analysis. Engineering Structures 182: 164–171.
Dubina, D. 2008. Structural analysis and design assisted by testing of cold-formed steel structures. Thin-
Walled Structures 46: 741–764.
Ungureanu, V., Both, I., Burca, M., Grosan, M., Neagu, C. & Dubina, D. 2018a. Built-up cold-formed
steel beams using resistance spot welding: experimental investigations. Proc. of the Eight International
Conference on Thin-Walled Structures - ICTWS 2018, Lisbon 24-27 July 2018 (e-Proceedings).
Ungureanu, V., Both, I., Tunea, D., Grosan, M., Neagu, C., Georgescu, M. & Dubina, D. (2018b). Experi-
mental investigations on built-up cold-formed steel beams using MIG brazing. Proc. of the Eight Inter-
national Conference on Thin-Walled Structures - ICTWS 2018, Lisbon 24–27 July 2018 (e-Proceedings).

1192
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Numerical investigation of built-up cold-formed steel beams with


corrugated web

V. Ungureanu
Politehnica University of Timisoara, Timisoara, Romania
Timişoara Branch, Romanian Academy, Timisoara, Romania

I. Lukačević
University of Zagreb, Zagreb, Croatia

I. Both & M. Burca


Politehnica University of Timisoara, Timisoara, Romania

D. Dubina
Politehnica University of Timisoara, Timisoara, Romania
Timişoara Branch, Romanian Academy, Timisoara, Romania

ABSTRACT: Built-up corrugated web beams (CWB) represent an assembly of multiple cold-
formed steel components of various thicknesses connected by means of screws or welding.
Recently, tests on such built-up beams have been performed within the CEMSIG Research Center
of the Politehnica University of Timisoara, in which the connections between the components
were made by spot welding. Following the validation of the numerical model, the paper investi-
gates the influence of several parameters, i.e.: the thickness of the flanges, the thickness of the cor-
rugated web, the thickness of the shear panel, the magnitude of the initial imperfections and the
number and position of the spot welds. The parametric study was conducted on a beam with the
same global dimensions as the tested one.

1 INTRODUCTION

Built-up cold-formed steel elements are efficient structural elements, very attractive due to the
material savings, but also for ease of construction. Moreover, such built-up beams add another
advantage to the list, namely the ease of handling due to the low weight of the components. The
connection between the built-up beam components can be easily obtained by screws, but the devel-
opments in the welding process also led to other solutions like spot welding.
A new technological solution of such a built-up beam was proposed at the Department of the
Steel Structures and Structural Mechanics of the Politehnica University of Timisoara, consisting
mainly of the lipped channel profiles as flanges, corrugated sheets for the web and flat shear
panels at both ends (Dubina et al., 2013). The parts were connected using self-drilling screws,
a common practice for creating built-up elements. An improvement for the initial solution is repre-
sented by the use of spot welding as a connecting technique, which eliminates the screws and
reduces the manpower.
Previous numerical studies, considering appropriate imperfections, material parameters, mesh
size and element type, have shown that finite element (FE) models can be used to accurately pre-
dict the load carrying capacity and post-buckling behaviour of built-up cold-formed steel beams
(Dubina et al., 2013; Ungureanu et al., 2018a). Different consideration of the initial imperfection
on the numerical analyses on corrugated web beams are presented in the literature leading to
more or less accurate results (Elgaaly & Seshadri, 1998; Gil et al., 2005; Nie et al., 2013).
The paper presents a calibrated numerical model based on the experimental results and the
influence of several parameters i.e.: (1) the initial imperfections, (2) the number and distance

1193
between spot welds on flanges, (3) the thickness of the flanges, (4) the thickness of the corrugated
web and (5) the thickness of the shear panel. From the parametric study, it results that the bearing
capacity of the corrugated web beams made of cold-formed steel components is highly affected by
the stability of the components and less affected by the configuration and the number of spot
welding. Especially in terms of rigidity, improvements were obtained if the spot welds are posi-
tioned in a stabilizing array.

2 EXPERIMENTAL DATABASE

The experimental results used for the validation of the numerical model were considered from
the tests performed at the Politehnica University of Timisoara (Ungureanu et al., 2018b)
where two CWBs were tested, having different thicknesses of the shear panel and a different
configuration of the spot welding fastening of the corrugated web panels.
A 6-point bending test configuration was used in order to simulate a uniform distributed
load in a planar rigid frame with both ends fixed to the frame. A 500 kN actuator loaded the
beams through a leverage system that evenly distribute the load in 4 points. A separate struc-
ture constrained the beam to vertical displacements in two positions.
The database of the testing includes the load force recorded by the actuator and the vertical
displacements at each quarter of the span monitored by LVDTs.
The built-up corrugated web beams have a span of 5157 mm, a beam height of 600 mm and
a corrugation height of 60 mm. The beams consisted of the following components: (1) corrugated
steel sheet for the web, with 0.8 mm at the mid-span and 1.2 mm at the beam ends; (2) additional
shear panels, i.e. flat plates of 1.0 mm (CWB SW-1) and of 1.2 mm (CWB SW-2); (3) two back-to-
back lipped channel sections for flanges, 2 × C120/2.0; (4) reinforcing profiles U150/2.0 used
under the load application points and (5) bolts M12 grade 8.8 for flange to end plates connections,
as shown in Figure 1.
The mechanical properties of base material have been tested and presented by Ungureanu
et al. (2018a). The 1.0 mm sheet for the corrugated web is classified as S250GD+Z, while the
1.2 mm and the 2.0 mm thicknesses for the shear panel and for the flanges respectively are
classified as S350GD+Z steel.
Figure 2(a) presents the force-displacement curves of the two tested beams, while Figure 2
(b) shows the qualitative deformations of the beam. The deformations include the buckling of
the shear panels, distortion of the web corrugations and failure of the spot welds.

3 NUMERICAL ANALYSIS

3.1 Validation
The general-purpose finite element program ABAQUS/CAE v.6.14 (Dassault Systemes,
2014), was used to carry out geometric and material non-linear analyses including the effects

Figure 1. Components of CWB specimens.

1194
Figure 2. (a) Force-displacement curves of the beams, (b) Failure modes during the experiment.

of initial imperfections (GMNIA). Finite Element (FE) models have been calibrated in
accordance with the experimental results based on the characteristics determined experimen-
tally such as material and lap joint specimen tests. Initial geometric imperfections, mesh size
and element type have been calibrated in accordance with the experimental results of tested
beams. Each part of the built-up beam was defined as a 3D shell element extruded according
to the shape of the part. Rectangular 4-node doubly curved thin or thick shell, reduced inte-
gration, hourglass control, finite membrane strains (S4R) were used to model the thin-walled
components. The global mesh size of 15 mm was used for the web, flanges and shear panels,
and 25 mm was used for the reinforcing profiles under the load application, as shown in
Figure 3. The mesh size was further reduced around the bolt holes to connect the shear panels
to the endplates.
General contact with the following parameters was used: normal direction - Hard Contact,
transversal direction - a friction coefficient of 0.1 and separation was allowed after the general
contact takes place.
While the support conditions were defined on the nodes of the holes provided for the
bolts that connect the beam to the end plate assembly as null displacements and rotations,
the loading of the beam was defined as a vertical displacement in a set of multipoint con-
straints MPC that forms a leverage system able to transmit the deflection to the 4 loading
points (see Figure 3). The link between the control points and the pressure surface was
defined by a Kinematic coupling constraint for all DOFs. RB3D2 elements were used as
a rigid body for load transfer and multi-point constraint beam (MPC beam) for DOF coup-
ling between groups of specified nodes.
The spot welds between different parts of the built-up beams were defined depending on the
tensile-shear tests of the simple specimens as follows. Attachment points were defined on each
part where SW was applied. The connection between the attachment points was defined using
Point Based Fasteners with the Connector response of the SW initially calibrated from the
results of the tensile-shear tests. The connector was considered by the Elasticity, Plasticity,
Damage and Failure parameters. Bushing connector elements were used to model the spot
welds. This type of the elements provides a connection between two nodes that allows inde-
pendent behaviour in three local Cartesian directions that follow the system at both nodes

Figure 3. Thicknesses of the CWB SW-1 beam by components: blue 2.0 mm (flanges), grey 1.0 mm
(shear panels), red 1.2 mm (webs near supports) and green 0.8 mm (mid-span webs).

1195
a and b and that allows different behaviour in two flexural rotations and one torsional rota-
tion (Dassault Systemes, 2014).
In order to obtain realistic results from the finite element nonlinear analyses, plastic
strains were included in the material definition, according to Annex C of EN1993-1-5
(2006). The measured stress-strain curves based on tensile tests on coupons cut from
the cross-section of component elements were included in the model. In the plastic ana-
lysis, the static engineering stress-strain curves obtained from tensile coupon tests were
converted to true stress vs logarithmic true plastic strain curves. The true stress and
true plastic strain were calculated using Equations (1) and (2) as follows:
 
σtrue ¼ σengineering 1 þ εengineering ð1Þ
 
εtrue ¼ ln 1 þ εengineering ð2Þ

where σtrue = true stress; σengineering = engineering stress; εtrue = true strain; and εengineering =
engineering strain.
For large deformations, the stress-strain points past yield were input in the form of true
stress and logarithmic plastic strain. The logarithmic plastic strain has been calculated with
Equation (3) as follows:

σtrue
εplastic
ln ¼ εtrue  ð3Þ
E

where εlnplastic = logarithmic plastic strain, σtrue/E is the elastic strain and E is Young’s
modulus.
The numerical model consists of two steps. In the first step, the initial imperfections are
modelled by performing static nonlinear analysis with the target displacements which results
in desired imperfection. In the second step the dynamic, explicit analysis is used to run the
load-displacement analysis of the beam based on geometry from previous static analysis for
imperfections and with all contacts and material nonlinearity included.
The calibration of the numerical models is affected by the magnitude of the initial
imperfections. The influence of the initial imperfections is investigated with three out of
web global imperfections with magnitudes of L/500, L/1000 and L/1500 (L - the span of
the beam) and local imperfections with a magnitude of approximately equal the sheet
thickness t.
Figure 4(a) compares the FEM and experimental load-displacement curves for CWB SW-1
beam, whereas Figure 4(b) compares the FEM and experimental load-displacement curves for
CWB SW-2 beam.
As shown in Figure 4, the bearing capacity of the beam is limited affected by the imperfec-
tion and, according to Nie et al. (2013), it is concluded that the shear buckling strength is
smaller than the shear yield strength.
Comparing the load-displacement curves of the experiments and the numerical analyses
a good correlation is observed.
For an initial imperfection of magnitude L/1000, the deformed shape of the numerical
model, Figure 5, replicates the phenomena encountered during the experiments i.e. shear
panel buckling, distortion of the corrugated web and the local buckling of the flange in the
load application points.

3.2 Parametric study


Considering the same beam dimensions, a parametric study was performed to investi-
gate the influence of the following parameters: the number, the position and the dis-
tance between the spot welds on flanges, the thickness of the flanges, the thickness of
the corrugated web and the shear panel.

1196
Figure 4. Experimental vs. FEM load-displacement curves: (a) CWB SW-1, (b) CWB-SW-2.

Figure 5. Qualitative deformation of the specimen.

3.2.1 Influence of the number of the spot welds


In order to investigate different arrangements of the SW, the model was modified to simulate the
effect of three spot welds per corrugation (instead of 2 as in case of experimental tests). Addition-
ally, the curve CWB SW-1 - FEM - SF in Figure 6, represents a reduced distribution of spot weld-
ing in the mid-span, according to the shear force distribution, i.e. has been investigated with 2SW
at every two corrugations in this area. As shown in Figure 6, the reduced number of SW in
the second third of the beam results in a decreased capacity and stiffness.
Without a significant influence on the capacity, the number of SW in the midspan can be
reduced at the cost of losing the initial rigidity. The results from Figure 6 show a small influ-
ence of 3SW in comparison with 2SW.
As observed during the experimental tests, one of the instabilities that occurred during load-
ing was the distortion of the corrugation. The phenomenon occurred since the discrete connec-
tion of the spot weld has an axis parallel to the corrugation direction. A more rigid connection
is achieved if the spot welds axis is horizontal, constraining the corrugation against distortion.
Figure 6 shows the increase in rigidity offered by this configuration of the spot welds. It is
to be mentioned that the capacity is less than in the initial case. In order to achieve a capacity
of similar magnitude as the initial case, four spot welds on two rows (horizontal spacing of
40 mm, vertical spacing 50 mm) can be assigned, leading to increased rigidity of the element.

1197
Figure 6. The influence of the number of spot welds between the flange and the web.

3.2.2 Influence of the distance between spot welds on the flanges


Influence of the distance between SW on flanges has been investigated for the beam CWB
SW-1, by changing the distance between SW on flanges according to Figure 7.
Although this is unrealistic to be done in practice for this particular case, the authors
wanted to see the influence of a larger distance between spot weldings. Consequently, the dis-
tance increases from 50 mm used in the experiment to 100 mm just to emphasise if such influ-
ence exists. The curves in Figure 7 shows a negligible influence.

3.2.3 Influence of the flange thickness


By considering the flange thicknesses of 1.2, 1.5, 2.0 (experiment) and 2.5 mm, the influence of the
flange thickness was assessed. A relatively small reduction in the thickness of the flange profile
(from 2.0 to 1.5 mm) causes a significant reduction in the resistance of the entire system. The rela-
tive increase of the flange sheet thickness compared to the reference case CWB SW-1 (from 2.0 to
2.5 mm) results in a small increase of the resistance as depicted in Figure 8.
It must be mentioned that the analysis shows that the flange thickness has a high impact also in
the initial rigidity of the beam.

3.2.4 Influence of corrugated web thickness


As the CWB can be formed of a single corrugated web, another case of analysis was set by the
fixed thickness of the flanges, 2.0 mm, and the shear panels, 1.2 mm, while the corrugated web
varies between 0.8, 1.0, 1.2 and 1.5 mm, having one thickness for the entire length of the
beam. The results are shown in Figure 9(a).
Although for small thicknesses, i.e. 0.8 mm, both the rigidity and the yielding are dependent on
the web thickness, for the thickness of the web above 1.0 mm the yielding is very similar but the
ultimate force is different. For the 1.5 mm corrugated web, the force-displacement curve shows
a continuous increasing force in the post-elastic stage, similar to the catenary effect. According to
Nie et al. (2013), this thickness can provide a shear buckling strength larger than the shear yield
strength.

3.2.5 Influence of the shear panel thickness


As observed from the experimental investigations, the thickness of the shear panel is
a parameter that may influence the rigidity of the specimen. In this direction, the analyses
were performed also for the case where the flange and the corrugated web are kept at the
same thickness as for the CWB-SW-1 specimen, while the shear panel thickness took the
values of 0.8 mm, 1.2 mm, and 1.5 mm.
The parametric numerical analyses showed that the influence in the rigidity is not signifi-
cant, but the resistance is affected by approximately 15%, (see Figure 9(b)). Over the entire
range of thicknesses, the influence of the shear panel thickness is smaller than in the case of
the web thickness from the previous case.

1198
Figure 7. The influence of the distance between spot weldings on flanges.

Figure 8. The enhancement of the flange thickness.

Figure 9. a) The influence of corrugated web thickness, b) The influence of the shear panel thickness.

4 CONCLUSIONS

Based on experimental results, the lightweight steel structures represent a sustainable solution in
structural engineering due to their material saving and ease of manipulation. For the built-up

1199
cold-formed elements with corrugated webs, the number of parameters is very large due to the
multiple parts involved and the number of possible configurations of the beam. Among the
parameters that may influence the response of the beam are the position of the spot welding, the
flange thickness, the web thickness and the shear panel thickness, all investigated in the present
paper. Their contribution influences the capacity and the rigidity at different shares.
The studied parameters which have a small influence on the results are the number of spot
welds on the flange and the distance between the spot welds.
From the numerical analyses of the studied beam, it results that the bearing capacity of the
corrugated web beams made of cold-formed steel components is highly affected by the stabil-
ity of the parts and less affected by the configuration and number of spot welds.
Nevertheless, the increasing thickness of the parts does not necessarily mean an increase of
the resistance but the most unstable parts limit the maximum force.
A numerical model considering an alternative arrangement of the spot welds on the flange,
i.e. aligned horizontally, can improve the response of the beam in terms of rigidity.
Overall, the rigidity of the beam is mostly influenced by the thickness of the flanges and by
the arrangement of the spot welds, while the contribution of the shear panels and the corru-
gated web is small except for the very thin web.

ACKNOWLEDGEMENT

This work was supported by a research grant of the Romanian National Authority for Scien-
tific Research and Innovation, CNCS/CCCDI-UEFISCDI, project number PN-III-P2-
2.1-PED-2016-1684/WELLFORMED - Fast welding cold-formed steel beams of corrugated
sheet web and by a grant of the Romanian Ministry of Research and Innovation, project
number 10PFE/16.10.2018, PERFORM-TECH-UPT - The increasing of the institutional per-
formance of the Polytechnic University of Timișoara by strengthening the research, develop-
ment and technological transfer capacity in the field of “Energy, Environment and Climate
Change”, within Program 1 - Development of the national system of Research and Develop-
ment, Subprogram 1.2 - Institutional Performance - Institutional Development Projects -
Excellence Funding Projects in RDI, PNCDI III.

REFERENCES

Dassault Systemes 2014. Abaqus 6.14 Documentation (Providence, RI, Simulia Systems).
Dubina, D., Ungureanu, V. & Gîlia, L. 2013. Cold-formed steel beams with corrugated web and discrete
web-to-flange fasteners. Steel Construction 6: 74–81.
Elgaaly, M. & Seshadri, A. 1998. Depicting the behavior of girders with corrugated webs up to failure
using non-linear finite element analysis. Advances in Engineering Software 29: 195–208.
EN 1993-1-5: 2006, Eurocode 3: Design of steel structures - Part 1-5: Plated structural elements, CEN,
Brussels.
Gil, H., Lee, S., Lee, J. & Lee, H.E. 2005. Shear buckling strength of trapezoidally corrugated steel webs
for bridges. In Transportation Research Board - 6th International Bridge Engineering Conference: Reli-
ability, Security, and Sustainability in Bridge Engineering, 473–480.
Nie, J.-G., Zhu, L., Tao, M.-X. & Tang, L. 2013. Shear strength of trapezoidal corrugated steel webs.
Journal of Constructional Steel Research 85: 105–115.
Ungureanu, V., Both, I., Burca, M., Tunea, D., Grosan, M., Neagu, C. & Dubina, D. 2018a. Welding
technologies for built-up cold-formed steel beams: experimental investigations. Proc. of the Inter-
national Conference on Engineering Research and Practice for Steel Construction 2018 (ICSC2018),
5-7 September, Hong Kong, China (e-Proceedings).
Ungureanu, V., Both, I., Burca, M., Grosan, M., Neagu, C. & Dubina, D. 2018b. Built-up cold-formed
steel beams using resistance spot welding: experimental investigations. Proc. of the Eight International
Conference on Thin-Walled Structures - ICTWS 2018, 24-27 July 2018, Lisbon, Portugal
(e-Proceedings).

1200
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Study on the out-of-plane stability of steel portal frames

M. Vassilev & N. Rangelov


Department of Steel and Timber Structures, UACEG, Sofia, Bulgaria

ABSTRACT: Within the current EN 1993-1-1 the verification of rafters of portal frames for
lateral-torsional buckling in the haunched portions loaded by hogging bending moments is
not clarified. Recently the authors have conducted an extensive theoretical analysis pro-
gramme on lateral stability of steel portal frames of hot-rolled profiles. Special software has
been developed for automatic modelling and applying both the GMNIA and the general
method for lateral and lateral-torsional buckling. To validate the numerical simulations, an
experimental programme has been conducted on three typical portal frames of hot-rolled pro-
files. It has been proven that the results by GMNIA comply well with the tests, therefore is
adopted as benchmark method of numerical analysis. An extensive parametric study is then
conducted on thousands of portal frames composed of hot-rolled profiles. An important find-
ing based on the comparative analysis is that the application of the general method to whole
portal frames appears non-conservative and therefore its use shall be limited to single mem-
bers. Based on the obtained results, simple design rules for practical application are suggested
by adapting the well-known equivalent compressed strut model.

1 INTRODUCTION

Despite the extensive application of steel portal frames for single-storey buildings, there
are still some aspects of their stability that require additional clarification. No codified
practical method is specified in EN 1993-1-1 for lateral-torsional buckling verification of
rafters in the haunched portions loaded by hogging bending moments. It seems that,
within the code, there are only two possible approaches: the general method for lateral
and lateral torsional buckling (§6.3.4 of EN 1993-1-1) and geometrically and materially
nonlinear analysis with imperfections (GMNIA). However, both methods seem quite com-
plicated and cumbersome for practical use.
The general method is clarified in details in Simões da Silva et al. (2010). The method uses
a Merchant-Rankine type of empirical interaction expression to uncouple the in-plane effects
and the out-of-plane effects, and has been successfully applied to single tapered members
(Marques et al. 2013, Marques et al. 2014). However, the lateral-torsional stability of the
frame rafters seems to be even more complicated problem, taking into account the haunched
portions, the negative (hogging) bending moments and the specific restraint conditions with
lateral restraints at the top (tensile) flange only (rafter ‘fly bracing’ is not considered herein).
Therefore it seems more appropriate to consider the portal frame stability as a whole. How-
ever, though the general method is declared applicable to plane frames, very few applications
are available in literature, e.g. Papp & Szalai (2011).
In this paper is summarised the research carried out by the authors in the above context on
lateral stability of single storey single span steel portal frames of hot-rolled profiles. The general
method for lateral buckling is discussed with emphasis on the specific issues of its application to
frame lateral stability. The application of the geometrical and material nonlinear analysis with
imperfections (GMNIA) is also discussed. The above methods are confronted with a well-known
equivalent compressed strut model for the haunch, with a view to propose simple and reliable
design rules for practical use. To validate the numerical simulations, an experimental programme

1201
has been conducted on three typical portal frames of hot-rolled profiles, which is also summarised
in the paper.
Special software has been developed for automatic modelling and applying the above
methods. An extensive parametric study is then conducted on thousands of portal frames
composed of hot-rolled profiles. Based on the obtained results, a modification of the equiva-
lent compressed strut model is proposed for simple practical applications. An important find-
ing based on the comparative analysis is that the application of the general method to whole
portal frames appears non-conservative and therefore its use shall be limited to single
members.

2 NUMERICAL ANALYSIS

The scope of the study is limited to single-span portal frames composed of hot-rolled profiles.
Purlins are adopted equally spaced at 1,5 to 2,0 m, and hinged lateral restraints only to the
top flange are considered at each purlin. The haunches are assumed with the same section as
that of the rafter, with length considered 10%, 15% and 20% of the frame span. At the ends of
the haunches, stiffeners to the rafter are always present. A typical portal frame of consider-
ation is illustrated in Figure 1.
Three types of analysis are performed. The first one is the conventional design approach
based on linear elastic analysis and the design methods of EN 1993-1-1. The well-known
equivalent compressed strut model (Koschmidder & Brown 2012) is used for the haunch buck-
ling verification (Figure 2). The equivalent strut is defined in Section 1 with cross-section com-
posed of the haunch flange and 1/3 of the compressed portion of the web. The axial
compression force for the strut is obtained based on the bending moment and normal force in
Section 1. Though no out-of-plane restraints are considered at the bottom flange, in this ana-
lysis the out of plane buckling length, Lcr,z, is assumed equal to the actual length of the
haunch flange.
The second analysis is based on the general method for lateral and lateral-torsional buckling
(§6.3.4 of EN 1993-1-1). The critical point when applying the method is the determination of

Figure 1. Typical portal frame made of hot-rolled profiles with points of lateral restraint.

Figure 2. The equivalent compressed strut model for the haunch.

1202
Figure 3. Typical frame failure mode obtained by GMNIA.

two load multipliers, αult,k and αcr,op, the former being the minimum load amplifier to reach
the characteristic resistance of the most critical cross-section, and the latter – the minimum
load amplifier to reach the elastic critical load with regards to lateral or lateral torsional buck-
ling without accounting for in-plane flexural buckling.
To obtain αult,k, an in-plane geometrically nonlinear analysis with imperfections is carried
out. The frame is modelled with beam/frame elements. The imperfection pattern is according
to §5.3.2 of EN 1993-1-1 and includes both local bow member imperfections, scaled appropri-
ately, and a global frame imperfection (1/200 of column height). Since the analysis is non-
linear, αult,k cannot be refined by simply scaling the load. Therefore an iterative procedure is
developed to find precisely the critical cross-section and the relevant load multiplier.
For αcr,op another model is automatically generated with shell elements to perform linear
buckling analysis. An algorithm is developed to exclude in-plane and local buckling modes
and to identify the first out-of-plane mode and the corresponding load multiplier αcr,op. Thus
the global slenderness can be determined. Finally, instead of one reduction factor χop, the
reduction factors χz and χLT are used as suggested by Marques et al. (2008) using the relevant
buckling curves for the critical cross-section.
The third type of analysis, GMNIA, is also carried out automatically. The model with shell
FE is generated and linear buckling analysis is initially performed. The first overall out-of-
plane buckling mode is used to obtain the initial imperfections pattern, scaled according to
§5.3.4 of EN 1993-1-1. A revised model is thus generated. Material nonlinearity is based on
bilinear constitutive law with isotropic strain hardening. The load-carrying capacity of the
frame is assumed to correspond to the ultimate state criterion ‘attainment of the maximum
load’. The stressed state and the failure mode are also analysed. The software used is
ABAQUS nonlinear FE software (Abaqus 2016). A typical picture at limit state is illustrated
in Figure 3.

3 EXPERIMENTAL STUDY

To calibrate and validate the models used in GMNIA and to demonstrate the actual behav-
iour of the studied frames under monotonic loading, a test programme has been conducted.

3.1 Scope and test set-up


Three representative test frames have been designed with 7.0 m span as illustrated in Figure 4.
The cross-sections are selected to represent some stiffness variety. Care has been taken to
avoid cases in which the joint web panel governs the frame capacity. The load application
points approximately correspond to the purlin locations of an actual frame; lateral supports
are provided at those points, while the out-of-plane displacements are measured at the mid-
points.
The test specimens, the loading system and the lateral restraints are carefully designed to
correspond to the model assumptions. Column bases are detailed as perfectly hinged.

1203
Figure 4. Dimensions and cross-sections of the tested portal frames.

Figure 5. Loading system.

Figure 6. General overview of the test setup (a) and load application and lateral restraint devices (b).

An interesting loading system has been developed to apply monotonic gradual force-
controlled loading. The latter is provided by a suspended water tank and is applied through
a system of rope and pulleys, thus guaranteeing a uniform load distribution (Figure 5). The
detailing provides centric and ‘hinged’ load application.
The bracing system is also designed to correspond to model assumptions. Lateral restraints
are provided at the loading points, which at the same time can move vertically. Care is taken
for proper detailing, as illustrated in Figure 6.
A total of 13 inductive displacement transducers are installed to record the global
behaviour of the frames. Seven transducers measure vertical displacements, four ones
measure the lateral displacements of the top flange, and two transducers measure the lat-
eral displacements at the bottom flanges near the haunch, where lateral-torsional buck-
ling phenomena are expected.
The load is continuously measured by a commercial water meter with ordinary precision
that has additionally been verified by geometrically measured volume of the water in the tank.
Such a precision is found sufficient for the tests. To synchronise the load measurement in
time, a computer is used to video-record the water meter with its clock synchronised with the
clock of the data-recording computer. Since the steel strength appeared well above the nom-
inal, in addition to the water, test weights are added to increase the experimental load.

1204
Figure 7. Steel test specimens and measured mechanical properties.

3.2 Material mechanical properties


The nominal steel grade of all profiles is S275. To estimate the real mechanical properties of
the steel, a total of 18 (3 from each profile) standard specimens has been made by water jet
cutting to avoid any additional thermal or mechanical influence to the steel (Figure 7). The
results for different profiles appear variable, especially for IPE80 and IPE120. Moreover, the
steel cut from IPE80 exhibited no yield plateau. The actual mechanical properties were used
to re-calculate the test frames and to estimate the expected ultimate load.

3.3 Behaviour of the tested frames and observed failure modes


All tested frames reached ultimate limit state by lateral-torsional buckling of the rafters at the
haunch ends, as predicted by GMNIA analysis. Accordingly, largest lateral displacements
were recorded there, depending on the actual initial imperfections of the test frames, which
unfortunately were not measured and recorded.
In the test of Frame 1, the lateral restraint mechanism exhausted its run and leaned on
the bracket. However, this took place well after reaching the limit state and did not
affect the final result. For Frame 2 the restraint detail was altered to provide larger dis-
placement capacity of bracing bracket. The observed ultimate behaviour was quite simi-
lar. For the test of Frame 3, the restraint details were additionally improved to avoid
any malfunction at large displacements. Together with that, the loading water tank was
modified to accommodate more test weights due to the expected higher load capacity of
this frame. Some illustrative photos representing the observed failure behaviour are
shown in Figure 8 for Frame 3.

3.4 Comparison with numerical analysis


To prove and validate GMNIA as benchmark analysis method, the three tested frames were
modelled with measured material properties of the material. Since the actual initial

Figure 8. Observed behaviour at failure of Frame 3.

1205
Figure 9. Experimental and analytical vertical and lateral displacements.

imperfections were not recorded, initially the analysis was performed using the most unfavour-
able imperfection patterns based on properly scaled first lateral buckling mode. The displace-
ments obtained at the gauge locations were monitored and plotted against the experimental
data in Figure 9, where the numerical results are plotted by dotted lines.
The comparison in terms of displacements shows a very good agreement between the analyt-
ical and the experimental results. As opposed to the numerical analysis, where a clear limit
state criterion is available, the determination of the experimental load-bearing capacity is based
on a careful examination of the recorded data to identify the moment when the rate of lateral
displacement increments at the relevant transducer (8 or 13 as indicated) becomes ‘significant’.
The predicted by GMNIA behaviour corresponds well to the experimentally observed one.
However, the theoretical ultimate loads appear somewhat lower by 24% for Frame 1, 33% for
Frame 2 and 18% for Frame 3. Indeed, these results provide a comfortable safety for the
numerical analysis and prove its adequacy, but an explanation of the discrepancies seems
noteworthy. In this regard, an additional numerical study on the effect of imperfections has
been carried out.

1206
Figure 10. The effect of imperfections to the numerical load-carrying capacity.

3.5 Effect of initial imperfections


To evaluate the imperfection sensitivity, the following approach is adopted. On one hand, two
boundary imperfection levels are adopted: ‘standard’ level with amplitude equal to 0,5e0,d, as
per §5.3.4 of EN 1993-1-1, ‘minimum’ imperfection level with 100 times smaller amplitude,
and the case without imperfections. On the other hand, various imperfection patterns corres-
ponding to different overall lateral buckling modes have been considered. The analysis dem-
onstrates a substantial effect of the initial imperfections as illustrated in Figure 10 in terms of
the obtained numerically total load-carrying capacity with various imperfections in compari-
son with the experimental capacity attained.
It is well seen in Figure 10 that the case of ‘standard’ imperfection level with imperfection
pattern corresponding to the first lateral-torsional buckling mode appears always on the safe
side. Thus, on one hand, a reasonable explanation for the discrepancies between the experi-
mental and GMNIA results is pointed out, and on the other hand, the safety of the adopted
numerical analysis procedure is proven. Indeed, these conclusions are drawn from only three
case studies, however, the authors believe that similar conclusion can be generalised.
All the above results validate the analytical model and prove GMNIA as a benchmark
method of analysis.

4 PARAMETRIC ANALYSIS AND RESULTS

Based on the procedure described in Section 2, using the specially developed software, an
extensive parametric study is carried out. Several thousands of frames are automatically ana-
lysed, varying the column and rafter sections, frame span (12 to 24 m), haunch length and
number of lateral restraints at the rafter top flange. By varying the different hot-rolled

1207
Figure 11. Comparison between the results of the general method with GMNIA.

profiles, a vast variation of flexural and torsional stiffness ratios is covered. The most repre-
sentative stiffness parameter for a member appears to be the following:
qffiffiffiffiffiffiffiffiffi
π2 EIw
K¼ GIt L2
ð1Þ

where Iw is the cross-section warping constant, It is the St. Venant torsional constant, L is the
length of the member, E is the modulus of elasticity and G is the shear modulus of the steel.
For a frame, more important appears the ratio:

ω ¼ KC =KR ð2Þ

where KC and KR are the values of K for the column and for the rafter, respectively.
Since the research is aimed at studying essentially the rafter out-of-plane stability, all irrele-
vant cases have been disregarded (e.g. when in-plane buckling or joint shear panel resistance is
governing). Generally, the analyses are grouped into series.
The main objective of the parametric study is to compare the results of the general method
(§6.3.4) and the simplified equivalent compressed strut model with those from GMNIA. The
comparative results are presented in terms of utilisation (demand-to-capacity) ratio FEd/FRd,
FEd representing the total load and FRd being the frame resistance according to the method
used.
The comparison between the results obtained by the general method and those from
GMNIA appears interesting and astonishing. It seems that if the general method is applied to
the whole frame, the obtained results appear non-conservative, i.e. on the side of unsafety.
This is illustrated in Figure 11, where the utilisation ratio is plotted against KR for a typical
series of frames.
On the contrary, the results obtained by the simplified equivalent compressed strut
approach as described in Section 2 appear always more or less conservative, as illustrated in
Figure 12 for two typical series of frames.
The analysis proves that the conservatism of the equivalent compressed strut method
mostly depends on the ratio ω, therefore, for the most typical case of 10% haunch length, the
following correction factor is proposed:

η ¼ 0:015ω þ 1:1: ð3Þ

1208
Figure 12. Comparison between the results by the simplified approach with GMNIA.

Thus the design resistance of the equivalent compressed strut may be obtained as:

Nb;Rd ¼ ηAfy χz =γM1 ð4Þ

with χz obtained for the out-of-plane buckling length as shown in Figure 2. In the above equa-
tion A is the area of the cross-section of the strut (Figure 2) and the rest of notation is obvious.

5 CONCLUSIONS

An extensive numerical and limited experimental research on steel portal frames of hot-rolled
profiles is presented, aimed at clarifying the out-of-plane stability of rafters in the haunched
portions loaded by hogging bending moments. The numerical analysis is carried out by spe-
cially developed software which generates the models and controls the analysis by three
methods: GMNIA, the general method of §6.3.4 of EN 1993-1-1 and the conventional design
approach combined with the equivalent compressed strut model.
The comparative analysis shows that if the general method is applied to the whole frame,
the obtained results appear well on the side of unsafety. Therefore, the use of this method
shall be limited only to single members.
The simplified equivalent compressed strut model is proven to be suitable and conservative
even without restraints to the bottom flange of the rafter. A correction to this method to rea-
sonably minimise its conservatism is proposed that fits well to the numerical results. This
approach is both simple and provides a good base for engineering judgement.

REFERENCES

Simões da Silva, L., Simões, R., Gervásio, H. 2010. Design of Steel Structures. ECCS.
Marques, L., Simões da Silva, L., Greiner, R., Rebelo, C., Taras, A. 2013. Development of a consistent
design procedure for lateral-torsional buckling of tapered beams. Journal of Constructional Steel
Research 89: 213–235.
Marques, L., Simões da Silva, L., Rebelo, C., Santiago, A. 2014. Extension of EC3-1-1 interaction for-
mulae for the stability verification of tapered beam-columns. Journal of Constructional Steel Research
100: 122–135.
Papp, F., Szalai, J. 2011. Theory and application of the general method of Eurocode 3 Part 1-1. Eurosteel
2011, Budapest, 31.08-02.09.2011.
Koschmidder, D.M., Brown, D.G. 2012. Elastic design of single-span steel portal frame buildings to Eurocode
3, SCI Publication P397, Ascot: SCI.
Marques, L., Simões da Silva, L., Rebelo, C. 2008. Numerical validation of the general method in EC3-1-1:
Lateral and lateral-torsional buckling of non-uniform members. Eurosteel 2008, Graz, 03-05.09. 2008.
Abaqus, 2016. Dassault Systems/ Simulia, Providence, RI, USA.

1209
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Warping transfer superelement model for bolted end-plate


connections subject to 3D loads

B. Vaszilievits-Sömjén
KÉSZ Holding Ltd., Budapest, Hungary

J. Szalai
ConSteel Solutions Ltd., Budapest, Hungary

R.M. Movahedi
Faculty of Architectural, Civil and Transportation Engineering, Széchenyi István University, Győr, Hungary

ABSTRACT: A simple beam element based modelling technique has been developed which
makes possible to analyze frames made of I sections with column-rafter bolted end-plate connec-
tions, subject to 3D loads, compatible with the thin walled beam theory with 7DOF beam elements.
The model previously developed by the same team for welded connections has been ex-tended with
the addition of linear spring elements to model the bolts located at the upper and lower beam flange
level. The spring stiffnesses are calculated based on the extension of the Eurocode component
method and verified by simulations performed with FEA software Abaqus.

1 INTRODUCTION

The goal of this research was to demonstrate how the out-of-plane flexibility of a bolted corner
connection can be considered in structural analysis based on 7DOF beam elements.
A superelement modelling concept has been developed and presented by the same team Vaszi-
lievits-Sömjén & Szalai (2019) and has been validated for welded connections. This modelling
technic has been extended with the out-of-plane flexibility resulting from the usage of bolts with
typical bolt patterns, instead of assuming a welded connection.
As the present edition of Eurocode 3 (EN 1993-1-8) doesn’t provide rules to calculate out-
of-plane stiffness of bolted connections, a simple extension of the rules will be recommended.

2 OUT-OF-PLANE STIFFNESS OF BOLTED CONNECTIONS

Eurocode 3 doesn’t propose rules to calculate stiffness of bolted-connections subject to out-of-


plane effects. Several researchers have proposals to handle this question. Neumann & Nuhic
(2011) and Neumann & Buzaljko & Thomassen & Nuhic (2012) have proposed a model on the
bases of existing Eurocode method for in-plane stiffness calculation and extended with additional
modes for cases specific for out-of-plane s tiffness problem. Romero (2010) has proposed for
I sections a simple 50%-50% division of standard T-stub stiffnesses to come to a practical value
applicable for a half T-stub consisting of 1 bolt only. Couchaux & Rodier (2016) have proposed
additional component to model the flexibility of compressed side of such connections.
All these researches have calculated a unique stiffness value corresponding for the whole
I section dealing mainly with symmetrical bolt patterns. Only the presence of Mz (out-of-plane,
weak axis) bending moment has been assumed, without any interaction of other internal forces
acting on the same bolted connection. Additionally, they have assumed, that as a result of the
application of the out-of-plane effects, the flanges of the connection will “close” at the most com-
pressed end. This assumption is valid if there is only Mz moment applied on the connection. In

1210
Figure 1. Effect of Mz and B shown with concentrated forces.

case of general 3D loading case this isn’t necessarily true. It depends on the ratio of all involved
internal force components (N, My, Mz and B, namely axial force, in+out-of-plane bending, bimo-
ment) and rigidity of the elements of the connection itself. The stiffness of the connection will
strongly depend on the way how contact works between opposite elements.
The proposed method of stiffness calculation for I sections eliminates some of the limita-
tions and extends the applicability to a wider range of connections and more general 3D load
application.
It is proposed to “cut” the I section into a separate upper and lower flange zone (see Figure 1)
and calculate separate stiffness values for these flange zones, respectively, by considering the bolt
rows located directly above and below the corresponding flanges only. Other bolts will be disre-
garded by this model. This separation makes the approach also compatible with an application
with the superelement introduced to handle welded corner connections in Vaszilievits-Sömjén &
Szalai (2019). This proposed separation of flanges also allows a straightforward consideration of
Mz + B interaction as these effects on flange level can be added together.
For sake of simplicity we concentrate on 2D frames made of I sections. We assume that such
a frame is loaded dominantly by in-plane gravity forces. There are usually no direct out-of-
plane loads on such a frame, such effects would be mainly eventual second order consequences
of applied imperfections with out-of-plane deformation components, using for example
a properly scaled eigenshape with out-of-plane displacement components (indirect loads). For
the interaction of opposite side flange zones, we define 3 possible simple scenarios:
In case of a dominant My bending moment, one of the flanges (on the compressed side) will
be “closed” and the other flange will be “open” or “partially open”. This assumption can be
simply checked by calculating the stress distribution from the actual 3D loading case. If there
are no direct out-of-plane loads and stresses available, this check is not possible. A safe assump-
tion might be in this case to assume an “open” condition for the flange zone in tension.
For each 3 scenario we define appropriate elastic springs values, to be applicable for
a flange zone.
– If a flange remains closed, there will be no elongation of the bolts and no resulting deform-
ation of the end-plates. For such case we assume that the elastic spring can be assumed as
perfectly rigid.

Figure 2. Basic scenarios considered.

1211
– If a flange is partially open, we refer to the calculation method presented by Neumann &
Nuhic (2011) and Neumann & Buzaljko & Thomassen & Nuhic (2012) and Romero (2010).
– For the case of “open” flanges we propose a new calculation model

2.1 Proposed stiffness calculation method for the “open” scenario


As stated above, the upper and lower flange zones will be handled separately. The proposed
model allows to calculate stiffness value for a flange zone considering an “open” scenario. As
it will not be assumed that the connection will “close” it is applicable to the general case when
the effect of in-plane bending moment has a dominant effect on the connection. Such
a situation exists when the tension force resulting from in-plane bending moment and/or other
axial force is large enough compared to the out-of-plane combined flange bending
moment Mz + B, so the end-plates around this bolt row are not in contact. In this case forces
are acting only through the bolts.
The rotational stiffness of a bolt row S(j,z) in this state can be calculated from analyzing
a unit Mz bending moment and the caused φ rotation angle.

Mz
Sj;z ¼ ð1Þ

A mechanical model can be made where the stiffnesses of all participant components are repre-
sented with a spring stiffness denoted by keq at both bolts.
This equivalent spring stiffness keq represents the stiffnesses of the following components: end-
plate in bending in one side of the connection, bolt under longitudinal loading and the end-plate in
bending on the other side of the connection. The stiffness coefficients of the individual components
can be calculated based on Eurocode EN 1993-1-8. From the stiffness coefficients spring stiffnesses
can be calculated by multiplying with the elastic modulus, therefore the equivalent spring stiffness is:

E
keq ¼ ð2Þ
1
k5;1 þ k110 þ k15;2

where
k5,1: stiffness coefficient for end-plate in bending in one side of the connection
k10: stiffness coefficient for bolt under longitudinal loading
k5,2: stiffness coefficient for end-plate in bending on the other side of the connection
Applying an Mz out-of-plane bending moment on this model translates into a compression
and a tension force FMz on these springs.

Mz ¼ FMz  z ð3Þ

The forces cause a δ deformation in both springs.

Figure 3. Mechanical model for the “open” scenario.

1212
FMz
δ¼ ð4Þ
keq

Since the model and the forces are symmetrical the center point of the rotation angle will be
midway in between the springs. Based on this the rotation angle can be calculated. If we
assume that ’ is small, then

FMz
δ k
’¼ ¼ eq ð5Þ
z=2 z=2

Substituting back into the formula for the rotational stiffness yields:

Mz keq  z2
Sj;z ¼ ¼ ð6Þ
’ 2

3 VERIFICATION OF THE PROPOSED STIFFNESS VALUE

In order to verify the correctness of the proposed analytic solutions for “open” and “partially
open” scenarios, we built numerical models using Abaqus FEA software. Although we target
the analysis of 2D portal frames, the correctness of the corner superelement for welded con-
nection has already been validated in Vaszilievits-Sömjén & Szalai (2019). In order to work
with simpler models, we created beam-to-beam models with 500-500 mm lengths with an
assumed bolted connection between them. The end of one of the beams has been fixed against
all displacements and at the other end of the other beam we applied loads appropriate for the
different scenarios and we checked the resulting force-displacement diagrams for the upper
and lower flange zones separately. For the models we used solid elements of C3D8R type. We
considered non-linear material models. We used the built-in Risk analysis based on arc-length
method. We considered 2 different bolt patterns (4 bolts and 6 bolts) with M16 8.8 bolts, with-
out pre-stressing. We considered 15 mm end-plate thicknesses.
In order to test the analytic spring value proposed by Neumann & Nuhic (2011) and Neumann
& Buzaljko & Thomassen & Nuhic (2012) and Romero (2010) for the “partially open” we made
a simulation with Mz bending moments incrementing up to failure. Mz is applied as 2 pieces of
concentrated bending moments shared equally between the flanges. We plotted the force-
displacement diagrams of the upper flange, considering both bolt patterns and compared with
the initial stiffness calculated with the proposed analytic method. Diagrams obtained with
Abaqus are drawn with intermittent lines and marked as N=0. Continuous lines are the analytic
values, calculated based on methods by Neumann & Nuhic (2011) and Neumann & Buzaljko &
Thomassen & Nuhic (2012) marked as “Neu” and by Romero (2010) marked as “Hei”.
Regarding the test for the proposed new “open” condition, we applied in step 1 a preload of
100 kN on both flanges to create tension along the flanges (no contact) and started with the
same incremental Mz application in step 2. The obtained initial stiffness values are compared
with the calculated analytic spring values. Diagrams obtained with Abaqus are drawn with
intermittent lines and marked as N = 100 kN tension. Continuous lines are the analytic values.
Two different bolt patterns have been considered as shown on Figure 5.
We can observe that the analytic values show agreement in tendency with the simulation based
results but give generally higher initial stiffness. The agreement in case of the six bolts pattern is
good, but for the four bolts pattern the analytic value clearly overestimates the simulated stiffness.
This might come from the fact, that in case of the four bolts configuration the end-plate has
a visible rotation and doesn’t start with a vertical tangent, as would be supposed by Eurocode. In
case of six bolts this assumption seems to be correct, as the bolt rows above and below the upper
flange result a vertical tangent of the end-plate. The tendency of Eurocode T-stub model to result
higher stiffness than the real one has also been reported by Wald (2016).

1213
Figure 4. Application of loads for “open” and “partially open” scenarios.

Figure 5. Four and six bolts patterns.

On the diagrams corresponding to the “open” condition there are points marked as “scen-
ario changing points”. They correspond to the load level of to the monotonic increasing out-
of-plane bending moment where it closes the gap at the compressed side and the scenario
changes to a “partially open” one.
For practical case we assume, that it is a generally acceptable safe side approach to consider the
“open” condition for a flange zone which is subjected to tension along the full width. For flanges
under fully compressed condition infinitely rigid spring values are used which is equivalent with
a continuous connection as appropriate for a welded connection. In intermediate cases, where the
“partially open” condition is deemed as justified, the corresponding higher connection stiffness can
be used.

4 EXTENSION OF THE PROPOSED SUPERELEMENT MODEL


TO INCORPORATE LINEAR SPRINGS

The corner superelement model presented in the previous paper was valid for welded connections.
In case of bolted connections, the flexibility of the connection must be considered. Following the
logic of Eurocode, this will be handled by the application of linear springs which represent the flexi-
bility of the connection. As written before the I section will be cut into flange zones and use two
appropriate spring values for the out-of-plane stiffness of the flange zones, instead of the usual one
spring value used by Eurocode to model the in-plane stiffness of bolted connections, as shown on
Figure 8.

5 APPLICATION OF THE EXTENDED MODEL

The application of the model extended with linear springs to represent the out-of-plane stiff-
ness of bolted corner connection is demonstrated on the following example. For simplicity it
has been assumed that the connection can be considered as rigid for in-plane deformations
and therefore no additional springs should be considered in this direction.

1214
Figure 6. Force-displacement diagram for the 4 bolts pattern.

Figure 7. Force-displacement diagram for the 6 bolts pattern.

Figure 8. Extension of the corner model with springs for each flange.

1215
We assume a simple 2D portal frame with 12 m span with 300 mm deep, symmetric welded
I sections.
The proposed analysis steps are carried out using ConSteel 3D structural design package. It
can be done in other software as long it is able to handle 7DOF beam elements considering
warping deformations for global structures:
1. Calculate internal forces assuming first welded corner models, without the application of
out-of-plane springs.
2. Calculate stresses from all 4 components of internal forces (N, My, Mz and B) at both
flanges. Assign the appropriate spring values for open or partially open cases or keep the
continuous connection for the closed case.
3. Rerun the analysis with the properly set springs.
4. Calculate stresses again from all 4 components (N, My, Mz and B) at both flanges and re-
check the validity of the assumption. If due to changed static system the stress distribution
falls into a scenario different from originally assumed, change the spring and repeat step 3.
5. First order displacements and internal forces are available.

5.1 Analysis models


In this paper we compare the results obtained from different analysis
1. First order analysis with using hinge at the tensioned flange as a conservative approach,
marked as “Model 1”. At the compressed side rigid connection is considered.
2. First order analysis with using hinge at the both flanges as a very conservative approach,
equivalent with the Eurocode 3 proposed approach of using isolated members with warp-
ing and out-of-plane moment releases. This is marked as “Model 2”.
3. The proposed method. First order analysis using the linear spring values at the tensioned
flange as proposed above, marked as “Proposed model”. At the compressed side rigid con-
nection is considered. Calculated spring values corresponding to four and six bolts pattern
are 1122 kNm/rad and 1632 kNm/rad, respectively.
4. First order analysis assuming that connections are fully rigid out-of-plane, marked as
“Model 4”. This corresponds to a welded connection.

Figure 9. Loads on 2D frames analyzed, with lateral supports at top flange level.

Table 1. Lateral displacements of the midspan section for different models and different bolt
patterns.
Type of model 4 bolts pattern 6 bolts pattern

Model 1 50.7 mm
Model 2 52.8 mm
Proposed model 37.7 mm 37.4 mm
Model 4 36.7 mm

1216
Table 2. Normal stresses in the beam flanges.
Four bolts pattern Six bolts pattern

Top flange Bottom flange Top flange Bottom flange


Type of model [MPa] [MPa] [MPa] [MPa]

Model 1 147.96 to 153.98 −178.85 to −140.87


Model 2 147.90 to 154.05 −157.46 to −162.27
Proposed model 76.85 to 225.09 −104.07 to −215.65 75.19 to 226.75 −102.08 to −217.64
Model 4 70.97 to 230.98 −97.86 to −221.87

5.2 Results of displacements


Displacements of the centerline of the middle cross section from the given loads are shown
below for all four static systems.
Visibly the hinged models (Model 1 and Model 2) result excessive lateral displacements. On
other side, considering the out-of-plane stiffness of any of the considered bolt patterns results
displacements very close to a welded connection assumption (Proposed model vs Model 4).

5.3 Results of stresses


The normal stresses of the first beam cross section right after the bolted connection are drawn
for all three static systems. Normal stresses are calculated from N, My, Mz and B internal
force components.
The stresses along the web are the same in cases therefore are not shown. Stress distribution
along the flanges are very different between the first two cases (hinged models) and the second
two cases (continuity assumed) The very conservative modelling proposed by Eurocode 3
gives an un-safe stress reduction from 225.09 MPa down to 153.98 MPa (30%!)

6 CONCLUSIONS

We have presented a simple method to estimate out-of-plane stiffness of bolted connections,


suitable for typical application in frames. With the help of the spring values obtained for the
upper and lower flanges we extended our corner superelement model and we demonstrated its
application for a 2D frames. The obtained results are in line with the expectations and based
on engineering judgements seem to be realistic. A proper validation of this method is under
preparation. The first results show that a connection designed to have enough stiffness for in-
plane bending moment demand might have enough stiffness for out-of-plane effects, therefore
an automatic consideration of a hinge at such locations seems to be excessively conservative.

REFERENCES

Vaszilievits-Sömjén, B. & Szalai, J. 2019. Simple superelement model of warping transfer in moment con-
nections between I sections Ninth International Conference on Advances in Steel Structures
(ICASS2018) 5–7 December, 2018, Hong Kong, China
Neumann, N. & Nuhic, F. 2011. Design of structural joints connecting H or I sections subjected to
in-plane and out-of-plane bending. Eurosteel Conference 2011, Budapest, 2011
Neumann, N. & Buzaljko, M. & Thomassen, E. & Nuhic, F. 2012 Verification of design model for out-of
plane bending of structural joints connecting H or I sections. Nordic Steel Construction Conference
(NSCC 2012), Oslo, 2012
Couchaux, M. & Rodier, A. 2016. Behavior of bolted end-plate connections of beams subjected to biaxial
bending moments and axial forces, Eurosteel Conference September 13–15, 2017, Copenhagen, 2016
Romero, E. 2010 Finite element simulation of a bolted steel joint in fire using Abaqus program, MSc
Thesis, Tampere University of Technology, 2010
Wald, F. & al 2016. Benchmark Cases for Advanced Design of Structural Steel Connections, CVUT
September 2016
ConSteel 12 2018. Structural analysis and design software, www.consteelsoftware.com

1217
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Tests of gusset plate connection under compression

J. Vesecký, K. Cábová & M. Jandera


Faculty of Civil Engineering, Czech Technical University in Prague, Prague, Czech Republic

ABSTRACT: Over the past years, research has shown, that behavior of connections with
gusset plates subjected to compression represents a complicated task. There has been developed
a few analytical methods since then, some of which are used in standards procedures, mainly in
Australia, New Zealand and Canada. However, in European standards, no design procedure is
described and EN 1993-1-5 offers only general rules for buckling of steel plates. This paper
introduces four analytical models used for gusset plates design – Whitmore, Thornton, Modified
Thornton and model proposed by Khoo, Perera and Albermani. Furthermore, results of new
experimental study on six full scale specimens are presented. Among other measuring methods,
3D digital image correlation was used for detection of planar strain of plates. Finally, experi-
mental results are compared to predictions of analytical models. It is shown, that all methods
based on dispersion angle are completely inappropriate for this type of connections.

1 INTRODUCTION

Gusset plates represent frequently designed component, which are used for joints of steel
structures. They can be typically found in trusses, connecting their individual members
together. Simultaneously, they are commonly used for connection of bracings to main load-
bearing elements.
Gusset plate connections are mostly subjected to axial force (tensile or compressive), or
sometimes to cyclic loading (dynamic and earthquake forces). Gusset plate can be connected
to structural member either centrically (e.g. by doubled cleat plate or by doubled structural
members with open section) or eccentrically (by slotted cleat plate). While concentric connec-
tion is better from statics point of view, it is also more expensive and difficult for fabrication.
Therefore, eccentric connection is often used instead.
New research, conducted at Faculty of Civil Engineering of Czech Technical University in
Prague, was focused just on eccentric connections. An experiment was executed on six full
scale specimens, representing bracing composed of CHS member with connection on both
sides consisting of cleat and gusset plate bolted together.
Axial and lateral deflection of specimens and plane deformation of plates, using 3D digital
image correlation (DIC) were measured during experiment. Results were later used for validation

Figure 1. Specimen C2 (before testing) with nomenclature used in this paper.

1218
of numerical model which was then used for parametric study (Vesecký et al., 2019). Measured
load bearing capacity was compared to calculated capacity using four different analytical models.
It was shown, that older analytical models based on dispersion angle greatly overesti-
mate capacity of tested connections. Therefore, these analytical models should not be used
in these cases.

2 ANALYTICAL MODELS

Even though gusset plates are used from the end of 19th century (i.e. Eiffel Tower), the
research of their behavior has begun much later, in the middle of 20th century. Whitmore
(1952) defined in his work concept of dispersion angle and effective width, based on his
observations of stress flow in plates. Thirty-two years later, Thornton (1984) improved
his work by introducing buckling length and critical load for gusset plate connections.
Yam and Cheng (2001) later proposed modification of Thorntons method by taking plas-
ticity into account. Finally, new series of tests on 12 specimens was conducted by Khoo
et al. (2009). They have proposed completely new analytical model, based on observed
collapse mechanism.
In this paper, design procedures for each analytical model are just briefly introduced, all
equations can be found in specific papers mentioned above or together in (Vesecký, 2019).

2.1 Whitmore method


Method is based on assumption, that stress is distributed under 30° angle from point load.
Therefore, Whitmore proposed that dispersion line should be considered from the first row of
bolts up to last row of bolts. Effective width is then determined as width of dispersed stripe
(see Figure 2).
Load bearing capacity of connection is then simply calculated by equation (1), while stabil-
ity, effect of gradual yielding and bending moment caused by eccentricity are not taken into
account.

Nu ¼ beff  t  fy ð1Þ

where beff = the Whitmore width; t = the gusset plate thickness; and fy = the yield strength of steel.

2.2 Thornton method


Thornton proposed a procedure considering the effect of buckling into account by introdu-
cing buckling lengths. He defined three nominal lengths (L1, L2 and L3) on the Whitmore
section and average value (Lavg). Based on specific gusset plate configuration, one of these
lengths is used.
Buckling length is then simply calculated as nominal length multiplied by buckling length
coefficient (βcr). Appropriate values have been proposed by Dowswell (2006), see part 2.5.

Figure 2. Examples of Whitmore effective section beff (grey area) for corner gusset plates.

1219
Figure 3. Left: Nominal lengths of gusset plate defined by Thornton. Right: Modified Thornton method.

With known buckling length, relative slenderness (λ) is calculated. It is then used for calcula-
tion of critical stress (σcr). Load bearing capacity is defined by equation:

Nu ¼ beff  t  σcr ð2Þ

2.3 Modified Thornton method


Yam and Cheng (2001) proposed modification of Thornton method, taking plasticity into
account. They have suggested using 45° dispersion angle instead of 30°, see Figure 3. For con-
nection with one bolt in each row it means 73 % increase in effective width and therefore load
bearing capacity. All equations from Thornton method remain unchanged.

2.4 Khoo, Perera, Albermani (KPA) proposal


Khoo et al. (2009) observed collapse mechanism with two plastic hinges (see Figure 4) during
their experiment on specimens with eccentrically connected gusset plates. Based on such obser-
vation, they have proposed a new analytical model. It takes into account stability, plasticity as
well as bending moment due to eccentricity. Load bearing capacity is computed in two steps,
using principle of virtual works. Original analytical model with limitation on gusset and cleat
plates with the same rectangular shape and same thickness has been later generalized
(Vesecký, 2019) by replacing plate width with new one – length of plastic hinge (Lpl), which
can be determined graphically as the shortest link between side edges of plates outside the
overlapping part.

Figure 4. Specimen at the end of test and analytical model KPA with two plastic hinges.

1220
Figure 5. Basic gusset plates configurations. From the left: Compact, Noncompact, Extended,
Single-Brace, Chevron-Brace.

Table 1. Critical length coefficients (βcr) and nominal lengths (L0) for differ-
ent gusset plate configurations.
Gusset plate configuration βcr L0

Compact * *
Noncompact 1,00 Lavg
Extended 0,60 L1
Single-Brace 0,70 L1
Chevron-Brace 0,75 L1

*compact gusset plates should never fail in buckling mode.

In the first step, collapse load (NI) can be calculated by equating external work (Wext = N·e·θ)
and internal work (Wint = (Mgpl,I +Mcpl,I)·θ):

g
Mpl;I þ Mpl;I
c
NI ¼ ð3Þ
e

where superscript “g” indicates a gusset plate; and “c” indicates cleat plate.
Then conservative estimate of squash load (Ny,min) and moment of inertia (Ic,min) is calcu-
lated, followed by elastic buckling load (Ncr), relative slenderness (λ) and critical buckling load
(Nc). The first step of the procedure is finished by calculation of ultimate load:

Nc
Nu;I ¼ ð4Þ
1 þ Nc =NI

In the second step, plastic moments are reduced, taking ultimate load and critical buckling
load from the first step into account. New collapse load (NII) is calculated and then, using
equation (4), the ultimate load (Nu,II) is determined. Ny, Ncr, λ and Nc remain unchanged from
the first step.

2.5 Nominal lengths and buckling length coefficients


Dowswell (2006) collected and evaluated results of all available experiments and numerical stud-
ies focused on connections with gusset plates. Then he recommended nominal lengths and buck-
ling length coefficients for specific configurations of gusset plates (see Figure 5 and Table 1).

3 EXPERIMENTS

Behavior of connections under compression has been investigated on six full-size experimental
specimens, representing bracing member connected on both sides by cleat and gusset plates.

1221
Figure 6. General geometry of specimen connections (type C and type D bottom supported end); ×
indicates position of displacement sensors with corresponding number.

Table 2. Average material properties obtained by tensile coupon tests.


component t [mm] E [MPa] fy [MPa] fu [MPa] A [%] εy=fy/E [-]

gusset/cleat plate 8,05 199 900 405,0 610,9 24,2 0,0020


CHS 102/4 3,65 201 100 358,2 598,4 23,0 0,0018

Two types of specimens were prepared – type C with perpendicular end of gusset plate and
type D with oblique end of gusset plate. Geometry is shown in Figure 6.

3.1 Tensile coupon tests


Three tensile tests for 8 mm thick plate and two tensile tests for 4 mm thick plate (made out of
CHS 102/4) were conducted, to measure real material properties of major specimen compo-
nents. Measured stress-strain diagrams were transformed onto true stresses (σtrue) and true
strains (εtrue). Values of modulus of elasticity (E), yield strength (fy), ultimate strength (fu) and
ductility (A) were evaluated. Table 2 summarizes obtained results.

3.2 Connection tests


Objectives of experiments were determination of load-deflection diagrams, ultimate loads, devel-
opment of lateral deflection of specimens, planar deformation of plates and mode of failure.
All specimens were made out of grade S355 steel, composed out of CHS 102/4, 8 mm thick
cleat and gusset plates and 20 mm thick end plate. Cleat plate was embedded into prepared
longitudinal groove at the end of CHS and then welded by 4 mm thick fillet weld. Cleat and
gusset plates were bolted by M20 8.8 bolts. Connections were situated symmetrically on both
ends of specimens. Figure 7 gives details of the tested specimens.
Free length of gusset plate (Lfree,g), overlapping length (Lol) and number of bolts (always
situated in two rows) were chosen as variable parameters of specimens. Table 3 summarizes
values of each parameter. There was one test for each specimen carried out.
Loading was displacement controlled. Deformation speed -0,3 mm/min was selected as
appropriate. The test took about 30 minutes on average per one specimen (ultimate load was
reached after about 18 minutes). Descending branch of the load-deflection diagram was
recorded for all specimens. Eight displacement sensors were used to record lateral deflection
of predefined points on specimens (seven situated at both connections, one in the middle of
CHS). Planar deformation of plates was recorded by two pairs of cameras. Recording was
later evaluated by DIC software. Prior to test, plates were painted in white and then sprayed
in black to create speckle pattern.

1222
Figure 7. Left: Frontal and side view of test set-up. Middle: Photo of specimen C2 attached to hydraulic
jack. Right: Close up photo of top connection of specimen D4 after test with sensors on the right.

Table 3. Geometry and parameters of specimens (nominal values).


α Lc Lol Lfree,g Lfree,c LCHS
Number of
Specimen [°] [mm] [mm] [mm] [mm] [mm] bolts

C1 90 210 135 55 20 2000 4


C2 90 245 170 55 20 2000 2
C3 90 210 170 20 20 2000 2
D1 45 248 135 93 20 2000 4
D2 45 298 135 143 20 2000 4
D4 45 333 170 143 20 2000 2

4 TEST RESULTS

Simply by comparison of measured ultimate loads with number and position of bolts, it’s
obvious, that their placement has no visible effect on capacity of connection (as long as bolts
does not fail on their own prior to plates buckling). On the other hand, with longer total
length of connection (Lc) and longer free length of gusset plate (Lfree,g) the ultimate load is
decreasing as expected.

4.1 Collapse mechanism


During testing of specimens, total of four phases of behavior were observed, differentiated by
sudden change in stiffness as can be seen in Figure 8.
In the first phase, ideally elastic behavior was observed. When axial load reached around 20
to 30 kN, bolt slip (within the hole tolerance) has occurred, followed by immediate drop in
stiffness. Later on, when bolt reached edge of the hole, stiffness increased again, almost to the
same value as prior to the slip. Loading force raised up to the ultimate load, which occurred

1223
Figure 8. Load-deflection (axial/lateral) diagrams of all specimens. Significant bolt slip for specimen C3.

between 90 to 110 kN. Finally, descending branch was recorded, creating approximately
hyperbolic shape on the load-deflection diagram.
It should be noted, that for type D specimens, bolt slip was eliminated before tightening.
Collapse mechanism with two plastic hinges, located at the free lengths of gusset and cleat
plate, was observed for all six specimens. Plastic hinges always occurred exactly when ultimate
load was reached. In all cases, only one connection buckled (Figure 9), while other one and CHS
member remained fully elastic during the whole loading process. There was even visible decrease
in lateral deflection of non-buckled connections, after reaching ultimate load of specimens.
For type C specimens, buckling always occurred at the bottom, supported end, while for
type D specimens, buckling always occurred at the top, loaded end.
Observed collapse mechanism is in agreement with the results obtained by Khoo et al. (2009).

4.2 Comparison of measured and calculated ultimate load


Using analytical models described in parts 2.1 to 2.4, predicted ultimate load was calculated for
all six specimens. For Thornton and Modified Thornton method, it was considered βcr = 0,70
and L0 = L1 (see Table 1). For KPA method, buckling length factor was considered βcr = 1,20,
according to (Khoo et al., 2009) and nominal length as length of the connection L0 = Lc. Plastic
hinge lengths (Lgpl and Lcpl) were determined graphically as the shortest link between side edges
of plates. All calculations considered real material properties, see Table 2. Following table sum-
marizes and compares reached results.

Figure 9. All six buckled connections. Right to left: C1, C2, C3, D1, D2, D4.

1224
Table 4. Comparison of measured and calculated ultimate loads of the specimens.
Test Whitmore Thornton Mod. Thornton KPA
Nu,exp Nu,W ρ Nu,T ρ Nu,MT ρ Nu,KPA ρ
Spec. [kN] [kN] [-] [kN] [-] [kN] [-] [kN] [-]

C1 106,2 400,2 3,77 372,7 3,51 512,9 4,83 85,8 0,81


C2 93,3 261,9 2,81 240,1 2,57 415,8 4,45 68,5 0,73
C3 112,5 261,9 2,33 251,9 2,24 436,4 3,88 87,0 0,77
D1 102,8 400,2 3,89 347,3 3,38 478,0 4,65 66,1 0,64
D2 101,7 400,2 3,93 307,2 3,02 422,8 4,16 49,1 0,48
D4 92,1 261,9 2,84 195,2 2,12 338,0 3,67 40,7 0,44

where Nu,i = measured or calculated ultimate load; ρ = calculated to measured resistance ratio.

It’s evident, that analytical models based on dispersion angle and effective width (Whit-
more, Thornton, Modified Thornton), are significantly overestimating the resistance of eccen-
tric gusset plate connections. They give unsafe results by hundreds of percent. The greatest
source of error can be associated with false assumption about influence of specific placement
and number of bolts. Other source of error is definitely neglection of bending moment arising
from eccentricity. The question remains, whether it’s appropriate to consider coefficient βcr as
low as 0,70, which is value for fixed-pinned member.
On the other hand, KPA method, based on collapse mechanism with two plastic hinges, is
suitable for such type of connections. Even this method could be improved as it gives conser-
vative results, especially for specimens with oblique end of gusset plate (type D).

4.3 DIC measurement


Deformation of connections was recorded by two pairs of cameras and then evaluated by 3D
DIC software VIC-3D. Accuracy of DIC was compared to values measured by displacement
sensors (LDVTs and laser sensors), see Figure 10. The chart shows a very good agreement of
the results. The same figure also shows deformed shape and isolines of the principal strain for
one specimen. By comparison of the maximal measured strain (0,0069) with yield strain
(0,0020) calculated in Table 2 it is clear, that gusset plate undergoes plastic deformation in the
area of free length.

Figure 10. Left: Lateral deflection of bottom connection C1 measured by sensors and DIC (sensors pos-
ition is shown in Figure 6). Middle and right: Final deformed shape and the principal strain (e1) for the C2
bottom gusset plate as measured by 2D and 3D DIC. Red color: e1 = 0,0069; violet color: e1 = -0,0002.

1225
5 CONCLUSION AND FINAL RECOMMENDATIONS

New tests of six specimens, representing bracing with eccentric gusset plate connections subjected
to compression were described in the paper. This type of connection is still widely used, despite its
disadvantages (additional bending moment due to the eccentricity). The main reason for use is
simplicity of fabrication when compared to more sophisticated concentric connections. For the
first time, configuration with oblique end of gusset plate (under 45° angle) was tested.
Collapse mechanism with two plastic hinges on buckled connection was observed during all six
tests. Other parts (second connection and CHS member) undergone purely elastic deformation.
Both connections were captured by cameras during tests and recordings were later evalu-
ated by 3D DIC, to obtain planar strain and spatial deformation of plates. Comparison with
values measured by sensors shows great accuracy of DIC.
Results of test revealed, that placement and number of bolts (resp. number of parallel rows)
has no visible influence on the resistance of connections. As expected, lower resistance was meas-
ured for specimens with larger total length or greater gusset plate free length of connection.
Four analytical models were described in this paper. Three of them (Whitmore, Thornton
and Modified Thornton), based on dispersion angle and effective width, were proven insuffi-
cient for eccentric gusset plate connections, giving greatly unsafe results (by hundreds
of percent). Only the most recent model KPA (Khoo et al., 2009), based on true collapse
mechanism, gives realistic predictions. However, it can be improved in the future as it tends to
be conservative for gusset plates with oblique end. Improvement can be done in more accurate
calculation of critical buckling load (moment of inertia, buckling length), in consideration of
rotational stiffness of bracing member or in inclusion of second order influence.
Despite the fact, that aforementioned procedure allows simple calculation of required thick-
ness of specific gusset plate, a correct design approach should follow these steps:
1. the shortest possible bolted joint between gusset and cleat plate;
2. minimization of free lengths for gusset and cleat plate;
3. calculation of required gusset and cleat plate thickness using KPA method;
4. decision about alternative ways to increase buckling resistance (e.g. stiffeners) when thick-
ness calculated in step 3 is too big;
5. if so, update of gusset and cleat plate thickness taking increased stiffness into account.

ACKNOWLEDGEMENT

Authors of this article would hereby like to thank to Technological Agency of Czech Republic
for its support of research by grant n. TJ01000045 “Advanced procedures of steel and com-
posite structure connection design and production”.

REFERENCES

Dowswell, B. 2006. Effective Length Factors for Gusset Plate Buckling. Engineering Journal (New York).
43: 91–101.
EN 1993- 1-5. 2013 Eurocode 3: Design of steel structures – Part 1-5: Plated structural elements. Euro-
pean Committee for Standardization (CEN). Brussels.
Khoo, X.E & Perera, M. & Albermani, F. 2009. Design of eccentrically connected cleat plates in
compression. Advanced Steel Construction. 6(2): 678–687.
Thornton, W. 1984. Bracing Connections for Heavy Construction. Engineering Journal. 21: 139–148.
Vesecký, J. 2019. Buckling Resistance of Gusset Plates. Diploma thesis. Czech Technical University in
Prague. Supervisor Jandera M.
Vesecký, J. & Jandera, M. & Cábová, K. 2019. Numerical modelling of gusset plate connections under
eccentric compression. (in press).
Whitmore, R.E. 1952. Experimental Investigation of Stresses in Gusset Plates. Knoxville: Engineering
Experimental Station, University of Tennessee.
Yam, M.C.H & Cheng, J.J.R. 2001. Behavior and design of gusset plate connections in compression.
Journal of constructional steel research. 58: 1143–1159.

1226
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Numerical modelling of gusset plate connections under


eccentric compression

J. Vesecký, M. Jandera & K. Cábová


Faculty of Civil Engineering, Czech Technical University in Prague, Prague, Czech Republic

ABSTRACT: Recently, new experimental study which was focused on behavior of bolted
connections with gusset plate under eccentric compression was carried out at CTU in Prague.
Buckling resistance and longitudinal and lateral deflection was measured for six specimens.
The specimens represented full scale tubular bracing connected to rigid load-bearing construc-
tion on both ends. Progressive 3D digital image correlation technology was used, to record
planar strain of plates. This paper describes complex numerical models based on aforemen-
tioned physical specimens. General FEA software Abaqus including geometric and material
nonlinearities (large deformations, plasticity, imperfections) was used. Numerical models were
validated on results measured during experiments. Later, an extensive parametric study was
performed with nearly 50 numerical models. Influence of total of eight free parameters was
examined. Besides other results, parametric study confirmed previous findings, that analytical
models based on dispersion angle are generally not accurate for design of eccentric gusset
plate connections.

1 INTRODUCTION

New experimental study focused on connections with eccentric gusset plates under compres-
sion is described in detail in paper (Vesecký et al., 2019). Based on these experiments
a numerical study had followed. Its aim was creation of accurate complex numerical model
(see Figure 1 and chapter 2) which was later successfully validated on results of aforemen-
tioned experiment.
An influence of input properties on bearing capacity of numerical models was studied in
a parametric study. Following parameters were considered in the study: inclination angle of
gusset plate side edges; inclination angle of gusset plate end edge; free length of gusset plate; free
length of cleat plate; thickness of plates; length of bracing member; number of bolts; steel grade.
Result of the parametric study are presented in chapter 4.

2 NUMERICAL MODELS

General FEA software Abaqus/CAE 6.14 was used to create complex numerical models. Real
dimensions of plates and tube measured prior the tests were used. Sensitivity study was per-
formed for the first numerical model to optimize accuracy and complexity of calculation. Besides
other findings, sensitivity study confirmed, that geometric and material nonlinearities must be
always taken into account. This finding was first published years ago by Chakrabarti (1987).
Analysis was performed in two steps. In the first one, normalized linear elastic buckling
eigenmodes were calculated. These were used in the second step as imperfections for nonlinear
static analysis. Amplitude of imperfection was determined in two ways – the first as a portion
of geometric fabrication tolerances according to EN 1993-1-5 (Annex C) and the second as
actually measured initial deformations by digital image correlation (DIC).

1227
Figure 1. Specimen C2. Top left: Test specimen. Top right: Numerical model with nomenclature.
Bottom left: First elastic buckling eigenmode. Bottom right: Deformed geometry with von-Mises stress.
Parameters of all six specimens (dimensions, number of bolts etc.) can be found in (Vesecký et al., 2019).

2.1 Model assumptions


Although no numerical models can fully capture reality, a maximal effort was made during its
creation to ensure good accuracy. Nevertheless, it should be noted, that few simplifications
were assumed to reduce computational complexity:
• general stress-strain diagram was replaced by trilinear elastoplastic one;
• bolt holes were modeled with the same diameter as bolts (i.e. 20 mm);
• no axial force (related to tightening) was applied to bolts;
• screw shank was modeled as completely smooth, without thread;
• end plates were replaced by boundary conditions (fixed connection);
• self-weight wasn’t considered (it was less than 0,5 % of the bearing capacity).
As the consequence of the listed simplifications, it is obvious, that bolt slip wasn’t simulated.

2.2 Sensitivity study


Specimens C2 and D4 were chosen for sensitivity study. The results for the investigated geom-
etries show that:
• plasticity has to be included, otherwise great computational errors occur;
• effect of hardening after yielding is completely negligible;
• when large deformations are not considered, results are about two times overestimated;
• bearing capacity of model without imperfections was about 20 % higher than the cap-
acity measured during tests;
• strength and stiffness of bolts have negligible effect on results;
• it’s necessary to model bolts with head and nut and perfectly hard contact with plates;
• a compromise between accuracy and calculation speed was achieved by using four node
shell elements with 8 mm mesh size for plate and 16 mm mesh size for bracing member
and by using four node tetrahedron elements with 4,5 mm mesh size for bolts.
Influence of material and geometric nonlinearities is shown in Figure 2.

Figure 2. Numerical models of specimens C2 and D4. Left: Influence of large deformations and imper-
fections on ultimate load. Right: Influence of material nonlinearities (plasticity) on ultimate load.

1228
2.3 Description of numerical models

2.3.1 Model parts


All models were assembled from four basic parts – gusset plate, cleat plate, CHS and bolt. Plates
were created as 3D planar shell, CHS as 3D extrusion shell and bolts as solids, see Figure 3.

2.3.2 Material properties


Real material properties, obtained by tensile coupon tests (see Vesecký et al. 2019), were used
for numerical models. Although sensitivity study showed hardening as negligible parameter,
trilinear elastoplastic stress-strain diagram was chosen to substitute general diagrams from
tests. Unlimited plastic strain was assumed after ultimate stress was reached, but during load-
ing of all models, such condition did never occur.
Material parameters of bolts 8.8 were assumed according to EN 1993-1-8. All three
assumed stress-strain diagrams and material properties are summarized in Figure 4 and
Table 1.

Figure 3. Individual parts of numerical models (different scale). Left to right: Gusset plate, cleat plate,
bracing member (CHS 102/4) and bolt (M20).

Figure 4. Stress-strain diagrams used for numerical models.

Table 1. Material properties used for numerical models.


Property Symbol Unit plates CHS 102/4 bolt 8.8

Young’s modulus E MPa 199 000 201 100 210 000


Poissons ratio ν - 0,3 0,3 0,3
Yield strength fy MPa 405,0 358,2 640,0
Ultimate strength fu MPa 610,9 598,4 800,0
Plastic strain at ultimate strength εpl - 0,242 0,230 0,250

1229
2.3.3 Sections
Gusset and cleat plates and CHS were all modeled by homogeneous shell section with real
thickness measured during tests. Simpsons integration was used with 11 points for plates and
5 points for tube. Section integration was performed for each step of the analysis.

2.3.4 Constraints and contacts


Fillet weld connecting cleat plate to CHS was replaced by simple rigid connection, due to its
four times higher capacity than gusset plate connection itself.
Contact between gusset and cleat plate and between head/nut of bolt and plates was con-
sidered completely frictionless in tangential direction and perfectly hard in normal direction.
Compressed parts of bolt hole were connected to shank by “shell to solid” constraint.

2.3.5 Boundary conditions


There were boundary conditions applied to the end edges of both gusset plates. Non-loaded
end was considered fixed in all directions, while loaded end was considered fixed with excep-
tion of free axial movement.

2.3.6 Finite elements and mesh size


Sensitivity study showed, that plates and CHS can be discretized by general shell finite elem-
ents S4R with four nodes, reduced integration, hourglass effect control and linear approxima-
tion. Mesh size was set to 8 × 8 mm for plates and CHS end parts resp. 8 × 16 mm for the
central part.
Bolts were discretized by solid tetrahedron finite elements C3D4 with linear approximation.
Average mesh size was set up to 4,5 mm to create three layers of FE for head and nut.

2.4 Analysis of numerical models


The calculation was performed in two steps. The first, linear buckling analysis was used to
obtain eigenmodes, which were then used as imperfections into nonlinear static analysis.

2.4.1 Buckling analysis


Unitary axial load was imposed on numerical models for linear buckling analysis. Critical
loads and corresponding eigenmodes were calculated by subspace iteration with limit of 30
iterations with 18 vectors.
Only the five lowest eigenmodes were calculated, as higher eigenmodes represent shapes,
that can never occur in reality (e.g. torsional failure of gusset plate).
Because of linearity of the problem, analysis was very fast, lasting approximately 1 minute.

2.4.2 Static analysis


During geometric and material nonlinear analysis (GMNIA) numerical models were loaded
by prescribed increments of axial deformation. That way, the execution of the physical experi-
ments was exactly simulated. Also, full Newton-Raphson method could be used to record
whole load-deflection diagram, including descending branch.
Imperfections were introduced as linear combination of few lowest eigenmodes so that ini-
tial deformation was applied for both connections and bracing member. Amplitude (e0) of
imperfection was determined in two ways:

1) based on measured initial deformation of plate using DIC, see Figure 5;


2) as geometric imperfection according to EN 1993-1-5 – for plates e0 = L/50 (L is
the connection length), for CHS as 80 % of fabrication tolerances (EN 1090-2), thus
e0 = 0,8·L/1000 = L/1250 (L is the bracing member length).

It took approximately 31 minutes for each nonlinear static analysis of numerical models.

1230
Figure 5. Initial shape of gusset plates as measured by DIC. Left: Bottom gusset plate of specimen C2.
Right: Bottom gusset plate of specimen D1.

3 RESULTS AND VALIDATION ON EXPERIMENT

All numerical models reached the same collapse mechanism – buckling of one connection with
two plastic hinges occurring at the free lengths of gusset and cleat plate. This collapse mechan-
ism is in complete agreement with the one observed during physical tests (see Vesecký et al.,
2019, Khoo et al., 2009 and Figure 6).
When plastic hinges occurred during static analysis, decrease in maximal stress of non-
buckling connection and CHS was observed. Development of plastic hinges continued, which
was signalized by exceeding of the yield strength (fy = 405,0 MPa), see Figure 7.
Axial deformation was limited to 12 mm, while buckling usually took place at around 2 to
4 mm. Ultimate strength of steel was never reached even at the maximal deformation (for com-
parison, see Figure 7: fu = 610,9 MPa, σmax ≤ 500,0 MPa).

Figure 6. Buckled connections of physical specimens. Left to right: C1, C2, C3, D1, D2, D4.

Figure 7. Deformed shape and von Mises stress for buckled connection of C2 (left) and D4 (right).
First picture: State at the ultimate load. Second picture: State at the maximal deformation. Yielding in
orange and red areas.

1231
Figure 8. Comparison of experimental and numerical load-deflection diagrams for specimens C2 (left)
and D4 (right). Diagrams shifted so that the ultimate load is reached at the same axial deflection.

Figure 9. Comparison of experimental and numerical ultimate loads for all six specimens.

3.1 Load-deflection diagrams


As stated before, bolt slip wasn’t considered in numerical models and some other simplifying
assumptions were made. Therefore, load-deflection diagrams obtained from numerical analysis
did not exactly fit the ones measured during tests. In all six cases, numerical models had higher
stiffness prior to buckling than physical specimens (see Figure 8). Descending branch of dia-
grams corresponded perfectly. Diagrams for all six specimens can be found in (Vesecký, 2019).

3.2 Ultimate load


Very good agreement between the test and the numerical results was achieved for the ultimate
load of specimens (see Figure 9). When imperfections determined from DIC measurements
were considered, the ultimate load of numerical models was approximately 8 % higher com-
pared to the tests. For imperfections according to EN 1993-1-5, the ultimate load of the
numerical models was approximately 7 % lower than for the tests. It can be assumed, that the
real amplitude of imperfections lies somewhere between those two limit values. However, both
predictions were very close to the test results and the model is accurate enough for its use in
a parametric study.

4 PARAMETRIC STUDY

A main goal of the parametric study was to determine the degree of influence of various
input values on the final load capacity and the first elastic critical load. Less attention
was dedicated to the load deflection behavior. Therefore, loading by force in combination

1232
Figure 10. Schematic connection drawings of specimens C2 and D4.

with full Newton-Raphson method was chosen because of its time savings. Consistency
with loading by deformation (until the limit point) was successfully verified.
Specimens C2 and D4 were chosen (see Figure 10) as reference models. All other models with
altered parameters were created by their modifications. Unlike for numerical models described
in chapter 2, material parameters for parametric study models were determined according to
EN 1991-1-1. Steel grade was S355 unless otherwise stated. Material parameters were:
E = 210 000 MPa, ν = 0,30, fy = 355 MPa, fu = 490 MPa, εpl = 0,25. Imperfections were deter-
mined according to EN 1993-1-5 and EN 1090-2 with amplitude e0 = L/1250 for CHS and
e0 = L/50 for plates that form connection. Examined parameters are summarized in chapter 1.
Always only one parameter was changed, while all others remained constant.

4.1 Charts

Figure 11. Results of parametric study. Green line = ultimate load; red line = critical load.

4.2 Influence of individual parameters


From the results presented above in form of charts, it can be derived that:
• inclination of side edges of gusset plate has significant influence on the both ultimate
and the critical load for small angles with diminishing influence for greater angles.
Decline can be explained by just partial yielding of wide gusset plates (see Figure 12).

1233
Figure 12. Partial plastic hinges (in blue) for wide gusset plates at ultimate load.

• Growing inclination of end edge of gusset plate causes small decrease in the ultimate
and critical load, which then stops and starts growing for greater angles.
• Growing free lengths of both gusset and cleat plate causes (for examined interval) approxi-
mately linear decrease in the ultimate load and hyperbolic decrease in the critical load.
• Ultimate load grows quadratically and critical load grows in cubic for growing thickness
of gusset and cleat plates.
• Growing length (and thus slenderness) has negligible influence on the ultimate load but
significant influence on the critical load. For high slenderness, buckling of bracing
member becomes decisive collapse mechanism.
• Number and placement of bolts have almost no influence on the ultimate and critical load.
• Growing yielding strength of steel causes almost linear increase in ultimate load.

5 SUMMARY AND CONCLUSIONS

5.1 Conclusions for the numerical models


Gusset plate connections under eccentric compression can be modeled by shell elements
(gusset plate, cleat plate and bracing member), solid elements must be used for bolts.
It is necessary to perform geometrically and materially nonlinear analysis (GMNIA), so that
large deformations and plasticity are all accounted. Hardening of steel has no effect on bearing
capacity and can be neglected (thus replaced by limitless plastic strain when yielding occurs).
For the examined models, neglection of the imperfections lead to 10 % to 20 % overesti-
mation of actual bearing capacity. It is appropriate to implement imperfections as combin-
ation of eigenmodes obtained by linear stability analysis. Amplitude of imperfection can be
conservatively determined as value for geometric imperfection according to EN 1993-1-5, or
as 80 % of fabrication tolerances according to EN 1090-2.
It’s convenient to load models by prescribed displacement increments. Then Newton-Raphson
solver can be used to get the whole load-deflection diagram including descending branch.
For better agreement with real load-deflection diagrams, it might be necessary to include
bolt slip, and bolt thread in the numerical model.

5.2 Conclusions for the parametric study


Parameters that increase bearing capacity:
• growing inclination of side edges of gusset plate; thickness of plates, grade of steel.
Parameters with little to no effect on bearing capacity:
• inclination of end side of gusset plate; length (slenderness) of bracing member; number
and placement of bolts.
Parameters that reduce bearing capacity:
• growing free length of gusset plate, growing free length of cleat plate.

1234
Results of parametric study confirmed, that analytical models based on dispersion angle and
effective width (Whitmore, Thornton, Modified Thornton) does not correspond with real behav-
ior of the examined connections. Number of bolts and its placement does not influence the bear-
ing capacity at all (as long as bolts does not fail on their own). On the other hand, bearing
capacity keeps rising even when gusset plate is expanded outside the effective width (beff).
Although length of a bracing member does not influence the bearing capacity (for examined
interval), buckling of bracing member is likely to prevail for greater member slenderness.

5.3 Practical application


Complex numerical models can very accurately simulate real behavior of gusset plate connec-
tions. However, their use is usually limited to scientific purposes, because of its time-
consuming process (preparation and computational difficulty).
In everyday practice, when designing structures, engineers are facing a lot of other prob-
lems, that need to be solved. As result, they need much faster and simpler tools, yet sufficiently
accurate.
One such solution is combination of analytical procedures (component method) with finite
elements method – CBFEM. Software IDEA Statica utilizes those principles. Its use for
design of gusset plate connections is described in detail by Vild et al. (2019).
Results obtained by complex numerical models (like ones described in this paper) can be
used as source of improvement for CBFEM analysis as well as for analytical models.

ACKNOWLEDGEMENTS

Authors of this article would hereby like to thank to Technological Agency of Czech Republic
for its support of research by grant n. TJ01000045 “Advanced procedures of steel and com-
posite structure connection design and production”.

REFERENCES

Abaqus 6.14 Documentation. Available in: http://ivt-abaqusdoc.ivt.ntnu.no:2080/texis/search/?query=wet


ting&submit.x=0&submit.y=0&group=bk&CDB=v6.14
Chakrabarti, S. K. 1987. Inelastic Buckling of Gusset Plates. Disseration. The University of Arizona.
EN 1090-2. 2018 Execution of steel structures and aluminum structures – Part 2: Technical requirements
for steel structures. European Committee for Standardization (CEN). Brussels.
EN 1993- 1-1 2nd ed. 2013 Eurocode 3: Design of steel structures – Part 1-1: General rules and rules for
buildings. European Committee for Standardization (CEN). Brussels.
EN 1993- 1-5. 2013 Eurocode 3: Design of steel structures – Part 1-5: Plated structural elements. Euro-
pean Committee for Standardization (CEN). Brussels.
EN 1993- 1-8 2nd ed. 2013 Eurocode 3: Design of steel structures – Part 1-8: Design of joints. European
Committee for Standardization (CEN). Brussels.
Khoo, X. E & Perera, M. & Albermani, F. 2009. Design of eccentrically connected cleat plates in
compression. Advanced Steel Construction. 6(2): 678–687.
Vesecký, J. 2019. Buckling Resistance of Gusset Plates. Diploma thesis. Czech Technical University in
Prague. Supervisor: Jandera M.
Vesecký, J. & Cábová K. & Jandera M. 2019. Tests of gusset plate connection under compression. (in
press).
Vild, M. & Kabeláč, J. & Kuříková, M. & Wald, F. 2019 Design of gusset plate connection with
single-sided splice member by component based finite element method. (in press)

1235
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Buckling of columns during welding

M. Vild & M. Bajer


Faculty of Civil Engineering, Brno University of Technology, Brno, Czech Republic

ABSTRACT: The paper presents experiments of columns loaded by compressive force during
welding along their longitudinal axis. A weld bead was being laid in the corner of the cross-
section from the bottom of the column to its mid-height at constant load. Then, still during
welding, the load was increased until the column failed by flexural buckling. Manual gas metal
arc welding with carbon dioxide was used. All the columns had the height of 3 m. The tested
cross-sections were HEA 100, IPE 120, IPE 160, SHS 100 × 5, and CHS 76 × 4. The measured
values were applied load, column deflections and changes in length, welding voltage, current
and speed, and temperatures. The experiments were performed at the laboratory of Department
of Metal and Timber Structures of Faculty of Civil Engineering of Brno University of Technol-
ogy. Material properties of steel are temporarily decreased by the high temperature caused by
the welding process. Analytical method is offered to design the buckling resistance of a column
during welding. The cross-sectional properties are assumed to be reduced at the most dangerous
height and the column is treated as stepped. The welds should be placed in short segments sym-
metrically. Otherwise, the deflection caused by the shrinkage of asymmetrically placed welds
must be also taken into account.

Keywords: column buckling, experiment, steel, welding

1 INTRODUCTION

Welding under load is often performed without consideration of load resistance of a member.
Welding of a gusset plate or a stiffener to an existing column might serve as an example. Such
column has temporarily decreased material properties due to the high temperatures in the
vicinity of the weld. While welding longitudinally to the column axis is relatively safe, trans-
verse welding is dangerous because a large portion of the column cross-section is weakened.
Another example, which is the core of the authors’ main research, is the strengthening under
load, i.e. welding of strengthening plates to an existing column without unloading, Vild &
Bajer (2016). In this case, long longitudinal welds are applied to transfer the shear flow
between an existing column and strengthening plates that are welded in different states of
stresses. The shear flow is not very large so there is no need for transverse welds.
The weld is formed by inducing very high temperatures to fuse two steel parts together, usu-
ally with added welded material. The steel near the welding rod is in a molten state. Gradually,
both in distance from the welding rod and in time, the temperature decreases. The temperature
distribution can be determined by simple analytical procedures, e.g. method of moving heat
source; see Figure 1, Masubuchi (1980), Rosenthal (1941), Rosenthal (1946).

2 ANALYTICAL METHOD

The analytical method was developed with a simple assumption that the steel heated to tem-
perature higher than 500°C is ineffective. At this temperature, according to EN 1993-1-2:2006,
the steel yield strength is reduced by the factor 0.78, proportionality limit by 0.36 and modulus

1236
Figure 1. Temperature field behind the welding point.

of elasticity by 0.6. This temperature was chosen as a safe value corresponding to more
detailed analysis in general cases, Huenersen, Haensch & Augustyn (1990). Welding and espe-
cially manual welding is full of uncertainties and more complicated assessment is not worth
the gain.
The buckling resistance of a simply supported column is the smallest when the heat source
is in the mid-height of the column. The part of the cross-section is weakened and the column
is treated as stepped column. The critical buckling load is the function of moments of inertia
of full cross-section and weakened cross-section, the length of the column, and the length of
the weakened part – ČSN 73 1401 (1998), Trahair & Kitipornchai (1971). The weakened
cross-section is determined by taking into account only the part where the temperature is
below 500°C. The temperature distribution can be calculated by Rosenthal equations –
Rosenthal (1941, 1946). The decrease in the buckling resistance of a weakened stepped
column compared to full, existing column was determined by numerical simulations in
ANSYS software. The results of numerical analysis were approximated by a formula of
increased buckling length Lcr;e of the stepped column:
   
Lcr;e Ltemp 0:6 I0 Ltemp 0:6
¼1 þ 
Lcr L Itemp L

where Lcr ¼ buckling length; Ltemp ¼ length of the weakened part; L ¼ column length;
I0 ¼ moment of inertia of the full cross-section; Itemp ¼ moment of inertia of the weakened
cross-section. Due to the increased buckling length, buckling load Ncr;e is decreased and
equivalent initial imperfection etemp increased:

Wel
etemp ¼ α  ðλtemp  0:2Þ 
A0

where α ¼ imperfection factor; λtemp ¼ relative slenderness of the stepped column; Wel ¼ section
modulus of the full cross-section; A0 ¼ area of the full cross-section.
Second order theory is used to determine the resistance of the column stability during welding.
Therefore, the stepped column deflection, wtemp , is necessary:

1 5 N1  Δw  L2
wtemp ¼  etemp þ Δw þ 
1  Ncr;e
N1 48 E  I0

1237
Figure 2. Comparison between the numerical simulation and proposed analytical procedure.

where N1 ¼ preload under which the welding is performed; Δw ¼ difference between the posi-
tions of centres of gravity of the full cross-section and the weakened cross-section. Finally,
maximum normal stress in the weakened cross-section σx;1 at mid-height is determined as:

N1 wtemp  N1
σx;1 ¼ þ  fy
Atemp Wel;temp

where Atemp ¼ area of the weakened cross-section; and Wel;temp ¼ elastic section modulus of
the weakened cross-section. The maximum normal stress must not exceed yield strength fy .
The procedure was verified by another set of numerical simulations; see Figure 2.

3 EXPERIMENTAL PROGRAM

There are very few experiments in the literature. The most relevant paper is by Suzuki & Hor-
ikawa (1984) comprising 8 specimens susceptible to buckling. Therefore, a new experimental
program was performed in the laboratory of Department of Metal and Timber Structures of
Faculty of Civil Engineering, Brno University of Technology. The tested cross-sections were
HEA 100, IPE 120, IPE 160, SHS 100 × 5, and CHS 76 × 4. There were three specimens for
each cross-section except for the CHS which contained only 2 specimens. All columns had the
length of 3 m. The boundary conditions were knife-edge bearings; pinned around the weaker
axis and fixed around the stronger axis. The column was inserted into the loading frame and
all the measuring devices were activated. The column was loaded by a high compressive force
near its limit calculated by the proposed analytical method. Then a weld bead was laid at the
corner/edge of the cross-section from the bottom to the mid-height. After the weld reached the
mid-height, the load was gradually increased while welding continued, until flexural buckling
of the column around the weaker axis occurred.
The temperature was measured by thermocouples in two height levels and in different dis-
tance from the weld bead. There were seven thermocouples in total, labeled T1–T7. The lateral
displacements were measured by four draw-wire sensors in three height levels, labeled wl,
wm1, wm2, and wu. At mid-height, there were two sensors to measure also possible rotation of
the column. The column was loaded by loading cylinder from the bottom. The force and the
displacement in longitudinal direction were measured by load cell and by LVDT, respectively.
The position of sensors is shown in Figures 3 and 4.

1238
Figure 3. Selected cross-sections of tested specimens with the placement of measuring devices.

Figure 4. Test set-up: Column IPE160-2 after failure by buckling (left), measuring devices across the
column height (right).

1239
Table 1. Welding speed v, heat input q, and preload magnitude N1 .
HEA1 HEA2 HEA3 SHS1 SHS2 SHS3 IPE120-1

v [mm/s] 2.05 2.73 3.85 2.7 2.92 2.54 3.45


q [J/mm] 878 661 469 668 618 710 464
N1 [kN] 160 160 160 290 290 290 48
IPE120-2 IPE120-3 IPE160-1 IPE160-2 IPE160-3 CHS1 CHS2
v [mm/s] 3.45 3.3 3.75 2.90 2.26 2.50 2.14
q [J/mm] 464 553 458 577 741 730 852
N1 [kN] 40 37 100 100 100 80 70

Manual gas metal arc welding with CO2 as a shielding gas was used. The parameters of
electric current, voltage and welding speed were designed for the rate of cooling from 800 to
500°C equal to 20 s to achieve high-quality weld – EN 1011-2:2003. However, due to manual
welding, the speed of welding varied; see Table 1.

4 RESULTS

Measured temperature distribution was slightly lower compared to the calculation according
to Rosenthal equations of the moving heat source. It seems that the calculation is the safe
assumption.
The applied weld was long and eccentric to the column axis. The heat of the welding is caus-
ing the steel to expand and then, when the steel cools, to shrink. Huenersen, Haensch &
Augustyn (1990) suggest that the column should first deflect in the direction to the weld as the
steel expands and then in the direction away from the weld as the steel shrinks. However, the
first phase – expanding – was not observed. Column deflection related to assumed weld length
can be seen in Figure 5. The subsequent increase in load, during which the deflection soared,

Figure 5. Lateral displacement at mid-height during welding.

1240
Figure 6. Lateral displacement at mid-height during welding.

is not shown. The specimens IPE120-1 and CHS2 are not shown because they buckled prema-
turely during welding and the deflection caused by buckling and welding cannot be distin-
guished. The expanding was effectively held by the column bending stiffness. On the other
hand, the weld shrinkage was significant and affected the buckling resistance. The column
started to deflect at the length of the weld from 300 to 600 mm corresponding to time of weld-
ing from 90 to 280 s. The columns with lower area of the cross-sections and lower moment of
inertia around the weaker axis started to deflect sooner.
The examples of load-displacement diagrams can be seen in Figure 6. Specimens with cross-
sections IPE 160 (labeled IPE1–3) and SHS 100 × 5 (labeled SHS1–3) are compared to the analyt-
ical method (labeled An). The initial deflections are included. The initial deflections of specimens
were determined using Southwell plot – Southwell (1932). The analytical method disregarded the
weld shrinkage and it can be seen that for such a long weld, the weld shrinkage should be taken
into account to keep the method safe. Preferably, it is convenient to divide the welds into seg-
ments up to the length of 300 mm and place them so that the weld shrinkage is in opposite
directions.

5 CONCLUSION

The analytical method to calculate the buckling resistance of compressed column during weld-
ing was outlined. The steel in the vicinity of the weld heated over 500 °C is assumed ineffect-
ive. The column is treated as stepped column. The lateral deflection is determined. Second
order theory is used to calculate the normal stress at the weakened cross-section and this
normal stress must not exceed the yield strength.
The method was validated by experimental program comprising cross-sections HEA 100,
IPE 120, IPE 160, SHS 100 × 5, and CHS 76 × 4. The columns were 3 m long and pinned
around the weaker axis. Gas metal arc welding with carbon dioxide as shielding gas was used.
The shortcoming was that manual welding was used and the welding properties, especially
welding speed, was not constant.
The single pass weld starts to shrink after 300 mm of the weld length or 90 s of welding. If
longer weld is applied, the shrinkage of the weld should be taken into account. If symmetrical
welds are designed, the welds should be divided into segments with the length of up to
300 mm and the welding sequence planned to place weld segments symmetrically to balance
the shrinkage. If unsymmetrical welds are designed, the full shrinkage should be added to the
lateral deflection in the calculation of the column buckling resistance.

1241
ACKNOWLEDGEMENT

The financial support of projects No. TH02020301 and No. LO1408 “AdMaS UP – Advanced
Materials, Structures and Technologies”, supported by the Ministry of Education, Youth and
Sports under “National Sustainability Programme I” is gratefully acknowledged.

REFERENCES

ANSYS Academic Research Mechanical APDL, Release 16.2, Help System, ANSYS, Inc.
ČSN 73 1401: Design of Steel Structures. Prague, 1998 (in Czech).
EN 1011-2 Welding – Recommendations for welding of metallic materials – Part 2: Arc welding of ferritic
steels. 2003.
EN 1993-1-2 Eurocode 3: Design of steel structures – Part 1-2: General rules – Structural fire design.
Prague: CNI, 2006.
Huenersen, G., Haensch, H. & Augustyn, J. 1990. Repair welding under load. Welding in the World 28
(9): 174–182.
Masubuchi, K. 1980. Analysis of Welded Structures: Residual Stresses, Distortion, and Their Consequences.
Pergamon Press, 642 pp. ISBN-13: 978-1483172620.
Rosenthal, D. 1941. Mathematical theory of heat distribution during welding and cutting. Welding Journal
20: 220–234.
Rosenthal, D. 1946. The theory of moving sources of heat and its application to metal treatments. Trans.
ASME 48: 848–866.
Southwell, R.V. 1932. On the Analysis of Experimental Observations in Problems of Elastic Stability.
Proc. Roy. Soc. London, Series A, 135: 601–616.
Suzuki, H. & Horikawa, K. 1984. Welding to Pipe Column under Axial Compressive Load. Transactions
of JWRI 13: 151–159.
Trahair, N.S. & Kitipornchai, S. 1971. Elastic Lateral Buckling of Stepped I-beams. Journal of the Structural
Division, Proceedings of the ASCE.
Vild, M. & Bajer, M. 2016. Strengthening of Steel Columns under Load: Torsional-Flexural Buckling.
Advances in Materials Science and Engineering 2016(1). ISSN: 1687-8434. DOI: 10.1155/2016/2765821.

1242
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Design of gusset plate connection with single-sided splice member


by component based finite element method

M. Vild & J. Kabeláč


Brno University of Technology, Czech Republic

M. Kuříková & F. Wald


Czech Technical University in Prague, Czech Republic

ABSTRACT: This paper describes the design of the single-sided gusset plate connection of
a truss steel member by a component based finite element method. The elements are analyzed
by geometrically and materially non-linear analysis. The proper behavior of components, e.g.
of bolts, anchor bolts, welds, is treated by introducing components representing its behavior
in term of initial stiffness, ultimate resistance and deformation capacity. Research oriented
finite element model is validated on experiments. The models are analyzed by geometrically
and materially non-linear analysis with imperfections. The research oriented model is com-
pared with simplified design one, which includes only the joint and equivalent horizontal dis-
ruptive force. Contribution shows the current trends in advanced modelling of connection
components and differences of the research oriented and design oriented finite element
models. The upcoming models with coupling of member and its joints are shown on this
single-sided gusset plate connection.

1 INTRODUCTION

Typical case which combines the behavior of member and connection and is difficult to design
separately is the single-sided gusset plate connection of diagonal steel member. Kitipornchai
(1993) predicts three types of collapse of eccentrically connected diagonals. The first possible fail-
ure is the collapse of the rod itself, which deflects in a sinusoidal shape. The other two modes of
collapse are the failure of the joint plates without swinging and the failure of the joint plates with
swinging, Figure 1.
The Kitipornchai’s approaches were applied in models studied by Wilkinson et al. (2010), who
proposed modifications in the load capacity calculation. The analytical design methods developed
based on this principle by Khoo (2010) and Lutz (2005) are conservative, but served as fast and
relatively accurate in current practice and are commonly used (Dowswell, 2005), (Rodier & Cou-
chaux, 2014), (CDS348:2014) and Bardot et al. (2017). Fang et al. (2014) studied yielding patterns
and failure modes for different plate geometries and improved the design approach also for HSS.
The global analyses of steel structures are today carried out by finite element analysis (FEA) and
all the traditional procedures are not used any more. Currently, fast development of software abil-
ity of connection design by FEA and thousands of experiments are available for the validation
process. In such situation, the verification process performed through benchmark tests gains cru-
cial importance. The source and the extent of such benchmark tests for the field of structural con-
nections is yet to be established. To achieve this goal a set of small benchmark tests that can be
used as a reference in the verification process of simulations was developed.
The recommendation for design by advanced modelling in structural steel is ready to be
used in Annex C of EN 1993-1-5:2005. Development of modern general-purpose software and
decreasing cost of computational resources facilitate this trend. The FEA of structural joints
is the coming step in structural steel design. As the computational tools become more readily

1243
Figure 1. Failure mode of a single-sided gusset plate connection (Vesecký, 2019).

available and easier to use even by relatively inexperienced engineers, the proper procedure
should be employed when judging the results of computational analysis.
FEA for connections is used from 70s of the last century as research-oriented FEA
(ROFEA). Their ability to express real behavior of connections is making numerical experi-
ments a valid alternative to testing and source of additional information about local stresses.
Material model for FEA uses true stress-strain diagram. Validation & Verification (V&V)
process of models is integral part of the procedure (Wald et al. 2014), and the FEA studies
are based on the researcher’s own experiments. During preparation of CM for EN1993-
1-8:2006, all basic components were deeply modelled (Bursi & Jaspart, 1997). Special atten-
tion was given to modelling of the T-stub, which represents the end plate connections of
beam to column joints, beam splices and column bases (Virdi et al.1999). Last generation of
FEA models of connections is utilized in studies focused on application of high strength
steel (Coelho, 2013) and bolts in the connections (Moze & Beg, 2011). Prediction of hollow
section joints is based on experimental evidence approved by FEA numerical experiments;
see e.g. Partanen et al. (2001). Due to large variety of geometry, some types were studied
only numerically, Fleischer et al. (2010). V&V of the FEA design models (DFEA) of steel
connection design is native part of its preparation (Wald et al, 2014). The detailed procedure
for verification of CBFEM was prepared (Wald et al, 2019).

2 COMPONENT BASED FINITE ELEMENT METHOD

The collapse mechanism by two plastic hinges at a gusset plate and a connecting plate well
represent the situation, when the member buckling resistance is higher than the resistance of
the connection, see Figure 1.
The CBFEM model comprises only the joint with stubs of connected members (Legner,
2019). The rotations and torsion of the member are restrained. The member is loaded by
normal force N and to simulate the bending of the gusset and connected plates, eccentricities,

1244
Figure 2. CBFEM model of a single-sided gusset plate connection with supports and schema of bending
moment diagram.

Figure 3. Von Mises stress on specimen C2 and plastic strain on specimen D4, deformation scale 3.

1245
Figure 4. Specimens of circular hollow sections connected with eccentric gusset plates (Vesecký, 2019).

and second order effects, shear force is added, see Figure 3. The applied shear force is in the
position of the center of bolt group, see Figure 2. The disruptive force has the magnitude of

V ¼ N=10 ð1Þ

Design resistances calculated by CBFEM were validated to experimental results (Cábová


et al, 2019) and research oriented finite element model (Jandera et al, 2019). The members
CHS 102 × 4 were connected by single-sided gusset plates and connecting plates with the
thicknesses of 8 mm; see Figure 4 (Vesecký, 2019). The research oriented finite element model
was made in Abaqus software and validated on the experiments. ROFEM - DIC is using the
measured imperfections obtained by digital image correlation to apply the real geometrical
imperfections; ROFEM - EN is using imperfections according to EN 1993-1-5:2005. The ana-
lytical models are labelled KPA1 and KPA2. Model KPA1 is Khoo-Perera-Albermani (Khoo
et al. 2010) with minimal lengths of plastic hinges and minimal moments of inertias of gusset
and connecting plates; model KPA2 is using average moments of inertia. The calculated and
measured resistances are summarized in Figure 5.
The ROFEM was further used for parametric study. The variables were free length of the
connecting plate, thicknesses of gusset and connecting plate, number of bolts, and steel grade.
The comparison of load resistances using ROFEM and CBFEM is plotted in Figure 6. The
CBFEM is usually conservative except for specimens D2 and D4. However, for these speci-
mens, the ROFEM is conservative compared to the experiments. The CBFEM model is much
simpler and uses geometrically linear analysis but the accuracy of its results is comparable to
ROFEM and exceeds analytical models.

1246
Figure 5. Validation of resistance by ROFEM - DIC (measured imperfection) and EN (standard imper-
fection), CBFEM, and analytical models KPA1 and KPA2 to experiments.

Figure 6. Verification of CBFEM to ROFEM.

1247
Figure 7. a) Joint-member-joint subsystem with loads and supports; b) the modelled single-sided gusset
plate connection.

Figure 8. The shape of the first and second failure modes.

3 COUPLING OF A MEMBER AND ITS JOINT DESIGN

The single-sided gusset plate connection is a typical joint which should not be checked indi-
vidually but as a part of the joint-member-joint subsystem. The coming solution is the FEA
modelling of complex model of member with its joints – see Figure 7. Figures 8 and 9 shows
the influence of the thickness of the gusset plates on the buckling mode shape in CBFEM
model including the member. Members of cross-sections CHS 114.3 × 5.0 and CHS

1248
Figure 9. Influence of the gusset and connecting plate thicknesses on the first and second failure modes.

76.1 × 4.0 have theoretical length of 3 m and are connected at both sides by the single-sided
gusset plates with the thickness of 4, 6, 8, and 10 mm to beam of cross-section HEB 200.
Four bolts M16 are used. The lengths of the gusset plate and the connected plate are
130 mm with the gap of 20 mm. In the case of slender gusset plates, the member remains
undeformed and failure occurs at the connections. Only when the thickness of the gusset
plates is greater, the member buckles with the expected sinusoidal deflection shape. The
compressive load resistance of such subsystem is increasing with the plate thickness until it
reaches the buckling resistance of the member. Then it is climbing only very slowly due to
slight decrease in buckling length.

4 SUMMARY AND ACKNOWLEDGMENT

The conservative analytical models are developed and used for joints. They do not fully repre-
sent the influence and limits of interaction of a joint and a member.
The CBFEM model of connection taking into account the eccentricity may include the
geometry of joint and angle of a diagonal. The disruptive force with the magnitude of
V = N/10 represents the imperfections and second order effects.
The compressive resistance of the subsystem of the member including its joints is more pre-
cisely determined by modelling of the whole subsystem and reveals the behavior of the subsys-
tem as a whole.
The work was prepared under the R&D project supported by Technology Agency of the
Czech Republic, project Advanced procedures of steel and composite structure connections
design and production, No TJ01000045.

REFERENCES

Bardot, C. Cábová, K. Kurejková, M. Wald, F. 2017. Behaviour of a gusset plate connection under
compression. Civil Engineering Journal: 26 (1).
Bursi, O. S. & Jaspart, J. P. 1997. Benchmarks for Finite Element Modeling of Bolted Steel Connections,
Journal of Constructional Steel Research: 43 (1–3), 17–42.
Cábová, K., Jandera, M. Vesecký, J. 2019. Tests of gusset plate connection under compression, in print-
ing, in SDSS 2019: International Colloquium on Stability and Ductility of Steel Structures, Prague.

1249
CDS348. 2014. Design of eccentric gusset plate connections. SCI. Calculation sheet.
Couchaux, M. & Rodier, A. 2014. Eccentric bolted gusset plate of tube, Models for resistance in
compression, in Proceedings of Eurosteel, Naples, 125–129.
Dowswell, B. 2006. Effective Length Factors for Gusset Plate Buckling. Engineering Journal, 2, 91–102.
EN 1993- 1-8. 2006. Eurocode 3: Design of steel structures – Part 1-8: Design of joints, CEN, Brussels.
EN 1993-1-5. 2007. Eurocode 3: Design of steel structures - Part 1-5: Plated structural elements, CEN,
Brussels.
Fang, Ch. Yam, M.C.H., Cheng, R. J.J., Zhang, Y. 2015. Compressive strength and behaviour of gusset
plate connections with single-sided splice members, Journal of Constructional Steel Research, 106,
166–183.
Fleischer, O. Puthli, R. Wardenier, J. 2010. Evaluation of numerical investigations on static behav-
iour of slender RHS K-gap joints, in 13th International Symposium on Tubular Structures,
Hong Kong, 75–83.
Jandera, M. Cábová, K. Vesecký, J. 2019. Modelling of the eccentric gusset plate connection, in printing,
in SDSS 2019: International Colloquium on Stability and Ductility of Steel Structures, Prague.
Khoo, X. Perera, M. Albermani F. 2010. Design of eccentrically connected cleat plates in compression,
Advanced Steel Construction, 6(2),678–687.
Legner, Š. 2019. Eccentric gusset plate connection of diagonal steel member, in Czech, diploma thesis,
ČVUT, Prague.
Lutz, D.G. & Laboube, R.A. 2005. Behavior of thin gusset plates in compression, Thin-Walled Struc-
tures, 43(5),861–875.
Moze, P. & Beg, D. 2011. Investigation of high strength steel connections with several bolts in double
shear, Journal of Constructional Steel Research, 67/3, 333–347.
Partanen, T. Niemi, E. Liukku, H. et al. 2001. Transverse and axial load capacities of the chord in
X-joints of square hollow sections due to the interaction of brace and chord loads, in 9th International
Symposium on Tubular Structures, Karlsruhe, 195–201.
Vesecký, J. 2019. Buckling resistance of gusset plates, in Czech, diploma thesis, ČVUT, Prague.
Wald, F. et al. 2019. Benchmark cases for advanced design of structural steel connections, Prague, Česká
technika.
Wald, F. Kwasniewski, L. Gödrich, L. Kurejková, M. 2014. Validation and verification procedures for
connection design in steel structures, in Proceedings of Conference on Steel, Space and Composite
Structures. Singapore, 111–120.
Wilkinson, T. Stock, D. Hastie, A. 2010. Eccentric cleat plate connections in hollow section members in
compression. In Tubular Structures XIII. CRC Press, 197–203.

1250
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Beam-to-column joints for slim-floor systems in seismic zones:


Numerical investigations and experimental program

C. Vulcu, R. Don & A. Ciutina


Politehnica University of Timisoara, Romania

ABSTRACT: The slim-floor building system is attractive to constructors and architects due
to the integration of steel beam in the overall height of the floor, which leads to additional
floor-to-floor space, used mostly in acquiring additional stories. The concrete slab offers nat-
ural fire protection for steel beams, while the use of novel corrugated steel sheeting reduces
the concrete volume, and can replace the secondary beams. Currently the slim-floor solutions
are applied in non-seismic regions, and there are fe w studies that consider continuous or
semi-continuous fixing of slim-floor beams. The current ongoing study was carried out with
the aim to develop reliable end-plate bolted connections for slim-floor beams, that can be
applied to buildings located in seismic areas. For this purpose, a numerical program was car-
ried out in two stages: (i) calibration of a FE model based on a four point bending test of
a slim-floor beam; (ii) case study performed for the investigation of beam-to-column joints
with moment resisting connections between slim-floor beams and columns. The current paper
presents the main findings of the study, an overview of the experimental program, the main
conclusions and future research activities.

1 INTRODUCTION

The structural solutions provided by the usage of composite elements are regarded as an
effective method of enhancing structural performance. A series of advantages emerge as con-
crete, steel and additional components are integrated into a more resistant and ductile
member – Arcelor-Mittal (2016). In particular, the slim-floor building system is attractive to
constructors and architects due to the integration of steel beam in the overall height of the
floor, which leads to additional floor-to-floor space, used mostly in acquiring additional stor-
ies. The concrete slab offers natural fire protection to the steel beams, while the use of novel
corrugated steel sheeting reduces the concrete volume, and replaces the secondary beams (for
usual spans).
The slim-floor solutions are currently applied mostly in non-seismic regions – Hauf (2010),
Braun et. al (2014) – and there are few studies that consider continuous or semi-continuous
fixing of slim-floor beams. It was shown by Malaska (2000), that the semi-continuous joining of
slim-floor beams improves the flexural stiffness of the slim-floor beams and allows the use of
shallower beam and floor sections, and better performance of beams in service conditions by
reducing cracking, deflections and vibrational problems. Wang et. al (2009) and Bernuzzi et. al
(1995) showed that in case of increasing gravitational loads the continuous fixing of the slim-
floor beams can lead to ductile plastic hinges in both beam-ends and middle spans. In contrast,
the usual seismic behaviour rely on increased frame lateral stiffness and failure mechanisms by
dissipation of seismic input energy by plasticization of dissipative elements or connections. Con-
sequently, in case of Moment-Resisting-Frames (MRF) or dual frame configurations consider-
ing MRF contribution, the beams or the beam-to-column joints of MRF will dissipate energy
through plastic hinges. Therefore, the application of slim-floor beam systems in seismic zones

1251
Figure 1. Slim-floor system.

should consider moment-resisting connection with columns, thus developing hogging bending
capacity, too. However, certain aspects characteristic for slim-floor systems should be con-
sidered (see Figure 1):
– the concrete slab encases the top steel flange and requires one layer of hogging reinforcement;
– the natural bonding and/or the concrete dowels contribute significantly to steel-to-
concrete connection. In many cases there is no need for additional connectors;
– bottom part of steel profile is larger than the top flange in order to accommodate the con-
crete supporting system: shallow decking or precast concrete slabs.
The present study investigates the possibility to develop reliable connections for slim-floor
beams, in view of application to buildings located in areas with seismic hazard. The paper pre-
sents the finite element numerical investigations and the outcomes of the study. In a first stage,
a finite element numerical model was calibrated based on a four point bending test of a simply
supported slim-floor beam. Further, a case study was developed in view of investigation of con-
tinuous slim-floor beam-to-column connections under both sagging and hogging bending.

2 FEM CALIBRATION OF A SLIM-FLOOR BEAM

The current numerical program on slim-floor beam-to-column joints – was initiated through the
calibration of a finite element (FE) model based on the experimental investigation as detailed in
Hauf (2010) on a four-point bending test of a slim-floor beam. The numerical study validated the
accuracy of the FEM models used for materials, contacts and boundary conditions further used
in modelling the beam-to-column connection models. Thus, the information on the behaviour of
a composite element, the steel-to-concrete friction coefficient, modelling procedure, importance of
concrete dowel connectors and reinforcement, meshing and interactions were derived through
calibration. Detailed outcomes of the study can be found in Vulcu et. al (2017, 2018).

3 PRE-TEST NUMERICAL INVESTIGATIONS OF SLIM-FLOOR BEAM-TO-


COLUMN CONNECTIONS

3.1 Configuration of the slim-floor beam-to-column joint


The joint assembly was conceived as an extended end-plate connection and reduced beam section.
The reinforced concrete slab assures the continuity of the reinforcement by extension beyond the
column. Figure 2 shows the configuration of the investigated external beam-to-column joint
assembly as well as the joint components, concrete slab reinforcement and steel elements. The cur-
rent technical solution resulted based on the outcomes of the previous study - Vulcu et.al (2017).
The steel column is a HEB340 profile, while the steel beam is composed by a bottom steel
plate (PL-20×380 mm) welded to half of an IPE600 profile. The column length is of 3930 mm,
while the beam length is 2680 mm. A supplementary web plate and continuity plates were con-
sidered in the column web panel in order to limit the panel deformation. The connection
between the slim-floor beam and the column is realized as bolted extended end-plate connec-
tion using four bolt rows of M36-HR.10.9 (see Figure 2). Within the lower steel plate,

1252
Figure 2. Configuration of the beam-to-column joint model: overall joint, slab reinforcement, bolt rows
and dog bone in lower steel plate.

a reduced cross-section was considered with the aim to force the development of the plastic
hinge in the beam and assure a prevailing elastic response of the connection.
The concrete slab, integrating the steel beam and the reinforcement (transversal, longitu-
dinal and inclined), was considered with a width of 1500 mm and a height of 145 mm. The
effective width computed according to the norm EN 1994-1 (2004) was of 1200 mm. In order
to assure a continuous reinforcing over the connection zone, the slab was extended beyond the
column with 600 mm. Additional to transversal and longitudinal bars, inclined reinforcement
was used for the concrete slab (see Figure 2). The continuity of the longitudinal reinforcing
bars is assured around the column. The reinforcement bars considered in the analyses satisfy
the reinforcement connection conditions required in the Annex C of EN 1998-1-1 (2004). Con-
sequently, the longitudinal reinforcing bars are included to contribute to the negative bending
moment capacity within the connection zone. Concrete in the bottom troughs has been
ignored in the analysis and consequently not modelled as additional FEM analyses have
shown that its influence is insignificant in the overall resistance of joint. Concrete dowels were
considered, i.e. reinforcement of 12 mm diameter passing through 40 mm holes in the beam
web. The centre-to-centre distances of the perforations is 125 mm.

3.2 Modelling procedure


The numerical investigations of the slim-floor beam-to-column joint assembly (see Figure 2)
were performed by using Abaqus v6.13 software (2013). Finite beam elements were used for
the reinforcement, and solid elements for other components (bolts, plates, concrete, etc.). The
material characteristics were defined for the following: concrete (C30/37), structural steel
(S355), bolts (HR.10.9), and reinforcement bars (S400), considering both elastic and plastic
properties. Figure 3a illustrates the true stress - true strain curves (excepting elastic deform-
ation) for bolts (HR.10.9), reinforcement (S400) and structural steel (S355).
The material model for bolts was defined based on a previous calibration of a T-stub char-
acterized by failure mode 3 (i.e. bolt failure), see Dubina et. al (2015). For all steel elements,
the elastic modulus for steel was taken as 210 MPa, and the Poisson coefficient was 0.3. The
input for concrete material model is detailed in Vulcu et. al (2017), considering: Young modu-
lus E = 32.5 MPa, and Poisson coefficient ν = 0.2. The global mesh size was adapted to differ-
ent FE: reinforcing bars/20 mm; concrete/18 mm; steel profile/14 mm; bottom steel plate/
15 mm; column/13 mm; end plate/10 mm; column web plate and stiffeners/12 mm; bolts/
8 mm. The discretization of the beam-to-column joint assembly and its components are illus-
trated in Figure 3b. The boundary conditions for the column and beam considered: (i) at the

1253
Figure 3. (a) True stress-strain curve: steel (measured S355 –), reinforcement (nominal S400), bolts
(Gr.10.9); (b) discretization of the beam-to-column joint assembly/connection.

top and bottom end of the column - a simple and respectively a fixed support; (ii) at the tip of
the beam the load was applied in displacement control, inducing sagging/hogging bending
moment within the connection.

3.3 Numerical results and parametric study


The numerical models of the beam-to-column joint assemblies were subjected to negative and
positive bending moment. Table 1 presents the overview of the studied joint models.
The graphic response, respectively the stress and strain distribution of the numerical investi-
gations are presented for the following configurations: (i) reference model (see Figures 4a–7a);
(ii) joint assembly without dog-bone in the lower steel plate (see Figures 4b–7b). The results
are presented in terms of moment-rotation curves under positive and negative bending (Figure
4a and 4b) and stress and plastic strain response (Figures 5–7).
As can be observed, the reference model presents a balanced behaviour under hogging
and sagging (see Figure 4a) both in terms of moment resistance and initial stiffness, with
a plastic descending behaviour but proving an important ductility. The failure mechanism
was characterized by the formation of the plastic hinge in the steel beam under both sagging
and hogging moment.
On the other hand, in case of the configuration without dog-bone (model M2), a small
increase of capacity was evidenced under both positive and negative bending moment (see
Figure 4b). However, under positive bending moment a brittle failure mode was recorded
by breaking of the bolts in tension (see Figure 7b), reducing the overall joint rotation
capacity.
In order to optimise the response of the joint, several parameters were considered, as
described in Table 1. The importance of each parameter is highlighted below:

Table 1. Investigated numerical models.


Model Description Loading

M1 Reference model M+ M–
M2 Influence of dog-bone, i.e. joint model without dog-bone M+ M–
M3 Partial interaction - reduction of the concrete dowels M+ M–
M4 Influence of the end plate type, i.e. flush plate at the top - M–
M5 Influence of the reinforcement amount (6/20 mm diameter) M+ M–
M6 Influence of the concrete class (C20/25; C40/45) M+ M–
M7 Influence of the concrete slab, i.e. joint model without slab (only steel) M+ M–

1254
Figure 4. (a) Reference model M1: moment-rotation curves; (b) Comparison of M1 and M2 models –
illustrating the influence of the reduced section in the lower steel plate of the beam.

Figure 5. (a) Reference model M1: response under negative bending moment (stress/plastic strain);
(b) Model without dog-bone M2: response under negative bending moment.

Figure 6. (a) Reference model M1: response under positive bending moment (stress/plastic strain);
(b) Model without dog-bone M2: response under positive bending moment.

1255
Figure 7. (a) Reference model M1: failure mechanism under negative and positive bending moment;
(b) Model without dog-bone M2: failure mechanism under negative and positive bending.

– the influence of the reduced section in the lower steel plate – proved to be positively
improve the joint behaviour, allowing the formation of the plastic hinge in the slim-floor
beam under both sagging and hogging bending moment;
– the steel-to-concrete connection degree was studied by lowering the considered number of
concrete dowels. Thus, the influence of the number of concrete dowels proved to be negli-
gible. The reduction of the concrete dowels (by 50% and 66%) did not affect the moment-
rotation curves for the considered planar T-shape joint assembly with slim-floor beam. How-
ever, in the case of a full length beam with end-supports, this observation is expected to be
different;
– the influence of the end-plate typology is important, as can be noticed in Figure 8.
Through the use of a flush plate at the top, the joint resistance to hogging bending was
directly influenced by the absence of the bolt-row in the extended end-plate (model M4),

Figure 8. Influence of the end plate type: (a) reference model with extended end-plate vs. joint model
with flush plate at the top; (b) plastic strain recorded for M4 joint model illustrating the failure mode.

Figure 9. Influence of reinforcement amount: (a) reference model (longitudinal reinforcement as:
2Φ20mm rebars adjacent to the steel beam +, 10Φ20mm rebars towards slab borders) vs. joint model with
12Φ6mm longitudinal rebars; (b) reference model vs. joint model with 12Φ20mm longitudinal rebars.

1256
Figure 10. Influence of the concrete class: (a) reference model (with C30/37 concrete) vs. joint model
with C20/25 concrete class; (b) reference model vs. joint model with C40/45 concrete class.

Figure 11. Influence of the presence of the concrete slab in hogging (a), and respectively in sagging (b).

and lead to a reduction of 30% in capacity. The stiffness was slightly affected as well. The
failure mechanism consisted in the brittle failure of the bolt row in tension;
– the influence of the longitudinal reinforcement diameter was observed to slightly affect
the response under negative bending moment (Figure 9). In particular, a reduction of
the reinforcing diameter to 6mm was accompanied by a resistance reduction of 7%,
while the increase of the bar reinforcing diameter from 12 to 20 mm, lead to an increase
of resistance of 3%;
– the influence of the concrete class – was observed to be insignificant under both positive
and negative bending moment (see Figure 10), as the global joint response was not
affected (differences were less than 2%);
– the presence of the concrete slab – was observed to be important for the response of the
beam-to-column joint assembly as shown in Figure 11. Under both, positive and negative
bending moment, the moment-rotation curves evidenced a significant reduction of the cap-
acity (25%) and stiffness (40%) – in case the concrete slab was not present. In spite of this,
the moment-rotation curves from the reference model and the bare-steel joint seemed to
converge for rotations higher than 0.11 radians. The gain in hogging resistance is mainly
due to the contribution of the reinforcing bars to the moments, while in sagging the gain is
due to the increased level-arm by shifting the centre of compression into concrete.

4 EXPERIMENTAL PROGRAM

4.1 Overview: Specimen configuration, test set-up, instrumentation, loading protocols


The experimental investigations are in preparation, thus only a brief overview is offered in
current paper. Similar to the FE models, the specimens are planar T-shape beam-to-column

1257
Figure 12. (a) Bare steel joint assembly; (b) overview of the reinforcement arrangement; (c) overview of
the specimen configuration and the experimental test set-up.

joints (see Figure 12). The technical solution adopted for the connection between beam and
column represents the model M1 and is shown in Figure 12a. The reduced section in the lower
steel plate is aimed for the development of the plastic hinge in the beam. A top view of the
slab’s reinforcement arrangement is shown in Figure 12b. The following reinforcing bars were
set: (i) longitudinal rebars (continuous around the steel column); (ii) transversal rebars (placed
on the upper beam flange); (iii) inclined rebars that assure the connection between slab and
the concrete region between steel flanges; (iv) rebars passing-through the beam, forming the
concrete dowels and assuring the composite action between steel beam and concrete slab.
Figure 12c illustrates the testing set-up and of the beam-to-column joint specimen. The load
will be applied at the tip of the column through a hydraulic actuator. A pinned support was
considered at the base of the column, respectively a pendulum with pinned supports was
adopted for the beam. In order to avoid the out of plane deformations during the test, an out
of plane system is conceived.
Global as well as local instrumentation will be considered in order to characterize the
response of the joint assembly during the tests in order to assure the measurement of: ▪ force
in the actuator; ▪ displacement at the tip of the column; ▪ displacement at the supports as well
as ▪ deformation of the column web panel; ▪ deformation in the dissipative zone (measurement
in the reduced beam section and concrete slab); ▪ relative displacement between steel beam
and concrete slab; ▪ force in the bolts through the use of strain gages placed in the bolts (in
drilled holes of 3 mm diameter).
In order to have both monotonic and cyclic response of the joints, two specimens are con-
sidered i.e. one monotonic and one cyclic. The cyclic loading procedure will consider increas-
ing amplitude cycles.

5 CONCLUSIONS

The presented study was performed with the aim to develop reliable connections between
slim-floor beams and columns – for application in steel structures located in seismic zones. In
a first stage, a numerical model was calibrated based on a four point bending test of a slim-
floor beam. In a second step, a case study was performed for the investigation of slim-floor
beam-to-column joint configurations. The FEM investigations lead to the following
conclusions:
– in seismic regions it is possible to rely on the full or semi-continuity of joints in the global
failure mechanism of MRF or dual steel structures with slim-floor systems;
– the failure mode of the joint configuration with reduced beam-section in the lower steel
plate is characterized by a ductile formation of the plastic hinge in the beam. In contrast,

1258
the configuration without dog-bone lead to the failure of bolt rows in tension (brittle fail-
ure). It is to be stressed that the design avoided a connection typology characterized by
brittle failure of the bolts. Consequently, the adopted failure criterion and the modelling
of the post-failure behaviour – developed based on the outcomes of Dubina et. al
(2015) – are not relevant. In this view, the bolts should be characterized by an overall
elastic response;
– the influence of concrete slim-floor slab is effective in sagging bending as it contributes to
the global increase of both stiffness and bending capacity. In hogging its influence is less
important and the connection characteristics are mainly based on steel components
including reinforcement;
– the presence of the reinforced concrete slab lead to an increase of capacity and stiff-
ness. The inclined reinforcement and the concrete dowels contributed to the load
transfer mechanism by connecting the concrete slab to the concrete within flanges.
A significant increase of longitudinal reinforcement will lead to higher capacity
under negative bending moment;
– the slope of the moment-rotation curves in the post-peak range slightly decreases due to
the reduction of the lever arm in the connection zone due to concrete crushing (i.e.
change of the compression centre location).
Based on the current investigation, it is proven that the slim-floor beams can be adapted to
seismic-resistant structures and the key aspect is related to the behaviour of slim-floor beam-
to-column joints. The ongoing and future research activities will involve: (i) experimental
investigations; (ii) calibration of numerical models and parametric study for improved slim-
floor configurations and (iii) structural numerical analyses for improving the applicability of
such systems.

REFERENCES

Abaqus v6.13 [Computer software]. Dassault Systèmes, Waltham, MA.


ArcelorMittal, 2016. Slim floor – An innovative concept for floors. Luxembourg: ArcelorMittal.
Bernuzzi, C., Gadotti, F., Zandonini, R. 1995. Semi-continuity in slim floor steel–concrete composite sys-
tems. 1st European Conference on Steel Structures. Eurosteel 1995.
Braun, M., Obiala, R., Odenbreit, C., Hechler, O. 2014. Design and application of a new generation of
slim-floor construction. 7th European Conf. on Steel & Composite Structures, Naples, Italy.
Dubina, D., et al. 2015. High strength steel in seismic resistant building frames. Final Report. Grant No.
RFSR-CT-2009-00024, RFCS Publications, European Commission, Brussels, Belgium.
EN 191994-1-1, 2004. Eurocode 4: Design of composite steel and concrete structures – Part 1: General rules
and rules for building. Brussels, Belgium.
EN 191998-1-1, 2004. Eurocode 8: Design of structures for earthquake resistance – Part 1: General rules,
seismic actions and rules for buildings. Brussels, Belgium.
Hauf, G. 2010. Trag- und Verformungsverhalten von Slim-Floor Trägern unter Biegebeanspruchung.
Malaska, M. 2000. Behaviour of a semi-continuous beam-column connection for composite slim floors.
Ph.D. Thesis, Espoo, Finland, ISBN 951-22-5224–4.
Vulcu, C., Don, R., Ciutina, A., Dubină, D. 2017. Numerical investigation of moment-resisting slim-
floor beam-to-column connections. 8th International Conference on Composite Construction in Steel
and Concrete (CCVIII 2017), July 30 – August 2, Spring Creek Ranch in Jackson, Wyoming (USA).
Vulcu, C., Don, R., Ciutina, A. 2018. Semi-continous beam-to-column joints for slim-floor systems in
seismic zones. 12th International Conference on Advances in Steel-Concrete Composite Structures
(ASCCS 2018). Universitat Politècnica de València, València, Spain, June 27–29. (Doi: http://dx.doi.
org/10.4995/ASCCS2018.2018.7199).
Wang, Y., Yang, L., Shi, Y., Zhang, R. 2009. Loading capacity of composite slim frame beams. Journal
of Constructional Steel Research 65.

1259
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

A reexamination of high strength steel Q690 plasticity model

Yuanzuo Wang, Yanbo Wang, Guoqiang Li & Yifan Lyu


Tongji University, Shanghai, China

ABSTRACT: The most generally used yield criterion of typical metals, the von Mises yield cri-
terion, shows that the effects of stress triaxiality and Lode angle on the steel plasticity model are
negligible. However, recently numerous experimental researches have shown that the influence of
the stress triaxiality and Lode angle on the plasticity model of some typical metals should be con-
sidered. In the present paper, experiments and finite element analyses are carried out to evaluate
the effect of stress triaxiality and Lode angle on the plastic behavior of Chinese high strength steel
(HSS) Q690, and the application of von Mises yield criterion of the is reexamined. Results dem-
onstrate that an appropriate yield criterion of HSS Q690 should consider effects of the Lode
angle and the influence of stress triaxiality is negligible. Based on experimental and numerical
results, a new yield function of HSS is proposed, to essentially simulate the experimental results.

Keywords: High strength steel, Yield criterion, Stress triaxiality, Lode angle

1 INTRODUCTION

The plastic behavior of metals is characterized by the constitutive model which is composed of
yield function, flow criterion and hardening rule. The importance of yield function for actual
engineering computations has been well noted. The von Mises yield criterion is generally adopted
to describe the plasticity of typical metals. There is one basic tenet of von Mises yield model: the
influences of hydrostatic pressure and Lode angle on metal plasticity are negligible. Based on this
assumption of the von Mises yield criterion, it is generally accepted that the yield and flow stresses
are identical in compression and tension. However, in a number of previous publications [1–10],
the yield strength of some real metal materials, is greater in compression than in tension at the
given strain, which is termed as strength differential effect. The strength differential effect is
induced by the dependence of the yield condition on the hydrostatic pressure and the Lode angle.
Therefore, the von Mises yield function is fail to describe this phenomenon and it has been well
demonstrated that the dependency of plasticity model of metals on the hydrostatic pressure and
Lode angle should be reexamined [11–15].
The influences of the stress triaxiality and Lode angle on plasticity model of soil and rocks has
been demonstrated in literatures [16–18]. Generally, the stress triaxiality controls the shape of the
yield locus on the meridian plane and the Lode angle controls the shape on the deviatoric plane.
For typical metals, Spitzig and Richmond first found the plastic behavior of aluminum alloys is
influenced by hydrostatic pressure, especially under high confining pressure. Richmond et al. [19]
found that the yield strength of four types of steels (4330, 4310, maraging steel and HY80) is
a linear function of hydrostatic pressure. Moreover, Richmond and Spitzig demonstrated that
a classical yield function proposed by Drucker and Prager is applicable for iron-based materials.
The classical plasticity model of geomaterials has long recognized the Lode angle effect and
Tresca yield function is a typical one. Recently a number of experiments have shown that the
influence of the Lode angle on the plasticity model of some typical metals should be considered.
The high strength steel (HSS) with a nominal yield stress not less than 460 N/mm2 has been
increasingly used in engineering structures, especially in high-rise buildings and bridges. Therefore,

1260
an accurate description of the HSS plastic behavior is significant for computational simulation
and structure design.
In the present paper, the applicability of von Mises yield criterion for Chinese Q690 HSS is
examined firstly. Based on experiments on five types of testing specimens, it is found that the
von Mises yield function, calibrated from the smooth round bar tensile test, fails to simulate
the real response of some other experiments. Results demonstrate that an appropriate yield
function of the HSS should consider effects of the Lode angle and the influence of hydrostatic
pressure is negligible. A new yield function of Chinese HSS Q690 are proposed and calibrated.

2 FUNDAMENTAL DEFINITION

The stress state can be described in terms of three stress tensor invariants and principal stresses
(σ1 , σ2 , σ3 ) are often adopted. An arbitrary stress state can be represented geometrically in the
Cartesian coordinates which take principal stresses as the axis, as shown in Figure 1. The Carte-
sian coordinates can also be transferred to cylindrical coordinates (σm , θ, σeq ) which defined by

σm ¼ I1 =3 ¼ ðσ1 þ σ2 þ σ3 Þ=3 ð1Þ


h i
σ Þ3
θ ¼ 1=3  arccos ðr= ð2Þ
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
h iffi
σ ¼ 1=2 ðσ1  σ2 Þ2 þ ðσ2  σ3 Þ2 þ ðσ3  σ1 Þ2 ð3Þ

where

r ¼ ½27=2ðσ1  σm Þðσ2  σm Þðσ3  σm Þ


1=3 ð4Þ

The parameter θ (0  θ  π=3) is often referred to as the Lode angle and σ is the von Mises
equivalent stress. The Cartesian coordinates can also be transferred to spherical coordinates
which adopt (η, θ, σ) as axis. The parameter η is often adopted to describe triaxiality of stress
state, defined by

η ¼ σ m =
σ ð5Þ

Figure 1. Geometry representation of Cartesian, cylindrical and spherical coordinates.

1261
3 APPLICABILITY OF VON MISES YIELD CRITERION FOR HSS Q690

The HSS Q690 is investigated in the present study. The experimental program, as summarized
in Table 1, is conducted to examine the applicability of von Mises yield criterion for HSS. In
addition, the corresponding initial values of stress triaxiality and Lode angle at the crack initi-
ation set are calculated using Bai’s formulas and listed in Table 1. All experiments were con-
ducted in Tongji University, Shanghai, China.

3.1 Smooth round bar tensile test


For the von Mises yield function, there is only hardening function σðεÞ needed to be cali-
brated. In this study, because two different thicknesses (10mm and 30mm) of steels are used to
manufacture specimens, we design two types of uniaxial tensile tests (flat specimen with rect-
angular cross-section for 10mm steel plate and smooth round bar tensile test for 30mm steel
plate) to obtain the hardening functions. The dimensions of the flat specimen and smooth
round bar are shown in Figure 2. All uniaxial tensile tests are performed on a MTS machine
with the loading speed is 0.3mm/min to achieve the quasi-static loading condition. The extens-
ometer with length of 20mm is assembled on the specimen to measure the elongation δ. The
typical load-elongation curves of Q690 steel are shown in Figure 3.
The engineering stress-strain curve is given by

σE ¼ P=A0 ð6Þ

εE ¼ δ=δ0 ð7Þ

Table 1. Overview of experimental specimens.


Loading type Specimen type ID Repeat η0 θ0

Smooth round bar SRB 3 1=3 0


Flat specimen FS 3 1=3 0
NRB-1 3 1.026 0
Monotonic uniaxial tension
Notched round bar NRB-2 3 0.739 0
NRB-3 3 0.556 0
Pure shear specimen PS 3 0 π=6
Monotonic uniaxial compression Cylinder specimen CS 3 -1=3 π=3

Figure 2. Dimensions of flat specimen and smooth round bar.

1262
Figure 3. A comparison of load-elongation curves between experiments and FE analyses.

where A0 is the initial area of the minimum cross-section and δ0 is the initial gauge section
length. Based on engineering stress-strain curve, the true stress-strain can be calculated by

σT ¼ σE ð1 þ εE Þ ð8Þ

εT ¼ lnð1 þ εE Þ ð9Þ

Because these two transformation equations are based on incompressible assumption, Eq. (8)
and Eq. (9) are no longer valid in the post-necking phase in which severe non-uniform deform-
ation occurs in the necking region. In order to solve this problem, a number of correction
methods are proposed in previous publications. In the present study, a power function is used
to fit the true stress-strain curve obtained from pre-necking phase, and then the approximate
true stress-strain curve after necking can be obtained by extrapolating the fitting curve. The
true stress strain curves for two different thickness steel plates are plotted in Figure 4. The
finite element analyses are carried out to verifies the hardening function by examining the dis-
crepancy between experimental data and simulation results of tensile tests. Due to the sym-
metrical shape of flat specimen and axisymmetric shape of smooth round bar, only 1/8 of the

Figure 4. True stress-strain curves of HSS.

1263
Table 2. Dimensions of notched round bar (Q690).
ID R0 (mm) D0 (mm) L0 (mm)

NRB-1 2.5 10.03 5.02


NRB-2 5.0 10.04 8.61
NRB-3 10.0 10.03 13.27

Figure 5. Dimensions of notched round bar.

full flat specimen is modeled using solid elements (C3D8R) and a quarter of full smooth
round bar is modeled using axisymmetric elements (CAX8R). After comparing different mesh
density, a fine mesh size is selected and the necking region is further refined with the minimum
mesh size of 0.2mm and 0.05mm, respectively. As illustrated in Figure 3, the load-elongation
curves of smooth round bar simulation agree with the corresponding experimental results
well, which implies the calibrated mechanical properties of HSS Q690 are valid.

3.2 Notched round bar tensile test


A group of notched round bar tensile tests are conducted to investigate the plastic behavior
of HSS Q690 in different triaxiality stress states. It is worth mentioned that the Lode angle θ of
notched round bar is identical with that of smooth round bar, but the corresponding ranges of
triaxiality are different. The range of triaxiality depends on the radii of the notch of the notched
round bar. In present study, three types of radii of the notches are machined: 2.5mm, 5.0mm and
10.0mm. Actual dimensions of notched round bars are listed in Table 2 and the other dimensions
are identical to Figure 5. These notched round bars are all manufactured from steel plates with
30mm thickness. Due to the symmetrical shape of the smooth round bar, only a quarter is mod-
eled by using axisymmetric elements (CAX8R) with finer mesh in the notched region with the
minimum mesh size of 0.05 mm.
The extensometer is assembled on the specimen to measure the elongation δ of the gauge sec-
tion with 20mm length. The typical load-elongation curves of Q690 steel are shown in Figure 6. It
can be observed that the analysis results by using von Mises criterion gives a well prediction of
the actual load-elongation response of the notched round bar tensile tests. The stress triaxiality on
the minimum cross-section of the notched round bar tends to be higher than that of the smooth
round bar, but there is no distinct discrepancy appears between the analysis results by using von
Mises criterion and the actual responses. It indirectly implies the influence of stress triaxiality on
the plasticity model of Q690 steel is negligible.

3.3 Cylinder specimen compression test


The von Mises yield criterion assumes that there is no difference between compressive and
tensile yield strength at the given strain. In order to check whether the Q690 steel shows
strength differential effect, another group of cylinder specimen compression tests are con-
ducted. The cylinder specimen compression test corresponds to stress state of θ ¼ π=3
and η 5 0. In order to prevent buckling failure of the specimen, the ratio of height H to diam-
eter D of the specimen is set as 2.0 as shown in Figure 7. Lubricant is applied to the machine-
specimen interfaces to reduce the friction effect. The extensometer with length of 10mm is
assembled on the specimen to measure the elongation δ. The FE model of the cylinder is

1264
Figure 6. Notched round bar load-elongation curves for Q690.

Figure 7. Dimension of cylinder specimen.

discretized using 4-node axisymmetric elements with the minimum mesh size of 0.05mm. The
numerical simulation results are compared to the experimental results in Figure 8.
It is observed that the FE results using von Mises yield function underestimates the experi-
mental response with 10% error, which means the strength differential effect exists for Q690
steel but von Mises yield function fails to characterize it. The strength differential can arise
from the dependence of yield condition on stress triaxiality or Lode angle. Because it has been
demonstrated the stress triaxiality independence of HSS Q690 based on notched round bar
tensile tests, the discrepancy between von Mises result and experimental result implies the
yield strength of the Q690 steel is greater in uniaxial compression with θ ¼ π=3 than in uni-
axial tension with θ ¼ 0 at the given strain.

3.4 Pure shear test


The specimen configuration for pure shear loading is shown in Figure 9. During the loading
process, an upward boundary condition is applied at the top and the bottom is fixed. It is
observed that severe deformation occurs in the gauge section and the crack initiates at the pure
shear section. The extensometer is assembled on the specimen to measure the elongation δ of
the gauge section with 20mm length. Corresponding numerical simulations, the pure shear spe-
cimen is modeled using 8-node solid elements with finer mesh at the pure shear section.
As shown in Figure 10, it can be observed that the simulation results by using von Mises criter-
ion overpredicts the actual load-elongation response of the pure shear test with 15% error. It is
worth mentioned that the Lode angle on the pure shear section of the pure shear specimen
(θ ¼ π=6) is different with that of the smooth round bar (θ ¼ 0). Because it has been demonstrated
that the influence of the stress triaxiality on plasticity model is negligible, the reason of this discrep-
ancy is differences between pure shear strength (θ ¼ π=6) and tensile strength (θ ¼ 0), which
means that the influences of Lode angle on the plasticity model of HSS Q690 is not negligible.

1265
Figure 8. Cylinder specimen load-elongation curves for Q690.

Figure 9. Dimension of pure shear specimen.

Figure 10. Pure shear specimen load-elongation curves for Q690.

4 A NEW PLASTICITY MODEL OF HSS Q690

According to the experimental results, it has been demonstrated that the Lode angle effect
should be considered in the plasticity model of HSS Q690 and influence of stress triaxiality is
negligible. Therefore, a new yield function which considers the Lode angle effect, is adopted
to predict plastic behavior of HSS Q690 and defined as

1266
h  π πi

f ¼ σ  σðεÞ Cs þ ðCθ  Cs Þ  θ   ð10Þ
6 6
8 π
>
< 1 for 0  θ 
6
Cθ ¼ ð11Þ
: Cc for π < θ  π:
>
6 3

There are two parameters Cc = 0.92 and Cs = 1.08 in the new proposed yield function. Accord-
ing to results of five types experiments mentioned in section 3, these parameters are calibrated
using the inverse method to minimize discrepancy between experimental and simulated load-
elongation curves. The numerical results using the new proposed plasticity model are plotted in
Figures 11 and 12. By examining the discrepancy between experimental data and simulation
results, the new plasticity model gives better prediction of plastic behavior of HSS Q690 with an
error less than 2%.

Figure 11. Cylinder specimen load-elongation curves.

Figure 12. Pure shear specimen load-elongation curves.

1267
5 CONCLUSION

Based on a series of experiments with 4 specimen geometries, the plasticity model has been
reexamined in case of Q690 steel. According to the experimental and numerical results, the
following conclusions can be drawn: The von Mises yield function, a classical metal plasticity
model which assumes the stress triaxiality and Lode angle have no effect on yield strength, is
no longer applicable well for high strength steel Q690, especially in shear and compressive
condition. The plasticity model of HSS Q690 should considers effects of the Lode angle and
the influence of hydrostatic pressure is negligible. The new proposed yield function with con-
sideration of Lode angle effect can give more accurate prediction of the actual response.

ACKNOWLEDGMENTS

Financial support by the National Key Research and Development Program of China (Project
No. 2018YFC0705505) is greatly acknowledged.

REFERENCES

[1] Chait, R., Factors influencing the strength differential of high strength steels. Metallurgical &
Materials Transactions B, 1972. 3(2): p. 369–375.
[2] Spitzig, W.A., R.J. Sober and O. Richmond, Pressure dependence of yielding and associated
volume expansion in tempered martensite. Acta Metallurgica, 1975. 23(7): p. 885–893.
[3] Meyer, L.W. and S. Abdel-Malek, Strain rate dependence of strength-differential effect in two
steels. Journal De Physique IV, 2000. 10(PR9): p. 63–68.
[4] Holmen, J.K., et al., Strength differential effect in age hardened aluminum alloys. International
Journal of Plasticity, 2017.
[5] Singh, A.P., et al., Strength differential effect in four commercial steels. Journal of Materials Sci-
ence, 2000. 35(6): p. 1379–1388.
[6] Spitzig, W.A., R.J. Sober and O. Richmond, The effect of hydrostatic pressure on the deformation
behavior of maraging and HY-80 steels and its implications for plasticity theory. Metallurgical
Transactions A, 1976. 7(11): p. 1703–1710.
[7] Spitzig, W.A. and O. Richmond, The effect of pressure on the flow stress of metals ☆. Acta Metal-
lurgica, 1984. 32(3): p. 457–463.
[8] Rauch, G.C. and W.C. Leslie, The extent and nature of the strength-differential effect in steels.
Metallurgical Transactions, 1972. 3(2): p. 377–389.
[9] Chait, R., The strength differential of steel and Ti alloys as influenced by test temperature and
microstructure. Scripta Metallurgica, 1973. 7(4): p. 351–354.
[10] Leslie, W.C. and R.J. Sober, The strength of ferrite and of martensite as functions of composition,
temperature and strain rate. ASM-Trans, 1967. 60(3): p. 459–484.
[11] Wilson, C.D., A Critical Reexamination of Classical Metal Plasticity. Journal of Applied Mechan-
ics, 2002. 69(1): p. 63–68.
[12] Voyiadjis, G.Z., S.H. Hoseini and G.H. Farrahi, A Plasticity Model for Metals With Dependency
on All the Stress Invariants. Journal of Engineering Materials & Technology Transactions of the
Asme, 2013. 135(1): p. 279–284.
[13] Yoshida, K., A. Ishii and Y. Tadano, Work-hardening behavior of polycrystalline aluminum alloy
under multiaxial stress paths. International Journal of Plasticity, 2014. 53(53): p. 17–39.
[14] Campanelli, F., A J2–J3 approach in plastic and damage description of ductile materials. Inter-
national Journal of Damage Mechanics, 2016. 25(2): p. 89–91.
[15] Keralavarma, S.M., A multi-surface plasticity model for ductile fracture simulations. Journal of the
Mechanics & Physics of Solids, 2017. 103: p. 100–120.
[16] Bardet, J.P., Lode Dependences for Isotropic Pressure-Sensitive Elastoplastic Materials. Journal of
Applied Mechanics, 1990. 57(3): p. 498.
[17] Menetrey, P.H., Triaxial Failure Criterion for Concrete and Its Generalization. Aci Structural Jour-
nal, 1995. 92(3): p. 311–318.
[18] Bigoni, D. and A. Piccolroaz, Yield criteria for quasibrittle and frictional materials. International
Journal of Solids & Structures, 2004. 41(11): p. 2855–2878.
[19] Richmond, O. and W.A. Spitzig, PRESSURE DEPENDENCE AND DILATANCY OF PLASTIC
FLOW. 1980.

1268
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Analysis of mechanical properties of cold formed high strength


steel in weld area

Martin Werunský & Jakub Dolejš


Faculty of Civil Engineering, CTU in Prague, Prague, Czech Republic

ABSTRACT: This research deals with mechanical properties of cold formed area of high
strength steel STRENX S960E, which was subjected to welding. Recent standard EN 1993-1-8
deals with welding of high strength steel up to the grade S700 and with a welding in a cold
formed area for only the mild steels. Main purpose of this paper is description of toughness,
hardness, stress-strain and metallography of 10 mm thick steel sheet made of high strength steel
STRENX S960E, which had been cold formed (bent) and subsequently steel sheet of the same
material has been welded to the cold formed area by MIG welding procedure. All above men-
tioned properties will be investigated with respect to bending radius, welding parameters and
filler material matching grade.

1 INTRODUCTION

In civil engineering steels with a yield strength higher than 550 MPa and tensile strength higher
than 700 MPa are considered as high-strength steels (HSS). Thanks to a specific manufacturing
process and an exact amount of alloys added in the steel during this process, a fine microstruc-
ture of the HSS is achieved. Thereby it is possible to produce HSS with high strength, excellent
toughness and adequate formability and weldability. However, weldability and formability
decrease with increasing strength of the steels generally. Utilization of HSS allows to construct
members with less dead weight, higher load capacity and thereby with lower purchase cost.
In Europe, HSS with yield strength up to 700MPa are used for long span bridges especially
and for high buildings. However, new design codes based on a recent research for HSS
(460-700 MPa) were published recently only. Before that, same mechanical properties of mild
steels were considered also for HSS. Therefore, there was no motivation for their application
in civil engineering, because it was not possible to fully utilize their potential and thereby cost
of the structures could not be sufficiently reduced.
One can conclude, that HSS (even with yield strength higher than 700 MPa) will be widely
used in civil engineering in Europe in near future, which proofs many intensive researches all
around Europe, which aim is to supress negative properties of HSS, in order to update recent
Eurocodes so that a full potential of HSS in civil engineering can be utilized. Logically, next
grade of HSS that is going to be implemented in Eurocodes is S960 grade steel.

2 STATE OF ART

Unlike for welding of conventional steels, welding of high strength steels requires to follow strict
manufacturing rules. Non-compliance of those rules, which are sometimes contraindicating each
other, leads to a deterioration of mechanical properties of welded joint or sometimes to its com-
plete failure. Final mechanical properties and microstructure of weld and heat affected zone
(HAZ) depends primarily on an amount of heat input, t8/5 cooling time, amount of added alloys,
on consumable matching, pre-heat temperature and on a welding process. All these parameters
relate to each other and cannot be determined separately (Tornlom, 2007; Brtnik, 2016).

1269
Generally, HSS is necessary to weld with the lowest possible heat input. Welding with the
heat input higher than prescribed by manufacturer leads to prolonging of t8/5 cooling time,
thus forming undesirable microstructure such as bainite or ferrite in the weld or/and in HAZ
and to its coarsening which causes the deterioration of mechanical properties – strength,
microhardness and toughness (Cui, 2016). High heat input also leads to extending the width
of HAZ and soft zone (Rodriguez, 2004; Erns, 2013). However, low heat input can cause lack
of fusion and initiation of hot cracking. It is worth to mention that the higher strength of the
HSS, more strict rules need to be satisfied (Brtnik, 2016).
In civil engineering the emphasis is put on the ratio of the yield strength of the base material
(BM) and the filler material (FM) the so called evenmatch, overmatch, undermatch. Utilization
of the filler material with the lower yield strength than the yield strength of the base material
(undermatch), leads to a deterioration of the overall strength of the welded joint (Brtník, 2016;
Gunter, 2013; Kuhlmann, 2014). On the other hand, its ductility increases, which is crucial to
fulfil the strict requirements of the Eurocodes for a plastic design. Utilization of undermatching
filler material has also a beneficial effect on residual stresses caused by welding, because the
lower the yield strength of the filler material is, the lower residual stresses are present in the weld
joint. Based on number of researches, it was determined that if multipass undermatch welds are
made, it is appropriate to make root welds with lower yield strength than the filler weld, which
leads to overall better ductility of the weld joint, but to its less strength limitation (Wang, 2017).
If welding HSS another important aspect that needs to be taken to account is an area of the
lowest hardness (strength) in HAZ called “soft zone”. The soft zone is defined as the width of
the HAZ where hardness values are up to 90% of the unaffected parent material and is created
during the welding process by tempering the martensitic microstructure in sub-critically HAZ
(SCHAZ), which was thermally affected by welding. However, some researches came up with
a finding, that for the specific ratio of the width of the soft zone and thickness of the steel
sheet, negative properties of the soft zone are negligible because of the constraint effect
(Rodriguez, 2004; Erns, 2013).
Despite the fact HSS perform lower elongation compared to mild steels, it is possible to subject
them to cold forming. Formability of HSS depends more on local elongation than on general
elongation, when maximal gradients of strain spread to adjacent areas (Ruoppa 2015, Ruoppa
2017). During the free-bending a plastic deformation develops form the neutral axis to the edges
of the steel sheet and thereby to strain hardening. The degree of strain hardening is possible to
measure by hardness test, when the most strain hardened area indicates the highest hardness value
(Ruoppa 2015). After unloading an elastic stress is relieved due to springback, which causes a zig-
zag shape of the residual stress along the thickness of the steel sheet (Weng, 1990; Anis, 2012).

3 EXPERIMENTAL STUDY

During the cold forming of the steel sheet, particularly on the inner and outer edges an alter-
ation of mechanical properties occurs, such as strain hardening and a deformation of the
grains of the microstructure. If a cold formed area is subsequently exposed to a welding pro-
cess, another alteration of mechanical properties and microstructure occurs. Main goal of this
experimental study is to determine toughness, hardness, alteration of the microstructure and
stress-strain properties of twice altered cold formed and welded area.
High strength steel STRENX S960E was chosen as a default material. Its mechanical and
chemical properties are in detail described in Tables 1 and 2. Steel sheet with dimension
10 × 2000 × 6000 mm was cut by laser to smaller sheets with dimensions 10 × 250 × 500mm
(S1) and 10 × 125 × 500 mm (S2). S1 were subsequently cold formed by free bending with the
inner radius of 30 mm (29 pieces) and 45 mm (29 pieces). For the bending of both radii,
a force of 120 tons had do be developed. Two different inner radii were manufactured so that
an effect of the inner radius on the properties of the steel sheet can be determined.
In the middle of the cold formed area on the outer side of the bend, a S2 steel will be welded
by MIG welding (see Figure 1).

1270
Table 1. Mechanical properties of STRENX S960E.
fy,Rp0.2 fu Elongation A5 Toughness

960MPa 980 – 1150MPa min. 12% 40J/ −40˚C

Table 2. Chemical composition of STRENX S960E [%].


C Si Mn P S Cr Cu Ni Mo B

0.20 0.5 1.60 0.02 0.01 0.80 0.30 2.00 0.70 0.005

OK Aristorod 69 and OK Aristorod 89 were chosen as an undermatching filler materials.


Two different filler materials were chosen in order to investigate a dependency of the welded
joint on their mechanical and chemical properties (refer to Table 3 and refer to Table 4). Mul-
tipass welds with different/same filler material for root welds and filler welds will be made so
that material properties of a such made weld joint could be investigated.
Different levels of heat input, t8/5 cooling times, preheat temperatures and combination of
filler materials will be chosen so that a wide spectrum of results will be achieved of which
a suitable welding procedure will be determined for welding in cold formed area of HSS.
10 mm wide test samples will be removed from finished welded members, on which Vickers
HV 0.1 hardness test will be performed according to EN ISO 9015-2 and EN ISO 6506-1 for
the weld material, HAZ and cold formed area.

Figure 1. Scheme of manufacturing of a test member (mm).

Table 3. Chemical composition of filler materials.


Filler material Mn Mo C Si Ni Cr

Aristorod 69 1.54 0.24 0.089 0.53 1.23 0.26


Aristorod 89 1.75 0.533 0.081 0.80 2.22 0.41

Table 4. Mechanical properties of filler materials.


Filler material fy fu Elongation A5 Toughness

Aristorod 69 715 MPa 805 MPa min. 17% 60J/ −40˚C


Aristorod 89 920 MPa 940 MPa min. 18% 47J/ −40˚C

1271
Figure 2. Left: complete test member with inner radius 30mm, right: complete test member with inner
radius 45mm.

Another test samples will be prepared so that a metallography according to ISO 17639 will
be performed with Carl Zeiss AxioObserver microscope with 500 x zoom in order to investi-
gate a microstructure of weld the material, HAZ and cold formed area.
2 cylinder test samples with 10 mm radius will be milled form the middle of the bent area on
which stress-strain properties will be tested according to EN ISO 6892-1. Similar way will be
used to extract two test samples for toughness tests. Toughness tests will be performed in −20˚C
and −40˚C.
Aforementioned test samples will be extracted from only cold formed steel sheet and form
cold formed and welded steel sheet so that a comparison of their mechanical properties could
be made.
Until today, two members were welded, one with inner radii 30 mm, second with inner radii
45 mm (see Figure 2). S2 steel sheet was welded with 3 multipass welds on both sides, when
filler and root welds were made with Aristorod 69 filler metal. Welding was performed in
a room temperature with parametres given in Tables 5 and 6. Heat input was then determined
according to equation (1), t8/5 time was then determined according to equation (2).

Table 5. Welding parametres for 30˚ bent steel sheet.


Weld Material U [V] A [I] v [mm/min] Q [kJ/mm] t8/5 [s]

L - Root weld Ar69 240 29,1 424.4 0,6 18,3


L - Filler weld Ar69 235 28,8 474.1 0.54 15.0
L - Filler weld Ar69 235 28.9 365.1 0.7 25.3
R - Root weld Ar69 240 29.1 459.6 0.55 15.9
R - Filler weld Ar69 235 28.9 518.1 0.49 12.5
R - Filler weld Ar69 235 28.8 395.9 0.64 21.5

Table 6. Welding parametres for 45˚ bent steel sheet – left welds.
Weld Material U [V] A [I] v [mm/min] Q [kJ/mm] t8/5 [s]

L - Root weld Ar69 210 25.2 461.5 0.73 27.5


L - Filler weld Ar69 210 25.2 587.7 0.55 15.9
L - Filler weld Ar69 210 25.2 524.5 0.62 20.0
R - Root weld Ar69 210 25.2 400.0 0.84 36.6
R - Filler weld Ar69 210 25.2 507.4 0.64 21.4
R - Filler weld Ar69 210 25.2 412.5 0.79 32.4

1272
Equation for the determination of heat input:

k:U:I:60
Q¼ ð1Þ
v:1000

where Q = heat input [kJ/mm], U = Voltage [V], I = Current [A], v = welding speed [mm/min],
k = thermal efficiency [-]
Equation for the determination of t8/5 time:
" 2  2 #
Q2 1 1
t8=5 ¼ ð4300  4:3T0 Þ  10  2 
6
  F2 ð2Þ
d 500  T0 800  T0

where t8/5 = time t8/5 [V], T0 = air temperature [˚C], d = plate thickness, F2 = 0,45 (shape
factor for two – dimensional heat flow)
After welding, test samples for hardness testing were extracted from both the pure cold
formed (see Figure 3) steel sheets and from welded members (see Figure 4).

Figure 3. Extracted test samples of only cold formed steel sheets (without welding).

Figure 4. Extracted test samples of completely welded members.

1273
4 CONCLUSION

Execution of all aforementioned tests on more than 40 test samples will take part in April and
May 2019. Test results will be used for validation of numerical model. Validated numerical
model will be used for a parametric study, in which different inner bend radiuses and different
materials (stronger) will be used. Results will serve as the support for an effective design of
structural members in civil engineering.

ACKNOWLEDGMENT

Current research is supported by grant TAČR – TJ01000045 „Advanced procedures of steel


and composite structure connections design and productions“.

REFERENCES

Anis A., Bjork, T., Heinilla S. 2012. Prediction of residual stresses in cold formed corners. Journal of
Advanced Science and Engineering Research Vol.2, No.4 December (2012) 252–264.
Brtník, T. 2016. Svary prvků z vysokopevnostních ocelí. Katedra ocelových a dřevěných konstrukcí,
ČVUT v Praze.
Cui, B. 2016: Effect of heat input on microstructure and toughness of coarse grained heat affected zone
of Q890 steel. ISIJ international, Vol.56 (2016), No. 1, pp. 132–139, 2016.
Erns, W., Vallant, R, Lozinger, N. 2013. Influence of the soft zone on the strength of welded modern
HSLA steels. Welding in the World, Le Soudage Dans Le Monde, 2013.
EN 1993-1-12: Eurocode 3: Design of steel structures – Part 1-12: Additional rules for extension of EN
1993 up to steel grades S700, European Standard, CEN, Brussels, 2005.
Günter, H.P., Rasche, C. 2008. High-strength steel fillet welded connections. Stuttgart University,
Germany.
Kuhlmann, U. 2014. Load bearing capacity of fillet welded connections of high strength steels. Stuttgart
University, Germany.
Rodrigues, D.M. 2004. Numerical study of the plastic behaviour in tension of welds in high strength
steels. International Journal of Plasticity 20, pp. 1–18.
Ruoppa, R. 2015. Bendability tests for ultra-high-strength steels with optical strain analysis and predic-
tion of bending force, Finland
Ruoppa R. 2017. Bending tests of very thick plates with advanced research equipment and techniques.
METNET Annual Seminar, Cottbus, 11th–12th October, 2017.
Törnlom, S. 2007. Undermathching butt welds in high strength steel. Departmens of civil and enviromental
engineering, Lulea University of technology.
Wang, Z. 2017. Effect of strength matching on mechanical properties of WELDOX 960 steel welded
joint. IOP Conf. Series: Materials Science and Engineering 248 (2017) 012018.
Weng, C. C., White, R. N. 1990. Residual stresses in cold-bent thick steel plates. Journal of Structural
Engineering, ASCE, 116: 1, 24–39.

1274
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Degradation processes in normalized mild- and low-alloy steel


building structures in service

W. Wichtowski & J. Hołowaty


West Pomeranian University of Technology, Szczecin, Poland

ABSTRACT: Material and strength tests on normalized steels from two building structures are
described; these are a railway bridge constructed from mild steel which has been in service for over
80 years, and a complex of five cylindrical tanks of 1500 m3 capacity each, constructed from low-
alloy steel, in service for 45 years. Chemical and mechanical tests were performed on a range of
micro and macro aged and normalized steels from these structures. The ageing and degradation
tests carried out on such steel grades have likely not yet been described in the technical literature.

1 INTRODUCTION

About 80 % of steel is presently cast using the continuous casting process. Output from this
casting method is up to 95% while casting into ingot moulds gives 80% (Blicharski 2002). Con-
tinuous-cast steel gives a better quality, higher cleanness and a more advantageous microstruc-
ture than mould-casted steel (Tasak 2002, Rykaluk 1999). These steels are totally killed. To
obtain such a steel about 0.03% aluminum is added to give a fine-grained steel, or silicon is
added to give a course-grained steel. Killed steels usually contain 0.15-0.50% Si, whose main
feature is to deoxidize the steel.
In order to obtain a variety of steel plates with specific properties, a number of control roll-
ing processes were designed and carried out (Figure 1). Standard control rolling uses steel
heated up to austenitic structure and rolling at different temperature (ranges I-III in Figure 1).
This thermo-mechanical process is described in detail in (Tasak 2002), as well as the SHT
(Sumitomo High Toughness) process of control rolling for low temperature service steels.
According to the literature, the pioneers of new structural steels were the German, French, Eng-
lish and American steel industries. The development and production of structural steels required
guaranteed and uniform material properties. As early as 1894-1895, the Wuppertal bridge

Figure 1. Comparison of classical and control rolling with microstructure change after rolling.

1275
Figure 2. Railway bridge of Thomas steel across the Wupper Valley near Müngsten (1897).

near Müngsten was constructed from early mild steel (Figure 2). The bridge was an arch construc-
tion with a span length of 170 m and height of 107 m. The steel grade later was designated as St37.
The minimum tensile strength of the steel was 37 kg/mm2 (370 MPa). The steel came into use in
1895 and later became the main steel material for the manufacture of steel bridges.
Later, steel St37∙12 was standardized by DIN 1612 (Schaper 1947, Mang 1977). Due to its
manufacturing processes, it was called Thomas-Stahl and SM-Stahl. It should be mentioned
that in Germany, Thomas convertors had been in use since 1880 and Siemens-Martin furnaces
since 1864. The first DIN 1612-09 was published in September 1924, but patents for structural
steels and methods of production to improve their properties also appeared. In 1928, “a way
of steel production with lowered possibility for cracking due to brittleness” was patented
(Figure 3). The invention involved the cold mechanical treatment of steel and then annealing
at a temperature of 700-900ºC.
The heat treatment of steel is used to change the microstructure and give improved properties.
Annealing involves refining steel grains while increasing yield strength and simultaneously lower-
ing brittle fracture transition temperature. Professor Gottwald Schaper (1949) defined the

Figure 3. Patent for production of normalized steel.

1276
Table 1. Properties of normalized steels St37·12 and S235.
Chemical
ReH Rm
composition, %
Steel C Mn Si P S P+S MPa MPa

Thomas max ~ max max max


0.4÷0.5
St37·12 0.16 0.01 0.09 0.06 0.13 max
370÷450
SM max ~ max max max 230
0.4÷0.5
St37·12 0.20 0.01 0.06 0.06 0.10
max max max max max max
S235J2 – 360÷510
0.17 1.40 0.025 0.025 0.05 235

properties of early mild steel St37∙12 (Normalgüte) according to the German standard DIN 1612.
For comparison, data for the contemporary non-alloy steel S235J2 according to EN 10025-2:2007
are also given.

2 PROPERTIES OF STRUCTURAL STEEL

Precise values for structural steel properties are taken directly from product standards or, with
some simplification, from Table 3.1. of EN 1993-1-1:2005. Over their service life, steel struc-
tures are influenced by functional and material degradation processes. The material undergoes
changes in mechanical properties caused by operational factors in the structure, chemical pro-
cesses in steel, atmospheric factors and so on. Assessing a structure for a further operational
period requires appropriate tests and precise analysis in order to characterize the material and
the magnitude of changes over a long service period (Hołowaty & Wichtowski 2013).
This paper gives a fragmentary analysis of the following normalized steels:
– St37∙12 mild steel, of German origin, from a railway bridge constructed in 1938,
– 18G2A low alloy steel, of Polish origin, from the sheeting of five cylindrical tanks of
V = 1500 m3 each, constructed in 1973.
Chemical and mechanical tests are described on a range of micro and macro aged and normalized
steels from these structures. Ageing and degradation tests on such steel grades are very rare and
they may be expected to model the ageing processes in contemporary European normalized steel
grades.
Testing steels from these structures before their refurbishment gave some astounding frac-
ture property test results. The Charpy impact energy results measured are given in Figure 4,
with the results the average values from three specimens. The tests were conducted on test bar
specimens of 10 × 10 × 55 with V-notch machining according to EN ISO 148-1:2017. The test
specimens were machined from samples taken from the web of the first bridge span of 26 m
length and from a 12-mm sheeting plate on the tank wall.
The fracture properties were assessed on two types of test pieces:
– naturally-aged A without any heat treatment,
– normalized N, annealing at temperature 930 ºC for an hour and then cooled in air.
The aged specimens with additional normalizing (N) are characterized by fine grain size and
mechanical properties similar to the properties at delivery. The “cosmic” values of KV for
N specimens ensured that the steels were normalized and additional microstructure tests were
undertaken.
The results in Figure 4 show an exceptionally large ageing effect on the steel from the rail-
way bridge which has been in service for almost 80 years. An ageing factor equivalent to the
ratio of impact energy KV after ageing to impact energy KV of a non-aged (normalized) piece,
at temperature -40 ºC and -20 ºC is: 0.056 and 0.054. Such low ageing ratio values had not
previously been detected, even in bridges with a much longer service period.

1277
Figure 4. Charpy impact energy for bridge steel (B) and tank steel (T).

3 THE STRUCTURES, AND THE MECHANICAL PROPERTIES OF THEIR STEELS

The single-track railway bridge spanning the Warta River is located in Gorzów Wielkopolski.
The bridge’s structural system and a current view is shown in Figure 5. The structure includes
eight riveted spans of length from 26.00 to 95.80 m with total bridge length 315.57 m. The
bridge was constructed for the German Railways in 1938. Low carbon mild steel designated as
St37 was made for the span structures. Here we are concerned with the seven riveted plate
girder through spans (Wichtowski & Woźniak 2015).
General views of the five cylindrical tanks and their geometric structure are shown in
Figure 6. The cylindrical tanks with a capacity of 1500 m3 were constructed for the storage of
phosphoric acid for a facility in Szczecin. For this reason, the tank sheeting and roof covers
were protected on the inside with a rubber lining of 5 mm thickness. The tank bases were
covered with protective acid-resisting slabs. Due to a change in technology production at the
facility, the tanks have been used to store garden liquid fertilizer for the past 45 years.
The steel walls of the tank of diameter DI = 13.0 m are constructed from six lower rings of
height 1.75 m and a top ring of height 1.50 m. The two lower rings utilize 12 mm steel plates, the
next two (III and IV) 10 mm plates and the three top rings (V, VI and VII) 8 mm plates. The roof
cover is constructed from 24 conical plates of thickness 6 mm strengthened by external roof
rafters.

4 STEEL PROPERTIES FOR PLATE GIRDER BRIDGE SPANS AND TANK WALLS

As mentioned above, the non-standard results of the Charpy impact tests required additional
explanation and further extended tests were undertaken, including:
– chemical composition,
– static tensile tests on rounded specimens,
– Brinell hardness,
– microstructure examination.
Chemical composition was identified by optical emission spectroscopy. Three samples from the
bridge structure were tested and one sample t = 12 mm from the tank wall. The bridge samples

1278
Figure 5. Railway bridge structure and its current side view.

are taken from the web of the first span of L = 26.0 m, from a vertical angle in the ballasted
deck of this span and from a plate in the top flange of the last span of L = 27.0 m. The chemical
composition for the 9 basic alloyed elements is given in Table 2. For comparison, the chemical
composition for mild steel used in the period 1888-1930 for bridge constructions as well as con-
temporary non-alloy steel of grade S355NL according to EN 10025-3 are also given in Table 2.
Analysis of the structural steel chemical composition shows that the plate girder bridge
spans were manufactured from low-carbon mild steel St37∙12 according to the German stand-
ard DIN 1612. It was the only normalized annealing steel of this standard, with ultimate

Figure 6. View of five tanks complex and tank schematic structure.

1279
Table 2. Chemical composition for plate girder bridge spans and tank walls.
Element content, %
Type of steel C Mn Si P S Cu Cr Ni Al

web1) 0.084 0.461 0.045 64 ppm 0.026 0.156 0.018 0.031 0.035
L-deck 0.031 0.309 0.043 0.036 0.044 0.021 0.071 0.016 0.007
Bridge top flange 0.035 0.351 0.042 0.027 0.013 0.045 0.012 0.030 0.010
0.030 0.040 0.00 0.004 0.004 0.110 0.007 0.030 0.010
mild steel
÷0.35 ÷0.75 ÷0.18 ÷0.16 ÷0.12 ÷0.14 ÷0.01 ÷0.04 ÷0.02
Tank wall plate 18G2A 0.166 2.055 0.395 0.035 0.027 0.066 0.066 0.019 0.083
max 0.90÷ max max max max max max min
S355NL
0.18 1.65 0.50 0.03 0.025 0.55 0.30 0.50 0.02
1)
all the presented test results are from the sample

strength Rm = 370 - 450 MPa and min A5 = 18-25 %. The characteristic feature of the steel is
a low level of carbon. This is not a good feature, however, as ageing speed increases when
carbon content is lower than 0.10 %.
The chemical composition of the steel from the tank wall is that of a low-alloy steel with higher
strength, grade 18G2A (ReH = 355 MPa, Rm = 490÷650 MPa) according to the former Polish
standard. It is the equivalent of the current steel S355NL. A much higher content of manganese
(2.055 %) negatively influences the steel weldability. At the same time, the 0.083 % aluminum con-
tent and 0.395 % silicon content increase ageing resistance. This is shown in Figure 4.
Static tensile tests allowed the mechanical properties of the steel in the structures to be
assessed. The test were carried out on rounded, fivefold specimens with a base diameter of
10 mm (bridge steel) and 8 mm (tank steel). The tests refer to:
– six specimens from steel St37∙12, including three specimens in a naturally-aged condition
(A) and three following normalized annealing (N) at a temperature of 930 ºC for an hour,
and air cooled,
– five specimens from low-alloy steel 18G2A, from the tank wall, including three in natur-
ally-aged conditions (A) and two specimens after normalization (N).
Figure 7 shows a sample diagram of the tensile tests on steel 18G2A, and in Table 2 the mech-
anical properties obtained are given. These are average values from the three samples except
for Re = min ReH.

Figure 7. Tensile curves for specimens A and N from tank wall.

1280
Table 3. Mechanical properties for bridge and tank steels.
Brinell
ReH ReL Rm A5 Z
hardness
Structure Type of RmB ReB
(steel) steel [MPa] [MPa] [MPa] [%] [%] HB5 α
[MPa] [MPa]

Bridge S 220 212 342 36 69 107 361 0.70 252


(St37·12) N 275 269 376 41 67 106 357 0.70 249
Tank S 361 366 536 29 64 163 547 0.65 355
(18G2A) N 387 380 526 32 72 165 553 0.65 359

When there is no possibility to obtain tensile specimens for testing, then the ultimate tensile
strength Rm may be determined from hardness test results. This is an approximate method, but
usually acceptable. For comparison purposes, hardness measurements of the steels were also
carried out. The results are given in the final columns of Table 3.
Hardness was measured using the Brinell method and a B3C-type tester. An indentation
steel sphere of 5 mm diameter was used with a testing weight of 7350 kN for a period of 15
s according to EN ISO 6506-1:2014. The coefficient values α = ReB/RmB are taken as 0.70 and
0.65 (Hołowaty & Wichtowski 2017). The obtained external difference for Rm and RmB are
+5.0 % and -5.0 %. The differences for ReH and ReB values are +14.5 % and -9.5 %.
Microstructural examination was also carried out under two conditions: naturally-aged
A and normalized N. The microstructures obtained are typical ferrite – pearlite phases for
structural steel. Under low magnification (×10), for the tank steel, a band structure is visible
which appeared during the plate rolling process (Figure 8). This is caused by dendrite segrega-
tion of Mn with a high content of 2.055 % during steel solidification. The microstructure con-
sists of two phases: dark pearlite and light ferrite grains in different shadows of grey.
At the same time, in a wider range, an analysis of impurities and inclusions in microsections
of aged – (A) and normalized (N) specimens from the bridge steel were carried out. The non-
treated structure and microstructure of specimens A and N using magnification ×70, ×350
and ×700 were observed (Wichtowski & Jasiński 2016).

Figure 8. Microstructure of aged (A) and normalized (N) tank steel (×10).

Figure 9. Microstructure of bridge steel: aged (A) and normalized (N).

1281
Grain size was determined by comparative method (using a master scale) according to the
requirements of EN ISO 634: 2013. As per metallographic investigations, grain size was deter-
mined for the steel specimens under two conditions, A and N.
The grain size for bridge steel in naturally-aged conditions S is suited to the G5 master with
average grain diameter of dm = 0.0625 mm (Figure 9), while for tank steel it is the G9 master
with average grain diameter of dm = 0.0156 mm. Under normalized annealing conditions N,
the grain size is smaller, for the bridge steel it is the G7 master with grains dm = 0.0312 mm
(Figure 9), and for the tank steel it is the G10 with grains dm = 0.0110 mm.

5 SUMMARY

The steel from the railway bridge constructed in 1938 is a mild steel designated as St37∙12.
The steel is a fine-grained material with G5 grain size number that has an average grain
diameter of dm = 0.0625 mm. The current steel yield strength is ReH = 220 MPa and ultimate
strength Rm = 340 MPa. These values are only 4.3 % and 8.1 % lower than the minimal values
recommended by the standard DIN 1612 for this steel grade. Degradation changes arising during
service were found in fracture properties (Figure 6). Their contribution to safety assessment and
the calculation of structure durability is impossible to determine.
The wall steel of tanks with a capacity of 1500 m3 which have been in service since 1973 is
a low-alloy steel, normalized with subgrade E. It is fine-grained steel with actual yield strength
ReH = 360 MPa and ultimate strength Rm = 535 MPa. The values are analogous to normative
values for this steel grade. The actual grain size for this steel is equivalent to the G9 scale with
an average grain diameter of dm = 0.0156 mm.
The degradation changes of tank wall steel:
– lower the values of ReH from 390 MPa to 360 MPa, that is by 7.7 %, with a small increase
in of 1.9 % Rm,
– lower the fracture properties – the Charpy impact energy KV; at temperature -40 ºC from
value 84 J to 29 J, that is by 65.5 %, and at temperature range from -20 ºC to +20 ºC by
a similar value of 51 J, that is by 34.5 %.
It is the authors’ opinion that these tests on such steel grades and their results are likely to
be the first published. There exists a significant probability that similar degradation processes
might appear in currently-manufactured normalized structural steels over their service life.

REFERENCES

Blicharski, M. 2004. Material engineering. Steel. Warszawa: WNT (in Polish).


Tasak, E. 2002. Weldability of steel. Kraków: Fotobit (in Polish).
Rykaluk, K. 1999. Cracks in steel structures. Wrocław: DWE (in Polish).
Schaper, G. 1949. Stählerne Brücken. Band I. Berlin: Verlag von Wilhelm Ernst & Sons.
Mang, F. 1977. Stähl in Altbau und Wahnungsbau. Stuttgart: Frauhofer IRB Verlag.
Albrecht, R. 1975. Richtlinien zum Brückenbau. Band 1. Stählerne Brücken einschließlich Stahlträger in
Beton und Verbund-Konstruktionen. Wiesbaden und Berlin: Bauverlag GmbH.
Hołowaty, J. & Wichtowski, B. 2013. Properties of structural steel used in earlier railway bridges. Struc-
tural Engineering International 23 (4): 512–518.
Wichtowski, B. & Woźniak, Z. 2015. Properties of a normalized mild steel of a railway bridge after 75
years in service. Welding Technology Review 87 (5): 110–114(in Polish).
Wichtowski, B. & Jasiński, W. 2015. Microstructural degradation processes in normalized mild steel
from a railway bridge. Welding Technology Review 87 (5): 94–99 (in Polish).
Hołowaty, J. & Wichtowski, B. 2017. Remarks on the material testing of historical railway bridges.
EUROSTEEL 2017, Copenhagen, 13–15 September, 2017.

1282
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Effect of the steel grade on equivalent geometric imperfections for


lateral torsional buckling

R. Winkler & M. Knobloch


Ruhr-Universität Bochum, Bochum, Germany

ABSTRACT: The equivalent geometric imperfections of EN 1993-1-1:2005 were developed


based on studies considering steel grades up to S460. For flexural buckling an extension of the
scope of the verification method to higher steel grades up to S700 is currently being evaluated
in the context of the revision and further development of Eurocode 3. For LTB, however,
similar studies, are still lacking.
This paper presents a comprehensive study on the development of equivalent geometric
imperfections for high-strength steel members subjected to lateral torsional buckling. The
study focuses on members with hot-rolled double symmetric I-/H-cross sections of steel grade
S700. The proposed imperfections and their application to verification methods are compared
to existing design approaches. In addition to normative rules applied in Europe, a novel
design approach is also taken into account, which is currently discussed as part of the further
development of Eurocode 3. Furthermore the paper provides a new proposal for LTB-
imperfections and the additional consideration of steel grades up to S700 using the partial
internal forces method. This proposal leads to design results that fit well with the results of
numerical simulations.

1 INTRODUCTION

The equivalent geometric imperfection method can easily be applied within the framework of
a computer-based analysis of steel structures using state-of-the-art computational tools.
Internal forces are computed according to 2nd order theory considering the effects of equiva-
lent geometric imperfections. These internal forces are used for a cross-section verification.
The equivalent imperfections take into account, among other things, the effects of geometrical
imperfections, residual stresses as well as stiffness degradation and load-redistribution due to
partial yielding of the structural members. The method is already often applied in design prac-
tice for a straightforward assessment of steel members and structural systems that may fail
due to flexural buckling. Increasingly computer tools are provided which facilitates the appli-
cation of the method also for the assessment of lateral-torsional buckling, e.g. [1]. The frame-
work for the structural analysis provided in prEN 1993-1-1:2018 also facilitates the
application of the equivalent geometric imperfection method for lateral-torsional buckling
verification.
In the context of the revision of EN 1993-1-1: 2010, Lindner et. al. [1] recommended to critic-
ally reflect the existing approaches [2, 3] for equivalent geometric imperfections. The cause and
background of this discussion were among other things: (i) the equivalent bow imperfections
were specified for a member that is subjected to a compression force only [4]. For flexural buck-
ling, however, the extended yielding range and stiffness reduction due to additional bending
moments should be taken into account. In case of lateral torsional buckling the divergent stabil-
ity behaviour between a member subjected to uniform compression and lateral torsional buck-
ling ought to be considered. (ii) As part of the further development of EN 1993-1-1, structural
steels with steel grades up to S700 are considered. Therefore, novel studies concerning equiva-
lent geometric imperfections should include these steel grades.

1283
Lindner et al. [5-7] published new studies on equivalent bow imperfections for the stability
case flexural buckling. The results were included in the proposal of the prEN1993-1-1: 2018 [8]
(so-called Final Document). The equivalent bow imperfections are determined as a function
of the buckling curve and the steel grade. In addition, a distinction is made between bending
about the strong and weak cross-sectional axis. While any plastic cross-sectional interaction
formula can be used for buckling about the weak axis (taking into account the limitation of
the plastic bending resistance to Mpl = 1.25 Mel), the cross-sectional check for buckling about
the strong axis is limited to the plastic linear plastic interaction. Winkler et al [9, 10] extended
the application of the approach for equivalent bow imperfections for flexural buckling to the
partial internal forces method (PIM) [11]. The extension enables a consistent design approach
for cross-section and stability failure. In addition, the method is commonly used in design
practice and easy to apply within the framework of the RUBSteEl-tools [12-14].
Lindner [15] proposed equivalent bow imperfections for the stability case lateral torsional
buckling that have been considered in [8]. These imperfections are based on a study carried
out by Kindmann et al. for steel grades S235 and S355, [16]. The results of this study were
considered in the German National Annex of EN 1993-1-1 [3]. The classification uses the ratio
h/b ≤ 2.0. The equivalent bow imperfection is calculated using a reference relative bow imper-
fection βLT and the material parameter ε to consider the steel grade. The reference relative
bow imperfection is 1/200 for h/b ≤ 2.0 and 1/150 for h/b > 2.0 and a plastic cross-section
verification.
This paper presents a proposal for equivalent bow imperfections for lateral torsional buck-
ling of a member in bending. The proposal is based on the results of a numerical simulation
study and avoids a limitation of the cross-section verification method. In conjunction with
imperfections for flexural buckling presented in [10, 9] a full set of equivalent geometric imper-
fections for second order analysis taking account of flexural and lateral torsional buckling is
available.

2 EQUIVALENT GEOMETRIC IMPERFECTION METHOD

2.1 Aim of equivalent bow imperfections


The simplified assessment approach “equivalent imperfection method” is based on the 2nd order
theory. The equilibrium is formulated on the deformed system, taking equivalent geometric
imperfections into account and the internal forces are determined according to 2nd order theory.
With these internal forces an elastic or plastic cross-section verification is performed. Thus, the
load bearing capacity according to this method is strongly affected by the size and shape of the
equivalent geometric imperfections as well as the cross-section verification approach.
A key objective for determining the amplitude of equivalent bow imperfections is that the
structural analysis according to second order theory and the subsequent cross-section verifica-
tion should cause the same design results as sophisticated numerical simulations.

2.2 Plastic cross-section interactions


Applying the equivalent geometric imperfection method for design purposes, the imperfec-
tions are linked to the cross section verification method. For lateral torsional buckling, the
cross-section verification must capture biaxial bending moments My and Mz axial forces
N and warping moments Mω. For this, the linear plastic interaction and the Partial Internal
forces Method (PIM) [11] are suitable.
The partial internal forces method can be used for the assessment of the cross section cap-
acity based on the plastic theory. A main advantage of the method is that it can easily take
into account all internal forces and moments of steel members of arbitrary cross-sections. In
principle, the allocation of the internal forces and moments on the individual partial elements
of the cross-section takes place, taking into account the equilibrium conditions. This results in
partial internal forces and moments. By considering the plastic capacity of the partial elements

1284
and interaction relationships, the load-carrying capacity can be verified. The method is based
on mechanical principles, promotes a fundamental understanding and is therefore very popu-
lar in education and engineering practice. The implementation of the method in software
tools, i.e. [12, 14, 13], also encourages its dissemination.

2.3 Method
A key objective of the paper is to develop required equivalent bow imperfections using the partial
internal forces method for rolled double-symmetric I/H cross-sections and to compare them with
existing approaches. The results are based on numerical simulations. For this purpose geometric-
ally and materially nonlinear investigations on imperfect members were performed (GMNIA).
The numerical simulations considered both structural and geometrical imperfections with the
common values and distributions (geometric imperfection v0 of the compressed flange v0 = L/1000
and residual stresses according to ECCS No. 33 [20]). The imperfections for the analyzed stability
case lateral torsional buckling included prerotations and predeformations. This is the common
approach but differs from the study made in [16]. The residual stresses were assumed to be 0.3 and
0.5 × 235 N/mm², respectively, irrespective of the grade of the considered steel grade, [20]. For
members with predefined LTB-slenderness the maximum bending capacities were determined.
This capacities were used in a next step to determine the required equivalent bow imperfections.
The RUBSteEl eEducation tool FE-STAB [14] was used to perform a 2nd order theory analysis.
For this the numerical determined capacities were applied and the bow imperfections e0,LT were
increased to reach a cross section utilization of 100 %. The required equivalent bow imperfections
are analyzed as j-values (j = L/e0,LT) to allow a direct comparison with the standard approaches.
The strict application of the described method resulted in disproportionately small j values
for specific cases. In order to avoid this, steel members outside the range of practical applica-
tion were excluded by means of a limit of rotation. The rotation according to 2nd order theory
was limited to ϑlim = 0.3 rad [16]. If the rotation limit was decisive, the j-value was set to 1000.
The numerical study considered various doubly symmetric cross-sections; and the param-
eter field was selected based on the objectives of the analysis:
– For an analysis of the geometric bow imperfections of members made of high-strength
steels, in addition to the steel grades S235 and S355, steels with the higher steel grades S460
and S700 were included.
– The h/b ratios of the considered cross-sections covered a large range of the commonly used
hot-rolled doubly symmetric cross-sections. The ratio was used for the classification of the
European buckling curves and thus enabled a comparison to common approaches found in
the literature.
– LTB slenderness ratios between 0.25 and 3.0 have been considered to check the consistency
of the approaches for compact to slender members.

2.4 Bending moment distribution


First, the influence of the applied moment distribution on the required equivalent bow imper-
fections was analyzed. The decisive load case was determined for further investigations. For
this purpose, Figure 1 shows the required equivalent bow imperfections for five different
bending moment distributions (LC 1 to 5) and three different cross sections (HEB 200 (left),
HEB 600 (middle) and HEAA 1000 (right)). Using the partial internal forces method the
imperfections are presented in terms of the j-value as a function of the LTB slenderness.
The constant moment distribution (LC 1, blue) and the parabolic moment distribution (LC 4,
purple) are decisive in terms of minimal j-values for the stability case lateral torsional buckling.
The following analyses are exclusively presented, based on the constant moment distribution. The
slenderness ratio influences the minimal j-values. For all cross sections the smallest value result
for a slenderness ratio of 0.75. For more slender members the j- value distinctively increases.

1285
Figure 1. Influence of the moment distribution on the j-values for the partial internal forces method for
an HEB 200, HEB 600 and HEAA 1000 with a steel grade S235 as a function of the slenderness ratio.

2.5 Cross section verification and steel grade


To analyze the influence of the cross section verification method on the required equivalent
bow imperfections, the j-values are determined using the partial internal forces method and
the linear plastic interaction. Figure 2 exemplarily shows the resulting j-values for the cross
section HEB 200 as a function of the LTB slenderness.
The j-values for the partial internal forces method and the linear plastic interaction differ
significantly. The minimal j-values according to PIM are approximately 130, while the use of
the linear plastic interaction lead to minimal j-values of 300. The j-values partially correct the
failure if the linear plastic interaction.
Figure 2 additionally shows the influence of steel grade on the equivalent bow imperfec-
tions. The various colors display the influence of the steel grade. The analysis considers the
steel grades S235 (dark blue), S355 (blue), S460 (green) and S700 (orange).
The influence of the steel grade differs with the considered cross section verification. Using
the partial internal forces method the influence is marginal. The minimal j-values were
between 132 (S235) and 150 (S700). For the linear interaction the influence is more pro-
nounced. The minimal j-values were between 300 (S235) and 423 (S700). In an approach for

Figure 2. Effect of the steel grade on the j-values as a function of the LTB slenderness ratio (PIM left
and linear interaction right) for a HEB 200 section and a constant moment distribution.

1286
Figure 3. Effect of various limits on the j-values for an HEA 1000 section with a steel grade S235 and
a constant moment distribution.

equivalent bow imperfections using the partial internal forces method the consideration of the
steel grade is less decisive.

2.6 Sensitivity study


The displayed results so far consider a rotation limit of ϑ < 0.3 rad and a cross section utiliza-
tion of 100%. The limitation of the rotation ϑ was chosen according to the calculations
described in [16]. The limit of 0.3 rad is equal to 17.4° and considers the practical application
range. Lindner et al. introduce two more limit criteria to avoid disproportional small j-values
for the imperfections for flexural buckling [6, 1]. The first criteria allows to exceed the cross-
section utilization by 3 %. For the stability case flexural buckling this small overestimation
leads to useful results, [1, 9, 10]. The second criteria limits the plastic shape coefficient αpl,z to
1.25. This limitation is decisive for buckling about the weak cross section axis and with this
also for lateral torsional buckling and was implemented in [21]. Both criteria build the basis
for the proposed bow imperfection approach for flexural buckling in [8].
Figure 3 shows the j-values for an HEAA 1000 section as a function of the relative LTB
slenderness ratio. The grey curve shows the results without any limitation. In this case the
limitation (continuous black curve) does not influence the minimal j-value. For compact cross
sections of slender members this criterion helps to consider only the practical application
range, [22] and thus it builds the basis for the present study.
With the overestimation of the cross section utilization as well as the limitation of the plas-
tic shape coefficient the j-values result higher. Without additional criteria a minimal j-value of
126 result. Considering 103% for the cross section utilization the value slightly increase to 134
and with an additional limitation of αpl,z the minimal j-value result to 160. The influence is
less distinctive than for flexural buckling.
To reach a consistent approach for all stability cases both criteria are considered in the devel-
opment of a new approach for equivalent bow imperfections for lateral torsional buckling.

3 PROPOSAL AND COMPARISON

The results presented in the last section are the basis for the development of an easy applicable
and consistent design approach based on the equivalent imperfection method using the partial

1287
Table 1. Values for initial bow imperfections, LC 1.
Required equivalent bow imperfections
based on numerical determined
Approach resistances
L [mm]
Cross for NA 100% CS 103% CS 103% and
section αpl,z Steel λLT ¼ 1:0 BC EC3 EC3 prEN utilization utilization αpl,z < 1.25

235 14596 200 148 196 236


c 300
HEM 355 9766 163 148 211 254
1,533 200
200 460 7628 400 143 149 218 262
b
700 5188 - 116 152 222 266
235 9264 200 132 150 175
c 300
HEB 355 6434 163 141 164 197
1,527 200
200 460 5195 400 143 150 177 213
b
700 3774 - 116 148 179 214
235 1818 200 124 140 161
b 400
355 1286 163 126 144 173
IPE 80 1,576 200
460 1051 500 143 126 145 173
a
700 778 - 116 119 146 176
235 9732 200 140 151 181
b 400
HEB 355 7243 163 138 154 185
1,542 200
600 460 6102 500 143 137 157 189
a
700 4706 - 116 135 159 191
235 8147 150 134 145 174
b 400
HEA 355 6329 122 130 145 174
1,557 150
800 460 5445 500 107 129 145 175
a
700 4308 - 87 123 140 169
235 5684 150 130 141 169
b 400
355 4448 122 124 137 164
IPE 600 1,577 150
460 3840 500 107 122 136 163
a
700 3050 - 87 115 130 156
235 6885 150 126 134 161
b 400
HEAA 355 5490 122 115 1251 150
1,604 460 4779 500 150 107 110 120 144
1000
700 3833 a - 87 101 113 135

internal forces method. In addition, existing approaches are compared with the analysis results
and j-values.
Table 1 shows the resulting j-values under consideration of the limit criteria described in
section 3.4 (PIM = 103% cross section utilization and αpl,z < 1.25). The j values for a constant
moment distribution (LC 1) are shown separately for all cross sections (sorted according to
their h/b ratios). In addition, the investigation was separately carried out for the different steel
grades. In addition, for a direct comparison, the table includes the j-values of the existing
standards and approaches described in section 2.3.
Both, the limitation of the plastic shape coefficient and a permitted cross section utilization
of 103 % lead to larger j values for all cross-sections than the analysis without limit criteria.
For compact cross-sections (h/b < 2.0) the j-values decrease slightly with increasing steel
grade, for slender sections the j-values increase. This results from two different effects. With
increasing slenderness of the cross section the influence of the steel grade on the member
length decreases (Tab. 1) which directly affects the stability risk of the member. The second
effect is the influence of the residual stresses. With increasing steel grade the influence of the
residual stresses decreases. Overall, however, the influence of the steel grade on the j-values is

1288
Figure 4. Comparison of different approaches for equivalent bow imperfections with FEM-results, LC 1.

low. A direct dependence on the j-values of the buckling curves (BC) as well as the h/b-ratios
cannot be determined.
The classification of the cross sections using the h/b-value leads to minimal j-values for both
groups and each considered limit criteria. Considering a cross section utilization of 103 % for
compact cross sections with h/b ≤ 2.0 a minimal j-value of 140 result. With the same limit cri-
teria for slender cross sections j = 113. With an additional consideration of the limit for αpl,z
the minimal values result to 161 (h/b ≤ 2.0) and 135 (h/b >2.0).
In the following, two approaches for members under LTB are proposed and compared with
the results of the numerical simulation study. Both approaches consider the minimal deter-
mined j-values with the cross section classification using the h/b-ratio and are set to fix values
without consideration of the steel grade.
1) j = 140 (h/b ≤ 2.0) resp. j = 113 (h/b > 2.0)
The first approach takes a cross section utilization of 103 % into account.
2) j = 160 (h/b ≤ 2.0) resp. 135 (h/b > 2.0)
The second approach additionally limits the plastic shape coefficient αpl,z to 1.25.

Figure 4 compares the limit load capacities in order of the maximum capacities per
approach My,EI,i in the related form my,i,lim = My,EI,i/My,pl (horizontal axis) to the numerical
limit load capacities My,FEM in related form my,FEM,lim = My,FEM/My,pl (horizontal axis). In

1289
addition to the proposals described above, the limit carrying capacities resulting from the
approach in prEN 1993-1-1 are compared with the carrying capacities from the numerical
analysis.
The two proposals correctly record the load capacities and lead to conservative design
results with little scatter. The value Δm describes the absolute mean deviation of the load cap-
acities according to the equivalent imperfection method compared to the numerical simula-
tions over the entire parameter range. Δm,S235 contains only the results for the steel grade
S235. The lowest mean deviation result for the imperfection approach according to prEN
1993-1-1 in combination with the partial internal forces method. However, the bow imperfec-
tions lead to a few slightly non-conservative results for steel grade S235 (dark blue symbols).
The prEN 1993-1-1 proposal is based on calculations described in [16]. The geometric imper-
fection of the numerical model did not consider a combination of predeformations and prero-
tations. Considering both leads to smaller capacities and to smaller required j-values.
Both proposals in this publication lead to conservative design results and are consistent to
the concepts for flexural buckling according to [8] and [9, 10].

4 CONCLUSIONS

Initial bow imperfections for doubly symmetric I-/H- cross sections subjected to lateral tor-
sional buckling have been determined based on a numerical simulation study. These bow
imperfections consider the well-suited partial internal forces method (PIM) for the cross-
section verification. For simplification, the bow imperfections have been proposed to be inde-
pendent of the steel grade and bending moment distribution for design purposes.
A comparative study has shown that the proposed bow imperfections lead to suitable design
results compared to numerically determined member capacities. Together with the previously
established equivalent geometric imperfections for flexural buckling, the proposal of this paper
forms a consistent approach for the structural design of compact and slender steel members.

REFERENCES

[1] Lindner, Joachim; Kuhlmann, Ulrike; Just, Adrian. Verification of flexural buckling using initial
bow imperfections e0/L of table 5.1 of EN 1993-1-1, CEN/TC250/SC3/WG1 N131, Meeting of the
Working Group 1-1 of CEN/TC250/SC3, Berlin, 12 October 2016.
[2] EN 1993-1-1:12.2010. EN 1993-1-1: 2010 Design of steel structures - Part 1.1: General rules and
rules for buildings”.
[3] EN 1993-1-1/NA:08/2015. Nationaler Anhang – National festgelegte Parameter – Eurocode 3:
Bemessung und Konstruktion von Stahlbauten – Teil 1-1: Allgemeine Bemessungsregeln und Regeln
für den Hochbau”.
[4] Lindner, J.; Scheer, J.; Schmidt, H. 1998. Stahlbauten, Erläuterungen zu DIN 18800, Teil 1 bis Teil
4. Steel structures, Commentary to DIN 18000 Part 1 to 4, Bd. 67. Beuth Verlag GmbH; Ernst &
Sohn.
[5] Lindner, Joachim. 2017. “Repräsentative Vorkrümmungen e 0 für das Biegeknicken - Ergänzende
Untersuchungen”. Stahlbau 86, Heft 8, S. 707–715, doi: 10.1002/stab.201710512.
[6] Lindner, Joachim; Kuhlmann, Ulrike; Just, Adrian. Nachweis des Biegeknickens auf der Grundlage
von Vorkrümmungen e0/L nach Tabelle 5.1 in DIN EN 1993-1-1. In: NA 005-08-16 AA.
[7] Lindner, Joachim; Kuhlmann, Ulrike; Just, Adrian. 2016. “Verification of flexural buckling accord-
ing to Eurocode 3 part 1-1 using bow imperfections”. Steel Construction 9, Heft 4, S. 349–362, doi:
10.1002/stco.201600004.
[8] prEN 1993-1-1:2018:2018. Eurocode 3 - Design of steel structures - Part 1-1: General rules and rules
for buildings”.
[9] Winkler, Rebekka; Niebuhr, Martin; Knobloch, Markus. 2017. “Geometrische Ersatzimperfektio-
nen für Biegeknicken um die starke Querschnittsachse unter Berücksichtigung des Teilschnitt-
größenverfahrens”. Stahlbau 86, Heft 11, S. 961–971, doi: 10.1002/stab.201710545.

1290
[10] Winkler, Rebekka; Knobloch, Markus. 2018. “Geometrische Ersatzimperfektionen zur Anwendung
des Teilschnittgrößenverfahrens für Biegeknicken um die schwache Querschnittsachse”. Stahlbau
87, Heft 4, S. 308–322, doi: 10.1002/stab.201810590.
[11] Kindmann, Rolf; Frickel, Jörg. 2017. Elastische und plastische Querschnittstragfähigkeit; Grundla-
gen, Methoden, Berechnungsverfahren, Beispiele. Online-Auflage 2017. http://www.ruhr-uni-
bochum.de/stahlbau/publikationen/buecher.html.de.
[12] QST I-plastisch. RUBSteEl eEducation Tool, Ruhr-Universität Bochum; Lehrstuhl für Stahl-,
Leicht- und Verbundbau. Version 10.2014. <www.ruhr-uni-bochum.de/stahlbau/software/index.
html.de>.
[13] QST-Kasten. RUBSteEl eEducation Tool, Ruhr-Universität Bochum; Lehrstuhl für Stahl-, Leicht-
und Verbundbau. <www.ruhr-uni-bochum.de/stahlbau/software/index.html.de>.
[14] FE-STAB. 2012. RUBSteEl eEducation Tool, Ruhr-Universität Bochum; Lehrstuhl für Stahl-, Leicht-
und Verbundbau. Version 7.2012. Bochum. <www.ruhr-uni-bochum.de/stahlbau/software/index.
html.de>.
[15] Lindner, Joachim. 7.3.3.2 second order analysis for lateral torsional buckling, CEN/TC250/SC3/
WG1 N240, Meeting of the Working Group 1-1 of CEN/TC250/SC3, Berlin, 21 March 2018.
[16] Kindmann, Rolf; Beier-Tertel, Judith. 2010. “Geometrische Ersatzimperfektionen für das Biege-
drillknicken von Trägern aus Walzprofilen – Grundsätzliches”. Stahlbau 79, Heft 9, S. 689–697,
doi: 10.1002/stab.201001347.
[17] Winkler, R.; Kindmann, R.; Knobloch, M. 2017. “Lateral Torsional Buckling Behaviour of Steel
Beams – On the Influence of the Structural System”. Structures, Heft 11, S. 178–188, doi: 10.1016/j.
istruc.2017.05.007.
[18] Snijder, H.; van der Aa, R. P.; Hofmeyer, H., van Hove, B. Design imperfections for steel beam lat-
eral torsional buckling. In: Proceedings of The International Colloquium on Stability and Ductility
of Steel Structures. Timisoara, Romania 2016.
[19] Snijder, H. H.; van der Aa, R. P.; Hofmeyer, H.; van Hove, B.W.E.M. 2018. “Lateral torsional
buckling design imperfections for use in non-linear FEA”. Steel Construction 11, Heft 1, S. 49–56,
doi: 10.1002/stco.201810015.
[20] European convention for constructional steelwork. ECCS TC 8. Ultimate Limit State Calculation of
Sway Frames with Rigid Joints, 1984.
[21] DIN 18800-2:11.2008. Stahlbauten - Teil 2: Stabilitätsfälle - Knicken von Stäben und Stabwerken”.
[22] Ebel, Rebekka. 2014. Systemabhängiges Tragverhalten und Tragfähigkeiten stabilitätsgefährdeter
Stahlträger unter einachsiger Biegebeanspruchung. System-dependent bearing characteristics and load
bearing capacities of steel beams under uniaxial bending. Dissertation, Ruhr-Universität Bochum.

1291
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Influence of collision damage on load-carrying capacity


of steel girder

E. Yamaguchi, Y. Tanaka & T. Amamoto


Kyushu Institute of Technology, Kitakyushu, Japan

ABSTRACT: An accident that a truck running underneath collides against an overpass bridge
happens occasionally. The influence of the damage on the safety of the bridge must be judged
right away. Yet it is not always an easy task, since the load-carrying capacity of a damaged girder
has not been studied much. The first author has been involved in the evaluation of a steel girder
overpass bridge damaged by collision. Based on the data obtained from this bridge, the load-
carrying capacity of the deformed girder is investigated numerically in the present study. To be
specific, the deformation of the main girder due to collision is reproduced by the finite element
analysis and the deformed steel girder is loaded to evaluate the load-carrying capacity. The result
indicates that as far as the damage is confined to the deformation of the girder, the collision does
not threaten the safety of the bridge even when the deformation is quite large.

1 INTRODUCTION

A truck running on a highway collides with an overpass bridge across the highway occasion-
ally. It is required to evaluate the influence of the accident on the bridge safety. However,
study on such influence appears very limited.
In Japan, the construction of railways preceded that of highways. Therefore, the clearances
under quite a few railway bridges do not satisfy the current requirement, which leads to colli-
sions of trucks against railway bridges. Some technical reports on collision damage in railway
bridges are available in the literature, for example Nieda & Suzuki (2000), Suginoue et al.
(2006) and Nakayama et al. (2008) among others. Most of those technical papers describe
only the damage and the first-aid measure employed without touching on the safety issues
such as the influence on the stiffness and the load-carrying capacity. The investigation by
Nakayama et al. (2008) is one of a very few studies on the influence of collision damage on the
load-carrying capacity of the damaged main girder.
Nakayama et al. (2008) first studied the characteristics of collision damage in the steel
railway bridge. They stated that out of 14 damaged bridges, eight bridges were subjected
to severe damage such as the deformation of track, the fall from the bearing and crack
in the main girder, which have resulted in the immediate closure of those railway bridges.
The remaining six girders underwent only the deformation of the main girders. The
major damages of those six girders were classified into three groups: local upward
deformation of the lower flange, horizontal deformation of the lower flange and the com-
bination of the two. Focusing on the damages of the first two groups, Nakayama et al.
(2008) have investigated the load-carrying capacity of the deformed girder experimentally
and numerically. To that end, they prepared three girder specimens. The span of each
girder was 5360 mm long and the difference between the three girder specimens lay in
the initial deformation: one had no deformation, another girder had a locally upward
displacement of the lower flange up to 78 mm and the lower flange of the other girder
was displaced horizontally up to 27 mm. These initial deformations were decided refer-
ring to the maximum values they observed in the actual railway bridges damaged by the
collision. Their research results have revealed that the damages they considered have

1292
Figure 1. Damaged overpass bridge.

Figure 2. Bridge cross section.

insignificant influence on the load-carrying capacity. They have observed the tendency
that the collision increased the capacity. They stated that a possible reason for this phe-
nomenon was the strain-hardening due to the deformation caused by the collision.
The first author has been involved in the safety evaluation of a steel overpass bridge dam-
aged by collision. The damage of the main girder was severer than the one investigated by
Nakayama et al. (2008). Yet no cracks in the girder and no significant damage around the
bearings were found. The main concern was therefore the load-carrying capacity of the dam-
aged girder. The present paper deals with this issue.

2 OVERVIEW OF BRIDGE

The damaged overpass bridge to be studied is a single-span steel girder bridge (Figure 1). The
bridge is 29.8 m long and the span is 29.0 m long. The superstructure consists of two main
girders, G1 and G2, and orthotropic deck. The main girder has transverse stiffeners while the
orthotropic deck has longitudinal stiffeners and cross girders. Lateral struts are also installed
to support two pipes. The dimensions of the cross section are shown in Figure 2.

3 DAMAGE

Only one of the main girders, G2, was found damaged. This is because the expressway below
the damaged part of the bridge is uphill. The lower flange and the web were displaced

1293
Figure 3. Schematic of deformation due to collision.

Figure 4. Measured residual displacement of lower flange in horizontal direction.

outward, as is shown schematically in Figure 3. The cross girders set in the upper part of the
web helped restrict the web deformation only to the lower part. The residual horizontal dis-
placements of the lower flange of G2 were measured, the results of which are presented in
Figure 4.
29 transverse stiffeners are welded to each web in addition to those at the locations of the
bearings. The cross section having the transverse stiffener is given the number; the section
closest to the abutment A2 is Section 1 and the section closest to the abutment A1 is
Section 29. The circled numbers in Figure 4 correspond to those section numbers.
Figure 4 indicates that the largest horizontal displacement was caused around Section 9,
which is 8.8 m away from the A2 bearing. A truck must have collided against around
Section 9. The displacement at Section 9 is 186 mm, which is 1/156 of the span length. Note
that the displacement of the girder studied by Nakayama et al. (2008) is 1/199 of the span
length and that the Japanese design specifications (Japan Road Association 2012) requires the
initial deflection to be less than 1/1000 of the member length.
The deformation of the main-girder web and the lower flange are not the only damage in
this bridge. In addition to them, some transvers stiffeners were bent and/or buckled; some
welded connections between the transvers stiffeners and the web were fractured, separating
the transvers stiffeners from the web; and some bolted connections between the transvers stiff-
eners and the lateral struts were broken, separating the lateral struts from the transverse stiff-
eners. The photos of those damages are presented in Figure 5.

4 COLLISON LOAD

It was observed that the lateral strut at Section 11 was separated from the transverse stiffener
and that it was held between the two main girders. Since the residual horizontal displacement
at Section 11 is 148 mm, the phenomenon can be created if and only if the maximum horizon-
tal displacement of the lower flange at Section 11 due to the collision had been equal to or
greater than 148 mm while the residual displacement would have been smaller than 148 mm if
not for the lateral strut at Section 11.
Herein the static load that can cause the above situation is estimated by the finite element
analysis that takes into account the material and geometrical nonlinearities. This load has the
effect virtually equivalent to the collision and so it is called the collision load in this study. For
this analysis, ABAQUS (Dassault Systemes Simulia Corp. 2008) is used. Shell elements are
employed for all the members except for the lateral struts that are modelled by beam elements.

1294
Figure 5. Damage in girder.

Since major deformation of the web is in the lower part and the primary role of the ortho-
tropic deck is to provide the main girders and the cross girders with the clamped edge, the
deck is modelled as a very stiff plate for the sake of simplicity.
Steel used for the bridge is SM490Y specified in Japan Industrial Standard. Young’s modu-
lus E is 205 GPa and Poisson’ ratio 0.3. Yield stress is 365 N/mm2 for the plate thickness t less
than 16 mm and 355 N/mm2 for 16 mm ≤ t < 40 mm. Beyond the yield stress, the material
stiffness is assumed to be E/100. The elastic-plastic behavior of von Mises type with the kine-
matic hardening rule is assumed.
As noted earlier, the largest horizontal displacement was found around Section 9 (Figure 4).
The coating on the bottom surface of the lower flange was scratched out near Section 9. It is
then legitimate to conclude that the collision had happened in this portion.
In the present study, based on the observation of the scratch on the coating of the lower
flange, the collision load is assumed to be the uniformly distributed load over 400-mm range,
as the red arrows in Figure 4 indicate. With this load, various analyses are conducted and it is
found that once the load is increased up to 2145 N/mm, the total removal of the load leaves
the displacement of 148 mm. The load-displacement curve A in Figure 6 represents this mech-
anical behavior. The curve A also shows that the displacement is 148 mm at 1650 N/mm.
Therefore, the collision load must be between 1650 N/mm and 2145 N/mm for the lateral strut
at Section 11 to be held between the main girders.
The lateral strut at Section 11 was found bent by 8 mm. The lateral strut is 3200 mm long
and the cross section is shown in Figure 7. The finite element analysis of this member under
axial load is conducted using shell elements and the axial load that causes the 8-mm deflection
is obtained as 45 kN.
The bridge is then analyzed again together with the point load of 45 kN, the reaction from the
lateral strut, at the lower flange of Section 11. It is found that the collision load of 2120 N/mm
leaves the horizontal displacement of 148 mm with the presence of the point load: the

1295
Figure 6. Horizontal load-displacement curve at Section 11.

Figure 7. Cross section of lateral strut.

load-displacement curve B in Figure 6 is the result with the point load of 45 kN at the
lower flange of Section 11. The computed horizontal residual displacement of the lower flange
is plotted together with the measured displacement in Figure 8. Fairly good agreement is
observed, confirming the validity of the present analysis.

5 INFLUENCE OF COLLISION DAMAGE

5.1 Load-carrying capacity of damaged girder


To evaluate the influence of the damage caused by the collision, the load-carrying capacity of the
damaged girder by itself is obtained by the nonlinear finite element analysis. The girder to be
analyzed here is taken out of the bridge deformed by the collision load in the previous section.

Figure 8. Residual displacement of lower flange in horizontal direction.

1296
Figure 9. Applied load.

Figure 10. Influence of damage.

The girder is simply supported. The out-of-plane displacement, the rotation around the vertical
axis and the rotation around the girder axis (torsion) are constrained at the top of the girder. As
shown in Figure 9, two point-loads are applied at the top ends of Sections 8 and 10, as the sever-
est damage is around Section 9. Point C is located at the bottom of the web at the center between
Sections 8 and 10. To quantify the influence of the damage, the intact girder is also analyzed.
The numerical results in terms of the load P and the vertical displacement at Point C are
presented in Figure 10. The figure shows that while the stiffness is reduced by about 8%, the
3% increase in the maximum load, from 243.0 kN to 253.0 kN, is observed. The intact girder
(NO DAMAGE) exhibits little nonlinear behavior, but the damaged girder (DAMAGED)
undergoes nonlinear response rather extensively.
Nakayama et al. (2008) conducted the similar study and observed the same phenomenon of
the increase in the load-carrying capacity due to the collision damage. It is noted that the
damage in the present study is 1.4 times larger than that assumed by Nakayama et al. (2008)
although their damage is just about the maximum they observed in their investigation of the
damaged railway bridges.

5.2 Influence of damage


Nakayama et al. (2008) have suggested that the increase in the load-carrying capacity is
caused by the strain-hardening associated with the plastic deformation due to collision. To
study the effect of strain-hardening, the numerical analysis is conducted, using different
material stiffness beyond the yield stress: in addition to E/100, 0 and E/50 are employed for
the second slope in uniaxial material behavior.
Figure 11 shows that the load-carrying capacity of the damaged girder becomes larger with
the increase of the second slope. But the load-carrying capacity has increased from 245.0 kN
to 246.7 kN even when no strain-hardening is assumed. This result indicates that the phenom-
enon of the increase in the load-carrying due to damage cannot be explained well from the
viewpoint of the strain-hardening.
The difference between the intact girder and the damaged girder lies in the residual stress
and the geometrical configuration. To see the influence of the configuration change solely, the
residual stress is eliminated from the damaged girder and then analyzed. The results are indi-
cated by GEOM in Figure 11.

1297
Figure 11. Influence of strain-hardening and configuration change.

1298
Regardless of the magnitude of the second slope, the difference in the initial stiffness
between DAMAGED and GEOM is very little. The change in the configuration is therefore
the main cause for the reduction in the stiffness. In the results of GEOM, the increase in the
load-carrying capacity is clearly observed, but the magnitude of the second slope doesn’t
make any difference. It can be confirmed that the configuration change must be responsible
also for the increase in the load-carrying capacity.
The residual stress influences the plastification quite much. With the presence of the residual
stress, plastic deformation starts in smaller load while the maximum load is reached with little
plastic deformation if not for the residual stress, which explains little influence of the second
slope on the load-carrying capacity in the results of GEOM.

6 CONCLUDING REMARKS

The influence of the deformation of the steel bridge girder due to collision was studied.
Through the nonlinear finite element analysis, the collision load was identified, which
gives the horizontal residual displacement of the lower flange caused by the collision
fairly well. Using the damaged girder reproduced by the finite element analysis, the influ-
ence of the damage due to the collision was investigated. The result indicated that while
the initial stiffness was reduced by 8%, the bending strength became 3% larger. This phe-
nomenon of the larger strength has been found attributable to that the damaged girder
behaved in a quite different way from that of the original girder because of the configur-
ation change.
From the present study, it may be concluded that the deformation of the girder doesn’t
necessarily threaten the safety of the bridge. However, it needs be noted that the other damage
such as crack, if present, could lead to considerable reduction in safety.

ACKNOWLEDGEMENTS

The financial support for the present study, Grant-in-Aid for Scientific Research (C)
(KAKENHI, No. 15K06184), is gratefully acknowledged.

REFERENCES

Dassault Systemes Simulia Corp. 2008. User’s Manual, ABAQUS Ver. 6.8.
Japan Road Association 2012. Specifications for Highway Bridges: Part 2 Steel Bridges. Tokyo:
Maruzen.
Nakayama, T., Kimura M. et al. 2008. Residual load carrying capacities of riveted steel girders subjected
to collision deformation. Structural Engineering, JSCE 54A: 68–79.
Nieda, H. & Suzuki, H. 2000. The pier overturning accident of inspect and restoration on the steel rail-
way bridge. Proceedings of Annual Conference of Japan Society of Civil Engineers 55(IV): 604–605.
Suginoue, T., Inoue, E., Imai, T. & Oka, Y. 2006. First-aid measures of steel girder deformed by
collision. Proceedings of Annual Conference of Japan Society of Civil Engineers 61(IV): 331–332.

1299
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Modelling of one-sided unstiffened beam-to-column joint

J. Zamorowski
The University of Bielsko-Biała, Bielsko-Biała, Poland

G. Gremza
Silesian University of Technology, Gliwice, Poland

ABSTRACT: In unstiffened beam-to-column joints and in end-plate joints of beams with a few
rows of bolts below the tension flange, as in (Kawecki et al. 2013, Kawecki & Kozłowski 2017),
the load capacity of the bolts in the row below the tension flange of the beam may be exceeded at
the value of the bending moment Mj,Ed ≤ Mj,Rd, where Mj,Rd – design moment resistance of these
joint determined according to the formula (6.25) in EN 1993-1-8. In such cases, there is a need to
reduce the design moment resistance according to point 6.2.4.2 of this standard, so that the forces
in the internal row of bolts or in the group of rows do not exceed the capacity of the T-stub
which imitate this row or group of rows. This work presents the full calculation model of the
unstiffened beam-to-column joint, together with numerical example, in which the need of reduc-
tion of design moment resistance of such joint, which is not taken into account in the examples
published in available literature as well as in computer programs, was demonstrated.

1 INTRODUCTION

In case of the group of bolts consisted of three or more rows number, the capacity of the
equivalent T-stub flange that imitate the inner row may be smaller than the capacity of flange
for the end rows. This may happen when the failure of the T-stub is determined by the load
capacity of the column flange or the end-plate of the beam. The difference in this capacity
may be substantial when a mechanism of failure with non-circular pattern of yield lines
occurs, and smaller when the mechanism with circular pattern is decisive. Such a variation of
the capacity is caused by the different effective length of the equivalent T-stub that imitates
(represents) the bolt rows –end and inner – see Figure 1.
In turn, the difference of the joint stiffness in the area of the inner and the end rows is
substantially smaller. This is due to the fact, that the flexibility of the joint is determined by
all its components, that is: unstiffened web of column in the tension zone, column flange in
bending, end-plate in bending and bolts in tension. Change of the flexibility of one of these
components affects the rigidity to a much lesser extent than in a case of the load capacity of
the row.
Conducted numerical and experimental tests indicate that in the case of the end-plate
joint of a beam to an unstiffened column with several bolt rows, the load capacity of
the bolts in the row below the tension flange of the beam may be exceeded at the value
of the bending moment Mj,Ed ≤ Mj,Rd, where Mj,Rd – design moment resistance of these
joint determined according to the formula (6.25) in EN 1993-1-8. Similar phenomenon
may occur in the case of the end-plate joint of beams with many rows of bolts –
Kawecki et al. (2013) and Kawecki & Kozłowski (2017). This occurs when the resistance
of the inner bolt row is determined by the resistance of the column flange or the end-
plate in bending, the effective length of which is small compared to the effective length
of the flange in the area of the end bolt rows.

1300
Figure 1. Effective lengths of the equivalent T-stub: a) non-circular pattern b) circular pattern.

The point 6.2.4.2 (3) of the standard EN 1993-1-8 contains a provision for an equivalent
T-stub, in which the designer is committed to ensure that the forces transmitted by the individ-
ual bolt rows and groups of these rows do not exceed the design load capacities of the row
and the group of bolt rows. Therefore it should be interpreted, that in a case when it turns out
that the force in any row or group of rows exceeds the computational load capacity of this
row or group, then the load capacity of the joint should be appropriately reduced. However,
the standard does not contain a description of the joint’s calculation model, on the basis of
which one could determine the forces in individual bolt rows. Such models are described in
literature, e.g. Coelho et al. (2005). Figure 2 shows a mechanical model of the beam-to-
column joint, in which the stiffness coefficients as in table 6.10 of EN 1993-1-8 were taken
into account. The stiffness of the individual bolt rows in the tension zone is the inverse of the
sum of flexibilities: the web of the column in tension, the flange of the column in bending,
end-plate in bending and the bolts in tension. According to the standard, in this zone the flexi-
bility of the beam web in tension is omitted. In compression zone, the influence of the flexibil-
ity of the web panel in shearing conditions and the influence of the flexibility of the column
web in compression were taken into account.
In the work an example of the unstiffened beam-to column joint, in which the need to
reduce the load capacity of this joint, according to the point 6.2.4.2 (3) w EN 1993-1-8, was
shown. In this example it turned out, that the forces in the second row of bolts, which were
determined on the basis of the calculation model as in Figure 2, are higher than the design

Figure. 2. Mechanical model of a joint with symbols like in EN 1993-1-8.

1301
load capacity of the T-stub imitating row 2 from the group of rows 1-2, at the design resist-
ance of node Mj,Rd determined from formula (6.25) in EN 1993-1-8.
Most of the examples in the literature present calculation algorithms for the joint of
the beam to the stiffened column, with two or three bolt rows in the tensile zone. In
such joints the need to reduce their load capacity usually does not occur, because the
lengths of the equivalent T-stub are comparable for all rows in the groups. If the joint
is shaped this way, creation of the groups of rows is clear. The rows placed above the
column’s stiffener and above the top flange of the beam in tension form one group of
rows, and below the stiffener and the beam flange - the second group. Difficulties in
grouping the rows of bolts appear in the case of the unstiffened beam-to-column joint.
In this case, groups 1-2 and 1-2-3 are created for the column, and for the beam, the row
located above the flange forms one group, and rows 2-3 forms the second group - see
Figure 3 and Publication P398 (2013).
In the case of such grouping (creation) of rows, the load capacity of rows 2 from group 1-2
is determined for the T-stub of the column flange. In turn, to determine the load capacity of
row 3, the group of rows 1-2-3 for the column and group 2-3 for the beam should be
considered.
The load capacity of row 2, in case of the column, is the difference between the load cap-
acity of the group of rows 1-2 and the load capacity of row 1. In turn, to calculate the load
capacity of the row 3, the load capacities of the rows 1 and 2 from the load capacity of the
group 1-2-3 set for the column should be subtracted. Then, from the load capacity of the
group 2-3 for the end-plate of the beam, the load capacity of the row 2 should be subtracted.
As a result, the value of the load capacity of the row 3 is assumed equal to the smaller value
from the load capacities calculated this way (Publication P398 2013).

2 CALCULATIONAL MODEL OF AN UNSTIFFENED BEAM-TO-COLUMN JOINT

2.1 General description of the model


Model based on the component method provided by EC3, described in the literature,
e.g. Kozłowski et al. (2009) and Publication P398 (2013), which is adopted for assessing
the moment resistances of joint, consists of several stages of calculation. At the first
stage, a minimum resistance of the joint components in compression and the column
web panel in shear is evaluated. The second stage consists of determining the resistances
of the rows, as an individual or as a part of the group. Next, the sum of the resistances
obtained in the rows from 1 to r is compared with the minimum load capacity of the
joint components in the compression zone and in the horizontal shear zone. If this sum
is greater than the minimum resistance of the tension zone and the compression zone,
the resistance of the row r should be reduced. Further reduction takes place in connec-
tions subject to impact loading and vibration. In the third stage, the design moment

Figure. 3. Groups of bolt-rows: a) in column, b) in end - plate.

1302
resistance of the joint from the standard formula (6.25) in EN 1993-1-8 is determined,
which resistance is considered reliable.
In the algorithm included in the presented model, there is no reference to the point 6.2.4.2
(3) of EN 1993-1-8, in which a requirement for the T-stub was introduced that the forces
transferred by the individual bolt rows and the groups of rows should not exceed design resist-
ances of these rows and groups of rows. Therefore, to the computational model of the
moment resistance of the joint the next stage should be introduced, that include determining
the values of forces in individual bolt rows and groups of these rows, at the load of the
joint Mj,Ed = Mj,Rd. Then, the obtained values of forces should be compared with the design
resistance of the rows and the groups of rows. In case it turns out that the forces in the bolt
rows or groups of these rows are greater than their design resistance, it would be necessary to
make the appropriate reduction of the load capacity of the joint defined by the formula (6.25)
in EN 1993-1-8.
This standard does not contain a computational model of the joint, on the basis of
which one could determine the forces in individual rows of bolts. Therefore, it is neces-
sary to use the models included in the literature, e.g. Coelho et al. (2005). For the pur-
pose of the example contained in this work, the model presented in Figure 2 was used.
It was necessary to determine the appropriate joint stiffness coefficients, in accordance
with Table 6.10 in EN 1993-1-8, and, on the basis of them, the flexibility of individual
bolt rows and the flexibility of the compression zone as well as the column web panel in
shear.
The detailed algorithm of the computational model of the load capacity of the joint,
together with the results of calculations that were obtained for the joint presented in Figure 4,
is included in the following tables in the section 2.3.

2.2 Characteristics of the analyzed joint


The analysis was performed on the example of a joint of a beam I 500 PE to a column HE 300
B, with the end-plate and pre-tensioned bolts M20-10.9. Geometrical characteristics are
shown in Figure 4. All joint components except the bolts are made of S235 steel. The analysis
included determining the load capacity and stiffness of the joint by use of the component
method according to EN 1993-1-8.

Figure. 4. Analyzed beam-to column joint with bolts M20-10.9.

1303
2.3 Algorithm and results of calculations of the joint capacity acc. to (6.25) in EN 1993-1-8

2.3.1 Stage 1 – design resistances of basic components in compression and shear

Table 1. Stage 1 - design resistances of basic components in compression and shear.


Column web panel in Column web in transverse Beam flange and web in
shear acc. to 6.2.6.1 compression acc. to 6.2.6.2 compression acc. to 6.2.6.7 Minimum value

Fc;v;min;Rd
8
< Vwp;Rd =β
>
¼ min Fc;wc;Rd
>
:
Fc;fb;Rd

Vwp,Rd/β Fc,wc,Rd Fc,fb,Rd


579.4 kN 588.1 kN 1068 kN 579.4 kN

2.3.2 Stage 2– design resistances of bolt-rows


Table 2. Stage 2.1 - design resistance of the bolt-row 1 considered individually.
Column web in transverse Column flange in transverse End-plate in bend- Design resistance*
tension acc. to 6.2.6.3 bending acc. to 6.2.6.4 ing acc. to 6.2.6.5 acc. to 6.2.7.2(6)

8
< Ft1;wc;Rd
Ft1;Rd ¼ min Ft1;fc;Rd
:
Ft1;ep;Rd

Ft1,wc,Rd Ft1,fc,,Rd Ft1,ep,,Rd


411.0 kN 266.5 kN 264.3 kN 264.3 kN

* Limiting the design resistance due to the compression and shear of the joint parts, acc. 6.2.7.2(7):
Ft1,Rd = 264.3 kN ≤ Fc,v,min,Rd = 579.4 kN.

Table 3. Stage 2.2 - design resistance of bolt-row 2 considered individually.


Column web in Column flange in End-plate in Beam web in Design resistance
transverse tension transverse bending bending tension of the row*
acc. to 6.2.6.3 acc. to 6.2.6.4 acc. to 6.2.6.5 acc. to 6.2.6.8 acc. to 6.2.7.2(6)

8
>
> Ft2;wc;Rd
<
Ft2;fc;Rd
Ft2;Rd ;ind ¼ min
>
> Ft2;ep;Rd
:
Ft2;wb;Rd
Ft2,wc,Rd Ft2,fc,,Rd Ft2,ep,,Rd Ft2,wb,,Rd
411.0 kN 266.5 kN 323.1 kN 624.2 kN 266.5 kN

* Limiting the design resistance due to compression and shear of the joint parts:
Ft2,Rd,ind = 266.5 kN ≤ Fc,v,min,Rd – Ft1,Rd = 579.4 – 264.3 = 315.1 kN

1304
Table 4. Stage 2.3 - design resistance of bolt-row 2 as a part of group of bolt rows 1-2.
Column web in transverse Column flange in transverse
tension bending Design resistance
acc. to 6.2.6.3 acc. to 6.2.6.4 of groups of rows*

Ft,1-2,wc,Rd Ft,1-2,fc,,Rd Ft,1-2,Rd = min (Ft,1-2,wc,Rd, Ft,1-2,fc,,Rd)


506.6 kN 348.5 kN 348.5 kN

* Limiting the design resistance of the bolt-row due to design resistance of the group of bolt-rows accord-
ing to 6.2.7.2(8): Ft2,Rd = min(Ft2,Rd,ind; Ft2,Rd,group), where Ft2,Rd,group = Ft,1-2,Rd – Ft1,Rd
Ft2,Rd,group = Ft,1-2,Rd – Ft1,Rd = 348.5 – 264.3 = 84.2 kN, Ft2,Rd = min(266.5, 84.2) = 84.2 kN.
If in connections exposed to dynamic actions and vibrations
Ft1,Rd > 1.9 Ft2,Rd then Ft2,Rd ≤ Ft1,Rd · h2/h1 – acc. to 6.2.7.2 (9) + polish National Annex
The connection is not exposed to impact and vibration.

Table 5. Stage 2.4 –design resistance of bolt-row 3 considered individually.


Column web in Column flange in End-plate in Beam web in Design resistance
transverse tension transverse bending bending tension of the row*
acc. to 6.2.6.3 acc. to 6.2.6.4 acc. to 6.2.6.5 acc. to 6.2.6.8 acc. to 6.2.7.2(6)

8
>
> Ft3;wc;Rd
<
Ft3;fc;Rd
Ft3;Rd ;ind ¼ min
> Ft3;ep;Rd
>
:
Ft3,wc,Rd Ft3,fc,,Rd Ft3,ep,,Rd Ft3,wb,,Rd Ft3;wb;Rd

411.0 kN 266.5 kN 303.7 kN 537.6 kN 266.5 kN

* Limiting of the bolt-row resistance due to components in bending and shear:


Ft3,Rd ≤ Fc,v,min,Rd – (Ft1,Rd + Ft2,Rd) – acc. to 6.2.7.2(7).
Ft3,Rd ≤ 579.4 – (264.3 + 84.2) = 230.9 kN, therefore Ft3,Rd = 230.9 kN.

Table 6. Stage 2.5 –design resistance of bolt-row 3 as a part of group of bolt-rows 1-3 and 2-3.
Column Column Beam web
web in flange in in
transverse transverse End-plate in tension
tension acc. bending acc. bending acc. acc. to Design resistances
to 6.2.6.3 to 6.2.6.4 to 6.2.6.5 6.2.6.8 of groups of rows*

n n
Ft13;wc;Rd Ft23;ep;Rd
Ft13;Rd ¼ min Ft13;fc;Rd Ft23;Rd ¼ min Ft23;wb;Rd
Ft,1-3,wc,Rd Ft, 1-3,fc,,Rd Ft, 2-3,ep,,Rd Ft, 2-3,wb,,Rd
688.1 kN 569.2 kN 538.0 kN 768.0 kN 569.2 kN 538.0 kN

* Limiting of resistance due to resistance of the group of rows – acc. 6.2.7.2(8):


Ft3,Rd = min(Ft3,Rd,ind, Ft3,Rd, group,c, Ft3,Rd, group,b),
where Ft3,Rd, group,c = Ft1-3,Rd – (Ft1,Rd + Ft2,Rd), Ft3,Rd, group,b = Ft2-3,Rd –Ft2,Rd.
Ft3,Rd, group,c = 569.2 – (364.3 + 84.2) = 220.7 kN, Ft3,Rd, group,b = 538.0 – 84.2 = 453.8 kN
Ft3,Rd = min(230.9, 220.7, 453.8) = 220.7 kN.
If in connections exposed to dynamic actions and vibrations
Ft1,Rd > 1.9 Ftx,Rd, where x = 1,2 then Ft3,Rd ≤ Ftx,Rd · h3/hx – acc. to 6.2.7.2 (9) + polish National Annex
The connection is not exposed to impact and vibration.
1305
2.3.3 Stage 3 – design resistance of the joint in bending Mj,Rd acc. to formula (6.25)
Distance of rows r from the compression zone are (see Figure 4): h1 = 526 mm, h2 = 442 mm
and h3 = 382 mm. According to the formula (6.25) given in EN 1993-1-1, the design moment
resistance is: Mj,Rd = Σ Ftr,Rd · hr = 264.3 · 0.526 + 84.2 · 0,442 + 220.7 · 0.382 = 260.5 kNm.

2.4 Checking the requirements contained in point 6.2.4.2 of EN 1993-1-8


In order to determine the values of forces in the individual bolt rows, a mechanical model of the
joint as in Figure 2 was used. Values of stiffness coefficients ki of the basic components of the
joint are collected in Table 7. Due to the fact that the resistance of the row 2 as a part of the
group of rows 1-2 was determined by plasticization of the column flange in bending, for that
row an elastic-plastic material model with a plastic plateau corresponding to their resistance was
adopted (Ft2,Rd = 84,2 kN). In turn, the resistance of the row 1 was determined by the end-plate,
which plasticization were associated with the destruction of the bolts, as in the failure mode 2 of
the T-stub. For this reason, for the row 1 the elastic model was adopted, because after breaking
the bolts, they will not affect the joint.
The obtained values of forces in the individual bolt rows and in the groups of these rows,
under the load MEd = Mj,Rd = 260,5 kN, are collected in column 2 of Table 8. Column 3 of
this table shows design resistances of these rows and their groups, and column 4 shows the
relation of the resistance to the force. As can be seen, the load capacity of the groups 1-2 and
the row 1 has been exceeded.
According to point 6.2.3.2 (3) of EN 1993-1-8, reduction of the design resistance of the joint
becomes necessary. The largest deficiency of the resistance was found for row 1, where wi = 0.915.
Therefore, the reliable load capacity of the joint should be Mj,Rd, red = 260.5 . 0.915 = 238.4 kNm.
A similar issue occurred during investigation on end-plate connections of beams subjected
to bending (Kawecki et al. 2013) and (Kawecki & Kozłowski. 2017), conducted in Rzeszow
University of Technology. As a result of the performed tests and calculations a significant
underestimation of the bolts resistance in the row near the flange and a significant overesti-
mation of the resistance of the further bolt rows and the whole joint was found.

Table 7. Stage 4.1 – determination of the force values in the individual bolt rows.
Basic component Parameter Bolt-row 1 Bolt-row 2 Bolt-row 3

Column web in transverse tension k3  103 ½m


4.620 2.116 4.267
Column flange in bending k4  103 ½m
44.683 20.466 41.272
End-plate in bending k5  103 ½m
25.129 17.328 13.810
Bolts in tension k10  103 ½m
6.374 6.374 6.374
The effective stiffness coefficient keff ;r  103 ½m
2.296 1.359 2.050
The equivalent lever arm zeq ½m
0.463
Column web panel in shear k1  103 ½m
3.894
Column web in compression k2  103 ½m
8.655

Table 8. Stage 4.2 – Comparison of the values of forces with the load capacity of the rows.
Bolt-rows Force [kN] Design resistance [kN] wi = (3)/(2)

1 2 3 4
1 288.7 264.3 0.915
2 84.2 266.5
3 187.0 230.9
1-2 288.7 + 84.2 = 372.9 348.5 0.935
1-3 372.9 + 187.0 = 559.9 569.2
2-3 84.2 + 187.0 = 271.2 538.0

1306
2.5 Rotational stiffness of the joint
The reduction of the design load capacity of the node also affects the rotational stiffness of the
joint – the relationship M – ϕ. According to EN 1993-1-8, at the range of the moment value in
node Mj,Ed ≤ 2/3 Mj,Rd, the initial stiffness Sj,ini, is applied, and after exceeding this value, the stiff-
ness Sj. The ratio of these stiffnesses is expressed in the standard using the coefficient μ.. The
value of this coefficient is determined by the normative formulas (6.28 a and b). Using the norma-
tive formula (6.27) together with formulas (6.28a) and (6.28b) the relationship M – ϕ, was deter-
mined, where ϕ, = M/Sj for the ranges: 0 ≤ Mj,Ed ≤ Mj,Rd,red – continuous line and 2/3Mj,Rd,red ≤
Mj,Ed ≤ Mj,Rd – dashed line – see Figure 5. When the reduction of the analysed joint load capacity
is included, the stiffness Sj of that joint for M = 2/3Mj,Rd is equal to approximately 78% of the
joint stiffness without this reduction.

Figure. 5. The relationship M - ø: 1 – for Mj,Rd, 2 – for Mj,Rd,red.

3 CONCLUSIONS

It is commonly recognized/accepted, that the fulfillment of the condition (6.23) in EN


1993-1-8, with Mj,Rd calculated from the formula (6.25) is sufficient to ensure the safety of the
end-plate beam-to-column joint. At the same time, the provision included in point 6.2.4.2 (3)
of this standard, which introduces the requirement that the forces in the individual bolt rows
and groups of these rows should not exceed the load capacity of rows and groups of these
rows, is omitted. In case of unstiffened beam-to-column joint, as a result of large differences
in effective lengths of the equivalent T-stub of the outer and inner bolt rows (located outside
and below the upper flange of the beam), plasticization of the column flange in the area of the
inner row may develop. Hence, the resistance of the external bolt row may be exceeded. This
results from the relatively small differences in the stiffness of the individual bolt rows com-
pared to the differences in their design resistances.
This paper presents a full algorithm for calculating of an unstiffened beam to column joint,
which also includes the provisions contained in point. 6.2.4.2 (3) of EN 1993-1-8. This algo-
rithm is illustrated by the results of calculations of such a joint, in which the need to reduce
the joint’s load capacity as a result of non-fulfillment of the conditions contained in these pro-
visions was taken into account. The values of forces in individual bolt rows were determined
using the mechanical model of joint shown in Figure 2, in which the stiffness coefficients of
the components of the joints according to the standard provisions were used. In case of inner
bolt-row, taking into consideration that its resistance was determined by the plasticization of
the column flange, an elastic-plastic model with a plastic plateau at the load-bearing level of
this row was adopted.
The obtained results indicate the urgent need to perform wider experimental tests of joints
with internal bolt rows in groups. This applies both to the beams-to-column joints as well as

1307
to the end-plate beams joints. Until such tests will have been carried out, as a result of which
the need for such a reduction of load capacity would be confirmed or ruled out, in the design
process of such nodes the calculation algorithm presented in this paper should be used.

REFERENCES

Coelho, A.M. & Silva, L.S. & Bijlaard, F.S.K. 2005. Ductility analysis of end plate beam-to-column
joints. In Hoffmeister B. & Hechler O. (ed.), Eurosteel 2005; Proc. 4th European Conference on Steel
and Composite Structures – Eurocodes – Practice, Maastricht 8-10 June 2005. Aachen: Mainz.
EN 1993-1-8: Eurocode 3: Design of steel structures. Part 1.8: Design of joints.
Kawecki, P. & Łaguna, J. & Kozłowski, A. 2013. Analiza nośności doczołowego styku belki dwuteowej
z wieloma szeregami śrub. Czasopismo Inżynierii Lądowej, Środowiska i Architektury. Journal of Civil
Engineering, Environment and Architecture. JCEEA 60(2): 117 –136.
Kawecki, P. & Kozłowski, A. 2017. Badanie rozkładu sił wewnętrznych w zginanych wielośrubowych sty-
kach doczołowych blachownic. Inżynieria i Budownictwo 73(6): 309 –312.
Kozłowski, A. & Pisarek, Z. & Wierzbicki, S. 2009. Projektowanie doczołowych połączeń śrubowych
według PN-EN 1993- 1-1i PN-EN 1993- 1-8. Inżynieria i Budownictwo 65(4): 193 –204.
Publication P398. (2013). Joints in Steel Construction: Moment-Resisting Joints to Eurocode 3. Ascot: The
Steel Construction Institute & London: The British Constructional Steelwork Association.

1308
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Modelling of roof bracings of single-storey industrial buildings

J. Zamorowski
The University of Bielsko-Biała, Bielsko-Biała, Poland

G. Gremza
Silesian University of Technology, Gliwice, Poland

ABSTRACT: The rules included in the standards EN 1090-2 and EN 1993-1-1 are not fully
compatible in terms of the industrial buildings roof bracings modelling, therefore they also do
not guarantee a reflection of the actual behavior of these bracings in proposed, normative calcu-
lation model. This work, by example of single-bay industrial building (hall) with vertical roof
bracings, presents the results of elastic spatial analysis of the structure with geometrical imper-
fections, obtained using various calculation models. These results were referred to the reference
model with global and local imperfections of the roof structure of the hall and compared with
state of forces that can be obtained from the model given in standard rules. This allowed to
conclude about the deficiencies of this model and to point out the need of verification of the
standard provisions in the field of roof bracing.

1 INTRODUCTION

Utilization ratio of the roof bracings in spatial industrial halls is determined by loads and geomet-
rical imperfections of the roof structure introduced during fabrication of steel members and their
assembly. The fabrication of the structure and its assembly should comply with the requirements of
EN 1090-2. This standard defines the permissible values of two types of deviations: essential -
important due to the load-bearing capacity and stability of the structure, and functional - important
due to the possibility of its fit up or appearance. Moreover, according to the point 11.1 of
EN-1090-2, special tolerances in terms of their type or value may be defined in specification.
Due to the distribution of forces in the roof structure of the hall, amongst the manufactur-
ing tolerances listed in EN 1090-2, the following essential deviations may be significant: from
straightness of a press braked profile or a flange of a welded profile Δ ¼ ±L/1000 (where
L – distance between restraints), the deviations of nodal (panel) points relative to the designed
straight line in lattice girders Δ ¼ ±L/500 but not less than /Δ/ ¼ 12 mm as well as deviation of
bracing from straightness Δ ¼ ±L/1000 but not less than /Δ/ ¼ 4 mm. In turn, amongst the
essential erection tolerances, the straightness of member subjected to bending or compression
on the length between restraints Δ ¼ L/750 as well as an angular misalignment Δθ ¼ ±1/500
rad in a full contact end bearing are important.
The range and the values of functional deviations presented in the EN 1090-2, that may
determine the fabrication and the assembly of the roof structure of the hall, are practically
analogous to those in the case of basic deviations, with one exception - in the case of func-
tional assembly deviations, a deviation from the straightness of the beam in the plan with per-
mitted value in class 1 L/500 additionally occur.
If the transverse roof bracings are considered as lattice girders, then for the upper chords of
roof trusses being the part of these bracings an initial deflection out of vertical plane with a value
Δ ¼ ±L/500, but not less than /Δ/ ¼ 12 mm, may be permitted. This deviation may be caused by
the initial deflection of the chord and/or the angular misalignment in the full contact end bearing,
at Δθ  ±1/500 rad. If under the transverse roof bracings the horizontal bracings between lower

1309
chords of roof trusses were installed, than these chords could be mounted with the same deviation
as the upper chords, wherein their initial deflection may have the opposite direction in relation to
the deflection of the upper chords. In turn, when horizontal bracings in the plane of the lower
chords are not installed, then the lower chords of all roof trusses may be kinked at the full contact
end bearing, at Δθ ≤ ±1/500 rad. Moreover, in case of lack of vertical bracings, if compression in
the lower chords might occur, then chords of all trusses might be deflected along their entire
length L, with a maximum value Δ ¼ ±L/750. In turn, if roof vertical bracings were present, the
lower chords could be deflected on the length between bracings with a deviation Δ  ±L/750,
where L - distance between these bracings. The upper chords of the remaining roof trusses could
be kinked in full contact end bearing, at Δθ  ± 1/500 rad. In turn, these chords could not be
deflected along the entire length, because EN 1090-2 does not foresee such a case.
The standard EN 1993-1-1 contains provisions for the calculation of bracings that ensure
lateral stability of beams or compressed members. In this provisions initial bow imperfections
e0 ¼ αm·L/500, at α ¼ [0,5(1+1/m)]1/2 have been taken into account, where m - the number of
members to be restrained, as well as the possibility of kinking of the compressed members
in the joint due to angular misalignment, assuming local impact on the bracing about the value
αm·NEd/100. This value results from kinking of a single chord of truss girder by an angle Δθ ¼ 1/
200 rad. This deformation can be taken into account omitting initial bow imperfection e0. On
the basis of the calculation model of roof bracings contained in the standard EN 1993-1-1, the
internal forces in the upper chords of trusses, purlins and diagonals of bracings can be determined.
However, this model does not provide an opportunity to assess the impact of initial deformation
of the roof on the state of forces in roof vertical bracings as well as in the lower chords, struts and
diagonals of roof trusses. This model also does not capture the influence of the number and loca-
tion of vertical roof bracings on the utilization ratio of the transverse roof bracings members.
In general, industrial halls with latticed roof structure are calculated by use of various models
based on the of the 1st or 2nd order theory. In the second order calculation models, in the equilib-
rium equations of compressed rod, the following influences are usually included: initial and elastic
bow deformation (P – δ effect) on their flexural rigidity and bending on an axial stiffness
(P – κ effects) and the influence of the change in the rod’s chord inclination on the value of nodal
forces (P – Δ effect for the rod). The change of axial stiffness of the initially and elastically
deflected, compressed rod results from its additional deformations along its chord due to the
bending caused by axial force and transverse load. It is usually included in calculations by intro-
ducing substitute (reduced) axial stiffness (EA)z, which is changed during the iterative solution
depending on the total deflection and axial force, wherein the total deflection results also from
the value of the bending moments in the nodes at the rod ends. The reduction of the bending
stiffness of the initially and elastically deflected rod, which is subjected to compression, results
from the additional bending deformations due to the axial force acting on the eccentric. In the
solution algorithm based on the displacement method, this reduction is taken into account by use
of an additional factor with hyperbolic functions that results from such a rod description. In the
case of the finite element method, construction of the equilibrium equation depends on the simpli-
fications made in the tensors of strains and stresses. In turn, in the case of the influence of
a change of the rod’s chord inclination on the values of forces in nodes, the influence of the axial
force on the value of transverse force is usually taken into account, as in the case of the standard
effect of P – Δ within the frames (Zamorowski 2013).
The above information indicate that the above-mentioned influences cannot be directly taken
into account in the linear model, however, it is possible to take into account the influence of the
P – Δ effect iteratively by adding to the co-ordinates of the nodes their displacements in subse-
quent iterative steps. Nevertheless, it is not possible to assess precisely the influence of elastic
deflection of the analyzed rod, e.g. by dividing it into several elements and adding elastic displace-
ments to coordinates of intermediate nodes, because these displacements would be obtained
assuming the initial bending and axial stiffness of these rod. Therefore, the method of calculating
of the equivalent load of bracings proposed in standard may give slightly underestimated results.
In the calculations of roof bracings, similarly as in trusses, different methods are used to take
geometric imperfections into account. These methods are mainly based on either the direct intro-
duction of changes in the geometry of the system and deformation of the rods, so-called IGI

1310
models - modeling of Initial Geometric Imperfections, or on the application of equivalent loads,
so called NHF models - application of Notional Horizontal Forces (Piątkowski 2017).
In the case of IGI models, it is necessary to consider all possible combinations of initial bow
bending of rods, what may be tedious in calculation. In order to limit the number of these
combinations, scaling of the first form of the loss of stability is used here, by the EBM
method - scaling of first Eigen Buckling Mode. Theoretically, the method of randomly select-
ing of initial deflection combinations from the allowable solution set could also be used, but
in this case it would be necessary to search the entire set. During selection of the combinations
of the rods’ initial bow imperfections, one must remember that it is necessary to choose their
arrangement in which the bracings are most susceptible to flexural deformation as a truss. For
example, a cross-section with different slenderness in the transverse bracings plane and in the
truss plane was chosen for the compressed chord: higher susceptibility of that chord will be
obtained assuming the initial bow deformation in the plane of its bigger slenderness.
Technical literature describes various models of bracings based on an equivalent load, in which
this load is either determined on the basis of internal forces in the chords of the roof trusses and
stiffness of the bracings expressed by elastic deflections (EN-1993-1-1, Biegus & Czepiżak 2017,
Giżejowski et al, 2008), or on the basis of external loads and tilting of the truss – initial and elastic,
similarly to the P – Δ effect in frame systems (Niewiadomski & Zamorowski 2017). Each of these
models may be more accurate if the solution is based on the second order theory. In this case,
during additional loads determination, only initial deformations should be taken into account.
There are also some publications in which both aforementioned models are used simultaneously,
whereby the obtained effects of global of bow imperfection in roof bracings are overvalued.
In the real structure of the hall of the imperfect type, additional forces resulting from
deformation and loads appear not only in the transverse roof bracings but also in the vertical
bracings as well as in diagonals and struts of the roof trusses. At the same time, bending of the
lower chord of the truss out of its plane occur. The mechanism of creating these forces is well
explained at work (Niewiadomski 2003). Configuration of the initial deflections of rods and
related change in rigidity of the side restraints of the roof trusses may also affect the coeffi-
cients of buckling lengths of the chords of the trusses subjected to compression when buckling
out of the truss plane (Krajewski & Iwicki 2015).

2 DESCRIPTION OF ANALYSED HALL

Spatial industrial hall consisting of seven transverse systems spaced at 6,0 m, with gable walls
located 1.2 meters away, were analyzed (Figure 1). The height of the main columns (IPE 360) was
8.3 m, and their axial spacing – 24.0 m. Trusses were designed with a height of 3.9 m with end-
plate joints in the middle of their span and a relatively large slope of the roof at 10°. The location
of the hall in the third snow and wind zone in Poland was planned. Chords of the trusses were
designed from ½ HEB 220, posts and diagonals from RHS 100x100x4 profiles. For purlins, IPE
180 is provided. Vertical roof bracings and wall bracings in the columns line were designed as
a cross made of L90  90  6 and 2L60  60 6 respectively. Transverse roof bracings, also type
X, were designed from L60 x 60 x 6. For the bottom horizontal struts of roof vertical bracings
2L60606 were adopted. The load-bearing element supporting the roof covering is a trapezoidal
sheet protecting the purlins from lateral torsional buckling and the purlins near eaves - addition-
ally against flexural buckling. The total weight of the roof covering was 0.45 kN/m2.

3 CALCULATION MODELS OF THE HALL

The work analyzed the spatial structure of the hall as in Figure 1. In the analysis, three nonlinear
models with geometric imperfections in the roof area and four models based on the first order
theory were used. As a reference model (model 1N), the system with global imperfections of the all
trusses chords and with local imperfections of compressed rods was adopted. Global imperfections
were adopted in accordance with EN 1090-2, in particular (Figure 2): bow imperfections of the

1311
Figure 1. Scheme of the analysed building (compressed rods of cross-bracings were removed).

Figure 2. Global imperfections according to EN 1090-2: a) identical to EC-3, b) considered differently


than in EC3 c) not included in EC3.

upper chords of the trusses being simultaneously the chords of the transverse bracings (e0 ¼ L/500),
deviations of the angle misalignment in ridge joints of the upper chords of the remaining trusses
(Δθ ¼ ±1/500 rad), deviations resulting from the angle misalignment in the joints of the lower
chords of all trusses (Δθ ¼ ±1/500 rad). In this model, local geometric bow imperfections of the
compressed rods were also introduced according to the buckling curves defined in the EN 1993-1-1.
An equivalent to this model (equivalent reference model) would be the perfect model with lat-
eral forces applied in all nodes of the upper and the lower chords of all trusses, resulting from the
change of the inclination of the rod chords relative to the vertical planes of the trusses and axial
forces in these rods, i.e. PΔ effects for rods - see Figure 3. In case of the 2nd order analysis, in the
equivalent perfect model, the lateral forces calculated this way should be taken into consideration,
at the displacement of the nodes resulting only from the initial global imperfections of the truss
chords. Analogous forces and displacements are then obtained as in the reference model.
In the second model (model 2N), in comparison to the reference model 1N, global imperfec-
tions of the lower chords of all trusses were omitted; these chords were straightened. In the third
model (model 3N), the global imperfections of the upper and the lower chords were replaced by
the ΔHg forces applied in the nodes of the upper chords. These forces were calculated on the basis
of the load of these nodes and tilting of posts out of the plane of trusses, analogous to the of

1312
Figure 3. Applying of equivalent forces in equivalent reference model and in model 3L.

P – Δ effect in frames. The arrangement of local bow imperfections remained as in the reference
model. In comparison to the equivalent reference model, in model 3N the lateral forces ΔHd in
nodes of the lower chords that results from the initial tilting of the rods out of the plane of the
truss due to the global imperfections of the chords, were not taken into account. Also not
included, in the upper nodes, the influence of horizontal components of forces in rods at initial
lateral displacements of nodes. Only the influence of vertical components of forces in the rods
and lateral displacement of the chords on lateral forces was taken into account.
In linear modeling, four cases were considered: model of a spatial hall without geometrical
imperfections and normative stabilizing loads qd – model 1L, model with stabilizing load qd
applied to the nodes of the upper chords of the end trusses in axes 1 and 7 and with reactions
ΣQd/2 that balance load qd in the end nodes of bracings – model 2L, model with a normative
stabilizing load qd distributed over the upper nodes of all trusses and the opposite directed
equivalent load distributed over the nodes of the lower chords of all trusses – model 3L – see
Figure 3 as well as in-plane 2D normative model, in which the force values introduced in the
upper chords of girders from the model 1L were obtained – model 4L.

4 ANALYSIS OF CALCULATION RESULTS

The obtained results of calculations are presented in Table 1 (markings of the rods are shown in
Figure 4). Columns 3 to 9 of the Table 1 present values of the axial forces obtained for individ-
ual members of the roof structure by use of the calculation models mentioned in section 3: in
trusses, transverse roof bracings and vertical roof bracings. In case of the trusses, besides the
extreme value of the force that was chosen for the rod from the seven roof trusses, after the
slash, the axis number of the truss in which it occurred, was also given. In turn, in case of rods
in roof bracings, also after the slash, the side of the hall at which the extreme force value
occurred – windward - n or leeward - z, was given.

1313
Table 1. Values of maximum forces (in kN) in members of the roof structure.
Nonlinear models Linear models
Members of Rod
the roof marking 1N 2N 3N 1L 2L 3L 4L
1 2 3 4 5 6 7 8 9
Truss upper 5 -149,9/6 -149,0/6 -149,1/6 -148,1/6 -151,0/6 -148,1/5 -152,1/6
chords 6 -210,2/5 -210,2/5 -209,7/5 -209,1/5 -209,1/5 -209,1/5 -209,1/5
7 -220,6/5 -220,6/5 -220,1/5 -220,9/5 -220,9/5 -220,9/5 -220,9/6
8 -204,4/5 -204,4/5 -203,9/5 -203,4/5 -203,4/5 -203,4/5 -203,4/5
Truss lower 14 154,2/4 154,2/4 154,3/4 153,8/4 153,8/4 153,8/4 -
chords 15 215,1/4 215,0/4 215,0/4 215,1/4 215,1/4 215,1/4 -
16 227,7/4 227,7/4 227,7/4 227,9/4 227,9/4 227,9/4 -
Truss 21 177,9/4 177,9/4 178,0/4 178,2/4 178,2/4 178,2/4 -
diagonals 23 77,0/4 77,0/4 77,0/4 77,0/4 77,0/4 77,0/4 -
25 18,6/2 18,6/2 18,5/2 19,5/7 20,2/7 21,1/7 -
27 -27,6/6 -27,6/6 -27,3/6 -27,6/6 -28,2/6 -28,9/6 -
Truss 22 -89,6/4 -89,6/4 -89,6/4 -89,3/4 -89,3/4 -89,3/4 -
struts 24 -45,8/4 -45,8/4 -45,7/4 -45,8/4 -45,8/4 -45,8/4 -
26 -11,9/2 -11,9/2 -11,8/2 -12,2/7 -12,7/7 -13,2/7 -
35 39,8/6 39,5/6 39,6/6 38,7/5 39,6/6 38,7/5 -
Roof braces 121 38,27/n 37,72n 37,48/n 35,33/n 38,36/n 35,88/n 49,30/n
123 14,61/z 14,24/z 14,43/z 14,16/z 15,12/z 12,97/z 17,49/n
Purlins 449 -13,48/n -13,31/n -13,57/n -13,09/n -13,07/n -13,23/n -46,02/n
450 3,49/n 3,55/n 3,60/n 0,34/z 1,24/z 0,32/z -0,35/n
-1,82/n -1,81/n -1,74/n -2,73/n -3,67/n -2,62/n
451 -28,61/n -28,36/n -28,36/n -27,73/n -29,02/n -27,33/n -33,77/n
452 -9,72/z -9,68/z -9,87/z -16,83/z -15,97/z -16,80/z -0,46/n
Vertical 453 5,84/z 6,13/z 5,87/z 9,88/z 9,18/z 10,30/z -23,74/n
bracings -24,10/n -23,94/n -24,02/n -24,23/n -24,67/n -23,38/n -0,55/z
611 -7,08z -7,31/z -6,95/z 7,27/n 6,61/n 3,92/n -
-5,59/z -5,66/z -7,13/z
610 8,34/z 9,09/z 7,66/z 10,31/z 10,04/z 15,05/z -
-8,10/n -7,46/n -8,03/n -4,73/n -5,00/n -2,68/n
609 5,76/n 5,51/n 5,72/n 3,71/n 3,82/n 2,92/n -
-0,48/n

The calculation results indicate that the values of the forces in the upper chords of the trusses
obtained from the in-plane 2D (flat) calculation model provided by the standard (col. 9) are in
practical accuracy consistent with the values of forces obtained using the reference model (col. 3).
Similar accuracy was obtained in case of the linear 2L model. A slightly better coincidence of
results in the upper chords of trusses was obtained using the non-linear model with ΔHg forces

Figure 4. Markings of the rods in the model.

1314
Table 2. Extreme values of bending moments in the lower chords (in kN).
Calculation models

nonlinear linear

Truss in axis 1N 2N 3N 1L 2L 3L
1 2 3 4 5 6 7
1 0,75 0,76 1,07 0,88 0,95 2,16
2 3,78 3,65 3,10 3,15 3,33 4,10
3 0,78 0,88 0,66 0,49 0,51 1,41
4 0,80 0,90 0,68 0,51 0,52 1,41
5 0,80 0,91 0,69 0,52 0,52 1,42
6 0,82 1,70 1,55 0,81 0,91 2,42
7 4,06 3,92 3,63 3,42 3,52 4,37

(col. 5), as well as the linear spatial model of the hall with the normative stabilizing loads qd
applied to the nodes of the upper chords of trusses in axes 1 and 7 and the equivalent load in the
form of ΣQd/2 forces in the nodes on the ends of bracings (col. 7). Practical agreement of extreme
values of axial forces, for all calculation models, also occurred in the rods of the lower chords, in
diagonals of the last and penultimate panels of trusses, as well as in struts, except the middle and
one the closest to the middle.
As the calculations shows, a lateral flexural deformation of the lower chords of the trusses
occurred in all spatial non-linear and linear models. This deformation is accompanied by bend-
ing moments with values as shown in Table 2. The order of models in this table follows the
order in the Table 1.
It can be noted that safe (higher) values of the bending moments in comparison to the refer-
ence model (col. 2) were obtained for the linear model 3L (col. 7). In this model, a stabilizing
load Qd,t was applied to the nodes of the upper chords of all trusses, where a – spacing of the
nodes in the projection. This load have been balanced by the opposite directed load Qd,b,
applied in the lower nodes of the all trusses – see Figure 3. Values of these loads were obtained
by adding the stabilizing load applied to the upper chords of trusses and then dividing them
into 7 trusses and spreading to 7 nodes of the lower chord, Qd,b ¼ Σqd.a/7, where a – spacing
of nodes along the bottom chord. The sum of these loads is zero.
In the actual roof structure, the nodes of the lower chords will be affected by the forces
resulting from the effect of P – Δ for the rods leaned out of vertical plane of the truss - see
description of the equivalent reference model.
In the model 3L, the extreme force in the strut below the ridge is about 3% smaller in rela-
tion to the force from the reference model and in the posts near to the symmetry axis of the
hall - higher about 11%. This model also underestimates the extreme forces in diagonals of
the transverse bracings up to 11.2%, in the purlins up to 4.5% and overestimates the values
of the tension forces in the vertical bracings. Due to its simplicity, the model 3L can be used to
estimate the forces in the bracings of the roof structure.
Flat (2D) calculation model may be useful only for determining the forces in diagonals and
purlins of the transverse roof bracings (being the members of these bracings).
The results best fitted to the reference model in all rods were obtained using the non-linear
model 3L, in which the global chord deformations were replaced by ΔHg forces resulting from
tilting of truss and nodal loads.

5 CONCLUSIONS

In engineering practice, various types of calculation models are used to design the roofs of indus-
trial halls. The simplest of them, the standard in-plane 2D model, only allows to estimate the
forces in members of the lateral transverse bracings, while the forces in vertical braces will not be

1315
known. The influence of global imperfections of trusses’ chords on the utilization ratio of their
bottom chords, diagonals and struts will also not be determined. In turn, when the normative sta-
bilizing load qd according to the EN 1993-1-1 would be introduced into the spatial model of the
hall calculated according to the 1st order theory, then the influence of tilting of the rods axes out
of the planes of the trusses on the loading of their lower chords will not be taken into account. To
eliminate this deficiency in the spatial model of a hall calculated according to the first order
theory, besides the stabilizing load qd applied in the upper nodes of the end trusses (or distributed
over all the trusses), equivalent to them lateral load should be introduced to the lower chords of
all trusses – see Figure 3. The total value of this load results from general equilibrium conditions,
and the sum of stabilizing loads in the nodes of the upper chords and the equivalent load in the
nodes of the lower chords should be zero (in known solutions presented in the literature, equiva-
lent load is applied at the end nodes of bracings). The value of the equivalent load applied to the
nodes of the lower chords may be approximated by summation the stabilizing load from the
nodes of the upper chord and then dividing them by the number of trusses and the number of
internal nodes of their lower chords.
It is also possible to calculate the spatial hall using the second order (or non-linear) theory
and stabilizing loads qd in the upper nodes of the end trusses (or in the upper nodes of all
trusses), and equivalent load in the bottom nodes of all the trusses. In such a case, these loads
should be determined from the standard formula (5.13) given in EN 1993-1-1, assuming δq = 0.
In order to more accurately estimate the value of the nodal equivalent load for a nonlinear
model or a model based on the second order theory, the out of the truss plane tilts of all the
rods converging at the node should be determined. Next, the values of lateral actions on the
nodes ΔHdi = Σϕi·NED,i should be added together ϕi = (uki – upi)/li, whereby uki, upi – horizontal
displacement of the end-and-start nodes of the i-th rod, and NED – axial force in the rod).
The most accurate calculation model is the model of the spatial hall with geometric imperfec-
tions – the global of the whole chords and the local of the individual rods. The limit values of
global initial imperfections that may occur in the properly assembled hall are defined in EN
1090-2. In turn, the EN 1993-1-1 standard provides equivalent values of geometric imperfec-
tions - local and global, which should be taken into account when using a non-linear or second
order theory. There is no full compatibility between these standards. It would be advisable to
specify which global imperfections and their values should be included in the spatial calculations
of the halls according to the second order theory. There are also no guidelines regarding the
values of equivalent (stabilizing) loads, which should be introduced to the lower nodes of trusses
when the hall is treated as a spatial structure and calculated according to the 1st order theory.

REFERENCES

Biegus, A. & Czepiżak, D. 2017. Obciążenia imperfekcyjne elementów wytężonych znakozmienną


wzdłużnie siłą osiową. Czasopismo Inżynierii Lądowej, Środowiska i Architektury Journal of Civil Engin-
eering, Environment and Architecture. JCEEA 64(3/I): 371–386.
Giżejowski, M. & Barszcz, A. & Ślęczka L. 2008. Projektowanie stężeń stalowych układów konstrukcyjnych
według PN-EN 1993- 1-1. Inżynieria i Budownictwo 64(11): 614–621.
Krajewski, M. & Iwicki, P. 2015. Analysis of brace stiffness influence on stability of the truss. International
Journal of Applied Mechanics and Engineering 20(1): 97–108.
Niewiadomski, L. 2003. Ocena wpływu imperfekcji ściskanych pasów dźwigarów kratowych na siły
wewnętrzne w stalowych elementach konstrukcji dachu. Zeszyty Naukowe Politechniki Śląskiej.
Seria Budownictwo 101. 287–294.
Niewiadomski, L. & Zamorowski, J. 2017. Wstępne imperfekcje łukowe w analizie połaciowych stężeń
poprzecznych. Zeszyty Naukowe Politechniki Częstochowskiej. Seria Budownictwo 23: 231–244.
Piątkowski, M. 2017. Metody uwzględniania imperfekcji geometrycznych w kratownicach stalowych.
Czasopismo Inżynierii Lądowej, Środowiska i Architektury. Journal of Civil Engineering, Environment
and Architecture. JCEEA 64 (4/I): 229–243.
Zamorowski, J. 2013. Przestrzenne konstrukcje prętowe z geometrycznymi imperfekcjami i podatnymi
węzłami. Gliwice: Wydawnictwo Politechniki Śląskiej.

1316
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Buckling assessment of cylindrical steel tanks with top stiffening


ring under wind loading

Ö. Zeybek
Department of Civil Engineering, Middle East Technical University, Ankara, Turkey
Department of Civil Engineering, Muğla Sıtkı Koçman University, Muğla, Turkey

C. Topkaya
Department of Civil Engineering, Middle East Technical University, Ankara, Turkey

ABSTRACT: A stiffening ring is commonly used at the top of the tank wall to increase its
strength against external pressure instability. Traditional design treatments generally consider
cylindrical storage tanks under uniform external pressure for sizing of the top ring. However,
cylindrical steel tanks under non-uniform wind loading have rather different and complex buckling
behaviour from those of tanks subjected to uniform external wind loading. In this study, the buck-
ling resistance of the cylindrical steel storage tanks with top stiffening ring under wind loading is
investigated using finite element analyses. The changes in the buckling capacity are studied in light
of the proposed stiffness ratio for a particular harmonic of wind loading. The results revealed that
the changes in the buckling capacity are closely related to the shell-top ring stiffness ratio. Further-
more, a generalized solution that shows buckling pressure ratio (qcr,w/qcr,D) is then developed.

1 INTRODUCTION

Cylindrical ground-supported storage tanks are widely used to store a great variety of liquids for
both short and long term purposes. A very common failure mode of such a storage tank is under
wind loading, where the tank wall may have the insufficient strength and stiffness to resist external
pressure. The wall thickness of a storage tank is normally chosen to resist only the internal pres-
sure from the stored product. Storage tanks are susceptible to buckling under wind load when
they are empty or at a low-level of filling (Ansourian 1992, Flores & Godoy 1998, Maraveas et al.
2015). One way to strengthen the tank wall is to use a stiffening ring placed at the top or near the
top of the tank (Figure 1). This ring also plays an important role in maintaining circularity when
the tank is subjected to wind loads. The circularity is particularly important when a floating roof
is used within the tank, and this condition also means that there is no structural roof to resist
buckling displacements at the top of the wall. Because the tank is very thin, it is very susceptible to
buckling under external pressure, and under wind this is exacerbated because the pressure varies
significantly around the circumference, flattening the wall locally and inducing stresses in different
directions (Maher 1966). Thus, the specific pattern of pressure variation around circumference is
of great importance in design against wind. Traditional design treatments generally consider cylin-
drical storage tanks under uniform external pressure for the sizing of the top ring. However, cylin-
drical steel tanks under non-uniform wind loading have rather different and complex buckling
behaviour from those of tanks subjected to uniform external wind loading (Chen & Rotter 2012).
The stiffening ring should be designed against buckling under external pressure. For
this purpose, this ring must have an adequate stiffness to fulfil its function. Buckling of
ring stiffened cylinders under non-uniform wind loading is a rather difficult problem in
shell analysis.
Blackler (1986) proposed an expression for the minimum stiffness requirement of top stiff-
ening ring under uniform external pressure as follows:

1317
Ir;min ¼ 0:048Lt3 ð1Þ

where L = length of the cylindrical shell; t = thickness of the shell wall, Ir,min = minimum
moment of inertia of the ring.
Schmidt (1998) also suggested a minimum moment of inertia for the top ring based on the
post-buckling behaviour under external pressure.

Ir;min ¼ 0:5Lt3 ð2Þ

Greiner & Guggenberger (2004) identified a different requirement for the limiting stiffness,
which this value was about 10 times the proposal of Blackler (1986).

Ir;min ¼ 0:48Lt3 ð3Þ

Eurocode EN 1993-4-1 (2007) requires both a strength and a stiffness requirement for the top
ring. It has a moment capacity requirement to address the effect of uniform external pressure
on a ring that is imperfectly round, as well as a further requirement for wind conditions.
A further requirement is for the flexural rigidity of the ring, intended to ensure that the ring
does not participate in the buckling mode if the shell wall buckles. The expressions for the
flexural rigidity of the ring about its vertical axis were given as follows:
EIr;min ¼ k1 ELt3 ð4Þ
pffiffiffiffiffiffi
EIr;min ¼ 0:08Cw Ert3 r=t ð5Þ

where Cw = the wind pressure distribution coefficient, r = radius of the shell, E = modulus of
elasticity, k1 = 0.1 is recommended.
According to this specification, the bending stiffness of the girder should be sufficiently
large to restrict the out-of-round displacement under wind action of the ring stiffened cross-
section to 2% of the radius (ECCS 2014). The above recommendations present a wide range
of required minimum stiffness values for the ring size.
There are two potential requirements for the top stiffening ring: strength to ensure that the ring
does not yield under the unsymmetrical loads that will be applied to it under wind, and the stiff-
ness to ensure that the buckling assessment of the tank wall under wind can be based on complete
radial restraint at the top. This paper addresses the latter using a more thorough analysis than any
found in previous studies. The buckling resistance of the cylindrical steel storage tanks with top
stiffening ring under wind loading that varies along the circumference is investigated using finite
element analyses. A shell-ring stiffness ratio is devised algebraically from the relative radial stiff-
nesses of the ring and the cylindrical shell under each harmonic of wind loading. The changes in
the buckling capacity are studied in light of the proposed stiffness ratio.

Figure 1. Rendering of tank and stiffening ring system.

1318
2 WIND PRESSURE ON CIRCULAR CYLINDERS

The wind pressure distribution around a cylindrical shell has been studied extensively using wind
tunnel tests (Maher 1966, Purdy et al. 1967, Resinger & Greiner 1982, MacDonald et al. 1988,
Uematsu et al. 2018) where significant amounts of experimental data have been collected, leading
to the characteristic pattern. The wind distribution for an isolated cylindrical structure depends on
many parameters like the Reynolds number of the wind flow, the cylinder aspect ratio and poten-
tially the shape of the roof (Maher 1966). The wind pressure varies both up the height and around
the circumference of a cylindrical shell structure. However, the vertical pressure variation on the
cylinder is usually assumed to be constant for tank structures because the aspect ratio is relatively
low, leading to a relatively small vertical variation in the pressure (Bu & Qian 2016, Shokrzadeh
& Sohrabi 2016) when compared with the major circumferential variation. Circumferential distri-
bution of wind pressure on a circular tank structure can be reasonably approximated by a Fourier
harmonic cosine series of the form as follows (EN 1993-4-1 2007):

X
4
Cp ðθÞ ¼ ai cosðmθÞ ¼ 0:54 þ 0:16ðdc =LÞ þ f0:28 þ 0:04ðdc =LÞg cos θ
m¼0
ð6Þ
þ f1:04  0:20ðdc =LÞg cos 2θ þ f0:36  0:05ðdc =LÞg cos 3θ
 f0:14  0:05ðdc =LÞg cos 4θ

where θ = the circumferential angle measured from the stagnation meridian, dc = the diameter
of the cylindrical shell, ai = coefficients of each harmonic, m = the harmonic number.
The range of wind pressure distributions considering different tank aspect ratios given in
the Eurocode standard EN 1993-4-1 (2007) is shown in Figure 2. The expression given in
Equation 6 considers significant effects of the aspect ratio of the cylindrical structures.
Small changes in the aspect ratio may considerably alter the wind pressure profile. But, this
pressure distribution changes little as the length increased for intermediate and tall cylindrical
structures.

3 SHELL-RING STIFFNESS RATIO

The wind loading is principally resisted by membrane shear in the shell which transmits the
translational force to the support. But because it contains loading components in higher har-
monics than m=1, it also induces bending in the top ring and in the shell according to their
relative stiffnesses. The ring must have an adequate stiffness to fulfil its function. A test for its
adequacy was developed by comparing the relative stiffnesses of the ring and the shell under
non-uniform load.

Figure 2. Wind pressure distribution around circumference for circular structures with different aspect
ratios (EN 1993-4-1, 2007).

1319
Kshell qr ðθÞ=ur;shell ur;ring
χ¼ ¼ ¼ ð7Þ
Kring qr ðθÞ=ur;ring ur;shell

where Kshell and Kring are radial stiffnesses of the shell and the top ring respectively; ur,ring
and ur,shell are radial displacements of the ring and shell respectively, qr(θ)= the circumferen-
tial variation of the non-uniform line load considering wind pressure ( q) as defined by
Equation 8.

qr ðθ Þ ¼ 
q  L  ai cos mθ ð8Þ

For any single harmonic loading, closed-form expressions were obtained for the radial dis-
placement of the shell and stiffening ring.

3.1 Ring beam stiffness


The Vlasov curved beam differential equations (Vlasov 1961, Heins 1975) were used to
study the response of the top ring. The equilibrium equations were first derived for the
curved beam element shown in Figure 3, where three orthogonal internal forces and three
internal moments develop at each cross-section. The six basic equilibrium equations can be
expressed as follows:

1 dQr 1 dQx
þ Qθ þ qr ¼ 0 þ qx ¼ 0 ð9Þ
r dθ r dθ

1 dQθ 1 dMr
 Qr þ q θ ¼ 0 þ Tθ  Qx þ mr ¼ 0 ð10Þ
r dθ r dθ

1 dMx 1 dTθ
þ mx þ Qr ¼ 0  Mr þ mθ ¼ 0 ð11Þ
r dθ r dθ

where Mr = bending moment in the ring about a radial axis; Mx = bending moment in the ring
about a transverse axis; Tθ = torsional moment in the ring; qx, qθ, qr = distributed line loads per
unit length in the transverse; circumferential and radial directions respectively; mx, mθ, mr = dis-
tributed applied torques per unit circumference about the transverse, circumferential and radial
directions respectively; Qθ = circumferential force in the ring; Qx, Qr = shear forces in the ring
in transverse and radial directions respectively.
The differential equation for bending of the ring in its own plane can be uncoupled from the
other two. For the case of wind loading, only radial loads qr are needed
(i.e. qθ = qx = mr = mθ = mx = 0), and the uncoupled differential equation of equilibrium becomes:

Figure 3. Differential curved beam element and sign conventions.

1320

1 d 3 Mx dMx dqr
þ ¼ ð12Þ
r2 dθ3 dθ dθ

Equation 12 can be solved for the loading condition (qr) defined in Equation 8 to arrive bend-
ing moment in the ring about a transverse axis:

ai
Mx ðθÞ ¼  
q L r2 cos mθ ð13Þ
ðm2  1Þ

The following force-deformation expression was used to obtain radial displacement (ur):
 
EIx d 2 ur
Mx ¼ þ ur ð14Þ
r2 dθ2

where Ix= bending moment of inertia of the ring about transverse axis.
The radial displacement of the ring can be found using Equation 13 as follows:

ai 
q L r4
ur ðθÞ ¼ 2
cos mθ ð15Þ
ðm2  1Þ EIx

3.2 Shell stiffness


The radial shear loading (qr(θ)) applied to the edge of the shell shown in Figure 4 is carried by
what Calladine (1983) terms an edge-string. This non-symmetric load can be transformed into
a tangential shear loading (Nxθ) at the top edge of the shell as shown in Figure 4.
The membrane theory of shells (Flügge 1973, Calladine 1983, Rotter 1987, Ventsel &
Krauthammer 2001) was adopted to obtain the radial displacements of the shell under circum-
ferential shear loading using the edge-string treatment.
Considering the cylindrical shell element shown in Figure 5, the membrane theory of equi-
librium equations (Rotter 1987, Ventsel & Krauthammer 2001) are:

∂Nx 1 ∂Nxθ ∂Nxθ 1 ∂Nθ


þ þ px ¼ 0 þ þ pθ ¼ 0 Nθ þ rpn ¼ 0 ð16Þ
∂x r ∂θ ∂x r ∂θ

Figure 4. Transferring applied non-uniform radial edge loading to tangential shear loading (Calladine
1983).

1321
where r = middle surface radius; Nx, Nθ, Nxθ = axial, circumferential, and shear membrane
stress resultants, respectively; and px, pθ, pn = external distributed pressures in the axial, cir-
cumferential and radial directions, respectively.
The membrane stress resultants (Ventsel & Krauthammer 2001) were given as:
Z  Z 
1 ∂Nθ 1 ∂Nxθ
Nθ ¼ rpn Nxθ ¼  pθ þ dx þ f1 ðθÞ Nx ¼  px þ dx þ f2 ðθÞ ð17Þ
r ∂θ r ∂θ

where f1(θ), f2(θ) = unknown functions of θ to be determined from two boundary conditions.
The displacements were found considering strain-displacement relationships as:
Z Z Z
Et ∂ux
Etux ¼ ðNx  Nθ Þdx þ f3 ðθÞ Etuθ ¼ 2ð1 þ Þ Nxθ dx  dx þ f4 ðθÞ ð18Þ
r ∂θ
∂uθ
Etur ¼ Et  rðNθ  Nx Þ ð19Þ
∂θ

where ux, uθ, ur = displacements in the axial, circumferential, and radial directions, respect-
ively; ν = Poisson’s ratio; f3(θ), f4(θ) = additional functions to satisfy the boundary condi-
tions on the edges x = constant.
The initial algebraic treatment here involves no surface loading on the shell (px = pn = pθ = 0)
but involves only tangential shear loading Nxθ at the top (Figure 4). Considering appropriate
boundary conditions which were given in Figure 4, the radial displacement can be found as fol-
lows at the top of the shell (x = L).

2
3r ð þ 2Þ þ L2 m2
ur ðθÞ ¼ m2 ai 
q L2 cos mθ ð20Þ
3Etr2

3.3 The shell-ring stiffness ratio


The ratio of the stiffness of the shell to the ring (χ) was found by combining the above expres-
sions. Inserting Equations 15 and 20 into Equation 7 yields the shell-ring stiffness ratio (χ)
which can be expressed as:

ur;ring 1 3r6 t
χ¼ ¼ ð21Þ
ur;shell m2 ðm2  1Þ2 LIx ½3r2 ð þ 2Þ þ m2 L2

Figure 5. Loading, displacements and stress resultants in an element of the cylindrical shell.

1322
4 BUCKLING BEHAVIOUR OF THE CYLINDRICAL TANK STRUCTURES

4.1 A brief assessment of each harmonic wind loading


The effect of each harmonic of loading recommended by EN 1993-4-1 (2007) was investigated
separately. A unit pressure was applied at the stagnation point. The commercial finite element
program ANSYS v12.1 (2010) was used to perform these numerical analyses. Numeric studies
used a constant radius-to-thickness (r/t=1000) ratio and height-to-diameter (L/dc) ratios of
0.25, 0.5, 0.75, and 1.0 considering different size annular plate ring. The uniform buckling
pressure was determined from Donnell theory as follows.

 t 2 pffiffiffi
rt
ffi
qcr;D ¼ 0:92E ð22Þ
r L

From each analysis, the stagnation pressure at buckling (qcr,w) was extracted using linear
bifurcation analysis. Then, these values were normalised by the buckling uniform pressure
obtained from Donnell theory (qcr,w/qcr,D). The buckling pressure ratios for each harmonic of
loading were plotted in Figure 6 as a function of the shell-ring stiffness ratio.
The comparison shows that the dominant harmonic term is m = 2 for all cases.

4.2 Linear bifurcation analysis for cylindrical steel tanks with stiffening ring under
wind loading
The linear buckling behaviour of the cylindrical steel storage tanks with top stiffening
ring under wind loading was investigated for the same geometries as in the previous
part, but r/t changes from 500 to 2500 for all cases. From each analysis, the stagnation
pressure at buckling (qcr,w) was normalised by uniform pressure obtained from Donnell
theory (qcr,w/qcr,D). The buckling pressure ratios for wind loading are plotted in
Figure 7 as a function of the shell-ring stiffness ratio considering dominant harmonic
term of m = 2.
The relationship between buckling pressure ratio and stiffness ratio can be represented by:

qcr; w
¼ f ðχÞ ð23Þ
qcr;D

Figure 6. Buckling pressure ratios with different aspect ratios under each harmonic wind loading.

1323
Figure 7. Buckling pressure ratios with proposed equations under wind loading.

Table 1. Coefficients used in


the Equation 24.
L/dc C1 C2

1.0 1.71 0.07


0.75 1.66 0.15
0.50 1.58 0.36
0.25 1.52 0.87

where f(χ) is a function that can be approximated by curve fitting to the data (Figure 7) as

0:8
f ðχÞ ¼ C1    r r 2 ð24Þ
1 þ Cχ2
L t

where the C1 and C2 = the coefficients given in Table 1.


The proposed expressions considering the shell-ring stiffness ratio are also shown in Figure 7.
Obviously, the recommendations given by Schmidt (1998) and Greiner & Guggenberger (2004)
are fairly close to each other for design of stiffening rings, but differ from results obtained by
Blackler (1986) and EN 1993-4-1 (2007).

5 CONCLUSIONS

This paper has presented a new stiffness requirements of the stiffening ring at the top of
the wall of a ground-supported cylindrical tank under wind loading. The buckling resist-
ance of the cylindrical steel storage tanks with top stiffening ring under wind loading
that varies along the circumference is investigated using finite element analyses. The
changes in the buckling capacity are studied in light of the proposed stiffness ratio that
represents the ratio of stiffnesses of the shell and the top ring for a particular harmonic

1324
of wind loading. The results revealed that the changes in the buckling capacity are
closely related to the shell-top ring stiffness ratio. Furthermore, a generalized solution
that shows buckling pressure ratio (qcr,w/qcr,D) is then developed as a function of the
shell-top ring stiffness ratio.

REFERENCES

ANSYS, Version 12.1 On-Line User’s manual, 2010.


Ansourian, P. (1992) On the buckling analysis and design of silos and tanks, Journal of Constructional
Steel Research., 23, pp. 273–294.
Blackler, M.J. (1986) “Stability of Silos and Tanks under Internal and External Pressure”, PhD Thesis,
School of Civil and Mining Engineering, University of Sydney, Australia.
Bu, F., and Qian, C. (2016) On the Rational Design of the Top Wind Girder of Large Storage Tanks,
Thin-Walled Structures, 99, pp. 91–96.
Calladine, C.R. (1983) Theory of shell structures. Cambridge University Press, U.K.
Chen, L. & Rotter, J. (2012) Buckling of anchores cylindrical shells of uniform thickness under wind
load, Eng. Struct. 41, pp. 199–208.
EN 1993-4. (2007), Eurocode 3: Design of steel structures, Part 4.1: Silos, Eurocode 3 Part 4.1, CEN,
Brussels.
Flores, F.G. & Godoy, L.A. (1998) Buckling of short tanks due to hurricanes, Engineering Structures, 20
(8), pp. 752–760.
Flügge, W. (1973) Stresses in Shells. Springer-Verlag, Berlin.
Greiner, R. & Guggenberger, W. (2004) “Tall cylindrical shells under wind pressure”, in Buckling of Thin
Metal Shells, eds J.G. Teng & J.M. Rotter, Spon, London, pp 198–206.
Heins, C.P. (1975) Bending and torsional design in structural members. Lexington Books, Lexington,
Massachusetts.
MacDonald, P.A., Kwok, K.C.S. & Holmes, J.D. (1988) Wind Loads on Circular Storage Bins, Silos
and Tanks: I. Point Pressure Measurements on Isolated Structures, Journal of Wind Engineering and
Industrial Aerodynmamics, 31, pp 165–188.
Maher, F.J. (1966) Wind loads in dome-cylinder and dome-cone shapes, ASCE Journal of the Structural
ivision, Vol 92 (5), pp. 79–96.
Maraveas, C., Balokas, G.A. & Tsavdaridis, K.D. (2015) Numerical evaluation on shell buckling of
empty thin-walled steel tanks under wind load according to current American and European design
codes, Thin-Walled Structures, 95, pp. 152–160.
Purdy, D.M., Maher, F.J. & Frederick, D. (1967) Model studies of wind loads on flat-top cylinders.
ASCE Journal of the Structural Division, Vol. 93, pp. 379–398.
Resinger, F. & Greiner, R. (1982) Buckling of Wind Loaded Cylindrical Shells — Application to Unstif-
fened and Ring-Stiffened Tanks. In: Ramm E. (eds) Buckling of Shells. Springer, Berlin, Heidelberg.
Rotter, J.M. (1987) “Membrane Theory of Shells for Bins and Silos”, Transactions of Mechanical Engin-
eering, Institution of Engineers, Australia, Vol. ME12 No. 3 September, pp 135–147.
Rotter, J.M. & Schmidt, H. (2014) Buckling of Steel Shells: European Design Recommendations. Brussels:
European Convention for Constructional Steelwork (ECCS).
Schmidt, H. Binder, B. & Lange, H. (1998) Postbuckling strength design of open thin walled cylindrical
tanks under wind load, Thin Walled Structures, 31, pp. 203–220.
Shokrzadeh, A.R., and Sohrabi, M.R. (2016) Buckling of Ground Based Steel Tanks Subjected to Wind
and Vacuum Pressures Considering Uniform Internal and External Corrosion, Thin-Walled Structures,
108, pp. 333–350.
Uematsu, Y., Yamaguchi T. & Yasunaga J. (2018) Effects of wind girders on the buckling of open-topped
storage tanks under quasi-static wind loading, Thin-Walled Structures, Vol. 124, pp. 1–12.
Ventsel, E. & Krauthammer, T. (2001) Thin plates and shells: theory, analysis and applications, Marcel
Dekker, NY.
Vlasov, V.Z. (1961) Thin-walled elastic beams. National Science Foundation, Washington, D.C.

1325
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Slim-floor beam bending moment resistance considering partial


shear connection

Q. Zhang & M. Schäfer


Research Unit in Engineering Science, FSTC, University of Luxembourg, Luxembourg

ABSTRACT: Having the advantage of flat lower surface, high stiffness and integrated fire
resistance, slim-floor composite beams are widely used and favoured in many design solutions.
The current bending design methods are mainly derived from plastic design methods for clas-
sical composite beam with consideration of the special features for slim-floor beams such as
the transverse bending of bottom flange when used as support for slabs. Alternatively, more
advanced strain-limited design method or FE-method can be used. In the case of full shear
connection, with deep position of neutral axis and great compression zone height, there is
a risk that plastic design method may overestimate the bending resistance of the cross section
compared to the strain-limited design method. In the case of the partial shear connection,
shear design diagram for slim-floor beams obtained by means of the strain-limited design can
also differ significantly from the one obtained by plastic design method, thus further research
on slim-floor beams is still necessary.

1 INTRODUCTION

Slim-floor beam (shallow floor beam or slim-floor construction) refers to a construction type
where the steel beam is partially or fully embedded into the concrete slab (Figure 2). Usually
prefabricated slab systems or special profile sheetings are directly supported on the bottom
steel flange of the asymmetric section to simplify the construction process. Compared to trad-
itional composite beam systems, in buildings with slim-floor beams structural height is greatly
reduced, allowing the installation of technical services without any beam openings due to their
flat lower surface. Composite behavior between the steel section and the concrete slab can be
attained by shear connectors in forms of headed studs or concrete dowels. Due to the integra-
tion of the steel beam into the concrete slab, only the bottom flange of steel section is directly
exposed to fire. In the case of fire, this bottom flange can be substituted by longitudinal
reinforcement bars, thus achieving high fire resistance without any passive protection. Because
of their high bearing resistance and stiffness, composite slim-floor sections can attain econom-
ical span length, e.g. 8-14m. If dry construction is employed by using prestressed concrete
hollow core slabs, the slabs are normally supporting span length of 6 to 12m while the slim-
beam is supporting the shorter span length of 5 to 7.5m (Schäfer et al. 2018).

2 SLIM FLOOR BEAM TYPES AND DEVELOPMENT

Composite slim-floor system originated from Scandinavia region in the 1970s and continu-
ous developed afterwards, the development are also summarized in Schleich 1997, ECCS
1995, Baehre & Pepin 1995, Lawson 1999, Schäfer & Braun 2019 and other documents. Ini-
tially the precast concrete hollow core slab was used to reduce the slab self-weight and to
allow fast erection. Through the use of asymmetric steel sections with larger bottom flange,
the support of the pre-fabricated concrete elements became possible. One of the first devel-
opment is the THQ profile, consisting of a welded box section. Some of the variations of
this system are the NSQ-profile, the TBB-Hut profile and the SWT-Profile. Since the 1990s,
a new type of slim-floor system called SFB (Slim Floor Beam) has been developed. SFB has

1326
a standard rolled profiled steel section welded with an additional wider steel plate at bottom
plate. Subsequently, the IFB Profile (Integrated Floor Beam) entered the market - it was
made by welding a half of the hot rolled profiled steel section onto a steel plate. Initially,
the composite action was not considered, with profiles being designed merely as pure steel
beams, which led to conservative results. Since then, composite effects have been more
closely researched and applied in order to attain economically optimal design. The special
rolled asymmetric steel beams (ASB) with wider bottom flange have special embossment of
the top flange surface with the purpose of activating the composite behavior between con-
crete slab and steel beam through friction mechanism. Several new systems such as the
FEDU profiles (Tschemmernegg 1996), UPE-profiles (Fries 2001) and DELTA sections
have also been developed since. These systems use different shear connectors: the FEDU
beam uses punched steel ribs as shear connector, the UPE beam uses headed studs and the
DELTA beam originally used the concrete dowels in the circular holes of the web to trans-
fer longitudinal shear forces, while the modern DELTA® Beam uses additional reinforce-
ment. Today also SFB sections are used with vertical or horizontal headed studs in order to
active composite effect (Figure 1). Compared to the open steel sections, the box-sections
(hat-profiles) provide higher torsional stiffness, resulting in the better structural performance
in the cases of asymmetric loading and edge beams. For traditional composite beam, the
headed studs are most commonly placed on the top of steel beam, while for slim-floor
beams this layout usually faces difficulties due to the limited concrete cover. As a result,
many innovative shear connectors have been developed.
Nowadays many new types of slim-floor system solutions are commercially developed by
companies. Some examples in Europe are: the USFB® (Ultra Shallow Floor Beam) from
Westok Limited, the DELTABEAM® slim floor structure from Peikko Group and the
CoSFB (Composite Slim-Floor Beam) from ArcelorMittal. The steel section of the USFB® is
fabricated by welding together the two different halves of the hot-rolled steel sections with
circular web-openings. CoSFB uses CoSFB-Betondübel as shear connector (Braun 2018) to
improve the composite behaviour and also the section resistance. These three types of beam
are shown in Figure 2.

Figure 1. Development of part of European composite slim-floor systems (also see: Schäfer & Braun 2019).

1327
Figure 2. Examples of modern Slim-Floor beam systems (Zhang 2016).

3 PLASTIC BENDING RESISTANCE OF SLIM-FLOOR BEAMS

3.1 Design with full shear connection


Current Eurocode 4 (EN1994-1-1 2004) does not provide specific design rules for slim-floor
beams. The design methods applied for slim-floor beams are mostly derived from the plastic
design methods for traditional composite beams with additional consideration of their special
features. Plastic design is based on the assumption that sections have sufficient rotation cap-
acity, implying that plastic strains can be reached at each fiber of the section. For the calcula-
tion of the plastic moment resistance Mpl;Rd full interaction is assumed between structural
steel reinforcement and concrete (Johnshon 2004) meaning that the section is assumed to
remain plain. Full shear connection is reached when an increase in the number of shear con-
nectors within the critical length does not lead to further increase in the moment resistance.
Additional design rules to Eurocode 4 are given for hot-rolled and box-sections in many
research works, some examples are Bode et al. 1997, Lawson & Brekelmans 1999, Hauf 2010,
Lam et al. 2015, Leskelä et al. 2015, Schäfer et al. 2018, much more other literature related to
different topics are to be found, here it is not possible to list all of them. The general design prin-
ciples are summarized as follow: In the case of the full shear connection (Figure 3) and sagging
bending moment, the compression force consists of the integral of concrete stress bock ð0:85fcd Þ
and the compression blocks in the steel part (fyd ), while concrete in tension is neglected. For the
concrete inside a box-section fcd can be used, due to the better hardening conditions and confine-
ment effects. Due to the interaction of vertical shear and normal stress from bending moment,
according to EN1994-1-1 section 6.7.3.2, the steel strength of the steel web is to be reduced to
ð1  ρÞfyd if Va;Ed 40:5Vpl;a;Rd (Figure 3). In the case of important web openings, secondary
bending moments due to Vierendeel-effects are to be considered for moment resistance and verti-
cal shear design (Schäfer 2015b). Furthermore, the effects from torsional moments can lead to
additional reduction of the moment and shear capacity. When the extended bottom flanges are

Figure 3. Plastic moment design of slim-floor section based on full shear connection (also see:
Schäfer et al. 2018).

1328
Figure 4. Plastic moment design of slim-floor section based on partial shear connection.

used as supports for the concrete slab, significant amount of transverse local bending moment
can be generated. The interaction of the compression stress and tensile stress in longitudinal direc-
tion reduces the design resistance. Thus, the design strength of the bottom flange needs be reduced
to fyd;eff (Schäfer 2015a). For the open sections, some researchers (Bode et al. 1997) have sug-
gested using a reduced steel plate thickness to reflect this impact.

3.2 Design with partial shear connection


Where the shear connector arrangement is controlled by detailing, such as the geometry of
profile sheeting, or the proof in SLS become decisive, often partial shear connection is real-
ized. In partial shear connection, the bending moment resistance MRd is limited by the longitu-
dinal shear resistance Nc . The Plastic Neutral Axis (PNA) in (reinforced) concrete part and
the PNA in the steel beam are separated. Their locations are controlled by the longitudinal
shear force transferred by shear connectors and the stress distribution.
The bending moment resistance of partial shear connection can be calculated by solving the
equilibrium equation according to plastic theory. A more practical way is to use the partial
shear diagram provided in Eurocode 4 (Figure 4). This diagram is developed using the plastic
design method, for which a nonlinear ABC curve or a simplified linear line AC can be adopted
(A: no shear connection, B: zone of partial shear connection, C: full shear connection). Here
the degree of shear connection is defined as η ¼ ðNc  Ns Þ=Nc; f where Nc is the total concrete
normal force, and Nc;f is the value with full shear connection. Design moment resistance in
the case of partial shear connection can be then easily obtained from the previously calculated
plastic bending moment of composite section Mpl;Rd and the pure steel section Mpl;a;Rd . If
solid slab with big amount of reinforcement is used the contribution of slab should also be
considered at least in SLS (Hauf, 2010). Considering the limitation of shear connector deform-
ation capacity, a minimum degree of shear connection is required, depending on the relation
of Mpl;Rd to Mpl;a;Rd , ductility of shear connectors, loading and overall moment allowance.

4 STRAIN-LIMITED DESIGN RESISTANCE OF SLIM-FLOOR BEAMS

According to plastic design method, concrete stress is represented by a rectangle stress block with
compression stress of 0:85fcd . For traditional composite beams with small compression zone
height, the plastic design is on the safe side (Hanswille 1996, Schäfer et al. 2019). However, for
slim-floor beams the position of neutral axis is deeper due to the non-symmetric steel profile with
enlarged bottom flange. The raised compression zone height usually brings reduction of the rota-
tion capacity and the bearing behavior. Premature concrete crushing failure may happen before
yielding of the steel section. In this case the plastic method can overestimate the design resistance.
This effect was also mentioned in many former research works, such as in Ansourian 1984, plastic
design required conditions related to rotation capacity and others were analysed and
a compression zone height smaller than 16% of cross-section height was suggested. For a bigger

1329
compression zone height until xpl =h ¼ 0:3, in Hanswille & Sedlacek 1996 the reduction of plastic
bending resistance for steel Grade S420 and S460 was analysis. Consequently, in the design pro-
cedure provided in the EN1994-1-1 section 6.2.1.2 (2) full plastic bending moment is limited by
a β factor in the cases when steel grade S420 or S460 is employed and when the relation of xpl =h
is over 0.15. However, this problem does not occur only when the steel grade is greater than S420
- it is more a question of the cross-section shape and the position of the plastic neutral axis. There-
fore a reduction function β 0 ¼ Msl;Rd =Mpl;Rd between plastic bending moment resistance Mpl;Rd
and strain-limited bending moment resistance Msl;Rd was defined by Schäfer et al. 2019, expand-
ing the design procedure to other steel grades as well.
Schäfer et al. 2019 pointed out the limitations of the plastic design for composite sections,
by comparing this approach to the strain limited design procedure. In strain limited method,
the slim-floor section follows Bernoulli hypothesis, for which a linear strain distribution in full
interaction is assumed. Considering the shear-lag effects in concrete slab, a simplification with
effective width can be applied, similar to the procedure used in the design of classical compos-
ite beams. The resistance is reached when the ultimate strain in any fibre reaches its strain
limitation. Similar to the the bending design of reinforced concrete beams according to Euro-
code 2 (EN1992-1-1 2004), a parabolic-rectangle stress-strain model can be used. The strain
limits of structural steel can be calculated according to stress-strain relationship of EN1993-
1-5, Annex C (EN1993-1-5 2006) considering strain-hardening. However other suitable mater-
ial models are not excluded.

4.1 Design with full interaction


Figure 5 illustrates the strain-limited design principal on an example of a slim-floor beam
under sagging bending moment. For determination of strain-limited bending resistance
(Msl;Rd ) full interaction is assumed, implying that there is no slip at steel-concrete interface.
With full interaction, the strain limits can be either the compression strain-limit of concrete
(εcu;2 ) at top fiber, the steel tensile strain-limit (εau ) at the bottom fiber (concrete in tension is
neglected) or the strain-limit of the reinforcement (εsu ). Due to the fact that steel is a ductile
material, in most cases the concrete strain will be the governing factor.
To calculate the location of neutral axis and strain stress distribution, the following two
requirements must be met:
● at least one critical fiber of the section should reach its strain limit, and all other fibers
stay below their limits. For sagging bending moment, it is usually the fiber located at con-
crete upper surface (εc ¼ εc;u ).
pure bending, the sum of normal forces of each fiber inside a cross-section
● in the case of P
shall be zero ( Ni0 ¼ 0);
By setting a strain-limit point and rotate the strain curve based on the point, equilibrium can
be found, further the cross-section resistances can be acquired. For many finite difference
method based numerical procedure, in order to calculate beam deflection, moment

Figure 5. Strain-limited design of slim floor beam based on full shear connection.

1330
redistribution or longitudinal shear force, the Moment-curvature (M-k) curve is important.
Such procedures can be found in Fabbrocino G. et al. 2000, Fries 2001 and many other publi-
cations. To acquire the M-k curve, the strain-limits can be step-wisely changed from zero to
the limiting values, and calculate out each moment and curvature value.

4.2 Design with partial interaction


In the case of sagging moment and cross-section class 1 and 2, with ductile shear connectors,
the theoretical partial interaction can be applied. In reality there will always be slip at the inter-
face of concrete and steel element as typical shear connector needs to experience some deform-
ation in order to attain its design resistance. In the case of partial interaction, it is usually
assumed that the concrete and steel part have the same curvature at each cross-section, which
can be represented by pair of parallel strain curves (similar to the widely accepted Newmark-
model (Newmark 1951), however the assumption of elastic shear connector is not used here, as
this paper mainly discuss on cross-section level). Within the small deformation domain, these
assumptions are reasonable. Due to slip there will be two neutral axes for concrete and steel
section each, (while for full interaction there is only one neutral axis). The neutral axes of the
two parts separate with distance of slip-strain εslip (Figure 6). The slip in the composite joint
depends on the flexibility of shear connectors and degree of shear connection and the longitu-
dinal shear force. This model was widely applied for tradtional composite beams for example it
was used in the work of Fabbrocino et al. 2000 to develop the numerical calculation method of
continuous composite beams consider partial interaction. General theoretical basics are also
explained in Oehlers et al. 1995 for the definition of degree of interaction.
When the slip occurs, critical fibers, with regard to the imposed strain-limits, are no longer only
the out-most fibers of the section. Due to the slip, the strain in the internal fibers may increase,
which can cause the steel section to reach its yield strains - reinforcement or steel upper flange can
attain their failure strain first. This happens especially in the cases when the degree of shear con-
nection is very low. In the case of partial shear interaction, there are three unknown factors in the
strain diagram: the strain limits at critical fibres εi;u , the slip-strain value εslip and the curvature χ
(Figure 6). Therefore, at least three equilibrium equations need to be established:
● at least one critical fiber of the section should reach its strain limit (εi ¼ εi;u ), and all other
fibers stay below their limits;
pure bending, the sum of normal forces of each fiber inside a cross-section
● in the case of P
0
shall be zero ( Ni ¼ 0);
● the integral of longitudinal shear force within the critical length ensured by the provided
shear connectors
R should be 0equal to 0 the normal forces inside either the steel or concrete
part by value ( VL dx ¼ ηNc;f ¼ Na ).
In order to calculate the bending resistance for different degrees of shear connections η by
strain-limited design method, a numerical program is necessary, the steps outlined in the
flow-chart in Figure 6 can be followed. In general, the curvature can be first calculated

Figure 6. Strain-limited design of slim floor beam based on partial shear connection (also see: Schäfer
& Zhang 2019).

1331
Figure 7. Comparison of partial shear diagrams.

based on internal force ηNcf0 given by the shear degree η and afterwards the stain distribu-
tion over the whole cross-section and resistances can be acquired.

5 PARTIAL SHEAR DIAGRAM OF SLIM-FLOOR BEAMS

The partial shear diagram, based on the plastic design method presented in Eurocode 4, is
explained in section 3 of this paper. Similarly, we can generate the partial shear diagram by
strain-limited method or FE-method. The strain-limited bending resistance and maximum lon-
gitudinal shear are different to the values obtained by plastic design. Thus, longitudinal shear
resistance obtained by plastic design method should be replaced with the strain limited resist-
ance in order to calculate the degree of shear connection η.
Figure 7 shows the partial shear diagrams according to the plastic method (ABC), the sim-
plified method diagram (AC), as well as by strain-limited method. For comparison purposes,
the partial shear diagram of strain-limited design is presented in relative values, by dividing
the full shear connection resistance values with the values obtained from plastic design.
For classical composite beams with small compression zone height (xpl =h50:15), strain-limited
design without steel strain-hardening gives similar results as plastic design, as shown in
Figure 7a. Parameter studies on more than 5,000 cross-sections confirmed this conclusion. Com-
parison of partial shear diagrams for a slim-floor cross-section is shown in Figure 7b. It shows
considerable difference in the results between these two methods of design: At full interaction, the
maximum bending moment resistance by strain limited design (b) is around 7% smaller than the
plastic bending resistance (a). The longitudinal shear force is also around 17% smaller than plastic
design suggests. From the stress curves for strain-limited design and plastic design at full inter-
action we can notice important differences. In the case of the strain limited design, the reinforce-
ment is not likely to yield nor the steel bottom flange and a great part of the steel web. This is due
to the fact that concrete compression strain is the governing factor here, with limited strain in the
steel due to the deep neutral axis. Thus the maximum total longitudinal shear force is much smal-
ler than full plastic resistance - the same applies for the bending resistance. The difference is more
significant when compression zone height is greater and high strength steel is used. Similar phe-
nomenal happens to over-reinforced concrete beams or in the case of high strength reinforcement.
Therefore, from all of the above, it can deducted that in the cases when slim-floor beams are
being used, due to the relatively deep neutral axis position, plastic design method and partial
shear diagram provided in the current Eurocode 4 is not always suitable for safe use.

6 CONCLUSIONS

In the recent years the use of the composite slim-floor beams is becoming more popular as
they are expanding the areas of their applications. However, there are still no universal

1332
standards for their design. Many of the current design methods are based on plastic theory,
which may overestimate the bending resistance in the cases of deep compression zones. In
these cases strain-limited design is a preferable method of analysis as it offers more accurate
results. With the aim of reducing the difficulty of design procedure and in order to allow for
more economical design solutions, research on new theoretical design approaches and partial
shear diagrams for slim-floor beams is currently in the process. This paper presents some of its
early results. Further investigation in the rotation capacity, moment redistribution and other
related topics of slim-floor beam system are still required to reach a safe and economic design.

REFERENCES

Ansourian, P. 1984. Beitrag zur plastischen Bemessung von Verbundträgern, Bauingenieur 59: 267–272.
Baehre, R & Pepin, R. 1995. Flachdecken mit Stahlträgern in Skelettbauten, Bauingenieur
(Springer-Verlag): 70.
Bode, H. et al. 1997. Untersuchungen des Tragverhaltens bei Flachdeckensystemen mit verschiedener
Ausbildung der Platte und verschiedener Lage der Stahltrager, Forschung fur die Praxis: 261.
Braun, M. 2018. Investigation of the Load-Bearing Behaviour of CoSFB-Dowels. Dissertation, Univer-
sity of Luxembourg
ECCS. 1995. Multi-story building in steel-Design Guide for Slim Floors with Built-in Beams, Report
No.83 Paris.
EN 1992- 1-1. 2004. Eurocode 2: Design of concrete structures - Part 1-1: General rules and rules for
buildings.
EN 1993- 1-5. 2006. Eurocode 3: Design of steel structures - Part 1-5: Plated structural elements.
EN 1994- 1-1. 2004. Eurocode 4: Design of composite steel and concrete structures Part 1-1: General
rules and rules for buildings.
Fabbrocino G. et al. 2000. Analysis of continuous composite beams including partial interaction and
boud. Journal of structural engineering 126(11): 1288–1294.
Fries, J. 2001. Tragverhalten von Flachdecken mit Hutprofilen. Dissertation, University of Stuttgart.
Johnson, R.P. 2004. Composite Structures of Steel and Concrete - Beams, Slabs, Columns, and Frames
for Buildings, Blackwell Publishing, third edition: 24–25.
Hanswille G. & Sedlacek, G. 1996. The Use of Steel Grades S460 and S420 in Composite Structrues,
ECCS-EUROFER improvements by TC 11 to EUROCODE 4 report.
Hauf, G. 2010. Trag- und Verformungsverhalten von Slim-Floor Trgern unter Biegebeanspruchung. Dis-
sertation. Unveristy of Stuttgart.
Lam, D. et al. 2015. Slim-floor construction - design for ultimate limit state. Steel Construction 8(2): 79–84.
Lawson, R.M. & Brekelmans, J.W.P.M. 1999. Design recommendations for shallow floor construction
to Eurocodes 3 and 4. European Commission technical steel research report: 11–12.
Newmark, N M. 1951. Test and analysis of composite beams with incomplete interaction. Proceedings of
society for experimental stress analysis 9(1): 75–92.
Oehlers, D.J, Bradford, M.A. 1995. Composite steel and concrete structural members: fundamental
behaviour. Oxford: Pergamon Press: 23–26.
Schäfer, M. 2015a. Zur Biegbemessung von Flachdecken in Verbundbauweise, Stahlbau - Ernst & Sohn
84(4): 231–238.
Schäfer, M. 2015b. Schubtragfäigkeit und M-V-Interaktion von Flachdecken mit integrierten hohlkasten-
fömigen Stahlprofilen, Stahlbau - Ernst & Sohn 84(5): 314–323.
Schäfer, M. et al. 2018. Flachdecken in Verbundbauweise-Bemessung und Konstruktion von Slim-Floor-
Trägern, StahlBau Kalender - Ernst & Sohn: 631–741.
Schäfer, M. et al. 2019. Plastic design for composite beams - are there any limits? 9th International Con-
ference on Steel and Aluminium Structures, Bradford,UK.
Schäfer, M. & Braun, M. 2019. Entwicklung der Slim-Floor Bauweise in Europa, Special edition Slim-
floor beams Stahlbau - Ernst & Sohn (88) 7.
Schäfer, M & Zhang, Q. 2019. Zur Momententragfhigkeit von Flachdecken in Verbundbauweise Stahl-
bau - Ernst & Sohn (88) 7.
Schleich, J.B 1997. Slim floor construction: why?, Outstanding Composite structures for Buildings, Com-
posite Construction Conventional and Innovative, Innsbruck IABSE reports, Zurich: 53–64.
Tschemmernegg, F & Huber, G. 1996. Flachdecken mit Stanzdubeln Bauingenieur, Springer-Verlag 71.
Zhang, Q. 2016. CET-Report I, Internal research report for Research on Longitudinal shear in Compos-
ite Structures (unpublished). University of Luxembourg.

1333
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Stainless steel SHS and RHS beam-columns

B. Židlický & M. Jandera


Czech Technical University in Prague, Prague, Czech Republic

ABSTRACT: Presented research deals with stainless steel slender square hollow and rectangu-
lar hollow section (SHS and RHS) members loaded by combination of uniaxial bending moment
and compressive axial force. Four tests of eccentrically loaded stainless steel pin-ended columns
were conducted. Data obtained from the tests were used for the numerical model validation.
A comprehensive numerical parametric study was made using finite element software Abaqus
employing geometrically and materially non-linear analysis with imperfections (GMNIA). All
three main stainless steel groups were considered in study, namely austenitic, ferritic and duplex.
The investigated variables were mainly cross-section slenderness, member slenderness and ratio
between compressive force and bending moment. Only uniform bending moment along the
member length was considered. Based on the numerical parametric study results a new design
approach for stainless steel SHS and RHS beam-columns has been derived. It is described and
evaluated in this paper together with recently published design procedure developed by Zhao.

1 INTRODUCTION

Stainless steel is a highly alloyed steel exhibiting high corrosion resistance together with great
mechanical properties, easy maintenance and aesthetic appearance. Many procedures given by
the stainless steel codes were adopted from the procedures for carbon steel. As the Eurocode
was developed almost 15 years ago, some of the rules for stainless steel were based on much
lower number of tests and numerical models than available today.
This paper is focused on stainless steel square and rectangular hollow section (SHS and
RHS) beam-columns. Only uniform uniaxial bending moment along the member length is
considered. There is a codified design approach in the Eurocode EN 1993-1-4 (2006) for stain-
less steel, see Equations 1 and 2.

N Ed
N b;Rd;y þky MMb;Rd;y
Ed
 1:0 ð1Þ

N Ed N Ed
ky ¼ 1:0 þ 2 ðλy  0:5Þ but 1:2  ky  1:2 þ 2 ð2Þ
N b;Rd;y N b;Rd;y

where NEd ¼ the axial compressive force; MEd ¼ the maximal I. order bending moment; Nb,Rd,y ¼
the flexural buckling resistance, Mb,Rd,y ¼ the bending moment resistance; ky ¼ the interaction
factor considering compression and bending interaction; and λy ¼ the member slenderness.
However, this procedure exhibits some shortcomings, namely: it does not consider moment
distribution along the member length which leads to over-conservative results in non-uniform
bending moment cases; the lower bound of the interaction factor ky = 1.2 makes the procedure
over-conservative in cases where bending moment is dominant.
Many investigations have been carried out in order to make the beam-column design
both accurate and safe. Most of them were focused on the improvement of the inter-
action factor calculation. Evaluation made by Jandera et al. (2017) shown that all of
them have some inaccuracies and a new method should be developed. A new approach
was recently developed by Zhao et al. (2016). It considers the same interaction formula

1334
Table 1. Interaction factor coefficients Di.
Stainless steel group D1 D2 D3

Austenitic 2.0 0.30 1.3


Ferritic 1.3 0.45 1.6
Duplex 1.5 0.40 1.4

as EN 1993-1-4 (2006), see Equation 1, with flexural buckling resistance Nb,Rd,y calcu-
lated considering new flexural buckling curves derived by Afshan et al. (2017), bending
resistance Mb,Rd,y,CSM calculated according to CSM method developed by Gardner and
published e.g. in Afshan & Gardner (2013) and interaction factor kCSM calculated
according to a new formula, see Equation 3.

kCSM ¼ 1 þ D1 ðλy  D2 Þ nb  1 þ D1 ðD3  D2 Þnb ð3Þ

where Di = the interaction factor coefficients given by Table 1; and nb = the compressive force
NEd to flexural buckling resistance Nb,Rd,y ratio.
As could be seen, there are several coefficients regarding the stainless steel group. The pro-
cedure was presented as a safe and accurate for the design of stainless steel SHS and RHS
members loaded by compression and bending combination. The evaluation of this approach
on numerical results is given in this paper below.

2 EXPERIMENTAL STUDY

Two sets of experiments were conducted. First consists of four SHS specimens carried out for
a numerical model validation. The recently conducted second set of 12 specimens extended the
previous tests for more slender Class 4 SHS members as well as both stocky and slender cross-
section RHS members. Data obtained from the second set of the tests will be presented at the
conference.

2.1 Tensile tests


In order to obtain material properties, tensile coupon tests were conducted. It was necessary to
made tensile tests for both flat and corner part of the cross-section as the corner material prop-
erties are significantly higher due to cold-rolling. Figure 1 presents location of the coupons.

2.2 Experiments
Together 20 experiments were carried out, namely 12 SHS members and 8 RHS members.
Specimens were loaded by an eccentrically applied compressive force. The eccentricity was the
same for both ends of the member, causing uniaxial uniform bending moment in the tested
member, see Figures 3 and 4. Pin-ended supports were made from two steel plates on both
ends of the specimen. One plate was equipped by the wedge and the second by the V-shaped
notch which ensured deflection in appropriate plane and the eccentricity was easily adjusted,
see Figure 2. Both global and local imperfections of all specimens were measured manually
before the test. Specimen information are summarized in Table 2.

3 NUMERICAL STUDY

Numerical model was created in finite element software Abaqus in order to obtain behaviour
of the tested members. 3D shell element model using GMNIA analysis (Geometrically and

1335
Figure 2. Pin-ended support.

Figure 3. Test set-up and photo of a tested member.

Table 2. Tested member geometry.


Cross- Measured Wall Member Member Global Local
section dimensions thickness length slenderness Eccentricity Imperfection Imperfection
mm mm mm mm - mm mm mm

80  3 79.74  79.74 2.8 2610 1.28 20 0.833 0.01


80  3 79.74  79.74 2.8 2620 1.29 40 0.767 0.0125
80  5 Not measured 5.07 2625 1.46 20 1.3 0.0125
80  5 Not measured 5.07 2575 1.43 40 1.233 0.01

Materially Non-linear Analysis with Imperfections) was used. Both global and local imperfec-
tion amplitudes were introduced through the appropriate (global and local buckling) eigen-
modes. Material properties for both flat and corner part of the cross-section were obtained
from the tensile tests. The corner region with increased strength was considered by the area of
the corner itself with no extension to the flat part.

3.1 Numerical model validation


The numerical model prediction was compared with the experimental data, namely the rela-
tionship between compressive loading force and deflection at the mid-span of the member
length. The comparison is given by Figure 4 and Table 3.

1336
Figure 4. Comparison between the numerical model (ABQ) and experiments (EXP).

Table 3. Comparison between the numerical model (ABQ)


and experimental (EXP) ultimate forces.
Cross-section Eccentricity FEXP/FABQ
mm mm -

80  3 20 0.911
80  3 40 0.988
80  5 20 1.005
80  5 40 0.967
Average 0.968
Standard deviation 0.035

As could be seen, the results obtained from the numerical model calculations are in good
agreement with the experimental data. There are some differences in the ultimate loading
forces, however, it is less than 10% in all cases. Therefore, the numerical model was considered
as accurate enough suitable for subsequent numerical parametric study.

3.2 Numerical parametric study


The comprehensive numerical parametric study was made using the numerical model
described above. In the numerical parametric study, the global imperfection amplitude was
considered as L/1000 (where L is the member length) and local imperfection amplitude was
calculated considering Dawson & Walker (1972) with Gardner and Nethercot´s (2004) modifi-
cation for stainless steel. The main three stainless steel groups common in structures were con-
sidered, namely austenitic, ferritic and duplex. Young modulus E0 = 200 000 MPa was
considered for all materials. One grade of each stainless steel group was considered, however,
with two values of strain hardening factor n (representing variation among the grades and
degree of cold-forming). Regarding the appropriate choice of the stainless steel grades, mater-
ials with low yield strength fy and low ultimate strength fu (ferritic grade), high fy and high fu
(duplex grade) and the greatest ratio between fy and fu (austenitic grade) were considered.
Material properties are summarized in Table 4. Both SHS and RHS cross-sections were inves-
tigated. Dimensions were considered as 80 mm for SHS and 100  40 mm for RHS cross-
sections (representing the highest h/b ratio common for sections). Thickness of the wall was

1337
set as a variable parameter and calculated regarding the considered material in order to cover
cross-section Classes 1 and 4 of SHS and cross-section Classes 1, 3 and 4 of RHS cross-
sections. Moment to axial force loading ratio was established with the idea of reaching some
levels of nb = NEd =Nb;Rd;y for the whole slenderness range. Section bending resistance for the
estimation of the loading was calculated according to Zhao et al. (2016) proposal using CSM.
The investigated nb ratios and member slendernesses are shown in Table 5.

3.3 Comparison of Zhao’s procedure


In this chapter, the design approach developed by Zhao et al. (2016) and described above is com-
pared with the results of the numerical parametric study. The comparison was done using flexural
buckling resistance according to Afshan et al. (2017) and bending resistance proposed by Zhao
et al. (2017) as for the planning of the study. The evaluation is made in terms of the product of
the whole interaction formula, where the results greater than unity indicates safe predictions
whereas lower than unity unsafe ones. Figures 5 and 6 provide comparison of the approach as
dependent on the member slenderness and nb ratio, respectively.
As could be seen in Figures 5 and 6, the approach developed by Zhao (2015) provides gen-
erally good prediction with some scatter and conservativeness for Class 4 result and slight un-
conservativeness in Class 1 cross-section cases. Very limited number of results for Classes 3
and 4, namely with increasing influence of the bending moment became little un-conservative
for austenitic and ferritic stainless steel grades. On the other hand, the procedure gives safe
results for duplex stainless steel.
It should be noted that Zhao et al. (2016) procedure was developed based on resistances calcu-
lated according to Afshan et al. (2017) and Afshan & Gardner (2013). Therefore a scatter in the
comparison, could be in some cases result of the scatter in section or column resistance prediction.

3.4 New proposal


A new procedure for the stainless steel SHS/RHS beam-column design described in this chap-
ter was developed based on the numerical parametric study. Both flexural buckling resistance

Table 4. Material properties considered in the parametric study.


E0 fy fu n

Material MPa MPa MPa -

Austenitic 200,000 210 380 4.5


Austenitic 200,000 210 380 14
Duplex 200,000 480 660 4.5
Duplex 200,000 480 660 14
Ferritic 200,000 220 520 4.5
Ferritic 200,000 220 520 14

Table 5. Member slenderness and nb = NEd/Nb,Rd values investigated in the parametric study.
SHS members RHS members

Member slenderness nb = NEd/Nb,Rd Member slenderness nb = NEd/Nb,Rd

0.2 0.05 0.5 0.05


0.3 0.3 0.8 0.1
1.0 0.5 1.0 0.2
1.5 0.7 1.5 0.5
2.0 0.8 2.0 0.8
3.0

1338
Figure 5. Comparison for the Zhao et al. (2016) procedure in terms of nb = NEd/Nb,Rd ratio.

Figure 6. Comparison for the Zhao et al. (2016) procedure in terms of the member slenderness.

and bending resistance were also obtained from the numerical model. The same interaction
formula as given by EN 1993-1-4 (2006) (Equation 1) was considered, but with different for-
mulae for the interaction factor itself. It is given by Equations 4 and 5.

ky;New ¼ 1:0 þ 1:5λy nβb for λy  1:0 ð4Þ

ky;New ¼ 1:0 þ 1:5λy nβb pffiffiffiffiffiffiffiffiffiffiffi


0:8 ffi
for λy > 1:0 ð5Þ
λy 0:36

 2
β ¼ MMb;Rd
el
ð6Þ

where Mel = the elastic bending moment resistance.


The comparison of the New proposal is made through the whole interaction formula calcu-
lation again and given by Figures 7 and 8. Furthermore, comprehensive statistical evaluation

1339
Figure 7. Comparison for the New proposal in terms of nb = NEd/Nb,Rd ratio.

Figure 8. Comparison for the New proposal in terms of the member slenderness.

according to Afshan et al. (2015), using resistances calculated according to EN 1993-1-4


(2006) considering both nominal and mean values of input parameters, including also
a proposal for the safety factor is provided by Table 6.
As could be seen in Figures 7 and 8, the New proposal predictions are safe and consistent
for both SHS and RHS stainless steel beam-columns. It is a general procedure for all stainless
steel groups with no dependency on member slenderness, cross-section slenderness (Class of
the cross-section) and loading case (compressive force to flexural buckling resistance nb ratio).
The accuracy is confirmed by statistical evaluation given by Table 5. The results exhibit low
scatter with slightly conservative results in the average. However, the average value conserva-
tiveness is caused by the complex statistical evaluation according to Afshan et al. (2015) condi-
tion where the partial safety factor γM should be lower than partial safety factor γM1 value
given by stainless steel code EN 1993-1-4 (2006) which is recommended as 1.1. Based both on
the graphical comparison and statistical evaluation, the New proposal is very general and
accurate procedure for both SHS and RHS stainless steel beam-column design.

1340
Table 6. Statistical evaluation of
the New proposal.
Average value 1.096
Standard deviation 0.069
γM 1.086

4 CONCLUSIONS

Presented paper is focused on behaviour of SHS and RHS stainless steel members loaded by
compression and bending moment combination. There are several approaches for stainless
steel beam-column design with a proposal of Zhao et al. (2016) describing the behaviour quite
accurately. However, a new proposal was developed and presented in this paper.
A 3D shell element model using GMNIA was created in finite element software Abaqus
simulating the stainless steel beam-column behaviour. Validation of the numerical model was
successfully made based on the data obtained from experiments. A comprehensive numerical
parametric study was made subsequently.
The recent procedure developed by Zhao et al. (2016) was evaluated first in detail. It was
shown that the approach exhibits slightly scattered results with slight un-conservativeness for
Class 1 cross-section cases.
A new proposal was presented and evaluated. The procedure provides accurate and consist-
ent predictions with similar safety and scatter for various member and cross-section slender-
ness, material properties and compressive force to flexural buckling ratio. Furthermore,
statistical evaluation was presented and confirmed the accuracy of the new proposal.

ACKNOWLEDGEMENT

The support of the Czech Science Foundation grant 17-247695 “Nonlinear stability and
strength of slender structures with nonlinear material properties” is gratefully acknowledged.

REFERENCES

Afshan, S. & Gardner, L. 2013. The continuous strength method for structural stainless steel design.
Thin-Walled Structures. 68: 42–49.
Afshan, S., Francis, O., Baddoo, N. & Gardner, L. 2015. Reliability analysis of structural stainless steel
design provision. Journal of Constructional Steel research. 114: 293–304.
Afshan, S., Zhao, O. & Gardner, L. 2017. Buckling curves for stainless steel tubular columns. Eurosteel
2017. Copenhagen, Denmark.
Dawson, R.G. & Walker, A.C. 1972. Post-buckling of geometrically imperfect plates. Journal of the
Structural Division ASCE. 98: 75–94.
EN 1993- 1-4. 2006. Eurocode 3: Design of steel structures – Part 1-4: General rules – Supplementary
rules for stainless steels. European Committee for Standardization (CEN). Brussels.
Gardner, L. & Nethercot, D. A. 2004. Numerical Modeling of Stainless Steel Structural
Components-A Consistent Approach. Journal of Structural Engineering, ASCE. 130: 1586–1601.
Jandera, M., Syamsuddin, D. & Židlický, B. 2017. Stainless steel beam-column behaviour. Open Civil
Engineering Journal. 11: 358–368.
Zhao, O., Afshan. S. & Gardner, L. 2017. Structural response and continuous strength method design of
slender stainless steel cross-sections. Engineering Structures. 140: 14–25.
Zhao, O., Gardner, L., & Young, B. 2016. Behaviour and design of stainless steel SHS and RHS
beam-columns. Thin-Walled Structures. 106: 330–345.

1341
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Calibration of parameters of combined hardening model using


tensile tests

C.I. Zub, A. Stratan & D. Dubina


Department of Steel Structures and Structural Mechanics, Politehnica University of Timisoara, Timisoara,
Romania

ABSTRACT: When performing finite element simulations on structural elements made of


mild carbon steel proper modelling of the cyclic response of material allow for numerical
simulations with a high level of reliability. Good predictions can be obtained using the com-
bined isotropic-kinematic hardening model to simulate the plastic behaviour of steel. As
recommended by Abaqus, the finite element software used within these simulations, the cali-
bration of the material parameters of the combined model requires cyclic test data. In com-
parison to tensile tests, the cyclic tests are more difficult to be performed. Therefore,
a simplified calibration procedure of the material parameters using only tensile tests is pre-
sented in this paper. The material parameters obtained in this way allow for predictions with
an acceptable level of correlation, for both monotonic and cyclic loading. These predictions
are validated against experimental tests performed on S355 mild carbon steels.

1 INTRODUCTION

When performing finite element simulations on structural elements made of mild carbon steel
proper modelling of the response of the material is necessary to obtain reliable simulations.
While for the simulations involving monotonic loading simple material models can be used, in
the case of the simulations under cyclic loading (e.g. low-cycle fatigue simulations) more com-
plex material models are necessary. This will allow to properly capture the behaviour of struc-
tural components (e.g. buckling restrained braces) undergoing large plastic deformations (Zub
et al. 2018).
It was shown by Lemaitre & Chaboche (1990) that proper modelling of metal plasticity
under cyclic loading can be achieved using the combined isotropic-kinematic hardening mater-
ial model. The combined model, consisting of a kinematic (main) and isotropic (optional)
component, was implemented in the Abaqus finite element package (Dassault, 2014). Also,
calibration procedures were proposed to obtain the input parameters for both hardening com-
ponents. As described in Dassault (2014), the implemented combined model has several limita-
tions which relate to its capability of simulating the metal plasticity under arbitrary loading
history with a high level of accuracy. However, Abaqus allows for advanced users to define
their own material models via UMAT/VUMAT subroutine.
The combined model was used by many researchers in their numerical studies that involved
metal components. Among others, Nip et al. (2010) used the built-in combined model from
Abaqus and the calibration procedure described in Dassault (2014) to perform numerical
simulations on structural components made of carbon and stainless steel under cyclic loading.
Zub et al. (2018) used the combined model to simulate the cyclic response of buckling
restrained braces (BRB) under quasi-static loading regime (low-cycle fatigue simulations). The
combined model showed good correlation with the experimental results, although the BRBs
developed large plastic deformations. On the other side, there are some researchers who devel-
oped their own material models (UMAT for Abaqus Standard and VUMAT for Abaqus
Explicit) in order to obtain more accurate numerical results on metal plasticity. Among

1342
others, Ucak and Tsopelas (2011) developed a material model (UMAT) to simulate the cyclic
behaviour of structural mild carbon steels (with yield plateau). Shi et al. (2012) developed
a UMAT for high-strength steel. Also, Hu et al. (2016) developed a UMAT, which will be
further discussed. In all cases additional software (Fortran) and programming skills are
required in addition to a good understanding of the theory of plasticity.
According to the current implementation of the combined model in Abaqus, for the calibration
of the parameters of the model experimental uniaxial cyclic tests on coupon specimens are
required. In comparison to the more common monotonic tensile tests, the cyclic tests are more
difficult to perform, as buckling of the specimen should be prevented. As an alternative to per-
forming cyclic tests, Hu et al. (2016) developed a user material subroutine (UMAT) for the impli-
cit integration procedure in Abaqus/Standard that requires only tensile test data for the
calibration.
The objective of this paper is to provide a calibration procedure for the parameters of the
combined hardening model, available in both Abaqus/Standard and Abaqus/Explicit pack-
ages, that uses only tensile test data. The approach is based on the procedure developed by
Hu et al. (2016), adjusted to the requirements of the built-in models in Abaqus. The material
parameters calibrated with the proposed procedure can be used for both monotonic and cyclic
loading regimes.

2 PROPOSED CALIBRATION PROCEDURE

Considering a monotonic tensile test as schematically presented in Figure 1 a), the proposed
calibration procedure uses as input the following mechanical properties, expressed as engineer-
ing values: yield strength fy, ultimate strength fu, rupture strength fr, yield strain ey, end of
plateau strain esh, ultimate strain eu, rupture strain er. For simplification reasons and based on
experimental observations, several assumptions were made: the rupture strength is fr = 0.8fu,
the yield strain is ey = fy/Es (Es = 210000 N/mm2 is the elastic modulus of steel), the strain
corresponding to ultimate strength is eu = 0.55er, the Poisson’s ratio is ν = 0.3.
As Abaqus requires the input parameters of the material model to be expressed as true
(Cauchy) stress σ and logarithmic strain ε, therefore, the engineering stress ( f ) and strain (e)
values must be transformed accordingly using the following formulas (Dassault 2014):

σ ¼ f ð1 þ eÞ ð1Þ

ε ¼ lnð1 þ eÞ ð2Þ

In the procedure proposed by Hu et al. (2016) several assumptions are made, and additional
simulations and calibrations of specific parameters are required. In this paper, all the

Figure 1. Schematic representation of the response of steel under monotonic tensile loading a) engineer-
ing stress-strain curve and b) true stress-strain curve.

1343
parameters describing the plastic behaviour of steel are determined based on the input data,
therefore no additional simulations or empirical assumptions are required.
The first assumption is the weighted average factor w (Figure 1 b), which in the procedure
of Hu et al. (2016) is determined after several iterations (simulations). Within the proposed
procedure, the w factor is obtained directly as a function of rupture strain er. Although equa-
tion (3) was obtained by train and error, it shows to be consistent for the study cases presented
in section 3 of this paper.

0:1
w¼1 ð3Þ
er
σj0 ¼ σy ð4Þ

σu  fy ð1 þ esh Þ γ1
C1 ¼  ð5Þ
2 3
γ1 ¼ 10  w  γ2 ð6Þ

σu  fy ð1 þ esh Þ γ2
C2 ¼   1:8 ð7Þ
2 3
σu  fy ð1 þ esh Þ 1
γ2 ¼  ð8Þ
2 5ðεu  εsh Þ

σu  fy ð1 þ esh Þ σu
C3 ¼  C4 εu   εsh  γ2  1:66 ð9Þ
2 E
esh
γ3 ¼ ð10Þ
ey
w  σu
C4 ¼  1:2 ð11Þ
1  wσ
E
u

γ4 ¼ 0:075  γ3 ð12Þ

where: σj0 is the true stress at zero plastic strain, Ck are the kinematic hardening moduli and
γk are their corresponding decreasing rate with respect to increasing plastic deformation.
The previously obtained kinematic parameters must be introduced in Abaqus using the
“Parameters” option of the combined hardening model:


KIN ¼ σj0 ; C1 ; γ1 ; C2 ; γ2 ; C3; γ3 ; C4 ; γ4 ð13Þ

The evolution of the isotropic component is more difficult to be determined due to: (1) specific
features (yield plateau, Bauschinger effect, cyclic hardening) of the mild carbon steel under
monotonic and cyclic loading (Zub et al. in press.); (2) limitations of the Abaqus built-in
cyclic (isotropic) hardening component, as the loading history dependence (Dassault 2014).
To capture the main features of the cyclic response of mild carbon steel (yield plateau,
Bauschinger effect, cyclic hardening), it is convenient defining the isotropic component using
tabular data type (Zub et al. in press.). The evolution of the size of the yield surface (σ0 ) over
the entire loading history can be specified directly by providing data pairs (σ0j ; εjpl ).
The calibration of the input data pairs for the isotropic component can be performed con-
sidering a fictional tensile test with a large plastic strain range (e.g. Δεpl = 2.0).
The equivalent stresses, σ0j , corresponding to the equivalent plastic strains, εjpl , can be
obtained by using an incremental procedure ( j = current increment):

1344
σ0j ¼ σ0j1 ; if εjpl ¼ 0
ð14Þ
σ0j ¼ σ0j1 þ ðRj  Rj1 Þ; if εjpl 40

where: Rj is the amount of isotropic hardening at increment j, for j ¼ 1; σ00 ¼ σj0 .


The variation of Rj with respect to the equivalent plastic strain εjpl is obtained using the
superposition of several isotropic hardening rules (as in the case of the kinematic component)
to properly capture the behaviour of the mild carbon steel under monotonic and cyclic
loading:

Rj ¼ R1j þ R2j þ R3j þ R4j þ R5j ð15Þ

where:
R1j – is the very-short-range nonlinear isotropic softening parameter in the plateau region,
activates immediately after yielding.
R2j – is the short-range nonlinear isotropic softening parameter in the plateau region, acti-
vates on the entire plateau region.
R3j – is the short-range nonlinear isotropic hardening parameter in hardening region, acti-
vates up to εu .
R4j – is the additional short-range or long-range (depending on the case of fitting) nonlinear
isotropic hardening parameter.
R5j – is the long-range linear isotropic hardening parameter. This hardening rule is used to
provide the combined hardening model with isotropic hardening even at large values of
equivalent (cumulative) plastic strain (e.g. ε pl ¼ 2:0).
For the isotropic softening rules (Rjk ¼ 1;2 ), the following formulas (Lemaitre & Chaboche
1990) are used to evaluate the amount of reduction of the yield surface at each increment j:
 h  i
Rkj ¼ Qk þ Rkj1  Qk  exp bk εjpl  εj1
pl
; if εjpl  εsh
ð16Þ
¼ Qk ½1  exp ð bk  εsh Þ
; if εjpl 4εsh

(for j ¼ 1; Rkj ¼ 0)
For the nonlinear isotropic hardening rules (Rkj ¼ 3;4 ), the following formulas (Lemaitre &
Chaboche 1990) are used to evaluate the amount of increase of the yield surface at each incre-
ment j:

Rkj ¼ 0; if εjpl  εsh


 h  i ð17Þ
¼ Qk þ Rkj1  Qk  exp bk εjpl  εj1
pl
; if εjpl 4sh

(for j ¼ 1; Rkj ¼ 0)
For the linear isotropic hardening rule (Rkj ¼ 5 ), the following formula is used to evaluate the
amount of increase of the yield surface at each increment j:

Rkj ¼ 0; if εjpl  εsh


ð18Þ
¼ Q0  εjpl ; if εjpl 4εsh

(for j ¼ 1; Rkj ¼ 0)
The parameters used by the hardening/softening isotropic laws are defined below. Some of
the parameters (Qk and bw, with k = 1.3) have the same meaning as the one from Hu et al.
(2016) but computed using other formulas which were determined by trial and error and

1345
showed proper fitting to the experimental study cases. Other parameters (Qk and bw, with k =
0.4) were also considered for closer fitting with the experimental results, and the proposed for-
mulas were determined by trial and error.
 
Q3 εsh  εy
Q ¼
1
ð19Þ
3  w  εsh

b1 ¼ 0:5  b2 ð20Þ

Q2 ¼  0:7  Q1 ð21Þ

Q1 1
b2 ¼  ð22Þ
εu  εsh 2:8
σu  fy ð1 þ esh Þ
Q3 ¼ ð23Þ
2:2
b2  εsh
b3 ¼  1:5 ð24Þ
εu þ εsh

Q4 ¼ additional ðuser inputÞ ð25Þ


 
b4 ¼ additional user input or ¼ 2b1 ð26Þ

2
Q0 ¼ Q1 ð27Þ
3

where Qk , bk are material parameters and represents:


Q1 and b1 – are the maximum decrease in size of the yield surface and its corresponding
rate, respectively, at very small plastic strains ð εsh ÞÞ.
Q2 and b2 – are the maximum decrease in size of the yield surface and its corresponding
rate, respectively, at plastic strains smaller than εsh .
Q3 and b3 – are the maximum increase in the size of the yield surface and its corresponding
rate, respectively, at plastic strains larger than εsh but smaller than εu (the range of equivalent
plastic strain εpl is not strictly defined). Since the isotropic component is loading history
dependent, two values can be used for b3 depending on the  type of
 loading (calibration): for
monotonic loading, the authors
 recommend
 that b 3
¼ b 2
 ε sh =ðεu þ εsh Þ  1:5, while for
cyclic loading b3 ¼ b2  εsh =ðεu þ εsh Þ  1:0.
Q4 and b4 – have similar meanings as Q3 and b3 and they are additionally used in cases
where using only Q3 and b3 the calibration is not properly achieved over the
range εsh 5εpl  εu .
Q0 – is the slope of the equivalent stresses σ0i on the hardening region.

3 APPLICATION

The above-presented calibration procedure was validated against experimental tests per-
formed for characterization of cyclic response of steel within the frame of research projects
IMSER (Zub et al. in press.) and EQUALJOINTS (Both et al. 2017). Mild carbon steels with
yield plateau (steel grade S355) were used for assessing the capability of the proposed analyt-
ical formulas in providing input parameters for the combined hardening material model that
yields reliable numerical predictions. The mechanical properties obtained from the tensile tests
are summarized in Table 1.
Figure 2 presents the influence of considering or not the isotropic component of the
combined hardening material model in predicting the monotonic and cyclic behaviour of

1346
Table 1. Mechanical properties from tensile tests of steels used for the validation of the proposed cali-
bration procedure.
test-1-S355 test-2-S355 test-3-S355 test-4-S355* test-5-S355*

fv , N/mm2 363 345 349 398 398


fu , N/mm2 525 522 534 509 509
fr , N/mm2 420 417 374 331 331
ev , mm/mm 0.0017 0.0016 0.0017 0.0019 0.0019
esh , mm/mm 0.0140 0.0170 0.0150 0.0180 0.0180
eu , mm/mm 0.1650 0.1746 0.1655 0.1634 0.1634
er , mm/mm 0.3000 0.3880 0.3246 0.3230 0.3230
*
test-4-S355 and test-5-S355 characterizes the same material. Since two coupon specimens were cyclically
tested, therefore, individual IDs were assigned to each specimen/cyclic test.

Figure 2. FEM predictions using parameters calibrated with the proposed procedure: a) kinematic
parameters only, b) kinematic and isotropic parameters.

mild carbon steel with yield plateau. From Figure 2 a) it can be observed that using only
the kinematic component the material model cannot predict the monotonic or the cyclic
behaviour of steel that was experimentally obtained, test-1-S355 (Zub et al. in press.).
For the monotonic case, it can only accurately predict the yield and the ultimate
strength, while for the cyclic loading the model underestimates the stress level and the
capacity of energy dissipation.

1347
Figure 3. FEM predictions using the combined hardening material model with parameters calibrated
using the proposed procedure.

1348
Using both isotropic and kinematic components of the combined model acceptable predic-
tions can be obtained for both monotonic and cyclic cases (Figure 2 b). For the monotonic
tensile simulation, the prediction is accurate up to the ultimate strength. Beyond this point,
the model cannot accurately predict the failure (necking region) due to the fact that the cali-
bration of the input parameters was performed to properly predict the cyclic response. In the
case of the cyclic test, close predictions are obtained at all strain ranges. If the monotonic case
is the target for the calibration, then the parameters
 describing
 the isotropic component
should be modified, as follows: b3 ¼ b2  εsh =ðεu þ εsh Þ  1:5 and Q0 ¼ Q1  x=3, where
x can take values from x ¼ 0:002  0:2.
In Figure 3 are presented the stress-strain predictions for the other tensile and cyclic tests
which were obtained within the frame of EQUALJOINT project (Both et al. 2017). Consider-
ing the same observations as in the case of the test-1-S355, in all cases an acceptable level of
correlation is observed with respect to the experimental results.
Based on these results it can be concluded that the simplified calibration procedure pro-
posed in this paper can yield numerical predictions with an acceptable level of accuracy for
both tensile and cyclic loading. The great advantage of this calibration procedure is that it
uses only tensile tests results for the calibration of input parameters which can be used for
both monotonic and cyclic loading histories. To allow for developing of a database with input
parameters for different steel grades (to be used for FEM simulations), the procedure should
be also validated against different European steel grades which exhibits yield plateau (S235,
S275, S460) or not (S690).

4 CONCLUSIONS

A procedure for the calibration of the parameters of the combined isotropic-kinematic harden-
ing model is proposed in this paper. The procedure requires only uniaxial tensile test data and
the parameters obtained can be used for both monotonic and cyclic simulations. The combined
model is implemented in both Abaqus Standard and Abaqus Explicit finite element packages,
thus both static and dynamic analyses can be performed using the calibrated parameters.
The combined material model can be used to simulate the structural components (i.e. buck-
ling restrained braces BRBs) under low cycle loading regime, which usually involves the devel-
opment of large plastic deformations.
The performance of the proposed calibration procedure is validated against experimental
results obtained on the structural mild carbon steel S355 with yield plateau. Both monotonic
and variable cyclic uniaxial tests were performed and predictions with an acceptable level of
correlation were obtained.
Proper calibration of the parameters defining the kinematic component of the combined
model should provide simulations with similar yield and ultimate strength as in the experimen-
tal uniaxial monotonic tensile test. For close fitting, the calibration of the isotropic component
should be performed considering the type of loading (monotonic or cyclic) that is applied to
the structural element (i.e. BRBs cyclically loaded) which is simulated using the combined
model. This is due to the fact that the current implementation of the combined model is load-
ing history dependent. The authors propose different formulas for some isotropic parameters
(b3 and Q0 ) depending on the type of loading (monotonic or cyclic).
Further investigations are required to validate the proposed calibration procedure for other
types of European structural mild carbon steels such as S235, S275 and S460, but also for
high strength steels S690.

ACKNOWLEDGEMENTS

The research leading to these results has received funding from the MEN-UEFISCDI grant
Partnerships in priority areas PN II, contract no. 99/2004 IMSER: “Implementation into
Romanian seismic resistant design practice of buckling restrained braces.”

1349
Partial funding was received from the European Community’s Research Fund for Coal and
Steel (RFCS) under grant agreement no RFSR-CT-2013-00021 “European pre-qualified steel
joints (EQUALJOINTS)”. This support is gratefully acknowledged.

REFERENCES

Both, I., Zub, C., Stratan, A. & Dubina, D. 2017. Cyclic behaviour of European carbon steels. CE/
Papers, Ernst& Sohn/ Wiley, Vol.1, Issue 2-3, 3173–3180.
Dassault 2014. Abaqus 6.14 - Abaqus Analysis User’s Manual, Dassault Systèmes Simulia Corp.
Hu, F., Shi, G. & Shi, Y. 2016. Constitutive model for full-range elasto-plastic behavior of structural
steels with yield plateau: Calibration and validation, Eng. Structures, 118, 210–227.
Lemaitre, J. & Chaboche, J.L. 1990. Mechanics of Solid Materials, UK: Cambridge University Press.
Nip, K.H., Gardner, L., Davies, C.M. & Elghazouli, A.Y. 2010a. Extremely low cycle fatigue tests on
structural carbon steel and stainless steel. J Construct Steel Res, 66(1), 96–110.
Nip, K.H., Gardner, L. & Elghazouli, A.Y. 2010b. Cyclic testing and numerical modelling of carbon
steel and stainless steel tubular bracing members. Eng Struct, 32(2), 424–441.
Shi, G., Wang, M., Bai, Y., Wang, F., Shi, Y., & Wang, Y. 2012. Experimental and modeling study of
high-strength structural steel under cyclic loading. Engineering Structures, 37, 1–13.
Ucak, A. & Tsopelas, P. 2011. Constitutive model for cyclic response of structural steels with yield
plateau. J Struct Eng, 137(2), 195–206.
Zub, C.I., Stratan, A. & Dubina, D. (in press). Modelling the cyclic response of structural steel for FEM
analyses. Proc. 1st Int. Conf. on Computational Methods and Applications in Engineering, Timisoara,
Romania, May (acc. for publication).
Zub, C.I., Stratan, A., Dogariu, A. & Dubina, D. 2018. Development of a finite element model for
a buckling restrained brace, Proc. of the Romanian Academy, series A, no. 19(4), 581–588.

1350
Stability and Ductility of Steel Structures 2019 – Wald & Jandera (Eds)
© 2019 Czech Technical University in Prague, Czech Republic, ISBN 978-0-367-33503-8

Author Index

Afshan, S. 911 Chiorean, C.G. 253 Feber, N. 402


Ahmed, H.S.S. 80 Ciutina, A. 863, 1251 Ferdynus, M. 622
Alhasawi, A. 269 Clayton, P. 96 Fieber, A. 3
Amamoto, T. 1292 Cordeiro, M. 1034 Filipović, A. 409
Amoush, E.A. 363 Corman, A. 564 Fiorino, L. 221
Aoyama, M. 963 Costanzo, S. 262 Fominow, S. 1095
Arha, T. 88 Couchaux, M. 269, 329, Forejtova, L. 402
Armijos-Moya, S.V. 96 981 Franco, J.M.S. 164
Arrais, F. 106, 115 Couto, C. 106, 277, 744 Fric, N. 409
Arrayago, I. 124 Craveiro, H.D. 286, 295,
Azizi, E. 1077 303 Garcea, G. 417, 735
Ádány, S. 71, 468, 499 Cristian, A.A. 1185 Gardner, L. 3
Červenka, P. 229 Gervásio, H. 699
Baddoo, N. 409 Ghosh, S. 80
Bajer, M. 1236 D’Aniello, M. 262 Gluhović, N. 427
Balázs, I. 133 D’Aniello, M. 1128, 1136 Gonçalves, P.B. 139
Barcellos, A.B.G. 139 Das, R. 312 Gonçalves, R. 434
Barros, R.C. 147 Degée, H. 321, 329, 476, González, A. 443
Basaglia, C. 1015 312 González de León, I. 124
Bastos, C.C.D.O. 155 Demonceau, J.-F. 564 Graciano, C. 708
Batista, E.M. 155, 164, 682 Dewangan, A. 337, 1042 Gremza, G. 1300, 1309
Becque, J. 173 Díez, R. 515 Guarracino, F. 452, 460
Ben Larbi, A. 269 Dinis, P.B. 345 Gurneian, H. 937
Beyer, A. 181 Dobrić, J. 409, 427
Bhatt, G. 337 Dolejš, J. 229, 1269 Haffar, M.Z. 71, 468
Bhattacharyya, S. Kr. 1042 Don, R. 1251 Hajdú, G. 1111
Bâlc, R.M. 1167 Doynov, N. 1059 Harringer, T. 596
Both, I. 1185, 1193 Du, X.X. 244 Heinisuo, M. 1145
Bours, A.-L. 171 Dubina, D. 629, 753, 1086, Helwig, T. 96
Budaházy, V. 197, 205 1185, 1193, 1342 Henriques, J. 286, 295, 303,
Burca, M. 1185, 1193 Duchêne, Y. 321 476, 1034
Bureau, A. 181 Dunai, L. 197, 539 Herrera, J. 483
Buru, S.M. 253 Hisazumi, K. 491
Ecker, A. 354 Hjiaj, M. 981
Calado, L. 329, 476 El Aghoury, I.M. 371, 929 Hoang, T. 499
Camotim, D. 345, 434, 954, El Aghoury, M.A. 363, Hoffmeister, B. 321, 329
1015 371 Hofmeyer, H. 1025
Campiche, A. 213, 221, 997 El Hady, A.M. 363, 371 Hołowaty, J. 1275
Casanova, E. 708 Eliasova, M. 379 Horáček, M. 508
Castiglioni, C.A. 311, 484 Elliott, M.D. 388
Cábová, K. 88, 1218, 1227 El-Mahdy, G.M. 394 Ibañez, J.R. 515
Ceh, O. 508 El-Serwi, A.I. 944 Ibrahim, S.M. 363, 929
Chacón, R. 235, 483 Engelhardt, M. 96 Idota, H. 638
Chen, X.W. 244 Égető, C. 539 Igawa, N. 531

1351
Ikarashi, K. 523, 531, 761, Li, Hai-Ting 16 Pichal, R. 906
1005 Li, Z. 556 Pires, D. 147
Inden, K. 963 Liguori, F. 735 Pournaghshband, A. 911
Ishida, W. 531 Liguori, F.S. 417 Pourostad, V. 921
Lišková, N. 88 Pravdova, I. 379
Jandera, M. 115, 402, 587, Lima, L. 1034
1051, 1218, 1227, 1334 Linguiti, A. 1136 Ramzi, A.A. 944
Jaspart, J.-P. 564 Ljubinković, F. 699 Rangelov, N. 1201
Jáger, B. 539 Loaiza, N. 708 Rasmussen, K.J.R. 898
Jäger-Cañás, A. 548, 556, Lopes, N. 106, 115, 744 Real, E. 124, 244, 570
882 Lorenzo, G. Di 262 Real, P.Vila 106, 115, 277,
Jiménez, A. 570 Lorenzo, G.M. Di 1128 744
Joó, A.L. 691 López, C. 515 Rocha, P.A.S. 147
Jörg, F. 578, 972 Lukačević, I. 1193 Rodriguez, A. 235
Jůza, J. 587 Lyu, Y. 1260 Rosson, B. 937

Kabeláč, J. 664, 1243 Ma, T. 717, 727 Sabau, G. 946


Kachichian, M. 539 Machacek, J. 906 Santana, M.V.B. 139
Kamada, M. 638 Macorini, L. 3 Santiago, A. 286
Kamocka, M. 614 Madeo, A. 417 Santos, W.S. 954
Kanno, R. 491 Magisano, D. 417, 735 Sato, A. 638, 872, 963
Kanyilmaz, A. 312, 329 Maia, É. 744 Sato, Y. 638
Kazmierczyk, F. 655 Manoleas, P. 946 Schaper, L. 972
Kettler, M. 596 Marçalo Neves, R. 434 Schäfer, M. 1326
Knobloch, M. 189, 972, Marković, Z. 409, 427 Seco, L. 981
1283 Martino, A. De 262 Selariu, M. 253
Kołakowski, Z. 614 Martins, J.P. 28, 295, 303, Serna, M.A. 443, 515
Kobashi, T. 605 699, 1120 Shakeel, S. 221, 989, 997
Kolakowski, Z. 655 Martínez, X. 235 Shimizu, N. 605
Kolarik, L. 402 Matsubara, G.Y. 164 Shinohara, D. 1005
Kollár, D. 205 Mela, K. 1145 Sierra, P. 235
Koltsakis, E. 946 Melcher, J. 133, 508 Silva, L.S. 303, 699
Kotełko, M. 622, 629 Michailov, V.G. 1059 Silva, da, L.S. 295
Koyama, Y. 638 Milosavljević, B. 427 Silva, da, T.G. 1015
Kraus, M.A. 1176 Mirambell, E. 124, 570 Silveira, R.A.M. 147
Kroyer, R. 646 Mitsui, K. 963 Simões da Silva, L. 28,
Kästner, T. 1103 Mittal, A.K. 1042 1120
Kuś, J. 673 Müller, A. 1151 Slein, R. 42
Kubiak, T. 614, 655 Morelli, F. 329 Snijder, H.H. 1025
Kuhlmann, U. 578, 921, 972 Movahedi, R.M. 1210 Sobrinho, K. 1034
Kumar, S. 1042 Sonkar, C. 337, 1042
Kuříková, M. 664, 1243 Neagu, C. 1185 Šorf, M. 1051
Neves, L.C. 981 Spremić, M. 409, 427
Lagerqvist, O. 946 Niko, I. 1151 Stapelfeld, C. 1059
Laím, L. 286 Nunes, D.L. 863 Stehr, S. 1068
Landesmann, A. 954 Stranghöner, N. 1068, 1077
Landolfo, R. 221, 262, 997, Okoń, K. 622 Stratan, A. 1086, 1342
1128, 1136 Oller, S. 235 Stroetmann, R. 1095, 1103
Launert, B. 1059 Ono, T. 963 Szalai, J. 1111, 1210
Leal, A.S. 682
Lemes, Í.J.M. 147 Papp, F. 1111 Taher, M.H. 468
Lendvai, A. 691 Pasternak, H. 556, 882, Takaki, S. 638
Leonetti, L. 417, 735 1059 Tanaka, Y. 1292
Li, Guoqiang 1260 Phan, D.K 898 Tankova, T. 28, 1120

1352
Taras, A. 646, 787, 839, Vallelado, L. 443 Williamson, E. 96
1151, 1159, 1176 Vassilev, M. 1201 Winkler, R. 189, 972, 1283
Tartaglia, R. 1128, Vaszilievits-Sömjén, B. 1210
1136 Vellasco, P. 1034 Xu, L. 717, 727
Teeuwen, P.A. 1025 Vesecký, J. 1218, 1227
Teh, L.H. 388 Vigh, L.G. 205 Yagi, S. 638
Tenchini, A. 1034 Vila Real, P. 106, 744 Yamaguchi, E. 1292
Theofanous, M. 911 Vild, M. 1236, 1243 Young, B. 16
Tiainen, T. 1145 Villalon-Camacho, T. 937 Yuan, H.X. 244
Titulaer, L.H.J.D. 1025 Vulcu, C. 1251
Toğay, O. 42 Zamorowski, J. 1300,
Toffolon, A. 1151, 1159 Wald, F. 664, 1136, 1243 1309
Tomăscu, I.C. 1167 Wald, F. 88 Zeybek, Ö. 1317
Topkaya, C. 1317 Wang, Y. 55, 96 Zhang, Q. 1326
Treib, C. 1176 Wang, Y. 1260 Ziemian, R. 937
Wang, Y. 1260 Ziemian, R.D. 55
Ungureanu, V. 629, 1185, Werunský, M. 1269 Zimbru, M. 1136
1193 White, D.W. 42 Židlický, B. 1334
Unterweger, H. 354, 596 Wichtowski, W. 1275 Zub, C.I. 1342

1353

You might also like