You are on page 1of 498

Connections in

Steel Structures V
Behaviour, Strength & Design

Proceedings of the Fifth International Workshop

held at

Radisson SAS Hotel


Amsterdam, The Netherlands
June 3-4, 2004

Edited by

F.S.K. Bijlaard
A.M. Gresnigt
G.J. van der Vegte

Delft University of Technology, The Netherlands

I
The information presented in this publication has been prepared in accordance with
recognized engineering principles and is for general information only. While it is believed to
be accurate, this information should not be used or relied upon for any specific application
without competent professional examination and verification of its accuracy, suitability and
applicability by a licensed professional engineer, designer or architect. The publication of the
material contained herein is not intended as a representation or warranty on the part of the
American Institute of Steel Construction, Inc. or the European Convention for Constructional
Steelwork, or of any other person named herein, that this information is suitable for any
general or particular use or is free from infringement of any patent or patents. Anyone
making use of this information assumes all liability arising from such use.

Copyright © 2005 by Bouwen met Staal, Boerhaavelaan 40, 2713 HX Zoetermeer, The
Netherlands (www.bouwenmetstaal.nl).

All rights reserved. No part of this publication may be reproduced without written permission.

ISBN-10: 90-9019809-1
ISBN-13: 978-90-9019809-5

II
FOREWORD

This book presents the Proceedings of the Fifth International Workshop on Connections in
Steel Structures: Behaviour, Strength and Design. The workshop was held at the Radisson
SAS Hotel in Amsterdam, The Netherlands, during the period 3-4 June 2004 under the
auspices of the Faculty of Civil Engineering and Geosciences of the Delft University of
Technology, The Netherlands.

The four prepreceding international workshops, alternately organised in Europe and the
USA, were held at (i) Ecole Normale Supérieure de Cachan, in Cachan, France, 25-27 May,
1987, (ii) Westin William Penn Hotel in Pittsburgh, Pennsylvania, 10-12 April, 1991, (iii) Villa
Madruzzo in Trento, Italy, 29-31 May, 1995 and (iv) Roanoke Convention and Conference
Center in Roanoke, Virginia, USA, 22-25 October 2000. Proceedings from the four preceding
workshops were published by Elsevier Applied Science Publishers (1988), the American
Institute of Steel Construction (1992), Pergamon (1996) and the American Institute of Steel
Construction (2002).

These five workshops on connections in steel structures express the importance of these
structural components. Connections have great influence on the structural behaviour and
safety of steel structures. Furthermore they determine almost 40% of the costs of steel
structures. Research in this field is therefore more than worthwhile. The success of these
workshops is also formed by the fact that the experts are selected and invited to present their
most recent research in the field of connections by the organising committee.

The American Institute of Steel Construction (AISC) and the European Convention for
Constructional Steel (ECCS) supported the idea of holding this fifth workshop. Financial
support for the workshop was generously provided by the sponsors as listed hereafter.

The Society Bouwen met Staal was the official host of the workshop. The effort put forth by
the staff at Bouwen met Staal, in particular Mrs Soraya van Beuzekom, made the workshop a
great success. This support is gratefully acknowledged. Further, the assistance by the
participants that served as session chairs helped the workshop to run smoothly. The
organising committee is indebted to Reidar Bjorhovde and Roberto Leon for their help in
selecting experts world wide. Special thanks are also due to Addie van der Vegte for his work
producing the proceedings in this book and on the CD.

The support and technical contributions of the 56 invited participants from 24 individual
countries all over the world resulted for the fifth time in a workshop of high quality and a
contribution to the profession. Without the efforts on the research, design and
implementation of steel connections worldwide, the workshop would not have been possible.
The continued efforts by the participants will no doubt result in another successful workshop
in the future.

The organizing committee:

Frans Bijlaard, Delft University of Technology,


Nol Gresnigt, Delft University of Technology,
Harry Evers, ECCS-BV,
Soraya van Beuzekom, Bouwen met Staal,
Rob Lutke Schipholt, Bouwen met Staal.

Delft, Zoetermeer, The Netherlands,


August 2005.

III
THE FIFTH INTERNATIONAL WORKSHOP ON
CONNECTIONS IN STEEL STRUCTURES

was sponsored by

Arcelor Sections Commercial, Luxemburg


A. Klaassen bv, Ridderkerk, The Netherlands
Corus Construction & Industrial, United Kingdom
Peiner Träger GmbH, Germany
Samenwerkende Nederlandse Staalbouw, The Netherlands
Stichting Bouwen met Staal, Zoetermeer, The Netherlands
Vereniging Bouwen met Staal, Zoetermeer, The Netherlands
Würth Nederland bv, Den Bosch, The Netherlands

IV
TABLE OF CONTENTS

Design Codes, Design and Education

Connection design in the 2005 AISC specification 1


Cynthia J. Duncan, (USA)

Deformation considerations for connection performance and design 11


Reidar Bjorhovde, (USA)

Evolution of shear lag and block shear provisions in the AISC specification 21
Louis F. Geschwindner, (USA)

The constructable structure in steel 27


Maatje, F., H.G.A. Evers, (The Netherlands)

Continuing education in structural connections 37


František Wald, (Czech Republic), David Moore, (United Kingdom), Milan Veljkovic,
(Sweden), Martina Eliášová, (Czech Republic)

Docking solution between a steel truss and a concrete tower at the ski jump in 45
Innsbruck
Aste, C., A. Glatzl, G. Huber, (Austria)

Modelling

Characterization of the nonlinear behaviour of single bolted T-stub connections 53


Ana M. Girão Coelho, Luís Simões da Silva, (Portugal), Frans S.K. Bijlaard, (The
Netherlands)

Experimental behaviour of aluminium T-stub connections 65


De Matteis, G., G. Della Corte, A. Mandara, F.M. Mazzolani, (Italy)

Experimental behavior of T-stub connection components for the mechanical 77


modeling of bare-steel and composite partially-restrained beam-to-column
connections
Clemente, I., S. Noé, (Italy), G.A. Rassati, (USA)

Numerical simulations of bolted connections : The implicit versus the explicit 89


approach
Van der Vegte, G.J., (Japan, The Netherlands), Y. Makino, (Japan)

Four-parameter power model for M-θr curves of end-plate connections 99


Kishi, N., M. Komuro, (Japan), W.F. Chen, (USA)

Elasto-plastic FE analysis on moment-rotation relations of top- and seat-angle 111


connections
Komuro, M., N. Kishi, (Japan), W.F. Chen, (USA)

V
Modelling, Deformation Capacity, Seismic

Modelling procedures for panel zone deformations in moment resisting frames 121
Finley A. Charney, William M. Downs, (USA)

Statistical evaluation of rotation capacity of moment connections 131


Beg, D., E. Zupančič, (Slovenia)

Rotation capacity of MR beam-to-column joints under cyclic loading 141


Grecea, D., A. Stratan, A. Ciutina, D. Dubina, (Romania)

A probabilistic evaluation of the rotation capacity of end-plate beam-to-column 155


steel joints
Luís Simões da Silva, Luís Borges, Helena Gervásio, (Portugal)

Performance-based seismic design of braced-frame gusset plate connections 167


Roeder, C.W., D.E. Lehman, J.H. Yoo, (USA)

Seismic

Effect of column stiffener detailing and weld fracture toughness on the 177
performance of welded moment connections
Dexter, R.J., J.F. Hajjar, (USA), D. Lee, (South Korea)

Analysis of bolted end-plate joints: cyclic test and standard approach 191
Dunai, L., N. Kovács, (Hungary), L. Calado, (Portugal)

Rotational capacity and demand in top-and-seat angle connections subjected 201


to seismic loading
Roberto T. Leon, Jong Wan Hu, Corey Schrauben, (USA)

Seismic performance of deep column-to-beam welded reduced beam section 211


moment connections
Ricles, J.M., X. Zhang, J.W. Fisher, L.W. Lu, (USA)

Bolted links for eccentrically braced steel frames 223


Stratan, A., D. Dubina, (Romania)

Connections in Thin Walled Sections, Eccentrically Loaded Bolt Groups

Structural properties of connections for rack structures 233


Carlos Aguirre A., (Chile)

Experimental behaviour modes of cold-formed frame-corners 243


Dunai, L., P. Fóti, (Hungary)

Preliminary component method model of storage rack joint 253


Kozłowski, A., (Portugal, Poland), L. Ślęczka, (Poland)

Strength, stiffness and ductility of cold-formed steel bolted connections 263


Dubina, D., A. Stratan, A. Ciutina, L. Fulop, Z. Nagy, (Romania)

VI
An alternative approach to design of eccentrically loaded bolt groups 273
Muir, L.S., W.A. Thornton, (USA)

Exploring the true geometry of the inelastic instantaneous center method for 281
eccentrically loaded bolt groups
Muir, L.S., W.A. Thornton, (USA)

Bolted Connections in Shear

Four-plate HEB-100 beam splice bolted connections: tests and comments 287
Zygomalas, M.D., C.C. Baniotopoulos, (Greece)

Bolted connections with hot dip galvanized steel members with punched holes 297
Valtinat, G., H. Huhn, (Germany)

High strength half round head and nut HV-bolts for ancient steel constructions 311
Valtinat, G. (Germany)

A unified approach to design for block shear 323


Driver, R.G., G.Y. Grondin, G.L. Kulak, (Canada)

Ductility requirements in shear bolted connections 335


Pietrapertosa, C., E. Piraprez, J.P. Jaspart, (Belgium)

An improved approach to compute block shear capacity of steel angles 347


Mohan Gupta, L.M. Gupta, (India)

Welded Connections, Fire

Behaviour of flange tip connections: preliminary testing and analysis 355


Snijder, H.H., J.C.D. Hoenderkamp, (The Netherlands)

Applicability of PJP groove welding to beam-column connections under 367


seismic loads
Yoshiaki Kurobane, Koji Azuma, Yuji Makino, (Japan)

Shear lag effects in fillet-welded tension connections of channels and 381


similar shapes
Matthew J.R. Humphries, Peter C. Birkemoe, (Canada)

Experimental behaviour of steel joints under natural fire 393


František Wald, (Czech Republic), Luís Simões da Silva, (Portugal), David Moore,
(United Kingdom), Aldina Santiago, (Portugal)

Tubular Connections

Strength and stiffness of RHS beam to RHS concrete filled column joints 403
Szlendak, J.K., (Poland)

Strength and stiffness of 3D plates to RHS column pin joints 411


Szlendak, J.K., (Poland)

VII
An effective external reinforcement scheme for circular hollow section joints 423
Choo, Y.S., J.X. Liang, (Singapore), G.J. van der Vegte, (The Netherlands)

The influence of boundary conditions on the chord load effect for CHS 433
gap K-joints
Van der Vegte, G.J., (Japan, The Netherlands), Y. Makino, (Japan), J. Wardenier,
(The Netherlands)

Shear lag in slotted gusset plate connections to tubes 445


Willibald, S., J.A. Packer, G. Martinez Saucedo, (Canada), R.S. Puthli, (Germany)

Review of tubular joint criteria 457


Marshall, P.W., (USA)

General Discussion and Conclusions

Conclusions 469
Stark, J.W.B., (The Netherlands)

Participants ECCS - AISC Workshop 483

Author index 485

Keyword index 487

VIII
CONNECTION DESIGN IN THE 2005 AISC SPECIFICATION

Cynthia J. Duncan, Director of Specifications,


The American Institute of Steel Construction, Inc., Chicago, IL

ABSTRACT

The American Institute of Steel Construction’s Committee on Specifications is


currently developing a new Specification for Structural Steel Buildings,
scheduled to be released in 2005. This document will unify the two design
methods presently used for steel design in the United States, Allowable Stress
Design (ASD) and Load and Resistance Factor Design (LRFD), into one
standard. In addition to this unification, the entire document is being
reorganized and updated. One area of the specification that continues to
evolve is connection design. The new standard will include several revisions
in the areas of both welded and bolted connection design.

INTRODUCTION

The American Institute Steel Construction (AISC) introduced the first specification for the
design and construction of structural steel buildings in 1923, for the purpose of creating a
standard for the steel industry in the United States. This original document was a mere nine
pages approved by a committee of five, and it has grown to exceed 100 pages, undergoing
numerous revisions based on experience gained over the years and research; both analytical
and test-based. Today, the AISC Committee on Specifications consists of 40 members
currently working on the 2005 Specification for Structural Steel Buildings (1), hereafter
referred to as the 2005 Specification. This new document has a new format unlike any
previous versions, as it will combine both load and resistance factor design (LRFD) and
allowable stress design (ASD) methods into one. More specifically, many of the provisions
have been revised and updated in Chapter J, Design of Connections, since publication of the
most recent AISC specification, the 1999 Load and Resistance Factor Design Specification
for Structural Steel Buildings (2), hereafter referred to as the 1999 Specification. Although
the specification is still in draft form, with two remaining ballots, there are many issues that
can be discussed at this time. Some of the general connection design topics that will be
addressed are compression members with bearing joints, splices in heavy sections, beams
copes and weld access holes, combining bolts and welds, and limitations on bolted and
welded connections. The welding areas that will be revised are effective area and limitations
on effective throat area of groove welds, as well as, effective area, terminations, and strength
of fillet welds. Finally, some of the changes expected for design with bolts and threaded
parts occur in provisions for: the types of fasteners allowed, combined tension and shear
strength, design of slip-critical connections, block shear, and shear lag.

NEW FORMAT AND DESIGN BASIS

Before discussing the new revisions to the connection chapter, it is important to understand
the format of the 2005 Specification. The original 1923 document was based on the
allowable stress design format (ASD), which gives capacities in allowable stresses with the
safety factor incorporated. In 1986, AISC introduced their first load and resistance factor

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 1


design (LRFD) specification (3). This design method is consistent with what had been used
world wide, as well as for the design of other materials, for example, cold-formed steel and
concrete. Since 1986, there have been two more versions of the LRFD Specification, in
1993 and 1999, and one revision of the ASD Specification in 1989. For various reasons, the
LRFD method of design has not gained in popularity among steel designers. After careful
consideration of the needs of the design community and observing how other standards
developers have handled the dilemma of incorporating two design philosophies into one
standard, AISC has embarked on the development of a “combined” or single specification,
incorporating both the ASD and LRFD methods. The design capacity will be given in a “side-
by-side” format throughout, which consists of a nominal strength for each limit state, followed
by an LRFD resistance factor and an ASD factor of safety. For example, for calculating
tensile yield strength, the new specification will read:

Pn = FyAg

φt = 0.90 (LRFD) Ωt = 1.67 (ASD)

where the design tensile strength is φtPn and the allowable tensile strength is Pn / Ωt. The
safety factors were determined based on a live load-to-dead load ratio of 3, which results in
1.5 as the target effective load factor for the load combination of 1.2D+1.6L. Therefore, in
most cases, the safety factor is calculated as 1.5/φ and it is given to 3 significant digits. The
required strength or available strength are based on ASCE 7, Minimum Design Loads for
Buildings and Other Structures (4) factored load combinations for either LRFD or ASD,
depending on the method used. This arrangement will result in greater clarity, uniformity and
efficiency when applying AISC specifications. In the final analysis, the only difference
between the LRFD and ASD method of strength design is on the required strength side.
LRFD is based on factored load combinations given in ASCE 7 and ASD is based on service
load combinations in ASCE 7. Chapter J, Design of Connections, begins by stipulating the
design basis, similar to the above followed by more definitive design provisions as discussed
in the following.

GENERAL REQUIREMENTS

Chapter J of the 2005 Specification contains the majority of the connection design provisions
in that document. The first section entitled “General Provisions” contains revisions to such
topics as compression members with bearing joints, splices in heavy sections, beam copes
and weld access holes, bolts in combination with welds, and limitations on bolted and welded
connections.

Compression members with bearing joints

The new provision permits that compression members, other than columns, be proportioned
for the less stringent of: 1. an axial tensile force of 50% of the required compressive strength
of the member or 2. the moment and shear resulting from a transverse load equal to 2
percent of the required compressive strength of the member. The application of this
transverse load should be at the splice location “exclusive of other loads that act on the
member. The member shall be taken as pinned for the determination of the shears and
moments at the splice.” This sub-section begins with a User Note that reminds the designer,
“All compression joints should also be proportioned to resist any tension developed by the
load combinations….” User Notes are a new feature of the 2005 Specification. They are
non-mandatory and are interspersed throughout the document to offer the designer concise
assistance with using the specification.

2 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


This provision is required to account for member out-of-straightness and to resist unexpected
lateral loads that may not have been considered in the design. In the past 40 years of the
AISC Specification, the only requirement that has existed required that splice materials and
connectors have a capacity of at least 50% of the required member strength. In the new
provision 1., the stipulation that these elements be designed for a tensile force provides a
more definitive way to address situations where compression on the connection imposes no
force on the connectors. Although this is a simple way to address this issue, it also can be
very conservative. Therefore, provision 2. was added offering an alternative that more
directly addresses the design intent of these provisions. The application of a lateral load of
2% simulates a kink at the splice, which could be caused by slightly out-of-square finished
ends or other construction conditions.

Splices in heavy sections

The special material toughness requirements for splices of heavy sections connected by
complete-joint-penetration groove welds have previously existed in the 1999 Specification.
The 2005 Specification will include clarification of these requirements. Shrinkage of large
welds between elements that are not free to move causes strains in the material adjacent to
the weld that can exceed the yield point strain. As the Commentary to the 2005 Specification
states, "In thick material the weld shrinkage is restrained in the thickness direction, as well as
in the width and length directions, causing triaxial stresses to develop…." and this can
prevent the steel from deforming in a ductile manner. Thus, special material toughness
requirements, and carefully prepared weld access holes and copes are required for heavy
tension members to prevent brittle fracture.

For both rolled and built-up shapes, special toughness requirements apply to shapes with
flanges or plates exceeding 2 in. (50 mm), when "used as members subject to primary tensile
forces due to tension or flexure and spliced using complete-joint-penetration groove welds
that fuse through the thickness of the member." The latter phrase was added to clarify the
extent of welding required for these provisions to be applicable. The verbiage in the 1999
Specification explaining how the impact test should be performed is replaced with a
reference to ASTM A6/A6M, Supplementary Requirement S30, Charpy V-Notch Impact Test
for Structural Shapes - Alternate Core Location (5). The impact test must meet a minimum
average value of 20 ft-lbs (27 J) absorbed energy at +70°F. The requirements do not apply if
the splices and connections are made by bolting, or if shapes with elements less than 2 in.
thick are welded to a heavy section, or to splices of elements of built-up shapes that are
welded prior to assembling the shape.” On the other hand, the provisions do "apply to built-
up cross section consisting of plates exceeding 2 in. that are welded with complete-joint-
penetration groove welds to the face of other sections."

Beam copes and weld access holes

When splicing hot rolled shapes with a flange thickness exceeding 2 in. (50 mm) and similar
built-up cross sections, special attention must also be paid to the formation of beam copes
and weld access holes. More detailed rules for the size of an access hole are given;
specifically, the height shall be 1 1/2 times the thickness of the material containing the
access hole, most likely the web thickness, but not less than 1 in. (25 mm) nor greater than 2
in. (50 mm). Room for weld backing must also be provided and no arc of the weld access
hole shall have a radius less than 3/8 in. For built-up shapes the access hole may
terminate perpendicular to the flange as long as the flange-to-web weld is held back at least
the weld size from the edge. The weld access hole details included in the 2005 Specification
are very similar to those in AWS D1.1, Structural Welding Code-Steel (6).

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 3


Bolts in combinations with welds

The design criteria for bolts in combination with welds in a joint are being completely revised
in 2005. Formerly, only bolts in slip-critical connections were permitted to share load with
welds. In the current draft, the provision reads as follows:

Bolts shall not be considered as sharing the load in combination with welds
except that connections with high-strength bolts installed in standard holes or
short slots transverse to the direction of the load are permitted to be
considered to share the load with longitudinally loaded fillet welds. In such
connections the strength of the bolts shall not be taken as greater than 50% of
the bearing strength of the bolts.

In other words, bolts in standard holes and short slots transverse to the direction of load can
share load with only longitudinally loaded fillet welds, but with a 50% reduction in the bearing
capacity of the bolts. This new provision is based on a recent research paper published in
the AISC Engineering Journal by Kulak and Grondin (7).

Limitations on bolted and welded connections

This section of Chapter J lists under what conditions pretensioned joints, slip-critical joints, or
welds are required. A similar section has existed in the AISC Specification for several
editions. For column splices, the height limitations and the language is being updated and
simplified, such that pretensioned joints, slip-critical joints, or welds are required in column
splices in all multi-story structures over 125 ft (38 m) in height. Formerly, the height limit was
based on the width of the building. The new provision is consistent with the height above
which connections of all beams and girders to columns are required to be pretensioned
joints, slip-critical joints, or welds. The remainder of the list remains unchanged, including
connections where live loads produce impact or reversal of stress and structures carrying
cranes over five-ton capacity.

Minimum strength of connections

A brief section on minimum strength of connections will be deleted. This section, also a
remnant of older versions of the AISC specification, stated a minimum factored load of 10
kips (44 kN) that all connections "providing design strength" should carry (2). The task
committee determined that these minimum loads have no technical basis and had the
potential of giving the designer the false idea that connections with this minimum design load
were adequate for fabrication and construction loads without further analysis.

WELDS

Weld provisions given in AWS D1.1 (6) apply under the 2005 Specification, with the
exception of those modified by the AISC sections listed in the preamble to Section J2. The
intention is for AISC to update their provisions to be consistent with the referenced version of
AWS D1.1. However, due to the different development cycles of the two standards, in some
cases, differences occur.

The most significant revisions to the weld provisions in the 2005 Specification occur in the
following areas: effective area and effective weld sizes for groove welds, and effective area,
terminations, and strength of fillet welds.

4 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Groove welds

In line with AWS D1.1, as well as more recent research, the tables for effective throat of
partial-joint-penetration groove welds and effective weld sizes of flare groove welds are being
updated. Table J2.1 shown below has expanded to include more combinations of welding
processes and welding positions for partial-joint-penetration groove welds (new portions are
highlighted). The terminology to describe the effective throat thickness has been revised
from "depth of chamfer" to "depth of groove."

Table J2.1 Effective throat of partial-joint-penetration groove welds.


Welding Position
F (flat), H (horiz.), Groove Type
Welding Process Effective Throat
V (vert.), (AWS D1.1 Figure 3.3)
OH (overhead)
J or U Groove
Shielded Metal Arc
All
(SMAW)
60° V
Gas Metal Arc (GMAW)
All
Flux Cored Arc (FCAW) Depth of Groove
J or U Groove
Submerged Arc (SAW) F
60° Bevel or V
Gas Metal Arc (GMAW)
F, H 45° Bevel Depth of Groove
Flux Cored Arc (FCAW)
Shielded Metal Arc Depth of Groove
All 45° Bevel
(SMAW) Minus 1/8-in (3mm)
Gas Metal Arc (GMAW) Depth of Groove
V, OH 45° Bevel
Flux Cored Arc (FCAW) Minus 1/8-in (3mm)

The minimum effective throat thickness of a partial-joint-penetration groove weld is tabulated


in the 2005 Specification with numbers identical to the 1999 LRFD Specification, except the
minimum weld thickness is based on the thickness of the thinner part joined. Previously, it
was determined based on the thicker part joined. The new Specification will read "Minimum
weld size is determined by the thinner of the two parts joined." This is again consistent with
AWS D1.1, where when low hydrogen filler metals or processes are applied, the provisions
are based on the thinner part joined. Due to the prevalence of A992 and other high strength
low alloy steels in construction today, the use of low hydrogen filler metals are required.

Effective weld sizes of flare groove welds are being increased based on a March 2003 report
by Packer and Frater (8) as shown in Table J2.2. This table applies when the flare groove
weld is filled flush to the surface of a round bar, a 90° bend in a formed section, or
rectangular tube. For flare groove welds filled less than flush, the values in Table J2.2 apply
minus the greatest perpendicular dimension measured from a line flush to the base metal
surface to the weld surface. Examples of Flare-V-groove and flare-bevel groove welds are
shown in Figure 1.

Effective throats larger than either Table J2.1 or J2.2 can be qualified by tests. For flare
groove welds the fabricator must establish by qualification the consistent production of such
larger effective throat thicknesses.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 5


Table J2.2 Effective weld sizes of flare groove welds.
Welding Process Flare Bevel Groove¹ Flare V Groove
GMAW and FCAW-G 5/8 R 3/4 R
SMAW and FCAW – S 5/16 R 5/8 R
SAW 5/16 R 1/2 R
General Note: R = radius of joint surface (Can be assumed to be 2t for HSS)
Note 1: For Flare Bevel Groove with R<0.375 use only reinforcing fillet weld on filled flush joint

Figure 1. Examples of Flare V-groove and flare-bevel groove welds.

Fillet welds

The important revisions expected to the fillet weld provisions relate to effective throat, fillet
weld terminations, strength when fillet weld groups are oriented both longitudinally and
transversely to the direction of applied load. Regarding effective throat, historically, an
increase was permitted for submerged arc welding only. This increase has not been found to
be conservative for all process settings. Therefore, the new language allows an increase in
the effective throat using any welding process if consistent penetration beyond the root of the
diagrammatic weld is demonstrated by tests.

In the 1999 Specification, specific criteria for fillet weld terminations were incorporated into
the specification proper. This material has been "tweaked" slightly by revising the language
for terminations where cyclic forces exist. The previous text read, "For connections and
structural elements with cyclic forces, … fillet welds shall be returned around the corner for a
distance not less than the smaller of two times the weld size or the width of the part." This is
now only applicable to "double angle connections and structural elements subject to cyclic
forces." Additionally, the special termination for "fillet welds joining transverse stiffeners to
plate girder webs" now only applies to plate girder webs 3/4 in. (19 mm) thick or less - a less
stringent requirement.

A new provision not yet balloted by the committee at the time of this paper, relates to the
strength of fillet welds when fillet weld groups are concentrically loaded and consist of weld
elements that are oriented both longitudinally and transversely to the direction of applied
load. The combined nominal strength of the fillet weld group shall be determined as the
greater of:

Rwl + Rwt or 0.85 Rwl + 1.5 Rwt

where,
Rwl = the total nominal strength of longitudinally loaded fillet welds
Rwt = the total nominal strength of transversely loaded fillet welds

This new provision follows two existing sub-sections that allow higher design capacities
based on the angle of loading with respect to the weld longitudinal axis. Recent research by
Ng et al. (9, 10) has demonstrated that where fillet welds exist in the same weld group that
are oriented both transverse and longitudinal to the direction of applied load, the existing
provisions are unconservative (2).

6 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


BOLTS AND THREADED PARTS

Similar to design of welds in the AISC Specifications, use of high-strength bolts conforms to
another referenced standard, entitled Specification for Structural Joints Using ASTM A325 or
A490 Bolts, as approved by the Research Council on Structural Connections, hereafter
referred to as the RCSC Specification (11). For instances where the AISC Specification
differs from this referenced document, the AISC Specification controls. The new revisions
that will appear in the 2005 Specification include an expanded list of bolts or threaded rods
permitted, more liberal use of short-slotted holes, and revised procedures for combined
tension and shear in bearing-type connections and design for shear in slip-critical
connections.

High-strength bolts

Sometimes there is a need for larger diameter (greater than 1 ½ in. (38 mm)) or longer than
usual (greater than 12 diameters) high strength bolts, such as anchor rods for fastening
machine bases. Due to the diameter and length limitations of the more commonly accepted
bolts types, such as high-strength bolts, ASTM A325, A490, and F1852 (twist-off type), the
2005 Specification will permit the use of other specified ASTM bolts or threaded rods. Bolts
or threaded rods conforming to the following ASTM specifications are permitted in this case:
ASTM A354 Gr BC, A354 Gr BD, (Quenched and Tempered Alloy Steel Bolts, Studs, and
Other Externally Threaded Fasteners) or A449 (Quenched and Tempered Steel Bolts and
Studs). For slip-critical connections, it is important that the geometry of these special
fasteners, including the head and nut(s) is equal to or (if larger in diameter) proportional to
that provided by ASTM A325 or A490 bolts. Installation must comply with the RCSC
Specification with modifications as necessary to account for the increased diameter and/or
length to provide the design pretension.

Size and use of holes

In addition to permitting other ASTM bolt types, the new specification will relax the
requirements for hole types permitted. Previously, only standard holes were allowed in
member-to-member connections without the approval of the engineer of record (EOR). It is
proposed that additionally short-slotted holes oriented transverse to the direction of load may
also be used routinely without any special approval. This is in response to what is common
practice in the fabrication industry. Short-slotted holes oriented parallel to the load,
oversized holes, or long-slotted holes still require EOR approval.

Combined tension and shear in bearing-type connections

Research has demonstrated that the strength of bearing fasteners subject to combined shear
and tension can be closely represented by an ellipse (12). Previous versions of the AISC
Specification have employed a straight-line representation of the ellipse as shown in Figure
2. In the 2005 Specification, the actual equation for the sloped portion of the approximation
is being given in the provisions, with the more exact elliptical equations given in the
commentary. The provisions read as follows:

The available tensile strength of a bolt subjected to combined tension and


shear shall be determined as φRn or Rn / Ω where φ = 0.75 (LRFD), Ω = 2.00
(ASD), Rn = Fnt′Ab, and Fn t′ = nominal tensile strength per unit area modified
to include the effects of shearing stress, ksi (MPa) defined as
F
Fnt ' = 1.3Fnt − nt fv ≤ Fnt (LRFD)
φFnv

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 7


ΩFnt
Fnt ' = 1.3Fnt − fv ≤ Fnt (ASD)
Fnv

Equation for the sloped line: ft

⎛ ft ⎞ ⎛ fv ⎞
⎟+⎜ ⎟ = 1.3
A
⎜ (LRFD)
⎝ φFt ⎠ ⎝ φFv ⎠
⎛ Ωft ⎞ ⎛ Ωfv ⎞
⎜ ⎟+⎜ ⎟ = 1.3 (ASD)
⎝ Ft ⎠ ⎝ Fv ⎠ Pt. A = φFt or Ft / Ω

Pt. B = φFv or Fv / Ω

fv
B

Figure 2. Straight-Line Representation of Elliptical Solution.

High-strength bolts in slip-critical connections

One contentious issue in the current draft of the 2005 Specification is how to handle slip
resistance of slip-critical connections. The 1999 Specification gave two procedures for
calculating slip resistance: one method using factored loads and the other based on service
loads. For consistency with the format of the “unified” specification, only one procedure is
being proposed that is purported to give substantially the same results for ASD and LRFD.
This proposed procedure provides “resistance to slip at service loads or resistance to slip at
factored loads with a reliability appropriate for serviceability criteria.” The draft criteria can be
summarized as follows:

The design slip resistance φRn and the allowable slip resistance Rn/Ω shall be determined as:
φ= 1.00 Ω = 1.40

Rn = 1.13µ hscTbNs

where:
µ = mean slip coefficient for Class A (0.35) or B (0.50) surfaces, as applicable, or
as established by tests
hsc = hole factor based on the hole type (standard, oversize, etc.)
Tb = minimum fastener tension, kips (kN)

AFFECTED ELEMENTS OF MEMBERS AND CONNECTING ELEMENTS

This section of Chapter J is applicable to elements of members at connections and


connecting elements, including strength of elements in tension, in shear, and block shear
rupture strength. The latter has gone through numerous revisions in the more recent
versions of the AISC specification and the design procedure is being revised yet another time
in the 2005 Specification, based on the latest research. The new provisions are based on
one equation instead of two as given in the 1999 Specification. The nominal strength,

Rn = 0.6FyAgv + UbsFuAnt

where:
Ubs = 1 when tension stress is uniform; 0.5 when tension stress is nonuniform
Agv = gross area subject to shear, in.2 (mm2)
Ant = net area subject to tension, in.2 (mm2)

8 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Anv = net area subject to shear, in.2 (mm2)
and 0.6FyAgv ≤ 0.6FyAnv

For a more detailed discussion of this limit state and its history, see Reference (13).

CONCLUSION

The 2005 Specification will not reach final approval until later in 2004, therefore the material
discussed in this paper is for information only and should not be applied until the final
document is announced. When that happens, the revised provisions in Chapter J for bolted
and welded connection design will be another step forward for the steel design and
fabrication industry in the United States. The AISC Committee on Specifications will
continue to work toward the goals of their mission statement:

Develop the practice-oriented specification for structural steel buildings that provides for
• life safety
• economical building systems
• predictable behavior and response
• efficient use

Based on new information from the areas of research and industry practice, the 2005
Specification for Structural Steel Buildings will allow for continued safe, as well as
economical and efficient steel building designs.

REFERENCES

(1) AISC, (2004). Specification for Structural Steel Buildings, Draft dated March 2004,
American Institute of Steel Construction, Chicago, IL.
(2) AISC, (1999). Load and Resistance Factor Design Specification for Structural Steel
Buildings, American Institute of Steel Construction, December 27, Chicago, IL.
(3) AISC, (1986). Load and Resistance Factor Design Specification for Structural Steel
Buildings, American Institute of Steel Construction, September 1, Chicago, IL.
(4) ASCE, (2002). Minimum Design Loads for Buildings and Other Structures, ASCE 7,
American Society of Civil Engineers, Reston, Virginia.
(5) ASTM, (2002), Standard Specification for General Requirements for Rolled
Structural Steel Bars, Plates, Shapes, and Sheet Piling, ASTM A6/A6M-02,
American Society of Testing Materials, West Conshohocken, Pennsylvania.
(6) AWS, (2002). Structural Welding Code -Steel, American Welding Society, AWS
D1.1/D1.1M:2002, Miami, Florida.
(7) Kulak, G.L. and Grondin, G.Y. (2003). “Strength of Joints that Combine Bolts and
Welds,” Engineering Journal, AISC, 4th Quarter.
(8) Packer, J.A and Frater, G.S., (2003). "The Effective Throat of Flare Bevel and Flare
V Groove Welds," Final Report to AISC and STI, March.
(9) Ng, A.K.F., Deng, K., Grondin, G.Y., and Driver, R.G., (2004). “Behavior of
Transverse Fillet Welds: Experimental Program,” Engineering Journal, AISC, 2nd
Quarter.
(10) Ng., A.K.F., Driver, R.G., Grondin, G.Y., (2004). “Behavior of Transverse Fillet
Welds: Parametric and Reliability Analyses,” Engineering Journal, AISC, 2nd
Quarter.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 9


(11) RCSC, (2002). Specification for Structural Joints Using ASTM A325 or A490 Bolts,
Research Council on Structural Connections, Chicago, IL.
(12) Kulak, G.L., Fisher, J.W., and Struik, J.H.A., (1987). Guide to Design Criteria for
Bolted and Riveted Joints, 2nd Edition, John Wiley & Sons, New York, NY.
(13) Geschwindner, L., (2004). “Evolution of Shear Lag and Block Shear Provisions in the
AISC Specification,” Proceedings of the Connections in Steel Structures V
Conference, Amsterdam, The Netherlands.

10 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


DEFORMATION CONSIDERATIONS FOR CONNECTION
PERFORMANCE AND DESIGN

Reidar Bjorhovde
The Bjorhovde Group, Tucson, Arizona, U S A

ABSTRACT

A tension test is used to represent the properties of steel, but it has no meaning
for the response of the material in a structure. The uniaxial tension test was
developed as a consensus solution, to have a standard by which similar
materials could be compared to a common base. It does not represent the
actual behavior of the steel in a structure, and was never intended to do so. The
paper addresses the properties of a range of structural steels, how these are
incorporated into design standards and how the standards define deformation
characteristics and demands for bolted and welded connections.

INTRODUCTION

As a construction material, steel has significant advantages over many others: it offers high
strength and stiffness, has adequate deformation capacity and stress redistribution ability for
many applications, it does not crack or otherwise fracture under normal service conditions, and
is available in several strengths and geometric forms. Finally, for most practical purposes it may
also be regarded as isotropic, with resulting benefits.

On the other hand, many structures will experience "non-normal" conditions many times during
fabrication, construction or service. A dynamically loaded structure such as a bridge will
experience fatigue; seismic events impose major deformation demands on structural
components and details; fabrication methods such as welding place very high demands for local
deformation ability of the steel in certain regions of the structure. The state-of-the-art of
computation technology is such that it is possible to incorporate many of these effects explicitly
in the analysis phase, and the quality of fabrication and construction continues to improve as
staff training and equipment are enhanced. However, much of the advanced software is not
suitable for design purposes, and most of this work therefore continues to be strictly research-
oriented.

It is a major problem that the material itself is not adequately understood by the professionals
who specify its use for structural purposes. This includes the complexity of its chemical and
metallurgical makeup, as well as the fact that the models that are used by codes to represent its
mechanical response bear little resemblance to what the steel will experience under actual
fabrication and service conditions. For one, it is known that steel is anisotropic, as a result of
production operations as well as other plastic deformation effects. Although the anisotropy
normally is of no particular consequence, it will affect the response of the steel in extreme
loading and deformation demand situations. For another, the behavior of steel is a function of
deformation history, to the effect that an otherwise ductile steel may respond as a high strength,
low ductility material, given the prior occurrence of large displacements.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 11


In brief, the complex nature of the response is not generally appreciated. The common
measures, namely, the data obtained from simple uniaxial tension tests, have little bearing on
the in-structure performance. The tension test was developed as a consensus solution, to have
the convenience of a performance standard, by which similar materials could be compared to a
common base. It does not represent the actual behavior of steel in a structure, and was never
intended to.

Many designers tend to consider the requirements of the materials standards as reflecting actual
performance ability. Two- and three-dimensional effects are not recognized, at least in part due
to the inability of design standards to correlate such effects with the elementary material
behavior models that are used. This is done in spite of the fact that multidimensional response
is the key to the behavior of some of the most important regions of the structure. In particular,
experience has shown that this is where problems tend to develop, much more than in any other
areas of the structure (1, 2, 3, 4).

Nevertheless, the focal point for designers continues to be the design codes. For rational
decisions and proper recognition of material abilities, it is essential to appreciate the
relationships between strength and deformation demands, and to assess which one that
governs the end result.

BASIC MATERIAL BEHAVIOR REPRESENTATION

The uniaxial tension test uses engineering stress and strain to define the response of the steel,
for convenience in measurement and because the test results are ultimately intended only for
use in comparison with other steels. With P = applied axial load, A0 = original cross-sectional
area of tension specimen, A = cross-sectional area at load P, l0 = original gage length, l = length
(= instantaneous length) at load P, and ∆l = change in length = (l - l0), the stress and strain are
given by the elementary expressions of Eqs. (1a) and (1b):

σ = P/A0 (1a)

ε = ∆l/l0 = l/l0 - 1 (1b)

and Eq. (1b) can be used to express the instantaneous length as

l = l0 (1 + ε) (1c)

Although convenient, and suitable as representations of the steel behavior up to and slightly
beyond yielding, these definitions do not recognize that the area changes as the load increases.
Based on the concept of an incompressible material, which is fundamentally correct for steel,
the following holds true

A0l0 = A l (2)

which can be used to define the true stress, σtr, as

σtr = P/A = Pl/Al = Pl/A0l0 = σ(l/l0) = σ(1 + ε) (3a)

Similarly. the true strain, εtr, recognizes that the length changes continuously. The strain
increment, dεtr, for an instantaneous length l is given by

dεtr = dl/l

and the total accumulated strain therefore is given by

12 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


εtr = ∫ (dl/l) = ln (l/l0) (3b)

For all practical purposes, the engineering and the true strains are equal for small strains, and
the stress-strain curves coincide up to and slightly beyond the yield stress. After this point the
two curves will diverge, to the effect that the true stress will continue to increase until material
rupture.

The most common ductility measure for steel is the elongation at fracture, εu, which is defined in
terms of engineering strain. Somewhat inconveniently, this is a function of the gage length, and
mill test reports must therefore report the magnitude of l0, as either of the commonly used 50 or
200 mm lengths. True strain is clearly a better ductility measure, since it reflects the total
accumulated strain at the point of failure. For equal levels of engineering and true stress, the
true strain is significantly smaller than the engineering strain. Conversely, for equal levels of
engineering and true strain, the true stress is significantly higher than the engineering stress.

UNI- AND MULTI-DIMENSIONAL RESPONSE CHARACTERISTICS

The preceding developments are based on one-dimensional material behavior, with no restraint
offered against deformations in the other orthogonal directions of the steel specimen. Once the
yield stress is reached, plastic deformation takes place, and necking occurs in the area of the
specimen where the failure ultimately will occur. For a more realistic assessment of the
response of the steel under multi-dimensional restraint conditions, it has been shown that the
true stress-true strain relationship is the appropriate representation, unless one proceeds to
utilize yield criteria for multi-dimensional states of strain. However, this is not practical,
especially in view of the need for fairly simple material definitions.

At the same time, the true stress and strain reflect the fact that the steel cannot supply the
amount of deformation indicated by the results of the simple tension test. Even under minor
restraint conditions, fracture takes place at strains that are significantly lower than those
specified by the materials standards.

Complicating the issues further is the fact that the yield ratio has been shown to play a major
role for the deformability of steel. Designating the yield ratio by Y, it is defined as

Yield Stress
Y = ------------------------ = Fy/Fu (4)
Tensile Strength

where Fy and Fu are the mechanical properties utilized by all common materials standards.
Examining a wide range of structural steels, Kato (5) showed that the deformability decreases
with an increasing value of Y, and the decrease is especially marked for Y-values in excess of
approximately 0.6. As an illustration, Table 1 gives the yield ratios and relevant mechanical
properties for a number of the current American structural steels. The standard numbers are
those of the American Society for Testing and Materials (ASTM); the relevant stress levels are
given in units of MPa (N/mm2). These steels have very similar counterparts within European
and Japanese steelmaking practices, to mention two major areas of steel production. Standards
published by the International Standards Organization (ISO) also reflect these types of
materials.

It is clear that adequate structural performance cannot be guaranteed by basing the material
choice only on the standards' basic properties. In fact, Kato (5) recommended that in order to
assure reasonable and reliable deformation capacity of steel members and connections, as a

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 13


minimum an upper limit of the yield stress and/or the yield ratio should be specified for each
steel grade.

Based on the above recommendation and substantial research work performed in the aftermath
of the 1994 Northridge earthquake, an enhanced 350 MPa steel grade is now produced by steel
mills in the United States, using a maximum Y-value of 0.85 (6). The specified minimum yield
stress is 350 MPa; the standard also requires a maximum value of Fy of 450 MPa, and the
minimum tensile strength is Fu = 450 MPa. Detailed chemistry requirements are provided, as is
an upper limit on the carbon equivalent, to ensure satisfactory weldability.

Table 1. Yield ratios for common American structural steels.

ASTM Steel Grade Yield Stress, Fy Tensile Strength, Fu Yield Ratio, Y

A36 250 410 - 550 0.62 - 0.45


A572 (50) 350 450 0.77
A588 (50) 350 480 0.71
A852 480 620 - 760 0.78 - 0.64
A913 (65) 450 550 0.81
A992 350 450 0.85 (max)
A514 (t ≤ 63 mm) 700 760 - 900 0.91 - 0.77

ADDITIONAL RESPONSE CONSIDERATIONS

The preceding issues are further complicated by the way tension tests are performed and
reported by steel producers. Specifically, the upper yield point is commonly given as the value
of the representative Fy. Although the use of this property is understandable, from a production
viewpoint, it is not a dependable, realistic value, since it relies heavily on the specifics of the
method of testing. As noted by Lay (7) - "the upper yield stress is not of relevance in design, as
it is lost if small overloads or misalignments occur". Specifically, utilizing dislocation theory as
the basis for yielding or plastic deformation in steel, the upper yield point mobilizes dislocations,
and the lower yield point maintains the "movement" of the dislocations. In essence, yielding is
caused by crystal structure (lattice) defects, in the form of dislocations.

These issues have not been addressed by a number of studies, among them numerous seismic
projects, their reports noting that - "...... 50 percent of the material actually incorporated in a
project will have yield strengths that exceed these mean values. For the design of facilities with
stringent requirements for limiting post-earthquake damage, consideration of more conservative
estimates of the actual yield strength may be warranted" (4). The same reference notes that
"Design professionals should be aware of the variation in actual properties permitted by the
ASTM specifications. This is especially important for yield strength. Yield strengths for ASTM
A36 material have consistently increased over the last 15 years......" (4).

Several other factors also play important roles, such as the steel production methods (e.g.
conventional and thermomechanical control processes (8, 9)). The changes that have taken
place over the past 15 years in North America in going from iron ore and coke-based ingot steel
to scrap-based continuous cast steel have resulted in materials that are significantly different,
but clearly improved in metallurgical and mechanical sense. Steel chemistry and weldability,
and especially the carbon and alloy contents, further emphasize the complex problems facing
the designer and the fabricator in the steel selection process. Some of these issues have now

14 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


been taken into account, specifically, in the development of the criteria for what is now the most
commonly used structural steel in the United States, the A992 steel (6).

PERFORMANCE INDICATIONS OF CURRENT STRUCTURAL STEELS

Current structural steels in the United States span the range from the mild, carbon-manganese
A36, to the high strength, quenched and tempered A514 and the quenched and self-tempered
A913. The traditional methods have largely given way to continuous casting; for structural
profiles in the US, current (2004) production is based entirely on this technology. As a result,
steel chemistry has changed perceptibly, such that steel now gains its strength less from carbon
and more from a variety of alloying elements. The current steels have significantly lower levels
of carbon than previous production runs. Values of C-content less than 0.10 percent are the
norm; this contrasts with a carbon content of 0.2 percent and higher for earlier steels.

The lower carbon and higher alloying elements contents result in steels with acceptable strength
and ductility characteristics, as defined by the material standards. Further, the lower carbon, in
particular, means that weldability is significantly improved. Fracture toughness is improved as
well, indicating that fatigue performance and resistance to brittle fracture should be enhanced
(9). Nevertheless, localized effects of cold straightening, for example, continue to affect the
performance of the steel, especially in connection regions. On the other hand, the issue of
through-thickness strength and ductility, which used to be regarded as critical for the
performance of seismic connections, for example, has been found to be unimportant (10).

The basic quality and variety of structural steels available to designers and fabricators have
therefore been improved significantly, yet problems persist. As demonstrated earlier in this
paper, this has at least partly been caused by misinterpretation of materials standards and what
they imply for the actual in-structure performance of the steel. Of equal importance are clearly
functions that are controlled directly by the designer and the fabricator: it is unrealistic to expect
the material to provide for all of the stiffness, strength and deformability that are needed by the
structure under all expected service conditions.

An evaluation of the ductility and deformability requirements of the current American structural
steel design specification (11) is provided in the following. It is emphasized that most of these,
where they exist, are implied rather than explicit. This is most likely the result of an engineering
tradition of focusing on stress and strength rather than strain and deformation.

DEFORMATION CRITERIA FOR SOME ELEMENTS AND CONNECTIONS

General observations

Between structural strength, stiffness and deformability, the first two are supplied relatively
easily, although improvements continue to be made through higher strength and improved
production methods. Further, many structures are controlled by the need for stiffness, in the
form of deflection or drift limits or dynamic response characteristics. For these cases the use of
higher strength steel is not advantageous. Framing system, high redundancy, and less reliance
on a limited number of structural elements are keys to successful performance.

Possibly of the greatest significance are the problems and solutions for the variety of connection
types and details that are utilized in structures. These are the regions where the material will be
exposed to the highest degrees of restraint, during shop fabrication and field erection, as well as
during high-demand service conditions. The connections influence local ductility demands and
framing performance, as evidenced by numerous examples from the Northridge earthquake (4).
Many reports of fractured welds and base metal details have been publicized. The basic

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 15


concept that cyclic loads above yield for low-ductility steel will cause fracture after a few cycles
(and conversely for a high-ductility steel) continues to be correct, but the problem is severely
aggravated under multi-dimensional degrees of restraint.

The following examples examine some of the primary American design criteria for steel
members; it focuses on certain elements only.

Tension members

Chapter D of the AISC Specification (11) details the strength criteria for tension members,
possibly the simplest structural elements, and the ones whose performance is closest to the
uniaxial conditions of the basic tension test. The limit states of gross cross section yielding and
effective net section fracture are well defined, although the reliability of the fracture case is less
than that of the overall yield. The reason for this is the greater variability of the tensile strength
(Fu) of the steel, as well as the influence of the geometry of the net section and the shear lag
associated with the cross-sectional shape and the placement of the end connection.

Ductility is recognized through the reference to strain hardening, stress concentrations, and the
importance of large deformations accompanying the yielding of the gross cross section. These
observations are based on various full- and reduced-scale tension member tests, but no data
are presented on actual deformation demands. However, in view of the relatively simple (other
than within the end connection regions) condition of these members and their satisfactory
behavior over the long term, it is generally accepted that ductility and deformation needs have
been assessed correctly. Deformation data are judged to be roughly comparable to tension
specimen tests, although specific results in support of this finding are not presented. However,
it is understood that the deformations that will occur in full-size tension members will be larger
than those of the material tests, primarily due to residual stress, initial crookedness and
eccentric application of the axial load.

Columns

Chapter E of the AISC Specification (11) gives the design criteria for columns and other
compression members. Since column buckling is primarily a stability phenomenon that is not
related to local or overall deformation demands, the issues of material performance are not
central to the issues at hand.

Beams

Chapter F of the AISC Specification (11) addresses the design criteria for laterally supported
and unsupported beams, and sections of Chapter B details local buckling and other
compactness issues. Chapter I gives the criteria for composite members; these will not be
examined here.

The overall behavior of beams is based on ultimate limit states involving in-plane or out-of-plane
failure. For example, for a laterally supported, compact beam, the ultimate limit state is the
development of a plastic hinge at the location of the maximum moment. The strength in this
case is therefore governed by the fully plastic moment, Mp, of the cross section. As another
example, for a laterally unsupported beam with an unbraced length larger than Lr, the ultimate
limit state is governed by elastic lateral-torsional buckling.

However, in all of these cases there is no clear indication of a required deformation or rotation
capacity. This is implied only through the criteria used to define compactness or the capacity of
the cross section to rotate after reaching the fully plastic moment. Specifically, flange and web
width-to-thickness ratios are established to allow full yielding in the cross section. In addition,
the beam has to be capable of rotating a certain amount beyond what constitutes the theoretical

16 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


full plastification rotation, θp, before local buckling or strain hardening occurs, at the ultimate
rotation value, θu. The deformation demand is therefore inelastic, and concentrates on the
ability of the compression flange to deform sufficiently longitudinally without buckling locally, and
of the tension flange to deform sufficiently longitudinally before strain hardening or fracture
occurs. The deformation capacity is therefore very much a function of the type of steel, or, in
other words, the level of the yield stress as well as the shape of the stress-strain relationship.
The rotation need also depends on the type of loading, to the effect that structures in seismically
active areas must be capable of supplying significantly larger inelastic rotations. A summary of
some of the key compactness criteria is illustrated in Table 2.

In Table 2 only the requirements for the flange of a W-shape are given, as an example. As an
aside, it is interesting to observe that the current seismic b/t-criterion was used for non-seismic
applications as recently as the 7th edition of the allowable stress design specification of AISC
(1970); the change of the constant from 0.30 to 0.38 was made in the 8th edition (1980).

Table 2. Compactness criteria and associated beam rotation demands.


Non-Seismic Seismic
0.38 √(E/Fy) 0.30 √(E/Fy)
Flange b/t-ratio
( = 65/√Fy) ( = 52/√Fy)
Rotation Demand Ratio θu/θp ≥3 7 to 9
Unbraced length, Lp 1.76ry√(E/Fyf) -----
Unbraced length, Lpd ----- 0.086ry(E/Fy)

In the table, the term ry is the y-axis radius of gyration; θu is the rotation developed before local
buckling or strain hardening occurs; and θp is the rotation developed as the fully plastic moment
is reached. The unbraced length criteria pertain to the maximum length that will allow the
development of the fully plastic moment for a laterally unsupported beam. Since plastic design
response characteristics are needed for seismic conditions, the requirements are much more
demanding.

The key data in Table 2 are given in the third line, as the rotation demand ratio, θu/θp. The
original data for the non-seismic ratio are based on numerous beam tests, as reported by Yura
et al. (12). The history of the seismic deformation demand is not as clear; the demand ratio
value of 7 to 9 is primarily based on studies by Popov and others, but specific references for this
work cannot be cited.

For beams in 250 and 350 MPa yield stress steel, the rotation demand ratios are governed by
the occurrence of local buckling or strain hardening in the compression flange. Limited studies
have been made of higher strength steel, but research work at the U. S. Steel Research
Laboratory in the 1960-s and early 70-s showed that for Fy = 700 MPa, tension flange fracture
governed the beam behavior (13). At the time, U. S. Steel was exploring the potential
development of hot-rolled shapes in A514 steel; this was unsuccessful as a result of the limited
rotation capacity.

Studies of beams with yield stress values from 380 to 550 MPa are very limited at this time, and
no definitive conclusions can be reached for such members. However, the practical utilization of
higher strength beams is questionable, especially in seismic areas. This is in part caused by the
"strong column, weak beam" concept, as well as the fact that beam size is frequently governed
by stiffness, rather than strength. Since the modulus of elasticity is independent of the level of
yield stress, using higher strength material beams is unnecessary.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 17


Connections

Welds and bolts are addressed in Chapter J of the AISC Specification and connection details
are covered in Chapter K (11). These are the most complex sections of the specification and
the attention given to strength limit states as opposed to deformations is very substantial. This
is done in spite of the fact that the deformation response often controls the actual ultimate limit
state.

In the following only some of the requirements will be examined. However, in view of the severe
deformation demands that are placed on many types of connections, it would seem important to
assess all of these specification criteria in detail, to gain a clear understanding of what is
expected of the material when the connections are designed according to the Specification.
This is especially important for many types of beam-to-column moment connections and some
welded tension member splices, for which localized material deformation demands can be very
high (2, 3).

Welded splices in heavy shapes

The criteria for welded splices in very heavy wide-flange shapes are qualitative, but clear
recognition is given to the fact that the combination of residual stress, localized high deformation
demand due to fabrication operations, high localized hardness, and low fracture toughness in
the core area of hot-rolled shapes have the potential for leading to cracks and propagation of
cracks (2, 3). The event that caused this change in the AISC Specification was the bottom
chord fracture in one of the trusses for the Orange County Civic Center in Orlando, Florida. The
core area problem is much less important now, since continuous cast shapes have smaller and
less pronounced cores. The shapes that cracked in the Florida structure were all ingot-based.

Overlap in fillet welded joints

Single lap welded joints will rotate when subjected to axial forces in the longitudinal direction,
due to the eccentricity of one plate relative to the other. The Specification recognizes the need
for a certain length of overlap between the plates or members in the joint, equal to five times the
thickness of the thinnest part, but not less than 25 mm. If this is satisfied, ". the resulting
rotation will not be excessive....." (11). No specifics are given as regards rotation magnitudes.

Short vs. long bolted joints

Short bolted joints generally deform in such a fashion that localized yielding allows for a
redistribution of the bolt forces, to load each bolt equally. This is not the case with long bolted
joints, for which the non-uniform strain distribution leads to significant differences in the actual
bolt loads. In particular, the outermost bolts will have the higher loads, leading to the potential
for an "unzipping" type of failure. The AISC Specification recognizes this behavior by reducing
the tabulated bolt strength values by 20 percent for connections longer than 1270 mm.
However, the actual deformation demand is only accounted for qualitatively.

Bearing strength at bolt holes

This is one of the few cases where deformation and strength limit states are explicitly
recognized. Research has shown that the hole deformation will increase beyond 6 mm when the
nominal factored bolt load exceeds 2.4dtFu (11). Under many circumstances this will be
unacceptable, due to the contribution of such deformations to overall connection deformations.
If the bolt load increases to 3dtFu, the limit state will be that of hole ovalization or "dishing".

18 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Beam-to-column connections

The criteria for the design of the details of beam-to-column connections are given in Chapter K
of the AISC Specification (11). The section provides extensive ultimate limit state criteria for
local flange bending, local web yielding, web crippling, sidesway web buckling, compression
web buckling, panel zone web shear, unframed beam and girder ends, additional stiffeners
requirements for concentrated forces, and additional doubler plate requirements for
concentrated forces. It is only in the treatment of panel zone web shear that strength and
deformation are explicitly recognized. It is formulated qualitatively, to the effect that different
equations are used to determine the nominal strengths if panel zone deformations are
considered in the frame stability analysis.

The Commentary for the Specification gives qualitative observations involving deformation
needs, such as "... that flange must be sufficiently rigid to prevent deformation of the flange ......"
(11). A detailed evaluation is provided for the panel zone behavior and the importance of its
deformation as regards the story and overall drifts of the structure.

SUMMARY

The paper has presented a discussion of issues related to performance demands for steel in
structures, especially under high restraint and high dynamic load conditions. It is shown that the
use of elementary materials standards requirements, which reflect uniaxial tension test
response, is unacceptable as a means to assess the response of the steel in the structure
during actual operating conditions.

Designers and fabricators must fully understand the material behavior. However, it is also clear
that the steel cannot assure satisfactory behavior by itself. Only together, through (1) material
choice, (2) local and overall structural design, and (3) shop and field fabrication techniques and
operations, will overall performance demands be met. In all cases strict adherence to specified
procedures is essential, for what good does a specification do if it is not used? Future
developments may see improved material standards, particularly if upper and lower limits are
placed on the specified yield stress values, and/or yield ratios are defined and required. The
criteria of the A992 standard of ASTM reflect a novel and very significant improvement.

REFERENCES

(1) American Institute of Steel Construction (AISC) (1973), "Commentary on Highly


Restrained Welded Connections", AISC Engineering Journal, Vol. 10, No. 3, Third
Quarter (pp. 61-73).
(2) Fisher, J. W., and Pense, A. W. (1987), "Experience with the Use of Heavy W-Shapes in
Tension", AISC Engineering Journal, Vol. 24, No. 2, Second Quarter (pp. 63-77).
(3) Bjorhovde, Reidar (1988), "Solutions for the Use of Jumbo Shapes", Proceedings, AISC
National Steel Construction Conference, Miami Beach, Florida, June 8-11 (pp. 2-1 to 2-
20).
(4) Federal Emergency Management Agency (FEMA) (2000), "Recommended Seismic
Design Criteria for New Steel Moment-Frame Buildings", Bulletin No. 350, FEMA,
Washington, D.C.
(5) Kato, Ben (1990), "Deformation Capacity of Steel Structures", Journal of Constructional
Steel Research, Vol. 17, No. 1 (pp. 33-94).
(6) American Society for Testing and Materials (ASTM) (2003), "Standard Specification for
Structural Steel Shapes”, Standard No. A992, ASTM, Conshohocken, Pennsylvania.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 19


(7) Lay, M. G. (1982), "Structural Steel Fundamentals", Australian Road Research Board,
Vermont South, Victoria, Australia.
(8) Barsom, John M. (1987), "Material Considerations in Structural Steel Design",
Proceedings, AISC National Engineering Conference, New Orleans, Louisiana, April 29-
May 2 (pp. 1-1 to 1-29)
(9) Barsom, John M. (1996), "High Performance Steels", American Society of Metals
Advanced Materials & Processes, No. 3 (pp. 27-31).
(10) Dexter, R. J. and Melendrez, M. I. (1999), “Through-Thickness Strength and Ductility of
Column Flanges in Moment Connections”, University of Minnesota Research Report to
the SAC Consortium.
(11) American Institute of Steel Construction (AISC) (1999), "Load and Resistance Factor
Design Specification for Structural Steel Buildings", 3rd Ed., AISC, Chicago, Illinois.
(12) Yura, J. A., Galambos, T. V., and Ravindra, M. K. (1978), "The Bending Resistance of
Steel Beams", Journal of the Structural Division, ASCE, Vol. 104, No. ST9, September
(pp. 1355-1370).
(13) Haaijer, G. H. (1989), "Private Communication", Pittsburgh, Pennsylvania.

20 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


EVOLUTION OF SHEAR LAG AND BLOCK SHEAR PROVISIONS
IN THE AISC SPECIFICATION

Louis F. Geschwindner, Vice President of Engineering and Research,


The American Institute of Steel Construction

ABSTRACT

The American Institute of Steel Construction (AISC) publishes two standards


for design of steel structures, an allowable stress specification, ASD, and a
limit states specification, LRFD. The AISC Committee on Specifications is in
the process of developing a new standard that will include provisions for
design according to both ASD and LRFD. For connection design, the
specification is evolving as provisions from the previous specifications are
integrated. In particular, the provisions for shear lag and block shear have
seen some evolution. This paper highlights the new provisions and the
changes that have taken place from previous editions of the specifications.

INTRODUCTION

The planned 2005 AISC Specification for Structural Steel Buildings (1) will include provisions
for both allowable strength design (ASD) and load and resistance factor design (LRFD). The
overriding principle guiding the development of these new provisions is that the steel does
not know what method was used in its design. Thus, there should be a single approach for
determining member strength and the modification of that strength to be consistent with the
ASD and LRFD loading provisions of the governing building codes. The AISC Committee on
Specifications and its Task Committees were charged with evaluating the existing ASD and
LRFD provisions and incorporating the best of both standards into the new standard. In
addition, research results that have become available since publication of the previous
standards, ASD in 1989 (2) and LRFD in 1999 (3) should be incorporated. Thus, this new
specification should be a step forward for each design approach.

The original LRFD Specification (4) was calibrated to the then existing ASD Specification (5)
for a live load to dead load ratio, L/D = 3, in order that the new specification produce designs
that were comparable to the existing provisions. The ASD and LRFD design philosophies
are stated in the 2005 Specification (1) draft for ASD as
Rn
( D + L ) = Ra ≤ (1)

and for LRFD as
(1.2 D + 1.6 L ) = Ru ≤ φ Rn (2)

where Ra and Ru are the required strengths determined from the ASD and LRFD load
combinations, Rn is the nominal strength, φ is the resistance factor for LRFD, and Ω is the
safety factor for ASD . For the load combination of (1.2D + 1.6L) and L/D = 3, the effective
load factor becomes 1.5. Thus, the relationship between φ and Ω can be determined by
solving Eqs. 1 and 2 for Rn and setting the them equal, thus

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 21


(1.2 D + 1.6 L ) = 1.5( L + D) = Ω( D + L) (3)
φ φ
which yields

1.5
Ω= (4)
φ

This relationship guides the development of factors of safety for the ASD provisions of the
unified specification, based on the LRFD resistance factors. It will be seen that the resulting
strength provisions for tension members in the 2005 Specification (1) will be the same as
they are in the current ASD (2) and LRFD (3) Specifications. The provisions for shear lag will
be modified slightly to account for some recent research results.

Block shear has seen several changes over the years since it was first introduced into the
specification. However, it will also be seen that the strength provisions for 2005 are
essentially the same as the ASD and LRFD provisions with a slight variation in the
controlling factor.

SHEAR LAG

Shear lag provisions were first introduced in the 1978 AISC ASD Specification (5). This was
to account for the research findings that the net section was not fully effective in providing
fracture strength when all elements of the tension member section were not attached to the
connecting elements The provisions of Section 1.14.2 simply stated that the effective net
area was to be taken as the net area times a reduction factor, thus

Ae = Ct An (5)

Three cases were identified for determining Ct :

1. W, M, or S shapes with flange widths not less than 2/3 the depth, and structural tees
cut from these shapes, provided the connection is to the flanges and has no fewer
than 3 fasteners per line in the direction of stress. Ct = 0.90.
2. W, M, or S shapes not meeting the requirements of subparagraph 1, structural tees
cut from these shapes, and all other shapes, including built-up cross sections,
provided the connection has not less than 3 fasteners per line in the direction of
stress. Ct = 0.85.
3. All members whose connections have only 2 fasteners per line in the direction of
stress. Ct = 0.75.

The commentary to Section 1.14.2 indicates that these values are reasonable lower bounds
for profile shapes and connection means described in the research of Munse and Chesson
(6). In that research they proposed an equation

Ct ≈ 1 − x l (6)

that, although not a part of the actual specification provisions, was included in the
commentary.

22 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


The 1986 LRFD and 1989 ASD Specifications continued the use of these specified reduction
factors, although the symbol was changed to U. In addition, provisions were made for
members connected through welds. For the 1993 LRFD Specification, the equation
developed by Munse and Chesson (6), with an upper limit of 0.9 added, was made a part of
the Specification and the numerical values that had been in use until this time were moved to
the commentary. The background for these changes is reported by Easterling and Girouk
(7). The three previously used cases were made available for designers continued use as
reasonable lower bounds that could be used unless a higher value was determined through
the provided equation. The same provisions were carried over for the 1999 LRFD
Specification (3).

The 2005 Specification (1) needed to consider how the combination of ASD and LRFD
provisions would impact tension member strength and how, if at all, the effective net area
provisions would need to change. Using the relation between φ and Ω presented in Eq. 4,
the design tensile strength, φPn, and the allowable tensile strength, Pn/Ω, for the limit state of
fracture can be taken from

Pn = Fu Ae (7)

φ = 0.75 (LRFD) Ω = 2.0 (ASD)

This is the same provision as given in the 1999 LRFD Specification (3) and a comparison
with the 1989 ASD Specification (2) shows that it provides the same allowable strength.
Since effective net area is not a function of design approach, there will be no impact there.

There are two changes being considered for the 2005 shear lag provisions. The first is the
removal of the upper limit on U. There does not appear to be sufficient research results to
warrant retaining this limitation. The work of Munse and Chesson (6) did not include a
recommendation that the shear lag reduction factor have an upper limit, although their
testing program included only a few cases where this might have come into play.

The second change is the addition of a requirement that single angles, double angles, and
WT’s be proportioned so that U is equal to or greater than 0.6. Alternatively, a lesser value of
U is permitted if the tension member is designed for the effect of eccentricity through the use
of the interaction equations. Recent work reported by Epstein and Stamberg (8) suggested
that for WT sections, a lower bound of 0.65 be placed on U.

Since the 2005 Specification (1) will incorporate the previous separate specifications for
single angles and for HSS, the number of shear lag cases has increased. Thus, a table is
being provided to simplify the determination of appropriate U-values.

BLOCK SHEAR

As was the case for shear lag, block shear provisions first appeared in the 1978 ASD
Specification (5). These provisions were the result of the work of Birkemoe and Gilmor (9)
that was directed at the coped beam connection. The provisions, as stated in Section
1.5.1.2.2, indicate that the shear at beam end connections where the top flange is coped,
and in similar situations where failure might occur by shear along a plane through the
fasteners, or by a combination of shear along a plane through the fasteners plus tension
along a perpendicular plane on the area effective in resisting tearing failure, the stress was
to be limited to 0.3Fv This would have resulted in an allowable force equation of

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 23


Va = 0.3Fu ( Anv + Ant ) (8)

The commentary provided an alternative where the tension and shear areas could be treated
separately as
Va = 0.3Fu Anv + 0.5 Fu Ant (9)

Obviously this would have provided an increase in allowable strength.

The 1989 ASD Specification (2) brought Eq. 9 from the commentary into the specification
and provided the direction that block shear should also be considered for welded
connections.

Block shear provisions are not as clearly articulated in the 1986 LRFD Specification (4). The
provision for shear fracture is given in the body of the specification but the discussion of
block shear is relegated to the commentary. In the commentary presentation, the possibility
of a combination of yielding on one plane and fracture on the other is introduced and the
following two equations are given

φ Rn = φ [0.6 Fy Agv + Fu Ant ] (10)

φ Rn = φ [0.6 Fu Anv + Fy Agt ] (11)

with φ = 0.75 and the largest value of φRn taken as the design strength. It is worth comparing
this first introduction of LRFD block shear provisions with the ASD provisions using the
relationship of φ and Ω presented in Eq.4. Since φ = 0.75, then Ω=2.0. Thus, dividing the
nominal fracture term from Eqs. 10 ands 11 by the safety factor of 2.0 and adding them,
yields Eq. 9.

The 1993 LRFD Specification (10) brought the two block shear equations forward into the
specification but altered the way that the controlling equation was selected. For these
provisions, the controlling factor was fracture. Shear fracture and tension fracture were to be
calculated and the larger fracture term was to be combined with the opposite yield term.
Additional research reported by Ricles and Yura (11) for coped beams and Hardash and
Bjorhovde (12) for gusset plates confirmed that the strength could be determined by the
summation of the shear and tension terms.

The 1999 LRFD Specification (3) continues to use the previously presented block shear
equations but sets an upper limit on the strength that is equal to the sum of the two fracture
terms and can be stated as

φ Rn = φ [0.6 Fu Anv + Fu Ant ] (12)

which is actually a return to the commentary equation of the 1978 ASD Specification (5),
since Rn from Eq. 12 divided by the factor of safety, 2.0, yields Eq. 9.

For the 2005 Specification (1), the currently proposed block shear provisions have
undergone another slight modification. For shear strength, either fracture or yield, the
relations remain unchanged. For tension strength, two revisions are recommended. The first
is the recognition of the influence of non-uniform tension, as would occur on the block shear
tension face for a coped beam with two rows of bolts, as identified by Ricles and Yura (11).
The second change is the use, for all conditions, of the tension fracture strength, rather than
either tension fracture or yield strength. The resulting provisions are

24 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


φ Rn = 0.75 ( 0.6 Fy Agv + U bs Fu Ant ) (13)
but not greater than

φ Rn = 0.75 ( 0.6 Fu Anv + U bs Fu Ant ) (14)

When the tension stress is uniform, Ubs = 1.0 and for cases where the tension stress is not
uniform, Ubs = 0.5. This is consistent with the recommendations of Kulak and Grondin (13).
Although the committee considered a more involved approach to the calculation of Ubs, it
was decided to simplify the term so as not to make its determination laborious.

CONCLUSIONS

An evolution of the shear lag and block shear provisions has taken place since their first
introduction in the AISC Specification in 1978 (5). Although the changes have been slight,
they reflect an improving understanding of the behavior that they are attempting to predict.
The intent has always been to provide specification provisions that are sufficiently accurate
so as to provide for safe and economical structures while at the same time providing design
methods that are simple and economical to apply.

The 2005 “unified” Specification (1) is scheduled to receive final approval by the end of
2004. This schedule has been set in order that it may be incorporated into the next revisions
of the NFPA and IBC building codes. Although the provisions presented here for the 2005
Specification are those under consideration at this time. The committee process allows for
changes to be made until the final Specification has received Committee on Specifications
approval. Thus, this description of what might be ahead for the designer should not be relied
upon for design.

NOTATION

Ae = Effective net area


Agt = Gross area subjected to tension
Agv = Gross area subjected to shear
An = Net area
Ant = Net area subjected to tension
Anv = Net area subjected to shear
Ct = Shear lag reduction factor from the 1978 ASD Specification
D = Dead load
Fu = Specified minimum tensile strength
Fy = Specified minimum yield stress
L = Live load
Pn = Nominal tensile strength
Ra = Required strength (ASD)
Rn = Nominal strength
Ru = Required strength (LRFD)
Va = Allowable shear based on block shear
x = Distance from shear face to center of gravity of section
φ = Resistance factor
Ω = Safety factor

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 25


REFERENCES

1. AISC, (2004). Specification for Structural Steel Buildings, Draft dated March 2004,
American Institute of Steel Construction, Chicago, IL.
2. AISC, (1989). Specification for Structural Steel Buildings – Allowable Stress Design and
Plastic Design, American Institute of Steel Construction, June 1, Chicago, IL.
3. AISC, (1999). Load and Resistance Factor Design Specification for Structural Steel
Buildings, American Institute of Steel Construction, December 27, Chicago, IL.
4. AISC, (1986). Load and Resistance Factor Design Specification for Structural Steel
Buildings, American Institute of Steel Construction, September 1, Chicago, IL.
5. AISC, (1978). Specification for the Design, Fabrication and Erection of Structural Steel
for Buildings, American Institute of Steel Construction, November 1, Chicago, IL.
6. Munse, W. H., and Chesson, Jr., E. (1963), “Rivited and Bolted Joints: Net Section
Design,” Journal of the Structural Division, ASCE, Vol. 89, No. ST1, February, pp. 49-
106.
7. Easterling, W. S., and Giroux, L. G., (1993), “Shear Lag Effects in Steel Tension
Members,” Engineering Journal, AISC, Vol. 30, No. 3, 3rd Quarter, pp. 77-89.
8. Epstein, H., I., and Stamberg, H., (2002), “Block Shear and Net Section Capacities of
Structural Tees in Tension: Test Results and Code Implications,” Engineering Journal,
AISC, Vol. 39, No. 4, 4th Quarter, pp. 228-239.
9. Birkemoe, P. C., and Gilmor, M. I., (1978), “Behavior of Bearing-Critical Double –Angle
Beam Connections,” Engineering Journal, AISC, Vol. 15, No. 4, 4th Quarter, pp. 109-
115.
10. AISC, (1993). Load and Resistance Factor Design Specification for Structural Steel
Buildings, American Institute of Steel Construction, December 1, Chicago, IL.
11. Ricles, J. M., and Yura, J. A., (1983), “Strength of Double-Row Bolted Web
Connections,” Journal of the Structural Division, ASCE, Vol. 109, No. ST1, January, pp.
126-142.
12. Hardash, S. G., and Bjorhovde, R., (1985), “New Design Criteria for Gusset Plates n
Tension,” Engineering Journal, AISC. Vol. 22, No. 2, 2nd Quarter, pp. 77-94.
13. Kulak, G. L., and Grondin, G. Y., (2002), “Block Shear Failure in Steel Members – A
Review of Design Practice,” Connections in Steel Structures IV, Proceedings of the
Fourth International Workshop, AISC, pp., 329-339.

26 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


THE CONSTRUCTABLE STRUCTURE IN STEEL
F. Maatje, ICCS bv, The Netherlands
H.G.A. Evers, ICCS bv, The Netherlands

ABSTRACT
Nowadays engineers still tend to optimize a structure to a minimum of
kilograms, which is the only hard criterion that is available for an engineer
during the design of the structure. In the end this criterion leads to structures
that are expensive and have poor quality.
In this article it is illustrated that selection of a deeper top- and bottom chord
member and the use of 3D-analysis software leads to a reduction of costs of
the total project. Furthermore, it results in improvement of the quality of the
structure as a whole.

THE ISSUE

For most of the structures it applies that they are established in the same way for many
years. An architect makes an architectonic design, a design engineer designs the main
structure and a steel fabricator takes care that the construction will be detailed, fabricated
and erected.

These phases are mostly sequential and do not have much overlap. Consequence is that the
expertise and involvement of the several parties is restricted to his/her own specialty.
Because of this strict separation, but also because of ignorance and incapability, the
engineer often disregards the phases that come next. Most of the times the result is that the
detailing of the structure becomes very complex and that the structure will be difficult to
fabricate.

These things do not have a positive effect on the project financially, constructively,
aesthetical, as well as systematically. By keeping the detailing of the steel into account in an
early stage, a better and more cost effective structure can be designed.
From their own disciplines, ICCS bv and IDCS bv are involved with a lot of steel structures
and they observe that the abovementioned issue frequently occurs. Furthermore, it increases
the last couple of years because the quality of a main design decreases strongly.

STARTING POINTS

Inspired by many examples from the daily practice, the issue will be illustrated by discussion
of a standard truss.

Based on drawings of the architect the design engineer designs a steel structure. This
structure consists of a large amount of trusses that span a large space in two directions.
Therefore, both in longitudinal and transverse direction the main structure is modelled by
trusses that lay on a concrete substructure. For the analysis the centre lines of the profiles
are modelled. Members that come together in a node, thus theoretically intersect at one
point. Result is that there is a simple transfer of forces and no secondary forces or moments

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 27


arise. In a design phase this is a correct and also convenient choice, as worries about the
detailing and fabrication arise in a later stage.

Figure 1a. Centre lines intersect (actual). Figure 1b. Centre lines intersect (analysis model).

Figure 2a. Eccentricities e = 100 mm (actual). Figure 2b. Eccentricities (analysis model).

ECCENTRICITIES AND FORCE LEAD-IN

If the structure is made of concrete, the abovementioned approach can be followed. Because
of weight reductions, costs, construction time and erect onsite most of the time the trusses
are made of steel.

But the problem arises when the engineer has realised an “optimum” design on the basis of a
minimum of kilograms. For certain if, as in this outlined situation, high forces are involved. A
small eccentricity, by moving the centre lines, leads to a considerable extra bending moment.
NEN 6772, article 11.6.1 mentions that at determining the joint capacities the eccentricities
may be neglected if:

e e
− 0,55 ≤ ≤ 0,25 or − 0,55 ≤ ≤ 0,25
d0 h0
e=0 e=negative e=positive

Figure 3. Definition of eccentricity.

28 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


When a section is loaded to its full capacity problems arise. The simple detail that the steel
fabricator had in mind, cannot be realised. The centrelines of the profiles thus have to go
through one point, as the profiles do not have the extra capacity to withstand the extra
moments due to eccentricities.

Moreover, problems are not only caused by shifting the gridlines, but the area of force
application also deserves attention. When the diagonals and bottom- and top chord member
have similar profiling, it is very difficult to realise proper force lead-in without stiffeners.

The maximum allowable normal force in a profile is many times higher than the maximum
allowable shear force. Moreover, this is being reduced considerable when it concerns a web
that is sensitive for local buckling. Proper force lead-in without stiffeners is almost impossible
when the maximum allowable normal force in the diagonal is higher than the shear force
capacity of the bottom- and top chord member.

So not only does the bottom- and top chord member need to have an capacity left in order to
withstand the extra moments due to eccentricities, but, to prevent the need of stiffeners, they
also need to have a shear force capacity that is higher than the normal force capacity of the
diagonal.

The engineer often seems not to be aware of this. In the design he does not, or too little, take
this into account. If he chooses profiles with little reserve capacity in the design stage, it
results in fairly complicated, laborious and therefore expensive details.

Figure 4. Concrete. Figure 5. Steel, complex details.

THE CONNECTIONS MAKE THE COSTS

The last couple of years the awareness that costs are caused largely by the connections has
sunk in. See former publications (1) regarding costs comparisons in the steel industry.
Additional stiffeners must be prevented because these components are labour-intensive.
Preference is given to simple connections with fin plates, angle cleats or endplates. The
recent price increase of the material is the reason that saving on material becomes an
important item. A good insight in total costs is and stays essential.

In order to come to a structure that is acceptable for all parties, there is a number of
solutions. Every party has its own preference for the ideal structure. An engineer does not
want (m)any eccentricities, a Steel fabricator wants the most simple fabrication possible
without a lot of labour and an architect wants an aesthetical justified structure, especially
when the steel structure remains visible. There is a number of methods to decrease the
amount of labour in the construction afterwards. Below you will find a few possible solutions.
Dealt with is the detail in the bottom chord member, but the same is applicable to
connections at the top chord member.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 29


Solution 1: Rotating profiles by 90 degrees

If the starting point remains that the centre lines have to go through one point, the profiles
can be rotated 90 degrees. If necessary, the thickness of profiles must be locally adjusted at
the joint because of the various profile dimensions.

In a truss it only concerns tension or compression members. By detailing the joint with two
plates on the flanges and not bothering the centre lines, no extra forces are introduced and
the connection can be realised with sawing and drilling as main labour. During the
processing stage the tolerance on the clearance between bolts and boltholes have to be
taken into account. This can be done by shortening the tension members or by use of
injection bolts. Furthermore the introduction of holes in the bars means a reduction of the
cross section area. If the tension members are used up till their full capacity, the weakening
due to holes lead to a reduction of the capacity, so that the profile will fail.

The thickening of the profiles, as mentioned above in connection with various profile
dimensions, gives as a side-advantage an increase of the tensile force capacity, so that the
profile complies again. Applying this solution mostly only make sense when the truss in the
nodes is directly loaded by purlins or floor beams.

If a floor or roof rests on the top or bottom chord member and this causes moments, it is
possible that the profiles do not meet the requirements anymore in connection with moments
about the weak axis of the profile.

Figure 6. Rotating 90 degrees.

Solution 2: Strengthen/deepen bottom chord member (and top chord member)

When the diagonals are shifted away from the node simple endplate connections are made
possible and the connection no longer overlap. To enable this, the bottom chord member in
the truss must be deeper than is strictly necessary for transfer of the normal- and shear
forces that appear. This solution cuts both ways: because of the overcapacity of the bottom
chord member it is easy to take up eccentricities. The top flange moves upwards so that the
eccentricities of the diagonals become smaller.

ho e ≤ ho/4

Figure 7. Deeper beam.

30 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Moreover, the web of the deeper profile has a higher capacity, so that use of additional
components in order to lead in the forces properly can be prevented. This will be illustrated
by an example.

Suppose: a truss with a span of 20 meter, completely built up from HE200A profiles both for
the bottom- and top chord member and the diagonals.

Figure 8a. Analysis model.

e = 100 mm

Figure 8b. Analysis model due to detailing.

If the diagonals are connected to the bottom- and top chord member via endplates, the
eccentricity, that occurs at the connection of the diagonal on the bottom chord member will
have to be somewhat like100 mm. Suppose: a maximum normal force (Npl HE200A) in the
diagonal of 1200 kN. This gives a vertical force of 1000 kN and thus a moment of 100 kNm.
The maximum moment a HE200A can transfer is 101kNm, without keeping into account the
buckling lengths and moments that are already present as a result of the direct load. So the
“extra” moment cannot be withstood by the bottom chord member. In this realistic situation
the maximal normal force is 600 kN and thus the vertical force is 500 kN. As a result of the
eccentricity this leads to an extra moment of 50 kNm on the beam.

When the Head engineer has totally utilized the bottom- and top chord member, this moment
too cannot be withstood. Not any “extra” moment can be absorbed, so that the steel
fabricator is forced to connect the diagonals in the centre lines, which results in a
complicated and above all laborious connection design.

But what is the maximum normal force that can be absorbed without application of stiffeners
in the bottom and top chord member ? According to article 14 of the NEN6770 the bottom
chord member should be checked for:

Yield of the web F = ( c + d )t f


u ;1; d 1 w y; d

⎧⎪ t f ⎛t ⎞⎛ c ⎞⎫⎪
Local buckling of the web Fu;2;d = 0.125 t w2 Ef y;d ⎨ + 3⎜ w ⎟ ⎜ ⎟⎬
⎪⎩ t w ⎜ t f ⎟ ⎜ h − 2t f ⎟⎪
⎝ ⎠⎝ ⎠⎭
⎧⎪ t f ⎛t ⎞⎛ c ⎞⎫⎪
Fu;2;d = 0.5 t w2 Ef y;d ⎨ + 3⎜ w ⎟ ⎜ ⎟⎬
⎪⎩ t w ⎜ t f ⎟ ⎜ h − 2t f ⎟⎪
⎝ ⎠⎝ ⎠⎭

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 31


N c ; s ;d
Global buckling of the web ≤ 1 with
met bef = h 2 + c 2
ω y;buc N c;u;d
In which
tf = the thickness of the flange fy;d = the value of the yield stress
tw = the thickness of the web h = the depth of the profile
c = the length over which the force is being exercised E = the value of the elasticity
modulus

From this follows a maximum normal force in the diagonal of 337 kN without the need of
stiffeners in the bottom chord member, whilst the maximum allowable normal force for this
diagonal is 860 kN (HE200A with a buckling length of 3.60m). By strengthening the bottom
chord member to HE240B the normal force in the diagonal becomes 700 kN without
application of the necessary stiffeners in the bottom chord member. A side-advantage is the
increase of the moment capacity of the profile. The moment capacity now becomes 247
kNm. The eccentricity that has become even smaller because of the deeper profile, now
does not give any capacity problems anymore. By only using a deeper profile for the bottom
chord member (and top chord member) the connections are much easier to fabricate.

Table 1. Comparison profile capacities.


Npl [kN] Vu;d [kN]
Maximum Mpl [kNm]
Profile bottom in diagonal with in bottom chord
allowable N in Bottom chord
chord member buckling length of member without
diagonal [kN] member
3.6 m (HE200A) stiffeners

HE200A 860 280 337 101

HE240B 860 582 700 247

Solution 3: Supplementary web plates

If the first two solutions are not possible, web plates can also be chosen. The shear force
capacity of the profile will then increase. The truss built up from HE200A-profiles is taken as
an example again.

Increasing the web thickness of the HE200A into 20 mm (original 6.5 mm), the extra moment
as result of eccentricities can be absorbed without problems, as a result of the eccentricity.
With the available analysis tools it is simple to check if the selected profile complies.

Figure 9. Web plates.

The search for alternative detailing afterwards can be avoided totally, if the engineer in
his/her main design takes account of the fact that he tries to design a simple constructible
structure. In other words: a structure that not only is optimum according to his own criteria,
but is also perfect for the steel fabricator as well as the client. Nobody, not even the client, is
waiting for a design that has an optimum weight, but has complicated details that make the
steel structure unnecessarily expensive.

32 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


For comparison, a calculation has been made of the truss including detailing, whereby as
starting point the connection with members joined in the centre with complex modelling has
been taken.

Table 2. (-- very complex, - complex, + simple, ++ very simple) (2).


Simplicity Welding/
Erection Material use Total costs
Connection Production
Starting point: join members in
-- 100 % 100 % 100 % 100 %
centre, complex details

Solution 1: Rotate 90 degrees - 99 % 117 % 51 % 87 %

Solution 2: Deeper bottom


++ 78 % 128 % 63 % 58 %
member
Solution 3: Web plate(s) on
+ 84 % 97 % 74 % 72 %
bottom chord member

The conclusion is that all three outlined solutions show a considerable reduction of the total
costs in comparison to the original detail. Despite an increase of using material, due to
solutions 1 and 2, the connection has become a lot less complex in comparison with the
original connection design.

DESIGN STAGES

By going through the design stage of a construction once again, there will be indicated how
such problems can be prevented.

The client and the architect set up a list of requirements for the building that has to be
constructed. In the light of this list in combination with the building regulations and the
building code a certain material or a combination of materials to execute the design will be
selected. Every material has its own features. Timber and especially concrete are suitable to
withstand compressive forces, whilst steel is preferably suitable to withstand tensile forces.

The choice of material defines the shape of the main structure it is important that the right
material is used on the right spot.
Points for attention must be:
• Is there taken care of the stability of the structure by means of rigid plates, portals or
bracings
• Are high vertical forces transferred, by trusses with compression-, tension or
compression and tension members
• Is there any concentration of forces, as is the case with structures with large spans.
• How are forces lead in into the structure below.

In daily practice it often seems as if during the main design no account is taken at all of the
fact that the members must be joined.

When the engineer has made the first calculation and thus has an idea about the profiles, he
has to think about how the several profiles must be joined. By choosing for example a deeper
(=stronger) profile a lot of (detailing) sorrow can be prevented. An engineer also can insert
the eccentricities in his calculation and base profiling there-upon. But this is not the usual
procedure for calculations of steel constructions; moreover it gives a considerable increase
of the shear forces and moments in the top chord member and bottom chord member. Also
separating the nodes is a rather laborious for the constructor.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 33


See below an example of a truss and the moment- and shear force diagrams.

3 2

Figure 10a. Model without eccentricities.

3 2 3 2

Figure 10b. Moment diagrams Figure 10c. Shear force diagrams


(Centre lines intersect). (Centre lines intersect).

3 2
excentriciteit is 100 mm

Figure 11a. Summary of truss with eccentric


connections of the diagonals, eccentricity = 100 mm.

3 2 3 2

Figure 11b. Moment diagrams Figure 11c. Shear force diagrams


(Eccentricities). (Eccentricities).

These diagrams show clearly that the maximum moments and especially the shear forces
increase considerably by introducing eccentricities.

A design engineer, who thinks before designing about the way the beams and colums have
to be connected, enables the steel fabricator to construct an efficient and constructible
structure with the use of simple connections. This results in benefit for all parties concerned.

CONCLUSION

Nowadays engineers still tend to optimize a structure to a minimum of kilograms, the only
hard criterion that is available for an engineer during the design of the structure. Finally this
criterion leads to structures that are more expensive and have less quality. Working together
in a design and built-team in which the fabricator also has to assist in the design stage can
prevent these problems.

34 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


This design and built-team is not always possible, that is why a design engineer has to be
aware that his design also has to be constructed. He has to think whether or not simple
connections are possible with the profiling that is being determined with a calculation
program. He has to start a discussion with the client why sometimes there has to be chosen
for a heavier profile with some overcapacity and why the aim to a minimum weight has to be
left. If, moreover, a 3D finite element program is used, considerable material saving can be
achieved, because of the insight in the spatial distribution of forces.

REFERENCES

(1) H.G.A. Evers, F. Maatje, ICCS bv, Cost based engineering and production of steel
constructions, Proceedings of the Fourth International Workshop, Connections in
steel structures IV, Behavior, strength and design, published 2002, Actas do II
Entrocontro Nacional de Construção Metálica e Mista 2, published 1999.
(2) F. Maatje, ECCS bv, Voorcalculeren met behulp van de computer (Cost estimation
with the computer), Bouwen met Staal 1989

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 35


36 Connections in Steel Structures V - Amsterdam - June 3-4, 2004
CONTINUING EDUCATION IN STRUCTURAL CONNECTIONS
František Wald, Czech Technical University in Prague, Czech Republic
David Moore, Building Research Establishment, United Kingdom
Milan Veljkovic, Luleå University of Technology, Sweden
Martina Eliášová, Czech Technical University in Prague, Czech Republic

ABSTRACT

The European project Continuing Education in Structural Connections


(CeStruCo) under Leonardo da Vinci initiative No. CZ/00/B/F/PP-134049 was
prepared by partners from seven European countries to disseminate the latest
results in research and standardization during the period of transferring the
European Pre-Standard into the European Standard. The project has started
by a collection of questions from the European practice. The answers to those
questions have been prepared in the form of textbooks in the project partners
national languages. The material is available as an easily accessible
Internet/CD (www.fsv.cvut.cz/cestruco) media, and includes video and audio
files, slides and worked examples.

INTRODUCTION

Education has always been seen as an essential part of the introduction and dissemination
of new methods for the design of steel connections. One of the first educational packages
specialised on connections was produced by Owens and Cheal (1) who prepared
educational material for structural connections. This material has been extended and is now
incorporated into a European educational package called the European Steel Design
Educational Programme, ESDEP (2). This programme is used today by educational
establishments throughout Europe. Other educational packages which build on the work of
ESDEP are available some of which include WIVISS (3) and (4), a set of lectures on CD,
SteelCall (5) and (6), a virtual steel designers office, and SSEDTA which consists of a set of

a) b)

Figure 1. Screens from collection of questions on project Internet page – a) in partners


languages; b) format of collection.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 37


basic lectures on PowerPoint for the design of steel and composite elements, see (7). The
project NFATEC - A New and Flexible Approach to Training for Engineers in Construction
(8), is transferring the successful lessons of SSEDTA into the Internet based lectures and
worked examples (9).

For more than twenty year the European Convention for Constructional Steelwork’s
Technical committee for structural connections (ECCS TC10) has supported the
development and implementation of a common set of design rules for steel connections. It is
therefore not surprising to find that one of this committee’s priorities is to facilitate the
transition of EN1993-1-8 from a European pre-standard; see (10), (11) and (12), to a full
Euro-norm (13). Part of this activity is the development of the necessary educational material
to encourage designers throughout Europe to adopt EN1993-1-8. Consequently, a
programme called “Continuing Education in Structural Connections” (CESTRUCO) was
formed under the European Commission’s Leonardo initiative to collect commonly asked
questions on the background, implementation and use of prEN1993-1-8 and to publish
expert answers to these questions. The CESTRUCTO project was developed from an idea
by Mr. Marc Braham (Astron, Luxembourg), Prof. Jan Stark (TU Delft, The Netherlands) and
Mr. Jouko Kouhi (VTT, Finland) to provide designers with more detailed information on the
background and implementation of the design methods given in prEN1993-1-8.

Figure 2. Project textbooks in project partner’s languages.

38 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


The project partners were Aristotle University of Thessaloniki, Greece; Bouwen met Staal,
Netherlands; Building Research Establishment Ltd., United Kingdom; Czech Constructional
Steelwork Association Ostrava, Czech Republic; Czech Technical University (contractor),
Czech Republic; EXCON a.s., Prague, Czech Republic; KREKON Design office, Rotterdam,
Netherlands; Luleå University of Technology, Sweden; University of Coimbra, Portugal;
Politechnica University of Timisoara, Romania. The project team consist of the following
people: Prof. C. C. Baniotopoulos (in charge of Chapters on Welding and on Aluminium);
Prof. F. S. K. Bijlaard; Ir. R. Blok (internal review), Mr. J. Brekelmans; Prof. L. S. da Silva
(Chapter on Fire design); Prof. D. Dubina (Chapter on Seismic design); Mrs. M. Eliášová;
Mr. H. G. A. Evers (Chapter on Good and bad detailing); Mr. D. Grecea (Chapters on Hollow
section connections and on Cold-formed connections); Ir. A. M. Gresnigt (Chapter on
Moment connections); Dr. V. Janata (internal review); Prof. B. Johansson; Mr. T. Leino;
Mr. T. Lennon; Mr. T. Měřínský (internal review); Dr. D. B. Moore (editor and Chapter on
Simple connections); Mrs. A. Santiago; Mr. R. L. Schipholt; Dr. Z. Sokol (Chapter on
Structural modelling, Windows help Internet lessons); Ir. C. M. Steenhuis; Dr. M. Veljkovic
(Chapter on Bolts); Prof. F. Wald (project promoter, editor and Chapter on Column bases).
The material was reviewed externally by Prof. D. Beg, Mr. M. Braham, Prof. J. P. Jaspart,
Dr. G. Huber, Mr. J. Kouhi, Prof. F. Mazzolani, Mr. A. J. Rathbone, Prof. J. Studnička,
Dr. F. Turcic, Dr. K. Weynand and Mr. N. F. Yeomans.

PROJECT WORKED PACKAGES

Collection of questions

The collection was based on publications in national journals and on local seminars. The
questions were collected by project internet page, Figure 1, and in paper form as well. The
questions located during the conversion of ENV 1993-1-1 (including Annexes J (11), K (12),
L (10)) into EN 1993-1-8 were taken into account. Together 632 questions were collected
related to the topic.

Answering

All obtained questions were very good. Answers to 101 selected questions were chosen in
the second part of work based on its educative contribution. The agreement between all
partners was not reached for all answers. This is the reason that some nice questions are
missing in final material. The review of answers was prepared in tree steps: locally between
partners, by delegated partners and externally by members of ECCS TC10.

Dissemination

Under the third part of the project, dissemination, were prepared the educational materials
Textbooks and Internet/CD lectures. To facilitate easy of use questions/answers are split into
the following Chapters: Introduction, Bolts, Welding, Structural Modelling, Simple
Connections, Moment Resistance, Connections, Column Bases, Seismic Design, Fire
Design, Hollow Section Joints, Cold-Formed Member Joints, Aluminium Connections, and
Design Cases. Each chapter starts with a brief over-view of the method use in prEN1993-1-
8. This is followed by the commonly asked questions together with their answers. The
materials were localised for use in the partner’s national languages, Figure 2: Czech (300
copies of textbook was printed), Dutch (200 copies), English (1500), Greek (200),
Portuguese (1000), Rumanian (300) and Swedish (200). The translations into Polish and
Spanish are under preparation (written in February 2004). The pilot Seminars of the project
were organised in project partner’s countries to test the material in local environment and to
start to disseminate the material.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 39


a) b)

c) d)

Figure 3. Example from Textbook Chapter 13 Design cases - a), c) bad solutions of beam to
column joint, b), d) good cases.

2,29°
0,04 rad

a) b)

Figure 4. Example from Textbook Chapter 13 - Design cases a) bad solution of column
base, b) good design.

40 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


The pilot Internet Seminar was included in the project. It showed the ability to use in an
efficient way the prepared educational material. The pilot Internet seminar was broadcasted
to seven project partners from Luleå University. Very robust software MARRATECH,
Figure 5 was used to establish an interactive communication between partners. The easy to
use and simple to managed tool consists of windows showing the view of camera from each
partner, a white board and a text messenger. The quality of connection was indicated at each
partner window. By clicking on camera window the particular partner’s picture is enlarged to
allow high quality projection. The white board allows all partners to draw by mouse and to
share with other participants in the session available material such as figures, slides, videos
and files (in MS PowerPoint and Word format). The text message window was very reliable
and helped partners to begin communication. The weak point of the Seminar was audio
quality which was strongly influenced by the internet traffic. This is an important issue since it
is essential part of the lecturing. We all have learned how sensitive is the audio quality,
especially in a case of simultaneous broadcasting to various partners. A communication
between two partners, using so called node to node communication is more robust and
easier to establish. However, our laboratories, well equipped by computers and cameras,
have to be upgraded with adequately equipment for video broadcasting. All the partners were
pleasantly surprised with the performance of the communication tool under routine Internet
connection.

a) b)

Figure 5. Window of MARRATECH software, a) white board with Microsoft PowerPoint


presentation and text connection; b) from preparation of Internet Seminar.

INTERNET / CD LESSONS

The Internet/CD version, Figure 6, of project materials is based on Microsoft Windows help
format, which is a robust tool for education. The material prepared by RoboHelp tool (14)
allows the easy navigation in partner’s languages and in German and French. The
Internet/CD lessons are equipped with worked examples, Figures 3 and 4, PowerPoint
presentations, slides, videos, worked examples (15) and (16), animations of design cases,
educational software for connection modelling, and design tools available round the Europe.
Partners were so kind to equipped us with their demo version to show how easy may be nice
connection checked by EN1993-1-8. 3000 copies of CD first version were printed at the end
of the project in December 2003. The second version was disseminated in June 2004 based
on evaluation of first ones. Third upgraded version is intended to be published based of
interest of structural steel practice in June 2005.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 41


a) b)

c) d)

Figure 6. Lesson in Microsoft Windows – a) Title page, b) example of Chapter Bolts,


c) example of Chapter Column bases, d) example of Chapter Aluminium connections.

On the CD may be found the NASCon (Non-linear Analysis of Steel Connections) program
that offers a computer user-friendly tool for the component method. The tool allows modelling
the nonlinear behaviour of different components; see (17) Figure 7. The video film, see
Figure 8, demonstrates the correct design of T-stub connections and bolted splices to avoid
a fatigue failure of bolts. The related to the fire design is equipped with the PowerPoint
lessons including video/audio sequences, see Figure 9.

a) b)

a) menu of program b) component behaviour


Figure 7. Program nonlinear analysis of steel connections NASCon (17).

42 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


a) b)

Figure 8. The flow of the stresses in the connection in video film Statically Stressed Bolts
in Dynamically Loaded Connections.

a) b)
Figure 9. a) Front page of PowerPoint lesson ”Connection Design for Fire Safety”,
b) a slide from the lesson “Heating and Cooling of Structure”.

ACKNOWLEDGEMENT

The authors wish to dedicate this work to Mr. Martin Steenhuis, our good friend, who worked
with us in the field of structural connections for many years, launched this project, and who
tragically died in the summer of 2001.

REFERENCES

(1) Owens G. W., Cheal B. D.: Structural Steelwork Connections, Butterworths, London
1988, ISBN 0-408-01214-5.
(2) ESDEP, European Steel Educational Programme, SCI, London, 1994, www.esdep.org.
(3) Plank R.J.: Wider Vocational Initiative for Structural Steelwork, J. Construct. Steel
Research, 46, (1998), pp 278-279, ISBN 0-08-042997-1.
(4) Chladná M. - Wald F. - Burgess I. W. - Plank R. J.: Contribution of the Structural
Steelwork Educational Programme WIVISS, v Proceedings of the Conference
Eurosteel ´99, Studnička J., Wald F., Macháček J. ed., Vol. 2, Prague,
26 - 29 May 1999, ČVUT Praha, 1999, s. 137 - 140, ISBN 80-01-01963-2.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 43


(5) Roszykiewicz C.,: Breakthrough in Electronic Education for Steel Construction, SCI,
London, 1997.
(6) SteelCal, Electronic Education for Steel Construction, steelcal.steel-sci.org.
(7) SSEDTA, Structural Steelwork Eurocodes: Development of Trans-national Approach,
fp.emberey.plus.com.
(8) NFATEC, A New and Flexible Approach to Training for Engineers in Construction,
http://www.svf.stuba.sk/kat/KDK/NFATEC_E.htm
(9) Muzeau J. P. - Bourrier P., Education for Constructional Steel design: APK - the French
Example, J. Construct. Steel Research, 46, (1998), pp 259-271, ISBN 0-08-042997-1
(10) ENV 1993-1-1: Design of Steel Structures, Eurocode 3, European Pre-Standard, CEN,
Brussels 1992.
(11) ENV 1993-1-1-A2: Design of Steel Structures, Annex J, Joint Design, European Pre-
Standard, CEN, Brussels 1998.
(12) ENV 1993-1-1/A1: Design of Steel Structures, Annex K, Joints in hollow section
structures, European Pre-Standard, CEN, Brussels 1994.
(13) prEN 1993-1-8. Eurocode 3: Design of Steel Structures, Part 1.8: Design of Joints,
European Standard, CEN, Brussels 2003.
(14) RoboHelp office, http://www.ehelp.com.
(15) Jaspart J.P., Renkin S., Guillaume M.L.: European Recommendations for the Design of
Simple Joints in Steel Structures, 1st draft of a forthcoming publication of the Technical
Committee 10 “Joints and Connections” of the European Convention of Constructional
Steelwork (ECCS TC10) prepared at the University of Liège, September 2003.
(16) Wald F. at al: Steel structures 10, Worked examples to Eurocode 3, Czech Technical
University in Prague, Prague 2001, p. 146, ISBN 80-01-02308-7.
(17) Costa Borges L. A.: Probabilistic Evaluation of the Rotation Capacity of Steel Joints
Dissertação apresentada à Faculdade de Ciências e Tecnologia para obtenção do grau
de Mestre em Engenharia Civil, especialidade de Estruturas, Coimbra, 2003.

44 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


DOCKING SOLUTION BETWEEN A STEEL TRUSS AND A
CONCRETE TOWER AT THE SKI JUMP IN INNSBRUCK

C. Aste, aste konstruktion, Innsbruck, Austria


A. Glatzl, aste konstruktion, Innsbruck, Austria
G. Huber, aste konstruktion, Innsbruck, Austria

ABSTRACT
The crucial challenge in MBT (Mixed Building Technology) are the connection
elements between the different construction systems. A typical example of an
MBT realisation is the ski jump tower in Innsbruck where a steel cage of hollow
section trusses is cantilevering up to 12,5 m from the central concrete tower in
a height of about 35 m above ground. The transfer of the localised truss forces
into the concrete box section was solved by special pre-stressed docking
devices.

INTRODUCTION

The ski jump in Innsbruck known for the famous annual New Year ”Four-ski-jump-tour“ was
renewed. The original jumping tower (built for the 1976 Olympic winter games) was pulled
down and a new landmark similar to a lighthouse was erected. Zaha Hadid (London) won the
international architectural competition for this significant building. The constructional
realisation was ordered from aste konstruktion and was honoured with the “Austrian State
Award for Consulting 2002”. A speciality of this MBT are the high tension docking devices
between the cantilevering steel cage and the concrete tower.

Figure 1. Panoramic view of “Bergisel” with the new jumping tower.

PROJECT OVERVIEW

“Bergisel” - a glacier hill located to the south of Innsbruck (Figure 1) – has a significant history:
Offering place of the Celts, path marking for the roman-german emperors moving to Rome,
battle field of the Tyrolean war of liberation, centre of the Olympic Winter Games in 1964 and
1976 with the two flame basins and – already since more than 50 years - scene of the
international Four-ski-jump-tour at New Year. In the years 2001 and 2002 these facilities were
fully renewed and extended by new buildings (Figure 2, 3):

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 45


• Concrete tower and cantilevering tower top: described in detail in the following paper
sections.
• Approach ramp bridge: sagging fish-bellied steel truss girder with suspension over a
direct span of 69 m, bend with a radius of 100 m, inclination of 35°, overall length
including the take-off building 98 m, approach lane of trapezoidal composite slabs,
erection with a temporary pier within four weeks.
• Take-off building and ski-jump platform: concrete abutment looking like the knee of a
ski jumper, length 24 m, fix support of the approach ramp bridge.
• Front building: three-storey concrete building with bent roof below the jump platform –
flown-over by the jumpers, technical equipment, power supply, common rooms.
• Landing hill: concrete fixing and border bead or retaining wall, transverse ribs, holding
devices for the snow nets and the plastic mats for summer jumping, mat sprinklers.
• Reporter cabin tower: four storeys for 31 cabins, steel tube frame.
• Coach platform: grate platform close to the take-off building, steel tube frame.
• Funicular railway: automatic cabin inclination corresponding to the slope.
• Judge tower: redevelopment of the old timber construction.

Figure 2. Project overview.

Office / Company City / Country Function Competence


Bergisel Management Assoc. Innsbruck / A client
Hadid London / UK architect ski jump incl. tower
aste konstruktion Innsbruck / A design office design calculations and detailing
Pichler Bozen / I steel construction approach ramp and tower top
IMO-Bau Leipzig / D sub-steel constr. erection of ramp and tower top
Vorspann-Technik Oberndorf / A bridge equipment pre-stressing, ramp suspension

Figure 3. Construction board (tower).

46 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


CONCRETE TOWER

With 49 m above ground the ridge height of the concrete tower reaches 791 m above sea
level (Figure 4). The foundation was solved with a plate of 20 x 20 x 1,0 m at a level of –11 m
below ground with three basement storeys. The standard cross section of 7 x 7 m and a wall
thickness of 40 cm rises up for about 40 m above the foundation, stabilised with wall
diaphragms to the base plate limits. It contains the two elevators, the stairway and the supply
pit. From 29 m above ground the cross section tapers to 3,7 x 7 m making place for the
jumping access stairway.

Also at this level of cross section change the support girder for the ramp bridge cantilevers
4,5 m with a height of only 1,45 m. This slenderness was necessary to hide this girder in
between the steel truss flanges of the bridge.

The demand of fair-faced concrete in combination with the difficult access and supply
conditions resulted in the choice of a climbing formwork. Concreting started in June 2001.

Figure 4. Concrete tower.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 47


STEEL-COMPOSITE TOWER TOP

The tower top is not ordinary - neither in architectural nor in constructional respect. A three-
level steel cap with a rescue level, a restaurant and an observation platform is docked to the
central concrete tower (Figure 5). Being 250 m above the city centre one has a fantastic view
on Innsbruck and the surrounding mountains.

The levels are cantilevering around the concrete core up to 12,5 m. Together with the steel
hollow section frames and the diagonal suspension tubes anchoring back to the concrete
core a steel cage is built (Figure 6). The horizontal stiffening to the core is realised by the
trapezoidal composite slabs. The transparency and elegance of the facade is supported by
the fact that diagonal bars within the front could be avoided and huge glass elements were
placed into the facade.

Figure 5. Tower top – an architectural challenge.

Figure 6. Tower top – sketch of the steelwork construction.

48 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


DOCKING DEVICES

The crucial challenge in MBT construction are the connection elements between the different
construction systems. Thanks to the common, material-independent design and safety
concept of the constructional Eurocodes the interface problems at the level of design
methods and internal forces lost its deterrent effect and MBT solutions become more and
more usual in daily design practice. The effect is a more economical use of different
materials related to their constructional benefits (strength, stiffness, weight, prefabrication,
strengthening, dismanteling,...) and more innovative architectural solutions.

For the actual case of the Bergisel Ski Jump the considerable docking forces between the
steel cap and the concrete core had been a crucial challenge which was solved by special
pre-stressed steel brackets (Figure 7). These elements of at maximum 550 kg weight were
integrated into the formwork with a tolerance of less than 1 cm.

tension
compression
prestress

Figure 7. Pre-stressed docking devices. Figure 8. Characteristic docking forces.

The load transfer into the concrete walls was handled with mutual pre-stressing cables
(interior tendons) from one docking point to the opposite one going through the tower.
Additional concentrated rebars in the local load introduction zones were provided to cover
the bursting forces and for crack distribution. The characteristic docking forces can be taken
from Figure 8.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 49


Pre-stressing was applied from the opposite side of the fixed anchor after hardening of the
concrete and before connecting the steel sections. Depending on the tension forces either
one or two strands were provided. The conduits were then filled with injection grout against
corrosion.

The resulting necessary welding length led to the geometry of these docking brackets. By the
use of four longitudinal ribs which were welded on site to the slotted push-over hollow
sections the total bracket length could be minimised. The eventual negative influence of the
high welding temperatures on the end anchorage of the pre-stressing strands could be
dispelled by a test specimen. The maximum heat increase was measured to be only 50
degrees.

Attention has to be paid to the fact that the tendon head is no more accessible after
positioning of the steel cage. Therefore this application type is limited to quasi-static loading.

Figure 9. Slotted hollow sections welded to the docking brackets.

Figure 10. Cantilevering steel cage during construction.

CONCLUSIONS

The new building at Bergisel proved to be an excellent combination of architectural shape


and constructional design. Fair-faced concrete, steel and glass together with the harmonious
longitudinal section and the top view are showing the worldwide appreciated style of Zaha
Hadid. Construction and erection were based on modern steel-concrete mixed building
technology: concrete core with climbing formwork, pre-stressed steel docking brackets for
the steel frame cage on the tower top, a pre-stressed very slender concrete cantilever as
upper support of the approach ramp, three-level widely cantilevering steel frame cage on top,
approach ramp in the form of an organic fish-bellied and suspended trough bridge – all in all
“Toccata and Fugue in major F” for a civil engineer and his orchestra.

50 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Figure 11. Illuminated structure at night

REFERENCES

(1) Aste, C., Glatzl, A., Huber, G. (2002). Ski jump „Bergisel“ – A new landmark of
Innsbruck. 3rd Eurosteel Conference, ISBN 972-98376-3-5, Lisbon.
(2) Aste, C., Glatzl, A., Huber, G. (2002). Schisprungschanze „Bergisel“ – Ein neues
Wahrzeichen von Innsbruck. Stahlbau Heft 3/02, S.171-177, ISSN 0038-9145,
Ernst&Sohn, Berlin.
(3) Aste, C., Glatzl, A., Huber, G. (2003). Steel-concrete mixed building technology at the
ski jump tower of Innsbruck, Austria. International Journal on Steel and Composite
Structures, ISSN 1229-9367, Technopress, Korea.
(4) Aste, C., Glatzl, A., Huber, G. (2003). Innsbruck ski jump: a triumph of mixed building
technology. Concrete Journal, The Concrete Society, Berkshire.
(5) Aste, C., Glatzl, A., Huber, G. (2003). Schisprungschanze „Bergisel“ – Ein neues
Wahrzeichen von Innsbruck. Zement + Beton, www.zement.at, Wien.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 51


52 Connections in Steel Structures V - Amsterdam - June 3-4, 2004
CHARACTERIZATION OF THE NONLINEAR BEHAVIOUR
OF SINGLE BOLTED T-STUB CONNECTIONS

Ana M. Girão Coelho, Polytechnic Institute of Coimbra, Portugal


Luís Simões da Silva, University of Coimbra, Portugal
Frans S.K. Bijlaard, Delft University of Technology, The Netherlands

ABSTRACT
A systematic analytical procedure for characterization of the monotonic defor-
mation behaviour of single T-stub connections is presented and discussed. The
model is based on the prying mechanism implemented in Eurocode 3 and in-
cludes the deformations from tension bolt elongation and bending of the T-
element flange. The results of a series of tests (numerical and experimental) on
T-stubs that were conducted at the Delft University of Technology and the Uni-
versity of Coimbra are used to validate the procedures.

INTRODUCTION

The T-stub model is widely accepted as a simplified model for the idealization of the behav-
iour of the tension zone of bolted joints, which is often the most important source of deform-
ability of the whole joint (figure 1a) (1, 2, 3). The equivalent T-stub corresponds to two T-
shaped elements connected through the flanges by means of one or more bolt rows. The
system is loaded by applying a tensile force at the top of the webs that induces deformation
of the flanges (figure 1b). The load-carrying behaviour of this simple assembly is inherently
nonlinear and hence its characterization is not easily open to simple analytical formulations.
Tests (both experimental and numerical) are frequently carried out to obtain the actual re-
sponse, which is then modelled approximately by mathematical expressions that relate the
main properties: initial stiffness, ke.0, post-limit stiffness, kp-l.0, full plastic resistance, FRd.0, ul-
timate resistance, Fu.0 and deformation capacity, ∆u.0.

From a practical point of view, the characterization of the T-stub behaviour by means of ei-
ther of the above methods is not expedited. A simple methodology for prediction of the con-
nection response up to collapse is desired. The collapse is governed by fracture of the bolts
and/or cracking of the T-stub material. Because of the emphasis placed on connection ductil-
ity, the methodology must be able to predict the response of the T-stub well into its plastic
and strain hardening range with a reasonable degree of accuracy.

This paper presents simplified methods for determining the monotonic deformation response
of T-stub connections. First, existing models from the literature are discussed. Next, a two-
dimensional model is proposed and calibrated against test results from a database compiled
by the authors (4, 5). Some recommendations for modifications are also given. Finally, con-
clusions are drawn.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 53


F
2

T-stub T-stub (end


(column side) plate side)

F
2

F ⎛∆⎞
2 ⎜⎝ 2 ⎟⎠
(C) (D)

(A)
B
Q
(a) T-stub identification in the particular case of an ex- (b) T-stub basic prying mechanism
tended end plate bolted connection (3). (one quarter-model).

Figure 1. Equivalent T-stub model of the tension zone of bolted joints.

PREVIOUS RESEARCH

The European code of practice for the design of structural steel joints in buildings, Eurocode
3 (6), provides design rules for the evaluation of ke.0 and FRd.0, based on elastic (3, 7, 8) and
pure plastic theories (1, 2), respectively. The post-limit stiffness, kp-l.0 is taken as zero, which
means that strain hardening and geometric nonlinear effects are neglected. Concerning the
component ductility, Eurocode 3 (6) presents some qualitative principles based on the main
contributions of the T-stub deformation: whenever the bending deformation of the flanges
governs the plastic mechanism, the ductility is infinite; should the bolt determine collapse, the
ductility is limited. These rules are, yet, insufficient. In fact, if simple plastic theory is applied,
the limit resistance and deformation are determined by yielding of the flange. Therefore, con-
sideration of strain hardening is crucial to carry out an ultimate analysis of the system up to a
fracture condition.

To fill in this gap, several theoretical approaches for the characterization of the overall behav-
iour of T-stubs have already been proposed in the literature (7, 9, 10). Essentially, they use
the same basic prying mechanism, which is also the model implemented in Eurocode 3 (6)
(figure 1b). The model is two-dimensional, i.e. the three-dimensional effects are not ac-
counted for. The system is statically indeterminate to the first degree. It is loaded by applying
a vertical force F/2 to the support (A), which is the critical section at the flange-to-web con-
nection. F is the applied force per bolt row. Only one quarter-model is taken into account due
to symmetry considerations. The contact points at the tip of the flange are modelled with a
pinned support and reproduce the effect of prying forces, Q. The T-stub flange behaves as a
rectangular cross-section of width beff and depth equal to tf. Such width, beff, represents the
flange plate width tributary to a bolt row that contributes to load transmission at pure plastic
conditions. This width accounts for all possible yield line mechanisms of the T-stub flange (1,
6) and cannot exceed the actual flange width, b. Despite these major simplifications, the

54 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


nonlinear analysis of this model is still complex and requires an incremental procedure. Thus,
it is not intended for hand computation unless some simplifications that reduce the model
complexity to a reasonable level are assumed.

Jaspart approximates the nonlinear T-stub behaviour with a bilinear response (7, 8). The
characteristics of this bilinear behaviour are summarized as follows:
(i) The initial elastic region has a slope ke.o that is evaluated as in Eurocode 3 (6).
(ii) The swivel point in the bilinear relationship represents the full development of the yield
lines and the correponding force is FRd.0. This plastic resistance is also computed
according to the Eurocode 3 procedures (6).
(iii) In the plastic region, above the swivel point, the effects of material strain hardening are
dominant. The slope of this second linear region is then given by (7, 8):
k p − l .0 = Eh E ke.0 (1)

whereby E: Young modulus of the flange material and Eh: strain hardening modulus.
(iv) The point of maximum force, Fu.0, is determined by formally equivalent expressions to
FRd.0, by replacing the plastic conditions (index Rd) with ultimate conditions (index u).
This means that these expressions are based on the same geometric characteristics but
the plastic moment of the flange, Mf.Rd, is replaced with:
Mf .u = 0.25tf 2fu.f beff (2)

(fu.f: ultimate stress of the flange material) and the bolt strength, Bu, is:
Bu = fu.b As (3)

being fu.b: ultimate stress of the flange material and As: tensile stress area of a bolt. The
deformation capacity is readily determined by intersection of the plastic region, with
slope kp-l.0, with the maximum resistance, Fu.0, i.e.:
∆u.0 = FRd .0 ke.0 + ( Fu.0 − FRd .0 ) k p − l .0 (4)

b b b

F1.Rd.0 F2.Rd.0 F3.Rd.0

Q Q Q Q
B B BRd BRd
BRd BRd
(=F1.Rd.0/2+Q)

Mf.Rd Mf.Rd ξMf.Rd


Mf.Rd ξMf.Rd

n m m n n m m n n m m n

(a) Flange plastic mechanism (b) Combined bolt/flange (c) Bolt mechanism (type-3;
(type-1). mechanism (type-2; ξ ≤ 1). ξ ≤ 1).

Figure 2. Collapse mechanism typologies of a single T-stub connection at plastic conditions


and distribution of internal actions (1, 6, 9).

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 55


Piluso et al. developed an analytical procedure based on the resemblance of the distribution
of internal forces at plastic and ultimate conditions (9). Figure 2 shows the plastic conditions
of a bolted T-stub for the three possible failure modes and the corresponding internal forces.
Their model disregards compatibility requirements between bolt and flange deformation and
assumes that cracking of the material occurs when the ultimate strain in the extreme fibres of
the T-stub flanges is attained. The plastic deformation of the flange is computed from integra-
tion of the moment-curvature diagram, which is obtained from simple internal equilibrium
conditions of the section and by assuming that the material constitutive law can be approxi-
mated by a quadrilinear relationship (in natural coordinates). This simple model yields a
quadrilinear approximation of the force-deformation (F-∆) curve. The ordinates of this curve
are determined according to the potential failure mode (figure 2). In particular, for the evalua-
tion of the force coordinates, they use formally equivalent expressions to the Eurocode 3 (6)
for plastic conditions.

0.90Bu Bolt fracture


0.85Bu
Plastic,
Yielding, 0.02Kb
0.1Kb

Elastic, Kb

δb
δb.1 δb.2 δb.fract

Figure 3. Bolt force-deformation model according to Swanson (10).

Swanson developed a prying model that includes: (i) nonlinear material properties, (ii) a vari-
able bolt stiffness that captures the changing behaviour of the bolts as a function of the loads
they are subjected to, (iii) partially plastic hinges in the flange and (iv) second order mem-
brane behaviour of thin flanges (10). The bolt behaviour is incorporated by means of an ex-
tensional spring located at the inside edge of the bolt shank. This spring is characterized by
a piecewise linear force-deformation, B-δb, response (figure 3). The bolt elastic stiffness, Kb,
is given by:
Eb
Kb = (5)
Ls Ab + Ltg As

being Eb: Young modulus of the bolt material, Ls: bolt shank length, Ab: nominal area of the
bolt shank and Ltg: bolt threaded length included in the grip. The bolt fracture deformation,
δb.fract, is computed as follows:

0.90Bu Ls ⎛ 2 ⎞
δ b.fract = + ε u.b ⎜ Ltg + ⎟ (6)
Ab Eb ⎝ nth ⎠

whereby εu.b: ultimate bolt strain and nth: number of threads per unit length of the bolt. The
flange mechanistic model assumes an elastic-yielding-plastic (with strain hardening) constitu-
tive relationship for the steel. Plastic hinges will develop at the flange-to-web connection and
at the bolt axis. Swanson derived the stiffness coefficients and corresponding prying gradi-
ents, qij,k = ∂Q/∂∆, by using the direct stiffness method. Both parameters are used in an in-
cremental solution technique. First, the initial stiffness and prying gradient are determined.
Next, several checks are made to determine which limit is reached first (bolt force or flange

56 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


internal stresses limits). Incremental deformations are then calculated for each of the poten-
tial limits with the smallest value governing. The F-∆ curve can yield up to nine linear
branches, with different stiffness, before failure. Swanson states that the strength and defor-
mation capacity of the flange are not always predicted accurately because of sensitivities of
the model to strain hardening parameters and bolt ductility (10). It should be stressed that
this model was not intended for hand calculations.

PROPOSED MODEL

The models described above afford some basis for the development of a model for the
evaluation of the deformation capacity and the load-carrying capacity of single T-stub con-
nections. The mechanical model is similar to that depicted in figure 1b and is represented in
figure 4. A model that uses geometrical and mechanical properties consistent with the Euro-
code 3 (6) prying model is desired so as to make implementation easier.

Fracture conditions

The two possible ultimate (fracture) conditions are: (i) fracture of the bolt and (ii) cracking of
material of the flange near the web, i.e. at section (A) (figure 4). Normally, M(C) ≤ M(A) (section
(C) is the flange section at the bolt line, figure 4). In the context of a two-dimensional model,
whereby the flange is modelled as a rectangular cross-section, the latter condition may be
too severe.

0.8r
r

F
2
(C)

(A*)
(A) B-δb

m n

⎧⎪r → HR T-stubs
m = d − 0.8 ⎨
⎪⎩ 2aw → WP T-stubs

n = min (1.25m; e )

Figure 4. Proposed T-stub model.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 57


This section (A) is defined at a distance m from the bolt axis, where the flange thickness is
higher than tf owing to the fillet weld or profile root that provide some extra material thick-
ness. Therefore, the imposition of cracking of the material should also be checked at the end
of this fillet (section (A*), hereafter), i.e. at a distance m* = d - r or m* = d - 2 aw for T-stubs
made up of hot rolled profiles (HR T-stubs) and T-stubs made up of welded plates, WP T-
stubs, respectively (d: distance between the bolt axis and the face of the T-stub web; r: fillet
radius of the flange-to-web connection; aw: throat thickness of a fillet weld).

Bolt deformation behaviour

The bolt elongation response is based on the Swanson’s proposals (figure 3). For the com-
putation of δb.fract, Eq. (6), however, the parameter nth is disregarded.

With reference to this equation, it was developed for short-threaded bolts (10). If a full-
threaded bolt is considered instead, this expression seems to overestimate the bolt fracture
deformation. Therefore, in this case, the evaluation of δb.fract by application of the above ex-
pression should be cautious.

Flange constitutive law

The flange material constitutive law is modelled by means of a piecewise stress-strain rela-
tionship that accounts for the strain hardening effects. This law is a true stress-logarithmic
strain relationship, i.e. it is defined in natural coordinates in order to capture the actual mate-
rial behaviour. Piluso et al., in fact, adopt the same approach in (9) since the prediction of the
plastic deformation capacity of compact sections is more accurate if natural stress-strain co-
ordinates are used. From a design point of view, the constitutive law can be idealized by
means of multilinear models that reproduce well the material strain hardening and fracture
ranges. Gioncu and Mazzolani (11) present alternative models for the stress-strain curve of
structural steels.

The above model is not proposed for a hand calculation. Instead, a numerical finite element
(FE) method is used to determine the structural response. Consequently, the piecewise con-
stitutive law may contain several branches. Figure 5 shows the stress-strain relationship im-
plemented for three different steel grades that will be referred to later in the text (4, 5). The
FE code Lusas (12) implements a beam element that belongs to the Kirchoff beams group
(with quadrilateral cross-section). Shear deformations are excluded in this element formula-
tion.

900
S690
True stress (MPa)

750
S355 (1)
600
S355 (2)
450

300

150

0
0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35
Logarithmic strain
Figure 5. True stress-logarithmic strain characteristics for different steel grades.

58 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Analysis of the model in the elastic domain

First, the proposed model is analysed and validated in the elastic domain by using the speci-
mens from the authors’ database (4, 5, 13) (table 1). The initial stiffness of a T-stub connec-
tion is evaluated and compared with the actual predictions, which correspond to experimental
or (three-dimensional) numerical values. It was found out that the model yielded a very stiff
response in the elastic domain if the actual Young modulus of the flange was employed (13).
It was also verified that in this domain, disregarding the shear deformability of the flange
would introduce an error of 15% in the predictions of the initial stiffness. Therefore, a reduc-
tion of the Young modulus, Ered, was proposed in order to make the necessary corrections
(13). Such value was defined as follows:
2 2
2⎛m⎞ ⎡ 3 ⎤ E ⎛m⎞ ⎡ 3 ⎤
Ered = 0.5E × ⎜ ⎟ ⎢ 1 + − 1⎥ = ⎜ ⎟ ⎢ 1 + − 1⎥ (7)
( m tf ) ⎥⎦ 3 ⎝ tf ⎠ ⎢⎣ ( m tf ) ⎥⎦
2 2
3 ⎝ tf ⎠ ⎢

so that the shear deformability was taken into account (13, 14). This reduction does not have
much influence at ultimate conditions. In fact, the effect of shear on the moment resistance of
the flange is apparently beneficial (15).

Table 1. Geometrical and mechanical characteristics of the specimens (Fl: flange; Bt: bolt;
Bts: Bolt stripping; Flange steel grade, fy.f: S355 (1), regular values; S355 (2), underlined val-
ues; S690: values in bold).
Test ID Fracture ele- Geometrical characteristics Mechanical characteristics
ment
Ac- Model beff tf m n φ fy.f εp.u.f fu.b δb.fract Kb
tual pred. (mm) (mm) (mm) (mm) (mm) (MPa) (MPa) (mm) (N/mm)
T1 Bt Bt 40.0 10.7 29.45 30.00 12 431.0 0.284 974.0 0.97 6.92×105
P1 Bt Bt 40.0 10.7 34.45 25.00 12 431.0 0.284 974.0 0.97 6.92×105
P2 Bt Bt 40.0 10.7 24.45 30.56 12 431.0 0.284 974.0 0.97 6.92×105
P3 Bt Fl (A*) 35.0 10.7 29.45 30.00 12 431.0 0.284 974.0 0.97 6.92×105
P4 Bt Bt 52.5 10.7 29.45 30.00 12 431.0 0.284 974.0 0.97 6.92×105
P5 Bt Bt 60.0 10.7 29.45 30.00 12 431.0 0.284 974.0 0.97 6.92×105
P6 Bt Bt 35.0 10.7 29.45 30.00 12 431.0 0.284 974.0 0.97 6.92×105
P7 Bt Bt 45.0 10.7 29.45 30.00 12 431.0 0.284 974.0 0.97 6.92×105
P8 Bt Bt 40.0 11.0 27.10 33.88 12 431.0 0.284 974.0 0.98 6.78×105
P9 Bt Bt 40.0 14.0 28.75 35.94 12 431.0 0.284 974.0 1.36 5.45×105
P10 Fl Fl (A*) 40.0 7.0 30.75 30.00 12 431.0 0.284 974.0 0.75 9.37×105
P11 Bt Bt 40.0 14.0 28.75 30.00 12 431.0 0.284 974.0 1.20 5.57×105
P12 Fl Fl (A*) 40.0 10.7 29.45 30.00 16 431.0 0.284 974.0 1.01 1.19×106
P13 Bt Fl (A*) 40.0 10.7 29.45 30.00 12 355.0 0.284 974.0 0.97 6.92×105
P14 Fl Fl (A*) 40.0 10.7 29.45 30.00 12 275.0 0.284 974.0 0.97 6.92×105
P15 Fl Fl (A*) 40.0 10.7 24.45 30.56 16 431.0 0.284 974.0 1.01 1.19×106
P16 Bt Bt 70.0 10.7 29.45 30.00 12 431.0 0.284 974.0 0.97 6.52×105
P17 Bt Bt; fl (A*) 70.0 10.7 29.45 30.00 16 431.0 0.284 974.0 1.09 1.11×106
P18 Fl Fl (A*) 70.0 10.7 29.45 30.00 20 431.0 0.284 974.0 1.09 1.73×106
P19 Bt Bt; fl (A*) 70.0 10.7 29.45 30.00 16 431.0 0.284 974.0 1.09 1.11×106
P20 Bt Bt 70.0 14.0 28.75 30.00 16 431.0 0.284 974.0 1.37 9.54×105
P21 Bt Fl (2) 92.5 10.7 29.45 30.00 20 431.0 0.284 974.0 1.37 1.49×106
P22 Bt Bt 70.0 15.0 32.34 30.00 20 431.0 0.284 974.0 1.52 1.40×106
P23 Bt Bt 70.0 18.9 31.59 30.00 20 431.0 0.284 974.0 1.98 1.14×106
Weld_T1(i) Bt Fl (A*) 40.0 10.7 37.43 30.00 12 431.0 0.284 974.0 0.97 6.92×105
Weld_T1(ii) Bt Fl (A*) 40.0 10.7 33.42 30.00 12 431.0 0.284 974.0 0.97 6.92×105
Weld_T1(iii) Bt Bt 40.0 10.7 30.14 30.00 12 431.0 0.284 974.0 0.97 6.92×105

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 59


Table 1. Continued
WT1 Bt; fl Fl (A*) 45.1 10.3 33.73 30.00 12 340.1 0.361 919.9 1.14 1.10×106
WT2A Bt; fl Fl (A*) 45.0 10.3 36.29 29.90 12 340.1 0.361 919.9 1.14 1.10×106
WT2B Bt; fl Fl (A*) 45.0 10.3 31.69 29.90 12 340.1 0.361 919.9 1.14 1.10×106
WT4A Bt Bt 74.9 10.4 33.69 30.00 12 340.1 0.361 919.9 1.14 1.10×106
WT51 Bt Fl (A*) 45.0 10.0 34.39 30.20 12 698.6 0.174 919.9 1.14 1.10×106
WT53C Bt Fl (A*) 45.1 10.1 34.34 30.00 12 698.6 0.174 968.4 4.00 9.14×105
WT53D Bt Fl (A*) 45.0 10.1 34.24 30.00 12 698.6 0.174 1166.0 0.98 1.08×106
WT53E Bt Fl (A*) 44.7 10.1 34.26 30.10 12 698.6 0.174 1196.4 2.80 9.15×105
WT7_M12 Bt Bt 75.6 10.3 33.87 29.90 12 340.1 0.361 919.9 1.14 1.10×106
WT7_M16 Fl Fl (A*) 74.9 10.3 33.89 29.80 16 340.1 0.361 919.9 2.60 1.65×106
WT7_M20 Fl Fl (A*) 75.2 10.3 33.81 29.70 20 340.1 0.361 919.9 2.60 2.57×106
WT57_M12 Bt Fl (A*) 75.0 10.1 34.11 30.20 12 698.6 0.174 919.9 4.00 9.14×105
WT57_M16 Bts Fl (A*) 75.3 10.2 34.26 30.10 16 698.6 0.174 919.9 2.60 1.65×106
WT57_M20 Bt; fl Fl (A*) 75.1 10.2 34.27 30.20 20 698.6 0.174 919.9 2.60 2.57×106

From the comparison of the model predictions for the above specimens and the actual initial
stiffness values, it was observed that the stiffness prediction was sufficiently accurate leading
to an average overestimation equal to 18% and to a coefficient of variation equal to 0.16 in
the case of HR T-stubs. For WP T-stubs, the results are further improved: estimation error of
1% and coefficient of variation of 0.18 (13).

Analysis of the model in the elastoplastic domain and comparison with the other
models

The proposed model for the F-∆ curve is compared with the actual curve in figure 6 for some
representative specimens. The comparison with experimental and numerical (three-
dimensional model) evidence (13) shows a good agreement of results in terms of stiffness
and resistance (approximate errors: underestimation of Fu.0 in 15%, for those specimens
whose cracking of the flange determines collapse and overestimation of circa 10% if collapse
is governed by bolt fracture). In terms of ductility, the model predicts the deformation capacity
accurately if the cracking of the flange is critical (see figures 6c,6d,6e). (For specimen WT1
(figure 6d), however, the bolt also fractured in the tests.) The exceptions in this case are the
specimens made up of steel grade S690. Apparently, if the cracking condition is imposed as
the attainment of εu at section (A*), ∆u.0 is clearly underestimated (figure 6h). For those
specimens whose plastic collapse is of type-2 and eventually the bolt fractures (specimen
P16, for instance – figure 6b), the prediction of ∆u.0 is good (average overestimation of 9%
with a coefficient of variation of 11%). At last, for those specimens exhibiting plastic mecha-
nism type-1 and fracture of the bolt at ultimate conditions, the agreement between actual and
predicted values is not that satisfactory (predicted values nearly twice the actual values) (fig-
ure 6a).

Figure 6g illustrates the above mentioned setback with the bolt model applied to full-threaded
bolts. In specimen WT57_M12 the flange plates are fastened by means of two full-threaded
bolts. At collapse, the bolt model estimates a fracture elongation of 4 mm (table 1), which
seems rather high. Since bolt governs fracture of this specimen, the post-limit behaviour pro-
ceeds until this deformation of 4 mm is attained, leading to an overall deformation 8.1 mm
and ultimate resistance of 174.5 kN, corresponding to 1.43 times the maximum resistance
from the tests.

60 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


140 140
120 120
Load, F (kN)

Load, F (kN)
100 100
80 80
60 Actual response 60 Actual response
Simplified response (Beam model) Simplified response (Beam model)
40 Jaspart approximation (actual Eh) 40 Jaspart approximation (actual Eh)
Jaspart approximation (nominal Eh) Jaspart approximation (nominal Eh)
20 Piluso et al. approximation
20 Piluso et al. approximation
0 0
0 2 4 6 8 10 12 14 16 18 20 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0
Deformation, ∆ (mm) Deformation, ∆ (mm)
(a) Specimen T1. (b) Specimen P16.
280 105
240 90

Load, F (kN)
Load, F (kN)

200 75
160 60
120 45 Actual response
Actual response Simplified response (Beam model)
80 Simplified response (Beam model) 30
Jaspart approximation (actual Eh) Jaspart approximation (actual Eh)
40 Jaspart approximation (nominal Eh) 15
Piluso et al. approximation Jaspart approximation (nominal Eh)
0 0
0 3 6 9 12 15 18 21 24 27 30 0 2 4 6 8 10 12 14 16 18 20
Deformation, ∆ (mm) Deformation, ∆ (mm)
(c) Specimen P18. (d) Specimen WT1.
160 180
140
150
120
Load, F (kN)

Load, F (kN)

120
100
80 90
Actual response Actual response
60 Simplified response (Beam model)
Simplified response (Beam model) 60
40 Jaspart approximation (actual Eh)
Jaspart approximation (actual Eh) Jaspart approximation (nominal Eh)
30
20 Jaspart approximation (nominal Eh) Piluso et al. approximation
0 0
0 2 4 6 8 10 12 14 16 18 20 0 2 4 6 8 10 12 14 16 18 20
Deformation, ∆ (mm) Deformation, ∆ (mm)
(e) Specimen WT7_M16. (f) Specimen WT7_M20.
180 280

150 240
Load, F (kN)

Load, F (kN)

200
120
160
90
Actual response 120 Actual response
60
Simplified response (Beam model) 80 Simplified response (Beam model)
30 40
Jaspart approximation (actual Eh) Jaspart approximation (actual Eh)
0 0
0 1 2 3 4 5 6 7 8 9 10 0 2 4 6 8 10 12 14 16 18 20 22 24
Deformation, ∆ (mm) Deformation, ∆ (mm)
(g) Specimen WT57_M12. (h) Specimen WT57_M20.

Figure 6. Prediction of the force-deformation response of some selected specimens.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 61


In the above graphs, the predictions from Jaspart (7) and Piluso et al. (9) are also included.
Generally speaking, the model from Piluso et al. does not provide an accurate modelling of
the F-∆ response. As for the bilinear approximation proposed by Jaspart, it reproduces well
the actual behaviour for those specimens made up of S690 (figure 6g, for instance; for
specimen WT57_M20, figure 6h, the predictions are particularly poor). For the remaining
cases the predictions are fine provided that the nominal value of the strain hardening
modulus (Eh.nom = E/48.2, (12)) is used. If the actual value of Eh is used instead, then the pre-
dictions are not so good.

Sophistication of the method: modelling of the bolt action as a distributed load

Jaspart has shown that a significant increase in the resistance of single T-stubs that fail ac-
cording to a plastic mechanism type-1 (figure 2a) can be expected due to the influence of the
bolt action on a finite contact area (7). This effect is also taken into account in the proposed
methodology, as a modification of the original model. The bolt is then modelled as a spring
assembly in parallel, as indicated in figure 7. The length of this assembly is the bolt diameter.
The behaviour of this spring assembly is the same as the original single spring, i.e. the spring
stiffness and force values are divided by the number of springs in the assembly.

The application of this modified model provides significant enhancement of results in terms of
resistance, particularly for the evaluation of Fu.0 and ∆u.0, as well. Figure 8 compares the two
modelling approaches in terms of overall behaviour for some of the former specimens.

0.8r
r

φ
F
2
(C)

(A*)
(A) B-δb

m n
Figure 7. T-stub model accounting for the bolt action on a finite contact area.

CONCLUSIONS

The proposed beam model yields an accurate prediction of the F-∆ response of bolted T-stub
connections, despite the simplifications inherent to a two-dimensional modelling of the be-
haviour. In the model, the dominant effects are the strain hardening of the flange and the bolt
elongation behaviour. The behaviour predicted by this model is rather good in terms of resis-
tance. With respect to ductility, it reflects an overestimation of test results that is within an
acceptable error (13).

62 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


140 105
120 90

Load, F (kN)
Load, F (kN)

100 75

80 60
Actual response Actual response
60 45
Simplified response (simple Simplified response (simple
40 30 beam model)
beam model)
Simplified response accounting 15 Simplified response accounting
20 for the bolt action for the bolt action
0 0
0 2 4 6 8 10 12 14 16 18 20 0 2 4 6 8 10 12 14 16 18 20
Deformation, ∆ (mm) Deformation, ∆ (mm)

(a) Specimen T1. (b) Specimen WT1.


160 280
140 240
120

Load, F (kN)
Load, F (kN)

200
100
Actual response 160
80 Actual response
120
60 Simplified response (simple
Simplified response (simple
beam model) 80
40 Simplified response accounting beam model)
Simplified response accounting
20 for the bolt action 40
for the bolt action
0 0
0 2 4 6 8 10 12 14 16 18 20 0 2 4 6 8 10 12 14 16 18 20 22 24
Deformation, ∆ (mm) Deformation, ∆ (mm)
(c) Specimen WT7_M16. (d) Specimen WT57_M20.

Figure 8. Comparison of the force-deformation response for the simple proposed model and
with the inclusion of the modelling of the bolt.

The applicability of the model was well demonstrated within the range of examples shown in
this paper. The modification for inclusion of the bolt action provides an enhancement of re-
sults but introduces an additional complexity. From a design point of view, the methodology
should be further simplified so that it can be used in an expedite way, as the Jaspart’s simple
proposal. This can be achieved by modelling plasticity phenomena in the flange by means of
rotational springs at the critical sections that capture the overall behaviour. Current work on
this particular subject is being carried out by the authors.

ACKNOWLEDGEMENTS

Financial support from the Portuguese Ministry of Science and Higher Education (Ministério
da Ciência e Ensino Superior) under contract grants from PRODEP and FCT (Grant
SFRH/BD/5125/2001) for Ana M. Girão Coelho is gratefully acknowledged.

REFERENCES

(1) Zoetemeijer P, (1974). A design method for the tension side of statically loaded bolted
beam-to-column connections. Heron; 20(1):1-59.
(2) Packer JA, Morris LJ, (1977). A limit state design method for the tension region of
bolted beam-to-column connections. The Structural Engineer; 55(10):446-458.
(3) Yee YL, Melchers RE, (1986). Moment-rotation curves for bolted connections. Journal

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 63


of Structural Engineering ASCE; 112(3):615-635.
(4) Girão Coelho AM, Bijlaard F, Gresnigt N, Simões da Silva L, (2004). Experimental as-
sessment of the behaviour of bolted T-stub connections made up of welded plates.
Journal of Constructional Steel Research; 60:269-311.
(5) Girão Coelho AM, Simões da Silva L, Bijlaard F, (200?). Finite Element Modelling of
the Nonlinear Behaviour of Bolted T-Stub Connections. Journal of Structural Engineer-
ing ASCE (submitted for publication 2003).
(6) prEN 1993-1-8:2003, (2003). Eurocode 3: Design of steel structures. Part 1.8: Design
of joints. Stage 49 draft, May 2003, European Committee for Standardization, Brussels.
(7) Jaspart JP, (1991). Etude de la semi-rigidite des noeuds poutre-colonne et son influ-
ence sur la resistance et la stabilite des ossatures en acier. PhD thesis (in French),
University of Liège, Liège.
(8) Jaspart JP, (1997). Contributions to recent advances in the field of steel joints – col-
umn bases and further configurations for beam-to-column joints and beam splices. Ag-
gregation thesis, University of Liège, Liège.
(9) Piluso V, Faella C, Rizzano G, (2001). Ultimate behavior of bolted T-stubs. I: theoreti-
cal model. Journal of Structural Engineering ASCE, 127(6):686-693.
(10) Swanson JA, (1999). Characterization of the strength, stiffness and ductility behavior of
T-stub connections. PhD dissertation, Georgia Institute of Technology, Atlanta, USA.
(11) Gioncu V, Mazzolani FM, (2002). Ductility of seismic resistant steel structures. Spon
Press, London, UK.
(12) Lusas 13 (2003). Theory manual. Finite element analysis Ltd, Version 13.5. Surrey,
UK.
(13) Girão Coelho AM, (2004). Characterization of the ductility of bolted extended end plate
beam-to-column steel connections. PhD dissertation, University of Coimbra, Coimbra,
Portugal (in preparation).
(14) Faella C, Piluso V, Rizzano G, (2000). Structural Semi-Rigid Connections – Theory,
Design and Software. CRC Press, USA.
(15) McGuire W, (1968). Steel structures. Prentice-Hall International series in Theoretical
and Applied Mechanics (Eds.: NM Newmark, WJ Hall). Englewood Cliffs, N.J., USA.

64 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


EXPERIMENTAL BEHAVIOUR OF ALUMINIUM T-STUB
CONNECTIONS

G. De Matteis, University of Naples Federico II, Italy


G. Della Corte, University of Naples Federico II, Italy
A. Mandara, Second University of Naples, Italy
F.M. Mazzolani, University of Naples Federico II, Italy

ABSTRACT
In the current paper the experimental results related to welded aluminium alloy
T-stub components under quasi-static monotonic and cyclic loading are
provided. The experimental study refers to 26 different welded specimens. In
particular, 2 boundary conditions, 4 T-stub geometry (varying in-plane
dimensions, plate thickness, number and location of bolts), 3 aluminium alloys
and 3 types of bolt have been considered. The obtained results allow the
influence of the above parameters on the observed failure plastic mechanism
as well as on the deformation and dissipative capacity of the examined joint
type to be pointed out.

INTRODUCTION

The correct prediction of joint structural response is an important issue in metal structures (1).
For aluminium structures, current standards provide rules to estimate the influence of joints on
the global structural behaviour through the joint classification (2), but the application of such
methods is related to the correct evaluation of the main joint mechanical properties. On the other
hand, it is recognized that the behaviour of connections is very complex, because it includes
many local effects and the system is highly indeterminate, giving rise to complicate stress
redistribution among the basic components. Therefore, for a reliable understanding of
connection behaviour and setting up of relevant calculation methods the direct experimentation
is required. While several laboratory experiments have been carried out for steel, allowing some
consistent calculation procedures to be set up (EC3 – Part 1.8), only few are concerned with
aluminium alloys (3). On the basis of the above considerations, a large research project dealing
with aluminium joints has been carried out at the University of Naples Federico II, sponsored by
the Italian Ministry of Education, University and Research (MIUR). It was mainly concerned with
the behaviour of bolted T-stub joint components, which has been deeply examined under
analytical, numerical and experimental points of view (4).

T-stub is one of the more important joint components. It may be regarded both as stand-
alone connection (5) and as a part of several, more complex bolted joint configurations (6). In
previous papers, the authors have already proposed an analytical method to determine both
the collapse mechanism type and the corresponding load bearing capacity of aluminium T-
stubs based on the existing formulations adopted into EC3-Part 1.8 (7). It is important to
remind that for steel T-stubs, three possible failure modes are recognized: type-1
mechanism, with complete yielding of the flange through the onset of four plastic hinges, two
of which are located at bolt axis location; type-2 mechanism, when the bolt failure takes
place together with the yielding of the flange at the sections corresponding to flange-to-web
connection; type-3 mechanism, due to bolt failure with the overall up-lift of the flange. Due to

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 65


the limited ductility of the material, for aluminium T-stubs, the intermediate failure mechanism
(type 2) has to be split into three different mechanisms: the first one (type 2a) is
characterized by the failure of the flange material with plastic engagement of bolts; the third
one (type 2c) is distinguished by the failure of bolts and plastic engagement of the flange
material; the second mechanism (type 2b) can be considered as a ideal limit case between
mechanisms type 2a and type 2c, where the failure is attained in both flange and bolts at the
same time. Besides, in order to account for the hardening of the material, a special
procedure has been defined for the evaluation of plastic strength of aluminium T-stub flange
sections and aluminium bolts.

The above method has been firstly calibrated on the basis of numerical FEM analyses only
(8). Then, monotonic loading tests have been carried out for several aluminium T-stub
specimens (9), allowing the above analytical model to be checked and re-calibrated (10).
Finally, in the present paper, experimental results of the same specimens previously tested
under monotonic loading are analysed in comparison with results obtained under cyclic
loading (11), so to evaluate the dissipative capacity of the considered specimens as well as
the relevant failure mechanisms under repeated deformations.

SPECIMEN FABRICATION AND TEST PROGRAM

The experimental study, which has been carried out at the Laboratory of the Department of
Structural Analysis and Design at the University of Naples Federico II, concerns the
behaviour of isolated aluminium T-stubs having different geometrical configurations and
material properties of the basic components. Both monotonic and cyclic loading tests have
been carried out. The test specimens were fabricated starting from T-shaped profiles, which
were previously pre-assembled by means of welding made at flange-to-web connection. The
basic plate elements constituting the isolated T-stub specimens are made of three different
wrought aluminium alloys (namely, AW 6061, AW 6082 and AW 7020), having different
thickness (namely, 10 mm, 12 mm and 12 mm, respectively), which were artificially aged
through the application of the T6 heat treatment. Nominal values of the main mechanical
properties of the above alloys are summarised in Table 1. All welds between plate elements
were previously made using standard automatic MIG welding process and were carried out
by the commercial fabricator ALCOA-Italia of Novara.

Table 1. Nominal properties of selected materials (according to EC9).


T-stub flange material Bolt material
Alloy f0.2 [Mpa] fu [Mpa] Class fy (f0.2 ) [Mpa] fu [Mpa]
AW 6061-T6 240 290 Steel 4.8 320 400
AW 6082-T6 255 300 AA 7075 (440) 510
AW 7020-T6 280 350 Steel 10.9 900 1000

By combining available flange and web plate elements, 12 different T-stub configurations
have been obtained (Fig. 1), corresponding to 4 different geometrical configurations (varying
flange in-plane dimensions and thickness) and, for each shape, to several flange-to-bolt
material combinations. In particular, in Figure 1, letters A, B and C refer to the three different
materials used for the T-stub flanges. Besides, it has to be noted that labels 'Type 2' and
'Type 3' have been applied to specimens having the same geometry for allowing the
execution of tests with two different coupling conditions. T-joints were made by a bottom Y
partial penetration weld, connecting the bottom part of the two flange elements to each other
and to the web plate element, and by two complementary top fillet welds (5 mm throat
section), connecting the upper part of the flanges to the web. This welding procedure has
been selected in order to reduce the out-of-straightness of flanges. Only for specimen Type-

66 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


1A, T-joints were made by a simple flange-to-web top fillet weld, with a continuous flange
plate.

It is worth noting that normal engineering procedures were adopted during fabrication and
assembly of specimens. In particular, in all cases, according to current practice, bolts were
preloaded to a torque corresponding to an initial tensile stress in the bolt equal to the 80% of
the bolt nominal yield resistance, the latter being evaluated according to EC9. Furthermore,
no special effort was taken to get perfect fit-up of mill cutting during specimen preparation
and to centralize the bolts in the bolt holes during specimen assemblage.

TYPE 1A TYPE 1B TYPE 1C


TYPE 1 12 12 12
142
AW 7020-T6 AW 7020-T6 AW 7020-T6
30 35 12 65

135

135

135
22 36 22

35 30 35 30 35 30
80

AW 6061-T6 AW 6082-T6 AW 7020-T6


10

12

12
Ø 11 mm 142 142 142

TYPE 2A-3A TYPE 2B-3B TYPE 2C-3C


TYPE 2-3
12 12 12
112
(2C) AW 7020-T6
20 30 12 50 AW 7020-T6 AW 7020-T6
(3C) AW 6082-T6

135
135

135
22 36 22
80

30 20 30 20 30 20
AW 6061-T6 AW 6082-T6 AW 7020-T6
10

12
12

Ø 11 mm 112 112 112

TYPE 4 TYPE 5
10 10
122
128
28 28 10 56 AW 6061-T6 AW 6061-T6
24 35 10 59
130
124

35
60

35 24
70

AW 6061-T6 28 28 AW 6061-T6
120

35

10
10
60

Ø 11 mm
122 128

Ø 11 mm

Figure 1. Nominal data for tested specimens.

The performed experimental tests are related to two different types of T-stub assemblage
(Fig. 2). In particular, in some cases, two isolated aluminium T-stub specimens are
assembled (coupled T-stub arrangement), while, in some other cases, the isolated T-stub
specimen is joined to a purposely de-signed steel support having negligible deformability
(single T-stub arrangement). The former configuration is typical of end-plate beam splice
joints, where the connected elements have the same flexural stiff-ness. Conversely, the latter
configuration is representative of joints where the connected plate elements are different
from each other, such as in the case of beam-to-column joints. Furthermore, in order to
investigate the influence of bolts on the stiffness as well as the failure mechanism of
aluminium T-stubs, different locations, number and types of bolts have been considered,
taking into account the possibility to use both aluminium bolts (type AA 7075) and steel bolts
(grade 4.8 and 10.9). In all the cases, the bolt diameter is 10 mm while the hole diameter is

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 67


11 mm. Moreover, both one-bolt row and two-bolt row configurations have been considered.
Altogether, 5 different series of T-stub specimens have been devised, with 4 different
geometry (see Figure 1), which together with the other selected influencing parameters
(flange material, bolt material and type of assemblage) give rise to a total number of 26
different specimen configurations tested under both monotonic and cyclic loading (4).

LVDT 3 LVDT 5 LVDT 1

T-stub flange
T-stub web

LVDT 4 LVDT 2
LVDT 6
a) b) c)

Figure 2. Adopted instrumentation for displacement measuring – a) general lay-out; b) single


T-stub assemblage; c) coupled T-stub assemblage.

Tests on selected specimens were performed by a Material Test System (MTS 810) 500 kN
universal testing machine, under displacement control, using displacement rate equal to
0.005 mm/s in the first elastic loading and 0.05 mm/s in the remaining part of the loading
process. For cyclic tests, the Complete Testing Procedure provided by ECCS
Recommendations (1986) has been used. For this purpose, the conventional limiting elastic
displacement used for the definition of the loading history of each sample has been
established on the basis of the results obtained from monotonic tests (9).

Tie forces were applied to the webs of the connected T-stub elements, which were tightened
by the jaws of the testing machine for a length of 50 mm, having the axial force applied
collinear with the midpoint of the connection. During the test, the applied load was
automatically measured by the testing machine, while displacements were measured by
using displacement transducers having accuracy of 0.01 mm. In particular, according to
Figure 2a, on each side of the joint, a 500 LVDT was used to measure the flange-to-flange
relative displacement at the centreline of the connection (LVDTs 5 and 6) while two 200
LVDTs were used to measure the flange-to-flange relative displacements at the bolt locations
(LVDTs 1, 2, 3 and 4). Both plate material and bolt tensile tests have been carried out to
characterize the mechanical behaviour (4, 9). For each alloy, average values of material
properties are reported in Table 2, while obtained results for bolts are given in Table 3. In
such a table, Steel 10.9 bolts are not considered, because they are never involved in
collapse mechanism of the specimen.

Table 2. Measured mechanical properties for flange materials (average values).


Bolt location – unaffected Web-to-flange connection – affected material
Type of material (Longitudinal test bars) (Transversal test bars)
specimen Type 1-2 A Type 1-2B Type 1-2C Type 1A Type 2-3A Type 1-2-3B Type 1-2-3C
(AW 6061) (AW 6082) (AW 7020) (AW 6061) (AW 6061) (AW 6082) (AW 7020)
f0.2 [Mpa] 294 310 285 196 190 168 186
fu [Mpa] 327 335 328 264 241 233 261
εu [%] 7 5.5 9.3 4.2 1.9 2.7 2.1

68 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Table 3. Measured mechanical properties for bolts.
Mean Standard
Bolt class Test no. 1 Test no. 2 Test no. 3 M-STD
value (M) dev. (STD)
f0.2 [MPa] 438 441 428 436 7 429
AW 7075
fu [MPa] 690 678 536 635 86 549
fy [MPa] 505 390 540 478 78 400
Steel 4.8
fu [MPa] 589 629 844 687 137 550

TEST RESULTS

General

In the following, for both monotonic and cyclic loading tests, the obtained results are provided
in terms of applied force (F) versus relative displacement (∆) curves. The relative
displacement (∆) between T-stub elements has been evaluated at the connection centreline,
as average of the two measuring points. In particular, in Figures 3-4, results are provided for
the single T-stub arrangement, while in Figure 5, the results related to the coupled T-stub
arrangement are presented. In addition, in Table 4 for each test specimen, the main
measured and determined characteristics of F-∆ curves are summarized, together with the
revealed types of failure mechanism. It is worth mentioning that for each T-stub type, the
sequential numbering for samples refers to the specimen identification stated in De Matteis
et al. (4), where, the actual geometrical properties together with the measured main
imperfection parameters (flange out-of-straightness and uplift) for each specimen are
supplied. In the above table, the peak strength (Fmax), the corresponding displacement level
(∆max) and the ultimate displacement (∆ult), the latter evaluated either as the value
corresponding to a sudden drop off of the bearing capacity of the specimen or as the value at
the end of the test in case the relevant failure was gradual and not clearly identified, are
given. Also, for monotonic tests, the conventional elastic strength (Fel), evaluated according
to ECCS Recommendations (1986), and the elastic displacement (∆el) associated to Fel are
specified.

The observed behaviour

The examination of Table 4 reveals that tested specimens gave rise to all the possible failure
mechanisms, showing important differences in terms of strength and deformation capacity
due to T-stub geometry, bolt type and flange material. In particular, as it could be expected,
the T-stub geometry has an important influence on the failure mechanism: T-stub Type 1 and
Type 2-3 present weak-flange failure mechanism, while T-stub Type 4 and T-stub Type 5 are
characterised by a weak-bolt failure mechanism. Furthermore, as a rule, the specimens
endowed with steel grade 10.9 bolts provide an ultimate strength remarkably higher than the
other cases. On other hand, specimens with bolt type AA 7075 and steel grade 4.8 present
similar strength, even though the former are generally characterized by a lower value of the
ultimate deformation capacity. On the other hand, the influence of the flange material on the
overall response of the examined structural component has to be related to the relevant
collapse mode, it appearing more important for specimens exhibiting a weak-flange failure
mechanism (failure mechanism type 1 and type 2a).

The influence of applied bolt

A deeper examination of results evidences that there is a strict dependence between the
failure mode and the type of applied bolt. In fact, all specimen type 1 and 2-3 equipped with
steel bolts grade 10.9 exhibited a type 1 failure mechanism, with limited deformation

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 69


occurring in the bolts and a notable plastic engagement of flange material (Figure 6a). In
particular, for such specimens, the collapse was due to fracture of the welded web-to-flange
connection, with progressive development of a cracking initiated at the weld toe and
propagating in the internal part of the throat section. As a consequence, the structural
component appears to be rather ductile, with a gradual strength and stiffness degradation at
collapse. Conversely, all specimen type 1 and 2-3 equipped with steel bolts class 4.8
exhibited a type 2a failure mechanism, with the final rupture of the bolts, but with significant
plastic deformation developed at the web-to-flange connection with visible cracks at the weld
toe (Figure 6b). Finally, all specimen type 1 and 2-3 equipped with aluminium bolts AA 7075
evidenced a type 2c failure mechanism, with a collapse mode ruled by the tensile failure of
bolts, with limited plastic deformations occurring in the flange and without any cracking in the
welded connection (Figure 6c).

Table 4. Main T-stub properties and results derived by tests.


Specimen Monotonic tests Cyclic tests
Flange
material Bolt type Fail. Fmax Fel ∆el ∆max ∆ult Fail. Fmax ∆max ∆ult
Mon. cyclic
mech. [kN] [kN] [mm] [mm] [mm] mech. [kN] [mm] [mm]
1A1 1A4 AW-6061 Steel 4.8 2a 77.7 26 0.09 3.69 12.11 2a 70.7 1.5 2.9
1A2 1A5 AW-6061 AA 7075 2c 79.8 32 0.08 2.68 2.68 2c 80.4 1.8 1.8
1A3 1A6 AW-6061 Steel 10.9 1 89.5 59 0.11 1.40 5.52 1 84.1 0.8 3.8
1B1 1B4 AW-6082 Steel 4.8 2a 90.6 44 0.16 2.56 6.97 2a 53.9 1.1 2.5
1B2 1B5 AW-6082 AA 7075 2c 101.4 50 0.06 1.48 1.48 2c 77.7 1.1 2.3
1B3 1B6 AW-6082 Steel 10.9 1 132.5 59 0.12 2.71 3.25 1 120.2 1.9 1.9
Single T-stub assemblage

1C1 1C4 AW-7020 Steel 4.8 2a 88.4 65 0.38 2.11 7.34 2a 81.9 3.4 4.0
1C2 1C5 AW-7020 AA 7075 2c 91.6 42 0.15 0.96 0.96 2c 93.5 1.2 1.2
1C3 1C6 AW-7020 Steel 10.9 1 145.2 64 0.13 2.46 2.77 1 132.2 1.6 2.6
2A2 3A2 AW-6061 AA 7075 2c 87.8 43 0.15 1.17 1.17 2c 72.0 1.1 1.1
2A3 3A3 AW-6061 Steel 10.9 1 118.7 67 0.10 1.94 2.38 1 94.1 1.1 1.9
2B2 3B2 AW-6082 AA 7075 2c 105.2 52 0.08 1.08 1.08 2c 90.6 1.2 1.2
2B3 3B3 AW-6082 Steel 10.9 1 154.2 78 0.09 1.43 1.66 1 117.4 1.1 3.1
2C2 3C2 AW-7020 AA 7075 2c 102.9 49 0.05 0.85 0.86 2c 82.0 1.5 1.5
2C3 3C3 AW-7020 Steel 10.9 1 153.2 78 0.14 2.08 2.24 1 128.2 1.2 2.5
4-1 4-4 AW-6061 Steel 4.8 3 44.9 33 0.49 1.47 2.93 3 47.1 1.6 4.0
4-2 4-5 AW-6061 AA 7075 3 50.5 23 0.09 0.63 0.63 3 56.4 1.4 1.4
4-3 4-6 AW-6061 Steel 10.9 2b 110.7 79 0.69 3.90 4.54 2b 103.9 3.5 5.4
5-1 5-3 AW-6061 Steel 4.8 2c 52.9 38 0.04 1.97 7.01 2c 46.8 2.4 5.0
5-2 5-4 AW-6061 AA 7075 2c 52.3 29 0.03 0.90 0.91 2c 66.3 1.8 1.8
2A4 + 3A4 + AW-6061 AA 7075 2c 88.2 72 0.21 1.56 1.56 2c 79.4 1.5 1.5
2A6 3A6
Coupled T-stub assemblage

2A1 + 3A1 + AW-6061 Steel 10.9 1 119.4 54 0.15 3.32 3.67 1 98.8 2.1 3.2
2A5 3A5
2B4 + 3B1 + AW-6082 AA 7075 2c 92 50 0.24 2.22 2.24 2c 85.7 2.2 2.2
2B5 3B5
2B1 + 3B4 + AW-6082 Steel 10.9 1 147.8 88 0.24 2.44 2.57 1 102.4 1.6 2.9
2B6 3B6
2C4 + 3C4 + AW-7020 AA 7075 2c 89.9 44 0.12 1.56 1.56 2c 95.0 1.8 1.8
2C5 3C6
2C1 + 3C1 + AW-7020 Steel 10.9 1 137 66 0.18 2.72 2.97 1 114.9 1.5 2.8
2C6 3C5

Obviously, the above circumstances has an effect also on the dissipative capacity of the
joint; in particular, specimens equipped with steel bolts 4.8 show a larger residual plastic
deformation and amount of energy dissipation per cycle with respect to the same specimens
equipped with aluminium bolts 7075. It is also important to observe that specimens endowed

70 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


with steel bolts 4.8 provided even larger ductility and energy dissipative capacity than the
ones endowed with steel bolts 10.9, which exhibit a type-1 collapse mechanism.

The difference between aluminium AA 7075 and steel bolt class 4.8 is also evident by
analyzing curves related to T-stub type 4 and type 5, the collapse mechanism being related
essentially to the bolt failure (Fig. 6d). On the other hand, it is evident that the response of
specimens with steel bolts is better than the one of specimens with aluminium bolts, the latter
exhibiting significant strain hardening but a limited ductility and dissipative capacity. Finally, it
is important to observe that the combination of T-stub type 4 with steel bolt class 10.9 gave
rise to an unexpected failure mode, which was characterized by significant yielding of the
flanges with yield lines developed in a similar way as in the non-circular pattern provided by
EC3-Part 1.8 but with final rupture of bolts under tension.

Figure 3. F-∆ curves for T-stub Type 1 and 2 (single T-stub assemblage).

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 71


Figure 4. F-∆ curves for T-stub Type 4 and 5 (single T-stub assemblage).

Figure 5. F-∆ curves for coupled T-stub assemblage specimens.

The influence of assemblage condition

In case of T-stub type 2-3, it is also possible to investigate the effect of different coupling
conditions (Fig. 5). Obtained results show that the qualitative behaviour is very similar in
case of single and coupled T-stub assemblage. Really, in some cases, a small reduction of
strength was noticed for the latter configuration, which could be ascribed to the following two
reasons: (1) due to the increment of global deformation, the flexural effects on the bolts are
more important, especially in case of type 2 collapse mechanism; (2) due to high
imperfections characterising the specimens, a statistical scatter among results concerning
equal specimen has to be expected, with a major probability that lower strength is of concern
for coupled T-stub assemblage.

The influence of flange material

In order to highlight the influence of flange material on the overall response of the joint, in
Figure 7, with reference to T-stub type 1 and 2, the relevant monotonic F-∆ curves are
grouped for each type of bolt. Obviously, such an influence is related to the type of failure
mechanism, and it appears more clearly for failure mechanism type-1, when the flange itself

72 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


rules completely the component response. Anyway, according to material coupon tests, AW
6061 generally provided the lowest strength and highest ductility, while the behaviour of AW
7020 and AW 6082 was quite similar to each other. Finally, it is worthy noticing that the
comparison of T-stub type 1A and T-stub types 1B and 1C emphasises also the influence of
welding type on the joint behaviour, showing that the former weld detail conferred higher
specimen ductility but a remarkably lower moment capacity.

a) failure mech. type 1 b) failure mech. type 2a

c) failure mech. type 2c d) failure mech. type 3


Figure 6. Typical observed failure mechanisms.

180 F [kN] Type 1 (bolt S 4.8) 180 F [kN] Type 1 (bolt AA7075) 180 F [kN] Type 1 (bolt S 10.9)

150 A (AW 6061) 150 A (AW 6061) 150 A (AW 6061)


B (AW 6082) B (AW 6082) B (AW 6082)
120 C (AW 7020) 120 C (AW 7020) 120 C (AW 7020)
C
90 B 90 B 90
B
A
60 60 A 60
C A
30 30 C 30
∆ [mm] ∆ [mm] ∆ [mm]
0 0 0
0 3 6 9 12 0 1 2 3 4 0 2 4 6 8

180 F [kN] Type 2-3 (bolt AA7075) 180 F [kN] Type 2-3 (bolt S 10.9) 180 F [kN] Type 2-3 (bolt S 10.9)
coupled
150 A (AW 6061) 150 150
C
B (AW 6082)
120 C (AW 7020) 120 120
B A B C A
90 90 90
A (AW 6061) A (AW 6061)
60 60 B (AW 6082) 60 B (AW 6082)
C (AW 7020) C (AW 7020)
30 A 30 30
C B
∆ [mm] ∆ [mm] ∆ [mm]
0 0 0
0 0.5 1 1.5 2 0 1 2 3 4 0 1 2 3 4

Figure 7. Influence of flange material: T-stub type 1 and type 2-3.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 73


The comparison between monotonic and cyclic results

Figures 3, 4 and 5 show also the comparison between monotonic and cyclic test results.
Preliminarily, it should be observed that for every tested specimen, the failure mode detected
for monotonic loading tests was observed also for cyclic loading tests (see Table 4). This
means that the loading type has not a significant influence on the collapse mode of the
specimen.

From the examination of the above figures, it can be noted that cyclic tests generally gave
rise to a structural performance lower than the corresponding one gained by monotonic
loading tests. This can be ascribed to the mechanical and geometrical damage of the
specimen accumulated during the sequential plastic excursions to which the specimen itself
is subjected. As a result, the ductility of specimens under cyclic loading is reduced with
respect to monotonic tests. This is particularly evident for specimens failing with a type 1 or
type 2a failure mechanism, which involves the plastic collapse of the flange at the weld
location. In such a case, it has been already observed that the collapse of the specimen was
essentially due to crack initiation in the weld. Obviously, the application of cyclic loads
favourites the crack propagation, anticipating the failure of the specimen. On the contrary,
the damaging effect due to low cycle fatigue is limited for specimens collapsing due to bolt
failure (type 2c or type 3 failure mechanism), which are mainly subjected to tensile forces,
exhibiting a not-ductile failure mechanism even for monotonic loads.

Finally, it should be observed that among analysed specimens it seems that there are some
anomalous cases. In fact specimen type 4 (Bolt S 4.8), specimen type 4 (Bolt AA 7075) and
specimen type 5 (Bolt AA 7075) present a performance (strength and ductility) for cyclic
loading higher than for monotonic loading. On the other hand, all these cases refer to
specimens failing due to bolt fracture, where the influence of loading history is not important.

SUMMARY AND CONCLUSIONS

The experimental investigation presented in this paper comprises 26 different aluminium


joints tested under monotonic and cyclic loading. The experimental program has been
planned in order to examine the main parameters influencing the response of welded T-stub
joint components connected to other structural elements by means of bolts. Therefore, 4
different T-stub types (varying in-plane dimensions, plate thickness, number and location of
bolts), 3 different aluminium alloys as flange material, 3 different bolt materials and two
different types of coupling systems have been considered. The obtained results, which are
given in terms of force versus displacement curves up to collapse of the joint, allow us to
point out the main influence of the above parameters on the observed failure plastic
mechanisms as well as on the ultimate strength, ductility and dissipative capacity of the joint.

In particular, as it could be expected, it is apparent that the T-stub geometry has an important
influence on the failure mechanism exhibited by the structural component. In fact, among the
analysed specimens, T-stub Type 1 and Type 2-3 presented a weak-flange failure
mechanism, while T-stub Type 4 and T-stub Type 5 were characterised by a weak-bolt
failure mechanism.

Furthermore, it has been pointed out that the specimens endowed with steel grade 10.9 bolts
provide an ultimate strength remarkably higher than the other cases. In such a case, a type 1
failure mechanism, with limited deformation occurring in the bolts and a notable plastic
engagement of flange material, was evidenced. On other hand, specimens with bolt type AA
7075 and steel grade 4.8 present a similar strength, even though the former are generally

74 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


characterized by a lower value of the ultimate deformation capacity. In particular, it has been
pointed out that the specimens equipped with steel bolts class 4.8 exhibited mainly a type 2a
failure mechanism, with the final rupture of the bolts, but with significant plastic deformation
developed at the web-to-flange connection. On the contrary, for specimen with aluminium
bolts AA 7075 the dominant collapse mode was based on type 2c failure mechanism, which
is ruled by the tensile failure of bolts, with limited plastic deformations occurring in the flange
material.

On the other hand, the influence of the flange material on the overall response of the
examined structural component has to be related to the relevant collapse mode, it appearing
more important for specimens exhibiting a weak-flange failure mechanism (failure
mechanism type 1 and type 2a).

Eventually, the comparison between monotonic and cyclic test results allows us to conclude
that the loading type has not a significant influence on the collapse mode of the specimen.
Nonetheless, cyclic tests generally gave rise to a structural performance (strength and
ductility) lower than the corresponding one gained by monotonic loading tests, due to low-
cycle fatigue. This is particularly evident for specimens failing with a type 1 or type 2a failure
mechanism, which involves the plastic collapse of the flange at the weld location. On the
contrary, the damaging effect due to low cycle fatigue is limited for specimens collapsing due
to bolt failure (type 2c or type 3 failure mechanism), which exhibited a not-ductile failure
mechanism even for monotonic loading tests.

ACKNOWLEDGEMENT

The current experimental research has been carried out in the framework of the Research
Project “Damage of Connections in Metal and Composite Constructions”, funded by the
Italian Ministry of Education, University and Research (MIUR).

NOTATION

f0.2 conventional elastic strength for aluminium flange and bolt material
fu ultimate strength for aluminium flange and bolt material
fy yielding strength for steel bolt material
εu ultimate (uniform) deformation for aluminium flange material
F applied T-stub tensile force
Fel T-stub conventional elastic strength derived from monotonic tests
Fmax T-stub peak strength measured from tests
∆ relative displacement between T-stub elements at the connection centerline
∆el T-stub elastic displacement associated to Fel
∆max T-stub displacement level corresponding to Fmax
∆ult T-stub ultimate displacement

REFERENCES

(1) Mazzolani, F.M. 1995. Aluminium Alloy Structures, E & FN SPON, London.
(2) Mazzolani, F.M., De Matteis, G., Mandara, A., 1996. Classification system for
aluminium alloy connections, in Proc. of the IABSE Int. Col. on Semi-Rigid Structural
Connections, Istanbul, Vol. 75, 83-94.
(3) Matusiak, M. 1999. Strength and Ductility of Welded Structures in Aluminium Alloys,

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 75


Dr. Ing. Dissertation, Dept. of Strct. Eng., Norwegian University of Science and
Technology, Trondheim, Norway.
(4) De Matteis, G., Landolfo, R., Mazzolani, F.M. 2001. Experimental Analysis of
Aluminium T-Stubs: Framing of the Research Activity, in Proc. of the 8th Int. Conf. on
Joints in Aluminium (INALCO 2001), Munich (Germany), 28-30 March.
(5) Swanson, J.A., Leon, R.T. 2000. Bolted Steel Connections: Tests on T-Stub
Components, Journal of Structural Engineering, 126 :1, 50-56.
(6) Faella C., Piluso V. and Rizzano G. 1998. Experimental Analysis of Bolted Connections:
Snug versus Preloaded Bolts, Journal of Structural Engineering 127:7, 765-773.
(7) De Matteis, G., Mandara, A., Mazzolani, F.M. 2001. Calculation Methods for Aluminium
T-Stubs: A Revision of Ec3 Annex J, in Proc. of the 8th Int. Conf. on Joints in Aluminium
(INALCO 2001), Munich (Germany), 28-30 March.
(8) De Matteis, G., Mandara, A., Mazzolani, F.M. 2000. T-stub Aluminium Joints: Influence
of Behavioural Parameters, Computers and Structures 78:1-3, 311-327.
(9) De Matteis, G., Della Corte, G., Mazzolani, F.M. 2001. Experimental Analysis of
Aluminium T-Stubs: Tests under Monotonic Loading, in Proc. of Giornate Italiane della
Costruzione in Acciaio, Venezia (Italy), 26-28 Sept.
(10) De Matteis, G., Mandara, A., Mazzolani, F.M. 2002. Design of aluminium t-stub joints:
calibration of analytical methods, In Proc. of the Third European Conference on Steel
Structures, Coimbra (Portugal), 19-20 Sept., Vol. II 1017-1026 .
(11) De Matteis, G., Della Corte, G., Mazzolani, F.M. 2003. Experimental Analysis of
Aluminium T-Stubs: Tests under Cyclic Loading, in Advances in Structures, Vol. 1, 467-
473, Balkema Publ., The Netherlands.

76 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


EXPERIMENTAL BEHAVIOR OF T-STUB CONNECTION
COMPONENTS FOR THE MECHANICAL MODELING OF BARE-
STEEL AND COMPOSITE PARTIALLY-RESTRAINED BEAM-TO-
COLUMN CONNECTIONS

I. Clemente, Universitá degli Studi di Trieste, Italy


S. Noé, Universitá degli Studi di Trieste, Italy
G.A. Rassati, University of Cincinnati, USA

ABSTRACT
The successful modeling of partially-restrained beam-to-column connections
requires less accurate and less complicated constitutive laws for the separate
elements when using mechanical modeling approaches, when compared to
traditional finite element approaches. Nevertheless, the adopted constitutive
law needs to conform to certain accuracy limits, lest the simulated behavior be
unsatisfactory. The paper presents some results from ongoing experimental
tests aimed at obtaining simplified constitutive laws for T-Stub connections
suitable for use within a previously published mechanical model. Also, the
paper investigates the behavioral differences between hot-rolled and built-up
T-Stubs. The experimental data is presented and discussed, and some
numerical comparisons are also discussed.

INTRODUCTION

The mechanical modeling approach for bare-steel and steel-concrete composite beam-to-
column connections allows for the utilization of simple constitutive laws for the single
components of the joint, by concentrating the deformation and stiffness response into a finite
set of non-linear springs, which need to simulate all significant contributions to the behavior
of the connection (1, 2). If for a bare-steel connection it would be theoretically possible to use
one single properly defined non-linear rotational spring to account for the monotonic and
cyclic behavior of a beam-to-column joint, sacrificing the completeness of the output results
for the ease of modeling, in the case of a steel-concrete composite connection such
modeling is not anymore a viable approach. In fact, it has been shown (3, 4) that for steel-
concrete composite connections the presence of the reinforced slab forces an interaction
between the two sides of a beam-to-column joint, which are not behaving independently
anymore. For this reason, the one-spring approach is not sufficient anymore for the
simulation of such complex behavior and modeling formulations capable of accounting for
that interaction are needed. One viable approach is the modeling schematization presented
in (2), which takes into account most of the deformability contributions, and which is suitable
for both static and dynamic analyses (Figure 1).

Such modeling approach offers the advantage of a limited number of aimed output results
(as opposed to a classic finite-element modeling, which is sometimes hindered by the sheer
amount of results to manage). Also, the considered modeling approach, provided that it takes
into account most of the deformability and stiffness contributions, is capable of simulating
quite closely the behavior of a beam-to-column connection, without the need for very
complicated constitutive laws for the single components.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 77


4 4
6

5 5

1 2 2 1

1 2 2 1

Figure 1. Mechanical modeling of steel-concrete composite connections (2).


200 200
F [kN] F [kN]
150 150

100 100

50 50

0 0
-0.09 -0.06 -0.03 0.00 0.03 0.06 0.09
-0.09 -0.04 0.00 0.04 0.09
-50 -50

-100 Modello numerico


Experimental Curves -100
Prova sperimentale
Numerical Model
-150 -150 Modello numerico
Experimental Curves
δ [m] Prova sperimentale
Numerical Model [m]
-200 -200
Figure 2. Numerical vs. Experimental F-δ Figure 3. Numerical vs. Experimental
comparison for cleated joints. F-δ comparison for welded joints.

Despite the lower threshold of acceptability for component constitutive laws, typical of
mechanical models, some issues have emerged concerning the simulation of end-plate and
welded connections. It is worthwhile to recall at this point that the approach followed in the
Annexes J to Eurocodes 3 and 4 (5, 6) for the evaluation of force and stiffness for a single
connection component is based on the evaluation of equivalent T-stubs, corresponding from
case to case to the portion of column subjected to the connection forces, or to the steel
connection, etc. It is therefore important within this framework to use an analytical expression
suitable for the simulation of T-stubs, whatever their fabrication process may be.

As opposed to cleated connections, the simulation of which by means of the mentioned


mechanical model is very successful (see Figure 2), the modeling of connections that require
the evaluation of an equivalent T-stub on the connection side presents some challenges. In
fact, the equivalent T-stub representing the column originates typically from a hot-rolled
section, whereas the equivalent T-stubs representing a welded connection, or a flanged
connection (i.e., full-, or partial end-plate joints) originate from different fabrication processes,
and it is therefore reasonable to hypothesize that they have different behaviors. As shown in
Figure 3, the simulated behavior of a fully-welded connection is less accurate than the
previous case. As far as the modeling process is concerned (5, 6) the only difference
between the two cases is represented by the different allowance for the equivalent T-stubs
simulating the steel connection.

The rationale for the ongoing experimental campaign herein presented lays in the need to

78 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


evaluate the differences between fabrication processes in T-stubs, which could then lead to
improved constitutive laws for the simulation of beam-to-column connections.

MODELING ISSUES

In order to be able to take correctly into account the mentioned behavioral differences among
different types of connections, various constitutive laws have been investigated for the
simulation of PR beam-to-column connections.

T-Stub, end-plate or cleated connections typically function as the means of transferring


forces from the beam flanges to the column; therefore, the correct assessment of their
stiffness, both in elastic and in post-elastic phases, is paramount for a satisfactory simulation
of the global behavior of the joint.

F F

δ δ

Figures 4a, b. Tri-linear constitutive law without and with stiffness degradation.

The simplest cyclic constitutive law considered was an elastic-plastic curve with constant
hardening, which provides the benefits of a simple definition at the price of a somewhat
rough reproduction of experimental data. The logical refinement of such constitutive law is
represented by a cyclic tri-linear curve, which is capable of more closely reproducing the
experimental behavior of components (Figure 4a). Such constitutive law is used in the
mechanical model described in (2) with satisfactory results. The advantages of the tri-linear
constitutive law are represented by a reasonable ease of definition associated with a
reasonably accurate simulation of the real behavior of the component. The simulation
accuracy tends to deteriorate during cyclic events, well within plastic range, when the
material starts showing damage, like stiffness and strength degradation. In this circumstance,
in fact, the tri-linear curve overestimates the strength values in ultimate conditions, eventually
leading to the prediction of values of stiffness and strength of the connections larger than the
actual ones. On the other hand, the tri-linear curve proved to be accurate for the simulation
of monotonic events, and to be able to provide good results in post-elastic range for
connections characterized by a stable behavior.

Considering that the scope of the simulations encompasses limit analysis of a structure (for
instance, in the framework of a performance-based design approach), it is also important to
accurately assess the behavior of structural components well into plastic range, in order to
be able to predict their failure parameters. To this end, a refined version of the tri-linear law
has been considered, which includes some consideration for stiffness degradation in post-
yielding range (Figure 4b). This constitutive law is based on the definition of the “classic” tri-

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 79


linear relationship; the main difference resides in the handling of unloading cycles. In fact,
some sort of kinematic hardening law is introduced after a predetermined plastic deformation
limit, so that the yielding range of the component is really depending on the maximum force
reached before unloading. Stiffness degradation can thus be assessed in a reasonably
simple fashion without introducing additional relationships for the calculation of the refined
constitutive law.

In order to test the effectiveness of the proposed constitutive law, and to investigate the
behavioral differences among differently fabricated specimens, a series of component tests
on rolled and built-up T-stubs, as well as on cleated connections has been conducted, and
the results have been compared to the simulations outcome.

EXPERIMENTAL TEST SETUP

The test program involves four different nominal T-stub sizes (original depth of the hot-rolled
HEA section equal to 120, 200, 240, and 300 mm, respectively) obtained in three different
ways: cut from hot-rolled shapes, built-up by welding plates together, and built-up by bolting
angles to a plate (cleated connection). The geometric dimensions of the three differently
fabricated specimens have been chosen in order to obtain easily comparable data. An
example of the typical specimen is shown in Figure 5. In this paper, only some results of the
T-200 and T-240 series will be reported, as the experimental campaign is ongoing.

HEA200

Figure 5. Typical hot-rolled specimen.


250
ε/εy

200

150

100

50

0
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30 32 34 36 38 40 42 44

Figure 6. Test frame. Figure 7. Load history.

80 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Figure 8. Specimen T240L ready for testing.

Figure 9. Specimen T240L after failure (front view).

Figure 10. Specimen T240L after failure (side view).

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 81


The specimens were subjected to predominant axial force in the stem, either monotonically
or cyclically, in order to obtain a force-deformation relationship against which the numerical
analyses should be compared. Table 1 summarizes the tests performed and those
scheduled.

Table 1. Test matrix.


T120 T200 T240 T300
L B S L B S L B S L B S
Test 1 X X X M M M M X X X X X
Test 2 X X X C C C X X X X X X
Test 3 X X X C C X X X X X X X
Test 4 X X X X X X X X X X X X
Test 5 X X X X X X X X X X X X
L=Rolled Tee; B=Bolted Tee; S=Welded Tee. M=Monotonic test; C=Cyclic test

The loading frame, shown in Figure 6 with a specimen installed, was made of two steel
plates (top 50 mm thick and bottom 30 mm thick with stiffeners), connected by means of four
high-strength steel 50-mm-diameter threaded rods. The specimens were directly bolted to
the top plate, and the stem was bolted to the loading device, consisting in a two-way
hydraulic jack (435kN in tension, 933 kN in compression, manually operated).

The force and the deflection of the specimens were measured by means of a 400-kN load-
cell and two linear motion position sensors, installed at each side of the T-stub.

The initial tests were conducted monotonically, with unloading cycles to verify the effects of
strain-hardening and eventual losses of strength or stiffness. For each set of specimens,
cyclic tests were also conducted. The loading history follows the spirit of the ECCS 1992 (7)
document; the load cycles are function of a theoretical yielding value, and are shown in
Figure 7. Figures 8 to 10 show specimen T240L before and after testing.

RESULTS

The obtained results can be classified under three different aspects: a) behavioral
differences among same-sized T-stubs fabricated with different methods; b) behavioral
differences among different-sized T-stubs with the same fabrication method; and c)
differences between the cyclic and the monotonic behavior of the specimens.

For what concerns the first aspect of the problem, Figure 11 shows the comparison of the
monotonic skeleton curve for the differently fabricated T-stubs. The elastic stiffness appears
to be the same in all specimens, as expected, whereas the ultimate behavior is remarkably
different: the welded Tee results weaker than the rolled one, due mainly to the smaller
amount of material in the k-zone of the profile. On the other hand, the bolted Tee is stronger
than the hot-rolled one, as a result of the thicker compound web, and of the larger k-zone.
Moreover, the offset of the two angles (8 mm) reduces the distance between the bolt-holes
and the theoretical location of the plastic hinge, therefore resulting in smaller stresses.

Figure 12 shows, on the other hand, a comparison of the monotonic skeleton curves for hot-
rolled T-stubs T-200 and T-240. It is evident that the ultimate strength of the specimens
increases proportionally to the size, as expected; it is nevertheless possible to notice how
there is little or no difference as far as elastic and post-elastic stiffness and yielding points
are concerned. This is mostly due to geometric differences between the two specimens, in
particular to the bolt-hole positioning.

82 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Finally, Figures 13 to 15 show the comparison between the cyclic curves and the
corresponding monotonic skeleton curves of the differently fabricated T-200 T-stubs. It is
evident how the components investigated are affected by the repeated loads, in that their
stiffness and, to a lesser extent, their strength tend to degrade towards the latter stages of
the experiments. On the other hand, the monotonic curve does not show any sign of
degradation of strength (stiffness degradation cannot be extrapolated from the monotonic
data). The hot-rolled profile (Figure 13) shows a very stable behavior up to large plastic
deformations, incurring in sensible strength degradation only in ultimate conditions. All
curves are characterized by different amounts of pinching, caused by localized plastic
deformations of the T-stub flanges, which tend to pull away from the plate surface; such gaps
need then to be closed at the inversion of load before any actual forces can be absorbed by
the component. The bolted Tee is by far the specimen experiencing the most pinching
(Figure 14). Also, the bolted Tee shows earlier signs of strength degradation, thus presenting
a less stable behavior than the hot-rolled profile. The welded Tee (Figure 15) shows early
signs of stiffness and strength degradation, mostly due to the formation of cracks in the
welds, suggesting that particular care needs to be put into the detailing of built-up T-stubs.
There is no apparent loss in ductility, nevertheless the loss of strength in ultimate conditions
may eventually cause an unacceptably low overall strength for the connection (in order to
qualify a connection for seismic zones, the AISC seismic provisions (8) require that 80% of
the strength be retained after a 40 mrad joint rotation).

Figures 16 to 18 show the comparison of the experimental results with some of the analytical
simulations outlined above, using either the basic or the modified tri-linear curve.

The tri-linear curve without deterioration seems to be more than suitable for the simulation of
monotonic behavior, and also of cyclic behavior, provided that the response of the
component remains stable. It is evident how in the later stages of the simulation, when
degradation of strength and stiffness are not negligible, the simulated curve deviates quite
noticeably from the measured curve.

On the other hand, the adoption of the modified constitutive law (Figure 18) seems to provide
a closer simulation of the plastic behavior of the investigated components. Both loading and
unloading curves follow the experimental data quite closely, and a visual inspection of the
cyclic curves shows how the total dissipated energy in the process seems to be predicted
quite well by the model. The diagrams deviate once the strength degradation starts evidently
affecting the experimental behavior.

350
F [kN]
300

250

200

150 T200L-M
T200B-M

100 T200S-M

50

δ [mm]
0
0 5 10 15 20 25 30

Figure 11. T-200 monotonic test comparison.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 83


350
F [kN]

300

250

200

T240L-M
150
T200L-M

100

50

δ [mm]
0
0 5 10 15 20 25 30

Figure 12. T-200L vs. T-240L monotonic test comparison.

400
F [kN]
300

200

100

-100

-200 T200L-M

-300 T200L-C

-400

δ [mm]
-500
0 2 4 6 8 10 12 14 16 18 20

Figure 13. T-200 hot-rolled Tee cyclic vs. monotonic comparison.

400
F [kN]
300

200

100

-100
T200B-M
-200
T200B-C
-300

-400

δ [mm]
-500
0 2 4 6 8 10 12 14 16 18 20 22

Figure 14. T-200 bolted Tee cyclic vs. monotonic comparison.

84 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


300
F [kN]

200

100

-100

T200S-M
-200
T200S-C
δ [mm]
-300
0 2 4 6 8 10 12 14 16 18 20 22 24 26

Figure 15. T-200 cyclic welded Tee vs. monotonic comparison.

350
F [kN]
300

250

200

T240L-M EXPERIMENTAL
150
T240L-M NUMERIC

100

50

δ [mm]
0
0 5 10 15 20 25 30

Figure 16. T-240L-M hot-rolled Tee numerical vs. experimental comparison.

300
F [kN]

200

100

-100

-200
T200L-C TRI-LINEAR WITHOUT DEGRADATION
T200L-C EXPERIMENTAL
-300

δ [mm]
-400
0 2 4 6 8 10 12 14 16 18 20

Figure 17. T-200L hot-rolled Tee cyclic numerical vs. experimental comparison.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 85


300
F [kN]

200

100

-100

-200
T200L-C EXPERIMENTAL
T200L-C TRI-LINEAR WITH DEGRADATION
-300

δ [mm]
-400
0 2 4 6 8 10 12 14 16 18 20

Figure 18. T-200L numerical vs. experimental comparison (with degradation).

CONCLUSIONS

Some results from a series of ongoing experimental tests on T-stub components have been
presented and discussed. The interest in the T-stub component resides in the conventional
use that Annexes J to Eurocodes 3 and 4 (5, 6) make of equivalent T-stubs to take into
account deformation contributions to the overall behavior of beam-to-column connections.
Furthermore, it has been noted that the behavior of connections realized by means of built-
up T-stubs results in some cases different from comparable connections that used hot-rolled
T-stubs.

A refined constitutive law has been proposed for the simulation of T-stub connections, which
takes into account degradation of stiffness in ultimate conditions, and should enable accurate
predictions of the ultimate behavior of PR beam-to-column connections under cyclic and
dynamic loading conditions. As soon as data for the remaining tests will be available, it will
be possible to verify the suitability of the new constitutive law for the simulation of the ideal T-
stub component in a steel or steel-concrete beam-to-column connection. Also, yet another
constitutive law will be investigated, namely a tri-linear curve with both strength and stiffness
degradation, in order to better simulate the ultimate behavior of the components.

ACKNOWLEDGEMENTS

This ongoing research project is part of the doctoral work of the first author. Funding has
been obtained through a grant from the Italian Ministry of University and Scientific Research.

REFERENCES

(1) Huber G., Tschemmernegg F., (1998), Modelling of beam-to-column joints, Journal of
Constructional Steel Research, Vol. 45, No. 2, pp. 199-216
(2) Rassati G.A., Leon R.T., Noè S. (2004). Component modeling of PR composite joints
under cyclic and dynamic loading, Journal of Structural Engineering, Vol. 130 (2):343-
351
(3) Amadio C., Benussi F., Noe’ S., (1994), Behaviour of unbraced semi-rigid composite

86 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


frames under seismic actions, STESSA ’94, Timisoara (Romania), 26 June-1 July
(4) Clemente I., Noé S., Rassati G.A. (2003). Influence of partially-restrained connections
on the overall seismic behavior of steel-concrete composite frames. A mechanical
modeling approach., in Hancock, G.J., Bradford, M.A., Wilkinson, T.J., Uy, B., and
Rasmussen, K.J.R. (eds.) ASSCCA 03 – Advances in Structures, Balkema
Publishers.
(5) CEN (European Committee for Standardization), (1994), ENV 1993-1-1 Eurocode 3:
Design of Steel Structures – Part 1: General Rules and Rules for Buildings, CEN,
Brussels, Belgium
(6) CEN (European Committee for Standardization), (1994), ENV 1994-1-1 Eurocode 4 -
Design of Composite Steel and Concrete Structures: Part 1.1: General Rules and
Rules for Buildings, CEN, Brussels, Belgium
(7) AISC (American Institute of Steel Construction), (2002), Seismic Provisions for
Structural Steel Buildings, AISC 341-02, AISC, Inc., Chicago, IL
(8) ECCS (European Convention for Constructional Steelwork), (1992), Recommended
Testing procedure for assessing the behaviour of structural steel under cyclic loads,
ECCS, Brussels, Belgium

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 87


88 Connections in Steel Structures V - Amsterdam - June 3-4, 2004
NUMERICAL SIMULATIONS OF BOLTED CONNECTIONS :
THE IMPLICIT VERSUS THE EXPLICIT APPROACH

G.J. van der Vegte


Kumamoto University, Japan / Delft University of Technology, The Netherlands
Y. Makino, Kumamoto University, Japan

ABSTRACT
Among finite element (FE) programs, two different categories can be
distinguished depending on the methodology to compute the nodal
displacements : implicit and explicit solvers. This study first briefly describes
and compares both the implicit and explicit approaches. Various issues such
as the solution strategy, computational time as function of model size, and
convergence aspects of non-linear contact analyses are addressed. In
addition, explicit FE simulations are made of two experiments on large-scale,
partially bolted beam-to-column connections. A comparison between the
experimental and numerical hysteresis loops is presented. In addition, the
advantages of the use of the explicit method relative to the implicit approach
are highlighted.

INTRODUCTION

High-rise steel structures in Japan usually consist of an assembly of rectangular hollow


section (RHS) columns and wide flange beams. The conventional beam-to-column
connections, which are most commonly applied in Japan, have diaphragms through the
columns at the positions of the flanges with the beam-ends being welded to these diaphragm
plates. However, the Kobe earthquake in 1995 revealed that this type of beam-to-column
connections is vulnerable to brittle fracture. In 1998, a new type of connection was developed
by Kurobane (1) with the beams being connected to the through diaphragm plates by bolts,
some distance away from the column face. Tests showed that the bolted connections did not
fail by fracture as observed in the conventional connections. In addition, the bolted
connections showed a much larger energy dissipation capacity than the conventional type.

Because of the high costs of experiments, the number of tests carried out is usually limited.
On the other hand, in the last decade, the FE method has developed as a cost-effective and
reliable tool to conduct research beyond the scope of experimental programmes. Numerical
simulations of bolted connections include a range of aspects such as material and geometric
non-linearities and “rigid body movements”. Implicit FE packages as ABAQUS/Standard (2)
are not able to analyse these problems effectively. In contrast, explicit solvers as
ABAQUS/Explicit (3) are well-suited to simulate multiple contact interactions between
independent bodies.

After having carried out numerical simulations on bolted beam splices, numerical research of
bolted connections by van der Vegte et al (4) then focused on the simulation of actual bolted
beam-to-column connections under cyclic loading. In 2002, Miura et al (5) reported the
results of three large-scale tests on partially bolted beam-to-column connections without

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 89


weld-access holes. Similar to the connections tested by Kurobane (1), the major aim of
Miura’s tests was to develop and evaluate a new design concept, significantly different from
the conventional type.

Because of the detailed description of the experiments conducted by Miura et al, these tests
were selected to serve as a reference to assess the reliability of FE models analysed with the
explicit solution technique. The present study describes the test specimens, the numerical
models and the comparison between the numerical and experimental results. However, as
an introduction to explicit approach, this study first briefly addresses the most important
features of the implicit and explicit solution techniques.

HISTORY OF NUMERICAL SIMULATIONS ON BOLTED CONNECTIONS

Bolted end-plate connections have been widely applied in steel structures due to the
simplicity and economy of their fabrication and assembly. Figure 1a shows an example of a
bolted end-plate connection. Because moment-rotation curves of partially restrained
connections are of great importance for designers, in the past, many research programmes
were conducted to generate such data. In general, most of the “older” studies consist of
experiments in combination with an analytical approach. Along with the introduction of
computers, in more recent publications, another tool has emerged to obtain the moment-
rotation behaviour of bolted end-plate connections : the FE method.

Because of limitations in computational capabilities, both in terms of software and hardware,


the first attempts to simulate the behaviour of bolted connections numerically primarily
involved two-dimensional (2D) models. In 2D models, each component of the connection is
modelled using shell elements, whereas displacements are assumed to be distributed
uniformly along the third direction. As bolted connections are 3D in nature, it is understood
that 2D models are not able to cover the 3D behaviour satisfactory.

Krishnamurthy who can be considered as one of the pioneers in the field of FE simulations of
bolted connections, reported his findings in several publications (6, 7). Because 3D FE
models were computationally expensive, Krishnamurthy and Graddy (6) analysed selected
benchmark connections by 2-D and 3-D models. The correlation between these results was
then used for the prediction of other 3-D models based on the results of corresponding 2-D
models.

Sherbourne and Bahaari (8, 9) also presented a number of publications evaluating the
moment rotation behaviour of bolted end-plate connections generated with the FE
methodology. Although the models were 3D, some simplifications were made. At first,
Sherbourne and Bahaari assumed a continuous connection between the nodes of the bolt
head and nut and the nodes of the contacting plates. As a result, relative motions between
bolt, column flange and end-plate were restrained. A further limitation of Sherbourne and
Bahaari‘s FE models was caused by the use of truss elements to model the bolt shank.
Because the interface between the bolt shank and the hole boundary was neglected, the
model would not be suitable if the bolts would go into bearing action.

As computers became more powerful during the last decade, a steady increase in the size of
the FE models became possible. This enabled the analyses of less simplified and thus, more
realistic 3D models.

As an example may serve the work of Choi and Chung (10), whose FE models, including the
bolts, were completely based on solid elements. Because bolted connections consist of
various components with complex interaction between the various entities, it is understood

90 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


that a more accurate description of the geometry of the components requires a more
complex solution procedure. The authors highlighted the iterative character of the solution
strategy to describe these highly non-linear processes, non-linear both in terms of material
behaviour (plasticity) and geometric response (contact).

Not only enhanced the capacity of the hardware, but also the capabilities of the software
improved. New solution techniques were derived, making numerical analyses much more
efficient in terms of computational time. The introduction of sophisticated contact algorithms
further widened the application of numerical methods. “Contact” in its simplest form can be
described by the use of so-called “gap elements” which impose displacement compatibility
between user-defined pairs of nodes. However, such elements can only be used when
friction can be ignored. In addition, modelling of such elements is a tedious and time-
consuming task. To overcome these problems, commercial FE packages developed more
user-friendly options, such as contact between surfaces and interface elements instead of
the node-to-node contact definition required by gap elements. The improved contact
algorithms also enabled the modelling of shear and friction between the contacting surfaces.

In 1997, Bursi and Jaspart (11, 12) presented their numerical results on isolated bolted tee-
stub connections. Unlike most of the previous studies, the bolt head and the flange plate
were now modelled as individual components and no longer connected through common
nodes, enabling the relative movement between these components.

The same approach was followed by Wheeler et al (13) who carried out numerical
simulations on bolted end-plate connections subjected to pure bending. Figure 1a illustrates
one of the configurations analysed by Wheeler. Prior to the FE simulations of the assembled
model, Wheeler first conducted a series of numerical analyses to determine the most efficient
type of element and corresponding mesh density for each of the components (i.e. end-plate,
bolts, weld beam section). Wheeler recommended the use of eight noded, hybrid elements
with four elements modelled through the thickness of the end plate. The same element type
was employed to model the bolts. To avoid problems associated with rigid body movements,
loading of the connection was carried out in five steps, such as pre-tensioning of the bolts,
applying moment to the tip of the beam section, etc.

(a) Example of bolted end-plate connection (b) FE mesh analysed by Willibald et al (14)

Figure 1. Bolted end-plate connections.

Another example of an experimental programme followed up by a numerical study is the


work reported by Willibald et al (14) who investigated bolted flange-plate connections for
square and rectangular hollow sections under pure tension. Except for the element type
employed for the bolts, the concept of the FE models analysed by Willibald closely followed

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 91


the approach adopted by Wheeler et al (13). Unlike Wheeler, Willibald used twenty-noded,
quadratic brick elements to model the bolts. An illustration of a quarter FE model generated
by Willibald is given in figure 1b.

Most of the references so far considered end-plate connections, in which the load transfer
primarily takes place through axial loading of the bolts. FE simulations become more
complicated and expensive in terms of computational time when bolted connections are also
subjected to shear load, simply because of the specification of additional contact surfaces.
An example of the use of shear bolts, including the difficulties encountered in the FE
analyses is described by Swanson et al. After having conducted a series of experiments (15),
Swanson et al (16) then carried on with FE simulations on bolted T-stub connections. The
configuration, illustrated in figure 2, consists of a T-stub section which was attached to a
column flange through tension bolts. Shear bolts were used to connect the beam flange to
the T-stub. The FE mesh, almost exclusively modelled by quadratic, solid elements,
contained approximately 7300 elements, resulting in 53,000 degrees of freedom. Although
the loading of the connection was merely monotonic, the authors reported that a full run on a
Pentium II - 450 Mhz took approximately 36 hours. In addition, convergence problems
caused by rigid body motions, had to be overcome by assuming artificial boundary conditions
on the T-stub section.

Column flange

T-stub

Beam

Figure 2. T-stub connections studied by Swanson et al (15, 16).

OVERVIEW OF EXPLICIT AND IMPLICIT SOLUTION STRATEGIES

As described in the previous section, in the last decade, the FE method has become a
popular tool to solve a wide range of structural mechanics problems. Two different types of
solution strategies can be distinguished : the explicit and the implicit solution procedures.
This section briefly describes and compares the following aspects of both approaches : the
solution strategy, computational time as function of model size, and convergence in a non-
linear contact analysis. Further information is presented by van der Vegte et al (4).

Solution strategy

IMPLICIT APPROACH : The implicit method is based on static equilibrium and is characterized
by the simultaneous (implicit) solution of a set of linear equations. For a model to be in static
equilibrium, the net force acting on each node must be zero. After a full set of linear
equations has been assembled, this set of equilibrium equations is then solved
simultaneously and the unknown nodal displacements are obtained. The assembly of a

92 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


global stiffness matrix is an important feature of the implicit solution procedure

EXPLICIT APPROACH : In the explicit procedure, the displacements, velocities and


accelerations of each node are advanced explicitly through time. This means that the state of
the model at the end of an increment (time t+∆t) is solely based on the displacements,
velocities and accelerations at the beginning of the increment (time t). The nodal velocities at
time t+∆t are obtained assuming that the accelerations are constant during a small time
increment ∆t. Similarly, the nodal displacements at the end of the increment can be
computed by adding the displacements during the time increment ∆t to the displacements at
time t. The explicit method lacks a global stiffness matrix.

As might be clear from the explicit solution algorithm, explicit simulations may yield results
significantly affected by dynamic effects. A possible tool to identify whether or not this occurs,
involves monitoring the various components of the energy balance throughout the loading
process.

Effect of model size on computational time

IMPLICIT APPROACH : Although for the implicit method, the computational cost as a function of
model size is rather difficult to predict, experience shows that for many problems, the CPU
time is approximately proportional to the square of the number of degrees of freedom.

EXPLICIT APPROACH : Using the explicit method, the computational cost is proportional to the
number of elements and roughly inversely proportional to the smallest element dimension.

Implicit
CPU time

Explicit

0
0

Number of degrees of freedom


Figure 3. CPU time versus model size for the explicit and implicit methods.

A qualitative comparison between CPU time and model size for both approaches is
illustrated in figure 3. Although for small models, the implicit method may be favourable, for
larger models the explicit approach becomes more attractive. This becomes even more valid
when disk space and memory requirements are taken into account. As result, for very large
models, the maximum size of an implicit model may be controlled by the available memory
and disk space instead of the required computational time.

Convergence in non-linear contact analyses

IMPLICIT APPROACH : Figure 4 shows a flow chart for one increment in a non-linear contact
analysis. From this diagram, it becomes clear that within each increment, two iterative loops
can be distinguished. Only after both criteria regarding contact constraints and equilibrium
are satisfied, the analysis proceeds with the next increment. In case of severe discontinuity, it

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 93


is possible that either one or even both criteria are not met and, as a result, no solution can
be obtained at all.

EXPLICIT APPROACH : In a non-linear explicit analysis including contact, it is first assumed that
no contact occurs between the various entities. If, at the end of the increment, overclosure
(i.e. penetration due to contact) is found, corrections for the kinematics of the contacting
bodies are calculated and adopted that would have been required to prevent penetration
from taking place. Unlike the implicit approach, no convergence criteria are considered and,
as a result, contact simulations proceed smoothly.

Start of increment

Determine contact status

Perform iteration

Contact failure No convergence

Check changes in
contact

No changes

Check equilibrium

Convergence

End of increment

Figure 4. Flow chart of non-linear, implicit analysis.

NUMERICAL SIMULATIONS OF LARGE-SCALE TESTS ON BEAM-TO-COLUMN


CONNECTIONS

Experiments by Miura et al (5)

In 2002, Miura et al (5) reported the results of three large-scale tests on bolted beam-to-
column connections without weld access holes. Because of the detailed description of the
experiments, these tests were selected to serve as a reference to evaluate the reliability of
the explicit FE predictions.

The connection details of two of the three T-shaped configurations are shown in figure 5. In
each of these specimens, a wide flange beam was partially welded and bolted to a cold-
formed square hollow section column. Specimen A contains an internal diaphragm plate
welded inside the column located at the bottom flange of the beam. The web and bottom
flange of the beam are bolted to a shop-welded shear tab and flange plate respectively. The
top flange of the beam is field-welded to the external diaphragm plate. In Specimen C, the
beam web is bolted to a shop-welded shear tab, whereas both flanges are field-welded to the
external diaphragm plates.

The tip of the beam is subjected to reversed loading with increasing amplitude, while both
column ends are fixed against displacements.

94 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Field welding Field welding

Shop welding Field welding


(a) Specimen A (b) Specimen C

Figure 5. Beam-to-column connections tested by Miura et al (5).

Numerical model

FE MESH : Because the configurations tested by Miura et al lack symmetry, a full scale FE
model had to be developed for each specimen. The FE meshes, primarily modelled with 8-
noded solid elements with reduced integration (ABAQUS element type C3D8R), follow the
mesh layout recommended by van der Vegte et al (4). Based on the numerical results of a
simplified connection between three plates and a single bolt, it was found that, in order to
obtain reliable simulation results (i) at least 24 elements should be modelled around the
perimeter of a bolt and bolt hole, and (ii) at least 4 layers of solid elements should be
considered through the thickness of each plate. In addition, the aspect ratio of the linear solid
elements should not exceed 10. Figure 6 illustrates the mesh adopted for the bolt and
washers. For clarity, only one of the washers is displayed. Figure 7 displays the FE mesh
generated for Specimen A. The FE mesh of Specimen C is similar. The FE mesh of
Specimen A (with 17 bolts) contains approximately 135,000 nodes.

Figure 6. FE mesh generated for bolts and washers.

CONTACT, FRICTION AND PRE-LOADING : Similar to the explicit analyses on the single-bolted
configuration by van der Vegte et al (4), each set of a bolt and two washers considers 6
“contact pairs”. In addition, the following interactions are modelled : (i) contact between the
beam web and the web plate and (ii) contact between the beam flange and the flange plate.
As a result, the FE model of Specimen A with 17 bolts, includes 104 contact definitions.

Friction is included, whereas the friction coefficient is taken as either 0.3 (for contact
involving the bolts) or 0.6 (for contact between rusted surfaces). Following the
recommendations of AIJ (17), the bolts are pre-loaded by a tensile stress of 520 N/mm2.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 95


(a) Overall mesh (b) Detail of mesh at bottom flange

Figure 7. FE mesh generated for Specimen A.

KINEMATIC HARDENING : The kinematic hardening model adopted in the current study
considers linear hardening only. For the hardening modulus, a value of E/100 is assumed.

LOADING PROCEDURE : In the experiments, the amplitude of the beam rotation was increased
as 2θp, 4θp, 6θp ... up to failure, whereas θp is a theoretical, calculated value of the beam
rotation at full plastic beam moment (1). For each displacement increment, the beam was
subjected to two cycles. However, in the numerical analyses, the beam is subjected to only
one cycle for each displacement increment, due to constraints in computational time.

Numerical results

Because of the large demand in computational time, the numerical simulations of Specimens
A and C were terminated when the beam tip rotation was between 6θp and -6θp, providing
sufficient data to assess the reliability of the FE model.

Figure 8 displays both the numerical and experimental results, summarized in hysteresis
loops. The moment in the beam (at column face) Mm, made non-dimensional by the full
plastic beam moment Mp, is plotted against the non-dimensional beam-end rotation θm/θp.
Figure 8 reveals that for both specimens, the outer points of the experimental and numerical
loops (i.e. points of load reversal) are close. The slopes of both sets of curves near the points
where the load reverses, also match well. However, the numerical loops overestimate the
stiffness of the specimens shortly after the load reverses. The divergence between the
experimental and numerical loops is likely caused by the relatively simple, linear kinematic
hardening model adopted in this study. These observations are well in line with the
conclusions made by Goto et al (18) who conducted numerical research into the hysteretic
behaviour of thin-walled columns.

As schematically illustrated in figure 3, for large FE models, the explicit method is


computationally more attractive than the implicit approach. On the 5-year-old HP Exemplar
V2250KS server used for the current analyses, a “large” implicit model contains
approximately 40,000 nodes. Full-scale FE model of Specimen A includes approximately
135,000 nodes. This implies that by employing the explicit approach instead of the implicit
method, the model size can be increased by a factor of 3.4. In addition, because of the lack
to assemble and invert a stiffness matrix, explicit analyses do not impose severe
requirements on internal memory and disk-space. For implicit analyses for which the

96 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


assembly and solution of the stiffness matrix is an essential feature, these requirements
would have made analyses of this size impossible.

Even more important, unlike the frequent occurrence of convergence failures observed in
implicit contact simulations, the explicit analyses of highly non-linear FE models A and C
proceeded smoothly without any computational difficulty.
1.5 1.5
Num erical Numerical
1.0 Experim ent 1.0 Experim ent

0.5 0.5
M m /M p

M m /M p
0.0 0.0

-0.5 -0.5

-1.0 -1.0

-1.5 -1.5
-6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6
θm / θp θm / θp

(a) Specimen A (b) Specimen C

Figure 8. Experimental and numerical hysteresis loops.

CONCLUSIONS

After having presented a brief overview of the implicit and explicit FE solution techniques,
explicit FE simulations have been conducted on bolted beam-to-column connections. Based
on the results, the following conclusions can be made :

a. The explicit solution technique is able to provide reliable predictions of the hysteretic
response of actual bolted beam-to-column connections, even when the simple, linear
kinematic hardening model is used. However, for further improvement of the shape of
the numerical hysteretic loops, a more sophisticated material model should be
considered.
b. The explicit method is a suitable tool to simulate the behaviour of bolted connections
effectively. Unlike implicit analyses of bolted connections, the explicit solution technique
does not require simplifications to be made in order to generate a reliable solution.
c. The explicit solution algorithm enables the numerical simulation of FE models much
larger than the model size attainable by the implicit strategy.
d. The explicit method requires a more intense evaluation of the results than the implicit
methodology. The explicit solution, if not correctly applied, may yield results significantly
affected by dynamic effects. For the explicit analyses, various components of the
energy balance should be monitored throughout the loading process. Hence, the use of
the explicit method is only recommended for advanced FE users.

ACKNOWLEDGEMENTS

The first author would like to thank the Japan Society for the Promotion of Science for the
opportunity and financial support to carry out the research reported.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 97


REFERENCES

(1) Kurobane, Y., (1998). Improvement of I-Beam-to-RHS Column Moment Connections for
Avoidance of Brittle Fracture. Proc. of the 8th International Symposium on Tubular
Structures, Singapore, Keynote Paper, pp. 3-17.
(2) ABAQUS/Standard, (2000). Version 6.1, Hibbitt, Karlsson & Sorensen, U.S.A.
(3) ABAQUS/Explicit, (2000). Version 6.1, Hibbitt, Karlsson & Sorensen, U.S.A.
(4) Vegte, G.J. van der, Makino, Y. and Sakimoto, T., (2002). Numerical Research on
Single-Bolted Connections Using Implicit and Explicit Solution Techniques. Memoirs of
the Faculty of Engineering, Kumamoto University, Vol. 47, No. 1.
(5) Miura, K., Makino, Y., Obukuro, Y., Kurobane, Y., Vegte, G.J. van der, Tanaka, M. and
Tokudome, K., (2002). Testing of Beam-to-RHS Column Connections Without Weld-
Access Holes. International Journal of Offshore and Polar Engineering, Vol. 12, No. 3,
pp. 229-235.
(6) Krishnamurthy, N. and Graddy, D.E., (1976). Correlation between 2- and 3-Dimensional
Finite Element Analysis of Steel Bolted End-Plate Connections. Computers and
Structures, Vol. 6, pp. 381-389.
(7) Krishnamurthy, N., Huang, H.T., Jeffrey, P.K. and Avery, L.K., (1979). Analytical M-θ
Curves for End-Plate Connections. Journal of Structural Division, ASCE, Vol. 105, No.
1, pp. 133-145.
(8) Sherbourne, A.N. and Bahaari, M.R., (1994). 3D Simulation of End-Plate Bolted
Connections. Journal of Structural Engineering, Vol. 120, No. 11, pp. 3122-3136.
(9) Sherbourne, A.N. and Bahaari, M.R., (1997). Finite Element Prediction of End-Plate
Bolted Connection Behaviour. I : Parametric Study. Journal of Structural Engineering,
Vol. 123, No. 2, pp. 157-164.
(10) Choi, C.K. and Chung, G.T., (1996). Refined Three-Dimensional Finite Element Model
for End-Plate Connection. Journal of Structural Engineering, Vol. 122, No. 11, pp. 1307-
1316.
(11) Bursi, O.S. and Jaspart, J.P., (1997). Benchmarks for Finite Element Modelling of
Bolted Steel Connections. Journal of Constructional Steel Research, Vol. 43, Nos. 1-3,
pp. 17-42.
(12) Bursi, O.S. and Jaspart, J.P., (1997). Calibration of a Finite Element Model for Isolated
Bolted End-Plate Steel Connections. Journal of Constructional Steel Research, Vol. 44,
No. 3, pp. 225-262.
(13) Wheeler, A.T., Clarke M.J. and Hancock, G.J., (2000). FE Modelling of Four Bolt,
Tubular Moment End-Plate Connections. Journal of Structural Engineering, Vol. 126,
No. 7, pp. 816-822.
(14) Willibald, S., Packer, J.A. and Puthli, R.S., (2001). Bolted Connections for RHS Tension
Members. CIDECT Report 8D/8E-10/01, University of Toronto, Canada.
(15) Swanson, J.A. and Leon, R.T., (2000). Bolted Steel Connections : Tests on T-Stub
Components. Journal of Structural Engineering, Vol. 126, No. 1, pp. 50-56.
(16) Swanson, J.A., Kokan, D.S. and Leon, R.T., (2000). Advanced Finite Element Modelling
of Bolted T-Stub Connection Components. Journal of Constructional Steel Research,
Vol. 58, Nos. 5-8, pp. 1015-1031.
(17) AIJ, (2001). Recommendation for Design of Connections in Steel Structures.
Architectural Institute of Japan (in Japanese).
(18) Goto, Y., Wang, Q. and Obata, M., (1998). FEM Analysis for Hysteretic Behavior of
Thin-Walled Columns. Journal of Structural Engineering, Vol. 124, No. 11, pp. 1290-
1301.

98 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


FOUR-PARAMETER POWER MODEL FOR M-θR CURVES
OF END-PLATE CONNECTIONS

N. Kishi, Muroran Institute of Technology, Japan


M. Komuro, Muroran Institute of Technology, Japan
W.F. Chen, University of Hawaii, U.S.A.

ABSTRACT
Four-parameter power model for predicting moment relative-rotation relation of
semi-rigid connections has been established which is composed of initial
stiffness, strain-hardening stiffness, reference connection moment, and shape
parameter. These parameters for total 168 experimental test data of the end-
plate type connections stored in the connection database were determined. An
applicability of the model using the parameters determined here to frame
analysis was investigated comparing with the numerical analysis results
obtained directly using experimental test data.

INTRODUCTION

To rationally perform structural analysis and design calculation for flexibly jointed frames, it is
very important to establish a simple model for estimating nonlinear moment relative rotation (M-
θr) relation for semi-rigid connections. The authors proposed the use of three-parameter power
model for predicting nonlinear M-θr curves of angle type connections. However, this model may
not be much applicable for estimating M-θr curves of the end-plate type connections because
clear-cut strain-hardening connection stiffness is indicated in these experimental test data. On
the other hand, it is well known that four-parameter power model can present M-θr curves of
semi-rigid connections with strain-hardening connection stiffness. This is one expanded model of
three-parameter power model.

From this point of view, in this paper, in order to rationally estimate nonlinear M-θr curves of the
end-plate type connections stored in the connection database, four parameters for each 168
experimental test data were decided, and an applicability of the model was investigated
comparing with the numerical analysis results for a flexibly jointed frame obtained using
experimental test data for the end-plate connections.

(a) Extended end-plate (b) Flush end-plate (c) Header plate

Figure 1. End-plate type connections.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 99


Table 1. List of collected experimental tests.
Connection type Reference (author, year) Number of tests
L. G. Johnson et al. (1960) 1
A. N. Sherbourne (1961) 5
J. R. Bailey (1970) 26
J. O. Surtees & A. P. Mann (1970) 6
J. A. Packer & L. J. Morris (1977) 3
S. A. Ioannides (1978) 6
R. J. Dews (1979) 3
Extended end-plate P. Grundy et al. (1980) 2 136
N. D. Johnstone & W. R. Walpole (1981) 8
A. Mazroi (1983, 1984) 24
Y. L. Yee (1984) 16
D. B. Moore & P. A. C. Sims (1986) 2
J. B. Davison et al. (1987) 1
R. Zandonini & P. Zanon (1987) 9
L. F. L. Ribeiro & R. M. Goncalves (1996) 24
J. R. Ostrander (1970) 24
Flush end-plate J. Phillips & J. A. Packer (1981) 5 32
J. B. Davison et al. (1987) 3

END-PLATE CONNECTIONS

The end-plate connections, usually shop welded and field bolted, are widely used in practice due
to its simplicity of fabrication and erection. Depending upon the relative length of the end-plate to
beam height, there are three types of the end-plate connections in practice: extended end-plate
connections (Figure 1a), flush end-plate connections (Figure 1b), and header plate connections
(Figure 1c).

In terms of classification of connections, extended end-plate connections and flush end-plate


connections fit in the mix category of rigid and semi-rigid connections where header plate
connections fall in the category of flexible connections (2, 3) because initial connection stiffness
is small and strain-hardening effects can not be expected. Kishi et al. (4) has demonstrated an
applicability of the three-parameter power model for header plate connections based on the
concept that the header-plate connections behaves similarly to double-web angle connections.

Observing a significant number of experimental M-θr curves of extended and flush end-plate
connections in the database, it is clearly understood that these connections possess
pronounced strain-hardening connection stiffness. It means that four-parameter power model
including strain-hardening connection stiffness rather than three-parameter power model should
be applied to estimate these M-θr relations.

Now, a total of 168 experimental test data for these types of connections are installed in the
database (5, 6). The references and corresponding number of tests are listed in Table 1.

FOUR-PARAMETER POWER MODEL

A four-parameter power model, which is proposed by Richard and Abbott (7) for modelling
elastic-plastic stress-strain relation, is applied. Using this power model, the connection moment
M and relative rotation θr, and tangent connection stiffness Rkt are represented as follows:

100 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


1
⎧ n
⎫n
(Rki − Rkp ) θr ⎪⎛ Rki − Rkp ⎞ n +1 ⎪ (1), (2)
M= + Rkp θr , θr = ⎨⎜ ⎟⎟ − 1⎬ θ0

( ) ⎪⎝ Rkt − Rkp ⎠
1
1 + ( θr θ0 )n n ⎪
⎩ ⎭

dM (Rki − Rkp )
Rkt = = + Rkp (3)
(n +1)
dθr
(1 + ( θr
θ0
n
) ) n

where Rki = initial connection stiffness; Rkp = strain-hardening connection stiffness; n = shape
parameter; θ0 = reference relative rotation [= M0/(Rki-Rkp)]; and M0 = reference connection
moment.

Figure 2 shows the M-θr curves of a connection when a magnitude of shape parameter n is
varied from small value to infinity. From this figure, it is seen that if shape parameter n is taken to
be infinity, the model is reduced to a bilinear one with the initial connection stiffness Rki and
strain-hardening connection stiffness Rkp. The model has the following merits:
1. The model is composed of only four parameters and its formulation is simple and
straightforward;
2. The model requires less computing time in a nonlinear structural analysis, since not only
connection moment M but also relative rotation θr, and tangent connection stiffness Rkt can
be represented using simple closed form equations as indicated in Eqs (1) to (3);
3. If three parameters of Rki, Rkp and M0 were determined using simple structural mechanics,
an empirical equation for shape parameter n might be determined by using the database
and M-θr relation for the end-plate connection with arbitrary parameters might be evaluated.

Figure 2. Four-parameter power model. Figure 3. Determination of connection stiffness.

DERIVATION OF FOUR PARAMETERS

Initial connection stiffness Rki

The initial connection stiffness Rki is defined as the linear relationship between moment M and
relative rotation of connection θr in the elastic region. In general, the experimental moment-
rotation curve is expected to be continuously smooth over the initial range of loading. In such
ideal case, initial connection stiffness Rki can be estimated using tangent connection stiffness at
the initial loading point. However, there may be some experimental M-θr data with discontinuous
even in the elastic region as shown in Figure 3. To rationally estimate initial connection stiffness
Rki for all cases including the above mentioned experimental test data, here, the maximum
secant connection stiffness among all data from initial to final loading is taken as the stiffness:

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 101


Mi
Rki = (4)
θi max

in which Mi and θi are the i-th experimental test data from initial loading.

Strain-hardening connection stiffness Rkp

The strain-hardening connection stiffness Rkp is evaluated taking tangent connection stiffness
obtains drawing straight line backward from the point of the maximum connection moment (Mk,
θk). The cases softened in the post peak region are excluded from the consideration because
these cases need more complicate procedure in frame analysis. Strain-hardening connection
stiffness Rkp for each experimental test data can be obtained using following equation:

Mk − M j
Rkp = (5)
θk − θ j
min
in which suffixes k and j are the point at the maximum connection moment and arbitrary point
less than point k.

Reference connection moment M0

The reference connection moment M0 is the value where the straight-line representing strain-
hardening connection stiffness Rkp intersects the moment axis in the moment and relative-
rotation diagram as shown in Figures 2 and 3. It is represented as:

M0 = Mk − Rkp θk (6)

Shape parameter n

The shape parameter n is introduced for fitting the curve obtained using the power model with
experimental test data under three parameters being determined as shown in Figure 2. Here, it
is determined numerically applying least-mean-square technique for the connection moment
obtained using Eq. (1) and experimental test data

List of four parameters for experimental test data

The four parameters determined based on the above procedures for each experimental test data
are listed in Table 2. Figure 4 shows the comparison among experimental test data and two
curves obtained using four-parameter power model proposed here and modified exponential
model proposed by Kishi and Chen (5). From these figures, it can be seen that M-θr curves
predicted using four-parameter power model agrees rather well with experimental test data, and
those from the modified exponential model represent precisely experimental test data.

APPLICATION OF THE MODEL IN SEMI-RIGID FRAME ANALYSIS

An applicability of the proposed four-parameter power model to semi-rigid frame analysis is


investigated by analyzing a semi-rigid frame example as shown in Figure 5. Beam and
column sections, floor height, and beam span are shown in this figure.

102 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


An applicability and accuracy of the four-parameter power model proposed here in frame
analysis are investigated comparing the numerical results obtained using the 168 sets of
experimental M-θr test data with those using the proposed model. Here, a modified exponential
model is used for representing experimental M-θr curves. A second-order elastic analysis

Table 2. List of four parameters derived from experimental data.


Connection Rki Rkp M0
Author Test Id. n
Type (kNm/rad) (kNm/rad) (kNm)
L. G. Johnson et al.
TEST 5* 28,700 1,270 108 2.81
(1960)
TEST A1 127,000 1,720 305 1.95
TEST A2* 277,000 0 592 1.27
A. N. Sherbourne
TEST A3* 304,000 3,790 544 1.38
(1961)
TEST B1* 298,000 1,020 554 1.07
TEST B2* 375,000 0 584 1.02
TEST A1-L 20,500 246 246 4.27
TEST A1-R 45,200 0 257 1.81
TEST A2-L 19,300 342 132 3.15
TEST A2-R 23,700 252 125 2.36
TEST A3-L 17,700 400 149 5.28
TEST A3-R 24,000 259 156 3.24
B4-L* 43,500 0 262 1.78
B4-R* 55,900 0 261 1.50
B5-L* 31,200 206 220 2.62
B5-R* 61,300 0 228 1.72
B6-L* 39,600 0 424 2.40
B6-R* 60,200 0 426 1.85
B7-L* 33,000 0 289 2.46
J. R. Bailey (1970)
Extended B7-R* 37,600 0 300 2.45
end-plate B8-L* 36,700 847 101 2.56
B8-R* 12,800 503 99.2 2.82
C9-L* 13,300 2,060 56.5 4.89
C9-R* 11,300 0 81.9 3.04
C10-L*, † 58,700 3,220 168 2.14
C10-R* 43,500 0 212 1.88
C11-L* 39,700 2,750 136 3.75
C11-R* 29,800 8,000 96.5 9.26
C12-L* 34,400 479 293 2.58
C12-R* 48,600 0 341 2.07
C13-L* 47,800 1,660 304 2.79
C13-R* 45,300 0 320 2.89
TEST C1 32,300 4,400 95.4 2.55
TEST C2 45,300 1,330 349 4.64
J. O. Surtees & TEST C3 99,100 0 399 1.98
A. P. Mann (1970) TEST C4 39,700 1,620 363 9.67
TEST C5 74,600 2,660 364 1.85
TEST C6* 113,000 1,440 564 4.30
TEST J1 7,410 0 96.5 1.54
J. A. Packer &
TEST J2 4,980 0 85.6 1.53
L. J. Morris (1977)
TEST J3 4,520 58.8 70.3 1.55
Note: * Case of column with stiffener, † This data is shown in Figure 4

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 103


Table 2. (Continued).
Connection Rki Rkp M0
Author Test Id. n
Type (kNm/rad) (kNm/rad) (kNm)
TEST 1 24,600 1,210 78.1 2.09
TEST 2 632,000 3,700 181 0.83
S. A. Ioannides TEST 3 226,000 3,320 269 1.32
(1978) TEST 4 76,200 992 110 0.95
TEST 5 77,400 1,980 219 1.13
TEST 6 1,420,000 8,610 264 1.08
TEST 1 6,340 526 89.2 2.68
R. J. Dews (1979) TEST 2 10,500 965 72.7 3.85
TEST 3 197,000 1,400 142 1.37
P. Grundy et al. T1 365,000 49,600 708 3.46
(1980) T2 387,000 23,200 855 4.31
TEST 1-L* 114,000 321 263 1.10
TEST 1-R* 108,000 0 312 1.04
TEST 2-L* 45,300 503 253 2.11
N. D. Johnstone & TEST 2-R* 63,700 0 261 1.21
W. R. Walpole (1981) TEST 3-L* 50,500 150 255 1.10
TEST 3-R* 28,600 0 238 1.82
TEST 4-L* 26,200 470 231 1.40
TEST 4-R* 15,800 430 180 2.24
EP3 1,490,000 30,200 801 1.81
EP4 2,850,000 175,000 728 1.74
EP5 2,830,000 506,000 702 2.33
EP6 5,650,000 279,000 814 1.60
A. Mazroi (1983)
EP7 2,830,000 258,000 975 2.05
Extended EP7 With Shim 6,280,000 90,300 1,450 3.37
end-plate EP8 3,370,000 268,000 1,040 1.68
EP8 With Shim 15,300,000 81,600 1,620 1.20
CF4-U12x87 800,000 16,200 634 1.39
CF4-U12x106 622,000 26,900 905 1.73
CF4-U12x120 1,020,000 37,600 920 1.90
CF5-U10x68 1,020,000 46,100 572 1.69
CF5-U14x61 714,000 36,400 381 1.75
CF5-U10x49 893,000 30,300 261 1.18
CF6-U12x96 1,240,000 56,900 598 1.39
CF6-U14x158 1,130,000 35,200 1,040 1.58
A. Mazroi (1984)
CF8-U14x159 791,000 29,200 1,360 1.78
CF3-S12x65* 904,000 162,000 315 1.67
CF4-S12x87* 1,670,000 67,400 891 1.90
CF5-S10x68* 1,360,000 190,000 537 2.15
CF5-S12x96* 1,470,000 69,400 1,080 2.16
CF6-S12x79* 1,700,000 43,100 774 1.35
CF8-S14x145* 1,700,000 60,800 1,280 1.68
CF8-S14x158* 1,700,000 156,000 1,180 2.00
K1B* 50,900 945 400 2.86
K9* 45,200 3,400 326 2.52
K4* 56,500 4,260 331 2.88
Y. L. Yee (1984) K7* 55,800 1,610 365 3.09
K5A* 67,800 797 379 2.53
K8* 47,100 2,940 320 2.41
K10* 56,500 1,060 382 1.89
Note: * Case of column with stiffener, † This data is shown in Figure 4

104 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Table 2. (Continued).
Connection Rki Rkp M0
Author Test Id. n
Type (kNm/rad) (kNm/rad) (kNm)
K11* 54,700 2,660 238 2.06
K12* 30,700 2,280 150 1.61
B1* 53,000 1,220 349 1.58
B2 40,500 2,270 171 1.94
Y. L. Yee (1984) B3 39,400 1,810 176 2.23
C1* 24,100 1,940 130 1.96
C2* 12,300 258 70.1 1.61
D1 22,700 1,440 128 2.00
D2 26,000 2,170 172 2.49
D. B. Moore & J3 13,300 444 46.7 1.00
P. A. C. Sims (1986) J4 16,600 682 54.3 1.19
J. B. Davison et al.
JT/13B 10,100 271 54.9 7.69
(1987)
EP 1-1 22,600 0.057 182 1.31
EP 1-2* 65,400 165 191 1.22
EP 1-3* 140,000 0 208 1.04
EP 1-4* 136,000 775 207 1.37
R. Zandonini &
EP 1-5* 226,000 775 211 1.21
P. Zanon (1987)
EPB 1-2 113,000 0 201 1.05
EPB 1-3* 208,000 2,430 162 1.35
EPB 1-4* 522,000 1,390 201 1.05
EPB 1-5* 279,000 694 215 1.48
Extended CT1A-1* 79,100 2,380 118 2.12
end-plate CT1A-2* 113,000 10,600 87.8 2.14
CT1A-3* 113,000 4,070 120 1.69
CT1A-4* 90,400 5,670 102 2.01
CT1A-5* 73,500 2,120 124 1.75
CT1A-6* 56,500 2,110 136 1.65
CT1B-1* 283,000 237 138 1.19
CT1B-2* 102,000 3,880 179 1.94
CT1B-3* 113,000 554 175 1.67
CT1B-4* 56,500 2,100 158 2.37
CT1B-5* 90,400 642 166 1.62
L. F. L. Ribeiro &
CT1B-6* 73,500 643 164 1.32
R. M. Goncalves
CT2A-1* 102,000 5,920 321 3.19
(1996)
CT2A-2* 147,000 5,340 324 2.07
CT2A-3* 79,100 1,750 376 2.63
CT2A-4* 170,000 5,530 238 1.43
CT2A-5* 113,000 1,400 404 1.79
CT2A-6* 113,000 2,200 371 3.55
CT2B-1* 113,000 912 359 2.77
CT2B-2* 90,400 1,500 356 2.84
CT2B-3* 113,000 6,230 338 2.30
CT2B-4* 102,000 2,050 350 2.47
CT2B-5* 147,000 2,060 362 1.56
CT2B-6* 56,500 1,150 378 2.42
TEST 1 9,660 203 50.8 1.62
Flush J. R. Ostrander
TEST 3 10,600 189 45.1 1.35
end-plate (1970)
TEST 4 5,310 241 21.1 1.97
Note: * Case of column with stiffener, † This data is shown in Figure 4

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 105


Table 2. (Continued).
Connection Rki Rkp M0
Author Test Id. n
Type (kNm/rad) (kNm/rad) (kNm)
TEST 9 21,200 215 55.2 1.30
TEST 11 13,300 385 51.7 1.60
TEST 12 21,200 318 74.7 1.55
TEST 13 13,300 601 77.1 1.85
TEST 17 13,200 76.5 59.7 1.18
TEST 18 8,170 138 67.0 1.54
TEST 19 13,300 300 57.6 1.51
TEST 23 21,200 894 79.7 1.83
TEST 2* 13,600 248 58.0 1.44
TEST 5* 10,600 159 59.5 1.70
J. R. Ostrander
TEST 6* 17,700 114 52.3 1.03
(1970)
TEST 7* 5,900 130 28.7 1.31
TEST 8* 10,600 226 24.0 1.17
TEST 10* 22,600 364 66.5 1.52
Flush
TEST 14* 7,910 251 67.4 1.71
end-plate
TEST 15* 26,600 730 78.0 1.42
TEST 16* 35,400 565 96.3 1.59
TEST 20* 10,100 238 54.6 1.75
TEST 21* 10,600 254 74.5 1.47
TEST 22* 23,600 384 69.5 1.34
TEST 24* 42,500 816 95.7 1.57
BM1* 33,900 272 105 1.43
BM2*† 81,700 545 116 1.19
J. Phillips &
BM3* 81,700 408 127 1.14
J. A. Packer (1981)
BM4* 81,700 408 132 1.33
BM5* 129,000 584 138 1.14
JT/11 15,300 723 39.0 3.67
J. B. Davison et al
JT/11B 43,400 0 50.5 1.78
(1987)
JT/12 8,450 303 17.9 2.34
Note: * Case of column with stiffener, † This data is shown in Figure 4

(a) Extended end-plate connection (b) Flush end-plate connection


(database no. 44) (database no. 29)

Figure 4. Comparisons of the prediction models with test data in connection database.

106 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Figure 5. Frame used in numerical analysis.

program (8) is used for this investigation which is capable of rationally analyzing flexibly jointed
frames with nonlinear connection stiffness. Three kinds of frame responses are used for this
investigation: 1) the nodal displacement under service load, 2) the fixed end moment of column
under factored load, and 3) the beam end moment under factored load. These response values
are normalized with reference to the corresponding values obtained using modified exponential
model as shown in Eqs (7) and (8).

Normalized displacement:

displacement obtained using four-parameter power model


d* = (7)
displacement obtained using modified exponential model
Normalized moment:

moment obtained using four-parameter power model


m* = (8)
moment obtained using modified exponential model

These normalized displacement d* and normalized moment m* are then plotted taking
logarithmic initial connection stiffness Rki .

Figure 6. Distribution of nodal displacement Figure 7. Distribution of fixed end moment


under service load at node no. 3. under factored load at node no. 1.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 107


Table 3. Summary of results for nodal displacement under service load.
% of connection data
Node No. Mean value µ Standard deviation σ
distributed in 5 % error
3 1.00 0.02 96 %
4 1.00 0.02 96 %
5 0.99 0.03 94 %
6 0.99 0.03 94 %
Note: Number of useful tests data: 168

Table 4. Summary of results for fixed end-moment under factored load.


% of connection data
Node No. Mean value µ Standard deviation σ
distributed in 5 % error
1 1.00 0.02 98 %
2 1.00 0.01 100 %
Note: Number of useful tests data: 168

Nodal displacement under service load

Figure 6 shows an example of normalized displacement d* versus log10 Rki obtained for node
3. In this figure, µ and σ indicate a mean value and standard deviation among all response
values respectively. In this case, since µ = 1.0 and σ = 0.02, it is seen that nodal
displacement under service load can be better predicted using four-parameter power model.
Table 3 summarizes the numerical results for normalized nodal displacement and shows that
all mean values µ’s are almost unity and the results for 94 % of connection data distribute in
the region less than 5% error.

End moments under factored load

Fixed end moment of column base


Figure 7 shows the distribution of non-dimensional fixed-end moment of column base (node no.
1 of elem. no. 1) under factored load against logarithmic connection stiffness (log10 Rki). From
this figure, since µ = 1.0 and σ = 0.02, it is seen that the results are concentrated around unity,
and the results obtained using four-parameter power model can better predict those obtained
using modified exponential model. The results for fixed-end moment of both column bases are
listed in Table 4. From this table, it is seen that both mean values µ’s are unity and the standard
deviation σ for windward column (node no. 1) moment is little bigger than that for leeward
column. The results for almost all the connections (98% to 100% of whole connection test data)
are distributed in less than 5 % error. Therefore, it can be concluded that fixed-end moment of
columns in semi-rigid frames with end-plate connections under factored load can be estimated
with reasonable accuracy using the four-parameter power model.

End moment of beam


Generally, bending moments at the windward node of beams under the combined loading of
wind and gravity loads are very small comparing with those of leeward beams and are largely
influenced by the connection flexibility because the connection moments caused due to these
two load components are different in sign each other. Therefore, since the connection moment
becomes small and its sign is depend upon the flexibility of semi-rigid connection, non-
dimensional connection moment m* becomes sometime negative. In this study, at node no. 3 of
elem. no. 5, the negative non-dimensional connection moment m* was obtained for two

108 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


connection test data.

Figure 8 shows the distributions of non-dimensional beam end moments for elem. no. 6. From
this figure, it is seen that the moment in the region of weak connection stiffness are scattered in
the area more than 10 % error. However, in the region of stiffer connection stiffness, the moment
has a tendency to concentrate in unity. The numerical results of beam end moment are listed in
Table 5. Standard deviations σ’s for corresponding beam end moments of elem. no. 6 are
distributed in the range of 0.01 to 0.05. More than 92% connection test data are distributed in
region less than 5 % error.

Standard deviation σ of non-dimensional beam end moment for node no. 3 of elem. no. 5 is very
large comparing with those of the other end moments. This may be due to the reasons that 1)
the moment is at windward beam end and becomes smaller than that of leeward one; and then
2) its sensitivity becomes higher than those in cases of the other columns. However, from the
engineering point of view, the windward beam end moment may not be much important. Indeed,
it has been confirmed that the windward beam end moments for all connection data are almost
1/80 – 1/2 those at the leeward node.

(a) Elem. no. 6, node no. 5 (b) Elem. no. 6, node no. 6

Figure 8. Distribution of beam end moment under factored load.

Table 5. Summary of results for beam end-moment under factored load.


% of connection data
Elem. No. Node No. Mean value µ Standard deviation σ
distributed in 5 % error
3 1.00* 0.11* 80 %
5
4 1.00 0.02 98 %
5 1.01 0.05 92 %
6
6 1.00 0.01 98 %
Note: *Number of useful tests data: 166

Then, from Figure 8 and Table 5, it can be verified that four-parameter power model for
estimating M-θr relation of end-plate connections can be applicable in semi-rigid frame analysis
in lieu of real experimental test data and/or modified exponential model.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 109


CONCLUSIONS

A four-parameter moment-rotation model has been proposed for estimating moment M and
relative rotation θr relation of the end-plate beam-to-column connections. The first three
parameters, initial connection stiffness Rki, strain-hardening connection stiffness Rkp, and initial
connection moment M0 were determined based on the experimental test data. The fourth
parameter, shape parameter n, is numerically determined applying the least-mean-square
technique for connection moments between the proposed model and experimental test data. It is
confirmed that the proposed model with four parameters decided here can better simulate M-θr
relation for experimental connection test data installed in the connection database.

An applicability of proposed four-parameter power model to semi-rigid frame analysis was


investigated analyzing a one-bay two-story semi-rigid frame. As the results, it was confirmed that
the response values of sectional forces and frame drift could be estimated with reasonable
accuracy by using proposed four-parameter power model in lieu of experimental connection test
data. The proposed power model is simple and easy to implement in nonlinear frame analysis
program.

NOTATION

d* normalized displacement Rki initial connection stiffness


m* normalized moment Rkp strain-hardening connection stiffness
M connection moment Rkt tangent connection stiffness
M0 reference connection moment θ0 reference relative rotation [= M0/(Rki-Rkp)]
n shape parameter θr relative connection rotation

REFERENCES

(1) Kishi, N. and Chen, W.F. (1990). “Moment-rotation relations of semi-rigid connections
with angles.” Journal of Structural Engineering, ASCE, 116(7), 1813-1834.
(2) Kishi, N., Chen, W.F., Goto, Y. and Hasan, R. (1997). “Study of Eurocode 3 steel
connection classification”, Engineering Structures, 19(9), 772-779.
(3) Hasan, R., Kishi, N., Chen, W.F., and Komuro, M. (1997). “Evaluation of rigidity of
extended end-plate connections”, Journal of Structural Engineering, ASCE, 127(12),
1595-1602.
(4) Kishi, N., Hasan, R., Chen, W.F. and Goto, Y. (1994). “Power model for semi-rigid
connections.” Steel Structures, Journal of Singapore Struct. Steel Society, 5(1), 37-48.
(5) Kishi, N and Chen, W.F. (1986). “Data base of steel beam-to-column connections.”
Struct. Engrg. Report No. CE-STR-86-26, Purdue Univ., West Lafayette, Ind.
(6) Hasan, R. (1996). “Evaluation of connection stiffness and modeling of semi-rigid
connections.” Thesis presented to Muroran Institute of Technology, in Partial
Fulfillment of the Requirements for the Degree of Doctor of Philosophy.
(7) Richard, R.M. and Abbott, B.J. (1975). “Versatile elastic-plastic stress-strain formula.”
Journal of the Engineering Mechanics, ASCE, 101(4), 511-515.
(8) Goto, Y. and Chen, W.F. (1987). “On the computer-based design analysis for flexibly
jointed frames.” Journal of Constructional Steel Research, Special Issue on Joint
Flexibility in Steel Frames (W.F. Chen Ed.), Vol. 8, 203-231.

110 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


ELASTO-PLASTIC FE ANALYSIS ON MOMENT-ROTATION
RELATIONS OF TOP- AND SEAT-ANGLE CONNECTIONS

M. Komuro, Muroran Institute of Technology, Japan


N. Kishi, Muroran Institute of Technology, Japan
W.F. Chen, University of Hawaii, U.S.A.

ABSTRACT
To establish a numerical analysis method for appropriately evaluating moment
relative-rotation relations of top- and seat-angle connections under monotonic
loading, three-dimensional elasto-plastic finite element analysis was performed
under following considerations: 1) all the connection components are modeled
using solid element; 2) a contact surface algorithm is applied between every
adjacent two components; and 3) bolt pretension is introduced. An applicability
of this analysis method for this type connections was discussed comparing
with the experimental results. From this study, it can be concluded that
applying the proposed analysis method, moment relative-rotation relations of
top- and seat-angle connections can be accurately estimated up to the ultimate
state.

INTRODUCTION

To perform the semi-rigid frame analysis and design, it is important to evaluate the moment
relative-rotation behavior of the beam-to-column connections. Many experimental studies
have addressed prediction of the moment relative-rotation characteristics of top- and seat-
angle type connections all over the world (1, 2, 3, 4). On the other hand, over the past few
years, several numerical simulations of angle type connections have been conducted by
means of elasto-plastic finite element (FE) analysis method. Kishi et al (5) and Citipitioglu et
al (6) investigated the effects of pretension of bolts and friction coefficient between
connection components on moment relative-rotation curves of top- and seat-angle type
connections in detail. In these studies, moment relative-rotation curves of the connections in
the initial loading area can be approximately predicted by using the FE analysis. However, in
the large deformation area including plastic deformation, connection moment relative-rotation
relation may not be appropriately estimated.

In this study, in order to establish a numerical analysis method for appropriately evaluating
moment relative-rotation relations of top- and seat-angle with/without double web-angle
connections in the finite deformation area under monotonic loading, three-dimensional
elasto-plastic FE analysis was performed under following considerations: 1) all the
connection components (beam, column, angle, etc.) are modeled using solid element; 2) a
contact surface algorithm is applied between every adjacent two components; and 3) bolt
pretension is introduced. Here, ABAQUS code (7) was used for numerical analysis of the
three kinds of connection specimens, in which dimensions of the web angles were taken as
variables. An applicability of this analysis method for this type of connections was discussed
comparing with the experimental results conducted by the author’s laboratory (4). Moment
relative-rotation behavior of the connection, strain relative-rotation relations of top and/or web

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 111


angle, and failure mode of the connection were used for this investigation.

EXPERIMENTAL OVERVIEW

Specimens

To investigate an influence of dimensions of web-angles on moment relative-rotation


behavior of top- and seat-angle type connections, three kinds of specimen were prepared.
The configuration and dimensions of the specimens are shown in Figure 1. Sections of
beam, top/seat angle and web angle are H400×200×13×8, L150×100×12 and L90×90×7,
respectively. F10T high strength bolts of 20 mm in diameter were used in this study, in which
the standard torque (480 N-m) was introduced in these bolts. Nominal name of each
specimen is designated using length of web angles, in which m of Wm means a length of
web angle (cm). Mechanical properties of the members are listed in Table 1, which were
obtained from the coupon test using JIS No. 6 specimens. Figure 2 shows the stress-strain
characteristics for each connecting member.

Test setup

Dimensions and geometry of the specimens are shown in Figure 3. The specimens were
attached to the column (H408×408×21×21), which was set on the steel bed and was
strengthened welding vertical stiffeners to keep from local buckling. The lateral load was
applied to the top end of cantilever beam through a swivel using a 500 kN screw jack. The
loading point was at a height of 1,500 mm from the column face. Two guide rails were used
to control the loading direction.

(a) W00 (b) W18 (c) W29

Figure 1. Configuration and dimensions of test specimens.

Table 1. List of material properties.


Young's Poisson's Yield Tensile
Connecting Elongation
Grade modulus ratio stress strength
member (%)
Es (GPa) νs fy (MPa) fu (MPa)
beam web 210 385 481 37.1
flange 210 325 463 39.4
SS400
top / seat angle 210 0.3 282 449 44.6
web angle 209 315 469 41.7
bolt F10T 212 1,060 1,098 19.7

112 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Figure 2. Stress-strain relation Figure 4. Measuring points for evaluation of
for each steel and bolt. relative-rotation of connection.

Figure 3. Test set-up.

The lateral load and displacement were measured by using a load cell (capacity: ± 300 kN)
and laser type Linear Variable Displacement Transducers (LVDTs) (maximum stroke: ± 250
mm), respectively. The vertical displacements at the four side-edges of beam flanges bolted
to angles were measured by using laser type LVDTs to evaluate the relative-rotation of
connection (Figure 3). Strains at some locations of angles were measured to investigate
strain and deformation behavior of angles. All of these measured data were continuously
recorded using digital data-recorders.

Lateral load was surcharged following a displacement control method. No axial load was
applied to the beam. To ensure the safety of experiment, monotonic load was increased until
displacement at the loading point reaches about 145 mm except the case of W00 specimen
which reached ultimate state due to bolts failure. Therefore, ultimate moment capacity of
connection for W18 and W29 specimens may not be precisely evaluated in this experiment.

Evaluation of relative-rotation

In order to precisely evaluate the relative-rotation of connection, the vertical displacements at


four side-edges of beam flange bolted to angles were measured by using the laser type
LVDTs (Figure 3). Assuming that angles deform such as shown in Figure 4, the relative-
rotation θr of connetion is estimated as follows:
δ1 − δ 2
θr = (1)
dl

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 113


where δ1 and δ2 = vertical displacements at tension and compression side edge of beam
flange bolted to angles, respectively; dl = the horizontal distance between two laser type
LVDTs set at tension and compression side flange of beam. Here, the relative-rotation θr was
evaluated using a mean value of rotations obtained from both sides of beam flange.

The connection moment M is represented as follows:

M = PH ⋅ h (2)

where PH = laterally surcharged load; h = a distance from the center of swivel to column face
(h = 1,500 mm).

Relative connection rotation θr and connection moment M discussed later was estimeted
using Eqs (1) and (2), respectively.

FE MODEL

In this study, ABAQUS code was used to estimate the connection moment relative-rotation
behavior of angle type connections. Figure 5 shows an example of FE model for W29
specimen. A half of specimen referring to the symmetrical axis was modeled. All connection
components (angle, beam, column, bolt) were modeled using eight-node solid elements.
Total number of nodal points and elements for W29 specimen are 55,810 and 33,729,
respectively. The bottom flange of column is assumed to be perfectly fixed according to the
experimental boundary conditions.

To accurately simulate the connection behavior, small sliding occurred between every
adjacent two components is explicitly considered applying a contact surface algorithm
prepared in ABAQUS code. These are 1) between bolt shank and bolt hole, 2) between bolt
head/nut and connecting components, and 3) between angles and beam/column flanges. All
components including bolts are completely independent from each other as assemblages in

Figure 5. FE model and boundary conditions for W29 specimen.

114 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Figure 6. Contact surface and Mises stress Figure 7. An example of stress-strain
contour after introducing pretension relation used in this analysis.
force into bolt.

(a) W00 (b) W18 (c) W29

Figure 8. Comparison of M-θr curves between numerical and experimental results.

real connection. Bolt holes which are 2 mm larger than the bolt size (D = 20 mm) were drilled.
178 kN pretension force was introduced into each bolt according to the experimental
conditions. Figure 6 shows the contact surface of top angle and Mises stress contour after
pretension being introduced into bolts. From this figure, it is confirmed that the pre-stress
was transfer from bolt to angle.

Stress-strain relation for each connection component is represented using multi-linear


constitutive model according to the material properties as shown in Figure 2 and Table 1.
Isotropic hardening rule with von Mises yielding criterion was applied for simulating plastic
deformation of connection. Figure 7 shows an example of stress-strain relation for top angle
used in this analysis.

Bending moment is surcharged to the connection by employing prescribed horizontal


displacement at the loading point as shown in Figure 5 according to the experimental
condition. In this study, automated loading increment scheme is preferred because ABAQUS
code can automatically decide an increment of loading so as to keep a high computational
performance.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 115


COMPARISON BETWEEN NUMERICAL AND EXPERIMENTAL RESULTS

Moment relative-rotation relation

Figure 8 shows comparison of moment relative-rotation relations between numerical and


experimental results. From these figures, initial connection stiffness for each specimen
obtained from numerical analysis agrees accurately with that obtained from experimental
results. Connection moment obtained from numerical analysis is a little bigger than that of
the experimental results in the region of 5 ≤ θr ≤ 30 mrad. It may be for this reason that the
center of bolt does not coincide to that of bolt hole. However, both numerical and
experimental connection moments better correspond to each other in the region of θr > 30
mrad. Moreover, the relative-rotation corresponding to the maximum connection moment
obtained from numerical analysis for W00 specimen is in good agreement with the
experimental maximum one. Therefore, moment relative-rotation relation of top- and seat-
angle type connections can be accurately estimated by using proposed numerical analysis
method regardless of longitudinal size of web angle.

Figure 9 shows examples of deformation configuration of W29 specimen. These figures


show the results for the three loading stages. From these figures, it is seen that angles are
gradually deformed with an increment of relative-rotation θr.

(a) θr = 23 mrad (b) θr = 40 mrad (c) θr = 96 mrad


Figure 9. Deformation configuration of W29 specimen for three loading steps.

Figure 10. Comparison of stress relative-rotation relation in top angle between numerical
and experimental results.

116 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Strain relative-rotation relation

Figure 10 shows comparison of strain relative-rotation relations in top angle between


numerical and experimental results in cases of W00 and W29 specimens. Here, the strains
near fillet of angle in both column and beam flanges were investigated. From these figures, it
is seen that at beginning of loading, the experimental strain of angle in beam flange is bigger
than that in column flange. Moreover, the former intends to increase from around 4 mrad of
relative-rotation at which connection is yielded. Although the numerical analysis cannot
estimate adequately the tendency of rapidly increasing of these experimental strains of angle
in beam flange, it is seen that the strain distributions obtained from numerical analysis are
qualitatively similar to the experimental ones.

Figure 11 shows the deformation configuration and equivalent plastic strain distribution of top
angle of W00 specimen. From this figure, it is observed that the toe of the fillet of angle in
beam flange and the area around bolt hole of angle in column flange have been yielded from
the initial loading stage, and these areas are gradually extended with an increase of relative-
rotation θr. Since a similar behavior was also observed from experimental results, the
proposed numerical analysis method can appropriately estimate the strain distribution of the
top angle.

Figure 11. Deformation configuration and equivalent plastic strain distribution of


top angle of W00 specimen.

Figure 12. Comparison of stress relative-rotation relations in web angle of W29 specimen
between numerical and experimental results.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 117


Figure 12 shows the comparisons of strains at fillet of web angle in case of W29 specimen
between numerical and experimental results. From the comparisons, it is seen that the
experimental results for angle in beam web are bigger than those of analytical ones as well
as strain distribution of top angle. Since both analytical and experimental results for strains of
angle in beam web are bigger than those of web angle in column flange, the strains of web
angle may be better estimated by using the proposed FE method qualitatively.

Figure 13 shows the deformation configuration and equivalent plastic strain distribution for
web angle of W29 specimen. From this figure, it is seen that the toe of the fillet of angle in
beam web is yielded first at θr = 11 mrad, and then yielded area is developed from tension
side (c1 section) to compression side (c5 section) with an increase of θr.

Figure 14 shows the deformation configuration and Mises stress distribution of top angle
including bolt at four loading stages. From these figures, it is observed that the heel of top
angle is apart from the surface of column flange at θr = 5 mrad, and then top angle is
gradually deformed with an increase of θr. At θr = 39 mrad, the area from edge of bolt hole
through angle’s heel is completely apart from the surface of column flange due to the prying
action of angle. Moreover, the yielded area is developed to the middle of bolt shank, in which
Mises stress reached up to more than 1,000 MPa. At θr = 85 mrad, yielded area is developed
to the whole of the bolt shank.

Figure 13. Deformation configuration and equivalent plastic strain distribution


in web angle of W29 specimen.

Figure 14. Deformation configuration and Mises stress distribution of


top angle of W00 specimen.

118 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Figure 15. Deformation and Mises stress distribution from FE analysis and a look of
failure bolt in experiment for top angle of W00 specimen.

Figure 15 shows the deformation configuration and Mises stress distribution of the top angle
including bolt at θr = 114 mrad of relative-rotation at which bolt failure was occurred in
experiment, and a look of failure bolt after the experiment. From this figure, it is observed that
the bolt shank was significantly deformed in numerical analysis, and the deformation
configuration of bolt is good correspond to a look of fractured bolt in experiment. Thus, it is
indicated that the collapse mode of the top- and seat-angle type connections may be
appropriately estimated using the deformations and Mises stress distribution of connection
components.

CONCLUSIONS

In this study, in order to establish a numerical analysis method for appropriately estimating
moment relative-rotation relation of top- and seat-angle with/without double web-angle
connections from initial stage to finite displacement stage under monotonic loading, three-
dimensional elasto-plastic finite element analyses were performed. The results obtained from
this study are as follows:
1) applying the proposed FE analysis method, moment relative-rotation relations of top- and
seat-angle type connections can be better estimated regardless of longitudinal size of
web angle;
2) strain distribution of top- and/or web-angle can be estimated by using the proposed
analysis method qualitatively, and
3) failure mode of connection may be estimated using the deformation and Mises stress
distribution of connection components.

ACKNOWLEDGEMENT

This research was supported in part by a grant from ‘The Japan Iron and Steel Federation’.
The support is gratefully acknowledged.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 119


NOTATION

dl horizontal distance between two laser type LVDTs set at tension and compression
side flange of beam
Es Young's modulus of steel
fu tensile strength of steel
fy yield stress of steel
h distance from the center of swivel to column face
M connection moment
PH lateral surcharged load
δ1 vertical displacement at tension side edge of beam flange bolted to angle
δ2 vertical displacement at compression side edge of beam flange bolted to angle
θr relative connection rotation
νs Poisson's ratio

REFERENCES

(1) Kukreti, A.R. and Abolmaali, A.S. (1999). Moment-rotation hysteresis behavior of top
and seat angle steel frame connections, Journal of Structural Engineering, 125(8), 810-
820.
(2) Calado, L., Matteis, G.D., and Landolfo, R. (2000). Experimental response of top and
seat angle semi-rigid steel frame connections, Materials and Structures, 33, 499-510.
(3) Azizinamini, A. (1985). Cyclic Characteristics of bolted semi-rigid steel beam to column
connections, PhD thesis, University of South Carolina, Columbia.
(4) Komuro, M., Kishi, N., and R.Hasan (2003). Quasi-static loading tests on moment-
rotation behavior of top- and seat-angle connections, Proceedings of the Conference
on Behaviour of Steel Structures in Seismic Areas, Naples, Italy, June 9-12, 329-334.
(5) Kishi, N., Ahmed,A., Yabuki, N., and Chen, W.F. (2001). Nonlinear finite element
analysis of top- and seat-angle with double web-angle connections., International
Journal of Structural Engineering and Mechanics, 12(2), 201-214.
(6) Citipitioglu, A.M., Haj-Ali, R.M., and White, D.W. (2002). Refined 3D finite element
modeling of partially restrained connections including slip, Journal of Constructional
Steel Research, 58, 995-1013.
(7) ABAQUS/Standard user’s manual (1998). Ver. 5.8, Hibbitt, Karlsson & Sorensen.

120 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


MODELING PROCEDURES FOR PANEL ZONE DEFORMATIONS IN
MOMENT RESISTING FRAMES
Finley A. Charney, Virginia Tech, U.S.A.
William M. Downs, Simpson Strong Tie, Inc., U.S.A.

ABSTRACT

Elastic and inelastic deformations in the panel zone of the beam-column joint
region of moment resisting frames are responsible for a very significant
portion of the lateral flexibility of such systems. This paper provides a brief
theoretical basis for computing panel zone deformations, and compares
results obtained from two simple mechanical models to each-other and to
those obtained using detailed finite element analysis. It is shown that the
simplest mechanical model, referred to as the Scissors model, produces
results that are comparable to those obtained from the more complex
mechanical model, and also correlates well with the results computed from
the detailed finite element model.

INTRODUCTION

The influence of panel zone deformations on the flexibility of steel moment resisting frames
is very significant. This is true for elastic response, and particularly for inelastic response
when yielding occurs in the panel zone. Structural analysis should always include such
deformations.

While the state of stress in the panel zone is extremely complex, the sources of deformation
can be divided into three parts: axial, flexural, and shear. For low to medium rise frames,
axial deformations are negligible, flexural deformations are minor but significant, and shear
deformations are dominant. This paper concentrates on the shear component of panel
zone deformation. See Downs (1) for a detailed discussion on modeling approaches for
axial and flexural deformations within the panel zone.

Mathematical modeling procedures for panel zone deformation are typically based on
simple mechanical analogs which consist of an assemblage of rigid links and rotational
springs. The principal challenge in the derivation of such models is the development of the
transformations from shear in the panel zone to rotation in the springs of the analog. Two
mechanical models were studied in the research reported herein. These are the
“Krawinkler Model” (2) and the “Scissors Model”. When properly used, the results obtained
from these models are essentially identical, even though the kinematics of the Krawinkler
model is significantly different than that of the Scissors model. Unfortunately the Scissors
model is often misused in practice because analysts tend to compute spring properties that
were derived for the Krawinkler model and use them in the Scissors model. A complete
description of the mechanical models is presented in the next section of this paper.

Results obtained from structures implementing the simple mechanical models were
compared with those computed from a detailed finite element model of a full beam-column
subassemblage. The detailed model was created using ABAQUS (3). Both elastic and
inelastic analysis was performed. It was found that good correlation was obtained between

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 121


the mechanical models and the finite element model. The results from the finite element
model were also compared to experimental results obtained during the SAC Project (4, 5)
and reasonably good correlation was obtained. The detailed analysis is briefly described in
the second main part of this paper.

MECHANICS OF BEAM-COLUMN JOINT

A typical interior beam-column subassemblage of a moment resisting frame is shown in


figure 1. The subassemblage is in equilibrium under the loads shown if the moments at the
mid-span of the girders and mid-height of the columns are zero. It is assumed that size and
span of the girders on either side of the column are same, and that a single column section
is used over the full height. The girders are welded to the column flanges. A doubler plate
may be used to reinforce the panel zone.

Terms α and β represent the ratios of the effective depth of the column to the span length,
and the effective depth of the girder to the column height, respectively. The effective depth
of a section is defined as the distance between the centers of the flanges. Use of these
terms in lieu of the actual physical dimensions greatly simplifies the derivation of the
properties of the models.
VC

VC H
VC / H
L

H βH

VVCC/HH
L
VC

αL
L

Figure 1. Typical interior beam-column subassemblage.

Total subassemblage drift

The total drift in the subassemblage, ∆, is defined as the lateral displacement of the top of
the column with respect to the bottom of the column under the load VC. Following the
procedure described by Charney (6), this drift may be divided into three components, one
for the column, one for the girder, and one for the panel zone.

∆ = ∆C + ∆G + ∆ P (1)

The column and girder displacement components are due to axial, flexural and shear
deformations occurring in the clear span region of the respective sections. The panel
contribution to displacement may also be divided into axial, flexural, and shear components:

∆ P = ∆ PA + ∆ PF + ∆ PV (2)

122 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


As stated earlier, this paper concentrates on the development of the panel zone shear
component of subassemblage displacement. It should be noted that this component
includes localized bending in the flanges of the column, but bending through the depth of
the panel is represented by ∆PF.

Panel zone participation in total subassemblage drift

If it is assumed that the moment in the girder at the face of the column is resisted entirely by
the flanges of the girder, it can be shown by simple statics that the horizontal shear force in
the panel zone is
VC (1 − α − β )
VP = (3)
β

The corresponding shear stress in the panel is

VC H (1 − α − β )
τP = (4)
∇P

This shear stress is uniform throughout the volume of the panel zone. The term ∇ P , which
represents the volume of the panel zone, appears repeatedly in the following derivations.

To determine the panel zone contribution to subassemblage drift, equal and opposite unit
virtual forces are applied in lieu of the actual column shears VC. The shear stress in the
panel due to the unit virtual shear force is

H (1 − α − β )
τ1 = (5)
∇P

The contribution of panel zone shear strain to subassemblage drift is obtained by integrating
the product of the real strains and the virtual stresses over the volume of the panel. The
uniformity of stress and strain over the volume of the panel simplifies the integration.

τ Pτ 1 VC H 2 (1 − α − β ) 2
∆ PV = ∫ dV = (6)
V
G G∇ P

The Krawinkler model and the Scissors model must have the same panel zone shear
contribution to displacement as given by equation 6.

Panel zone shear strength

Research performed by Krawinkler (2) has shown that the strength of the panel zone
consists of two components; shear in the panel itself, and flexure in the column flanges.
The larger of these components is the panel zone shear, which is resisted by the web of the
column acting in unison with the doubler plate, if present. If it is assumed that the yield
stress in shear is 1 / 3 ≅ 0.6 times the uniaxial yield stress and that the column and doubler
plate are made from the same material, the yield strength of the panel in shear is

0.6 Fy ∇ P
VYP = 0.6 FyαLt P = (7)
βH

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 123


The second component of strength arises from flexural yielding of the flanges of the column.
This phenomenon, which is most significant for W14 and W18 columns with very thick
flanges, has been observed from tests (2), and may be computed using the principle of
virtual displacements. The computed shear strength due to column flange yielding in the
joint region is
2
F y bCf t Cf
VYF = 1.8 (8)
βH

where the 1.8 multiplier is a calibration factor based on test results.

Force-deformation response

The assumed force-deformation behavior of the beam column joint is illustrated in figure 2.
In the figure the deformation is the racking displacement over the height of the panel. The
moment-rotation aspect of figure 2 is used later.

Shear, V
Moment, M
VYP
Total
βH
VYP MYP
Panel
MYP
δY

VYF βH
Flange θY
MYF

δY 4δY Shear Displacement δ


θY 4θY Spring Rotation θ

Figure 2. Force-deformation relationship for beam-column joint.

The total response is equal to the sum of the response of the panel and the column flanges.
Following Krawinkler (2) it is assumed that the flange component yields at four times the
yield deformation of the panel component. It should be noted that figure 2 shows that the
flange component of the resistance is effective immediately upon loading. Krawinkler
assumes that this component of resistance does not occur until the panel yields in shear.
We have used the relationship shown in the figure as it simplifies the implementation of the
mechanical models without compromising accuracy.

The Krawinkler and Scissors models must be proportioned such that yielding is consistent
with figure 2.

The Krawinkler model

The Krawinkler model is shown in figure 3. The model consists of four rigid links connected
at the corners by rotational springs. The springs at the lower left and upper right corners
have no stiffness, and thereby act as true hinges. The spring at the upper left is used to
represent panel zone shear resistance, and the spring at the lower right is used to represent
column flange bending resistance. A total of twelve nodes are required for the model (there
are two nodes at each corner). The number of degrees of freedom in the model depends

124 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


on the use of nodal constraints or slaving. The minimum number of DOF required to model
the panel is four in a planar structure. The maximum is twenty-eight.

Rotational Spring
for Panel Shear
Rigid Link

Real Rotational Spring


Hinge for Column Flange
Bending

Figure 3. The Krawinkler model.

The properties of the springs in the Krawinkler model are easily computed in terms of the
physical properties. Looking at only the panel spring, for example, the moment in the spring
is equal to the panel shear times the height of the panel. (See the diagram at the right of
figure 2.) The rotation in the spring is equal to the shear displacement in the panel divided
by the panel height. Hence,

M P , K = VP βH (9)

VP βH 1 V βH
θ P ,K = = P (10)
GαLt P βH G∇ P

Note that the “K” subscript in the above expressions refers to the Krawinkler model.

The stiffness of the rotational spring representing the panel in the Krawinkler model is the
moment divided by the rotation;

M P ,K
S P ,K = = G∇ P (11)
θ P ,K

The yield moment in the spring is simply the panel shear strength times the height of the
panel. Using equation 7,

M YP, K = VYP β H = 0.6 F Y ∇ PZ (12)

As seen in figure 2, the stiffness of the flange bending component of the Krawinkler model
is equal to the yield moment in the flange bending component divided by 4.0 times the yield
rotation of the panel component. The yield rotation of the spring representing the panel
component is

M YP, K FY
θ YP, K = = 0.6 (13)
K YP, K G

The yield moment is equal to the yield strength times the panel height;

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 125


M YF , K = VYF βH = 1.8FY (bCf )(tCf ) 2 (14)

and the resulting stiffness is

M YF , K
S F,K = = 0.75(bCf )(t Cf ) 2 G (15)
4θ YP, K

In summary, Expressions 11 and 12 and 14 and 15 are all that are needed to model the
panel spring and the flange spring, respectively, in the Krawinkler model. If desired, a strain
hardening component may be added.

The Scissors model

The Scissors model is shown in figure 4. This model derives its name from the fact that the
model acts as a scissors, with a single hinge in the center. Only two nodes are required to
model the joint if rigid end zones are used for the column and girder regions inside the
panel zone. The model has four degrees of freedom. As with the Krawinkler model, one
rotational spring is used to represent the panel component and the other is used to
represent the flange component of behavior.

The properties of the Scissors model are determined in terms of those derived previously
for the Krawinkler model. First, consider the displacement participation factor for panel
shear as derived in equation 6. Noting that the denominator of this equation is the same as
the panel spring stiffness for the Krawinkler model, equation 6 may be rewritten as

VC H 2 (1 − α − β ) 2
∆P = (16)
S P ,K

For the Scissors model, the moment in the spring under the column shear VC is simply VCH.
If the Scissors spring has a stiffness SP,S, the rotation in the spring is VCH/SP,S. The drift over
the height of the column is the rotation times the height, thus for the Scissors model,

VC H 2
∆ P ,Scissors = (17)
S P ,S

As this displacement must be identical to that given in equation 16, it is evident that the
relationship between the Krawinkler spring and the Scissors spring is as follows:

S P ,K
S P ,S = (18)
(1 − α − β ) 2

Similarly, when the moment in the Krawinkler spring is VPβH, the moment in the Scissors
spring is VCH. Using equations 3 and 9

M P , K = VP βH = VC H (1 − α − β ) (19)

M P,K
M P,S = (20)
(1 − α − β )

126 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


The relationships given by equations 18 and 20 hold also for the column flange components
of the models:
S F ,K
S F ,S = (21)
(1 − α − β ) 2

M F ,K
M F ,S = (22)
(1 − α − β )

Rotational Spring
For Panel Shear Rigid
Link

Real Rotational Spring


Hinge For Column Flange
Bending
Boundary of
Panel Zone

Figure 4. The Scissors model.

As an example, consider the case where α and β are 0.1 and 0.2, respectively, the Scissor
spring must be approximately twice as stiff and 1.43 times stronger than the Krawinkler
spring. Many analysts erroneously use the springs derived for the Krawinkler model in the
Scissors model. This will produce models that are more flexible than the true structure, and
that prematurely yield in the panel zone regions.

Comparisons between the Krawinkler and Scissors models

One should note from equations 18 and 20 that while the properties of the Scissors models
are dependent on the quantities α and β, those of the Krawinkler model are not. Since it
was explicitly assumed that the columns and girders on both sides of the joint are of equal
height and span, and these terms are reflected in α and β, the Scissors model may not be
used when this condition is violated. There is no such restriction on the use of the
Krawinkler model.

The deformed shape of the Krawinkler and Scissors models are shown in figure 5. In this
figure all of the deformation is assumed to be in the panel, with the girder and column rigid.
The most striking difference in the behavior between the two models is the offset in the
centrelines of the columns and girders in the Krawinkler model, which are not present in the
Scissors model.

A series of analyses were carried out using DRAIN-2DX (7) to determine the effect of the
kinematic differences on the pushover response of a series of assemblages and planar
frames which had yielding in the panel zone and at the ends of the girders. A variety of
girder spans were used, but the column height remained constant. Analysis was performed
with and without gravity load, and with and without P-Delta effects. For simple
subassemblages analyzed using the Krawinkler and the Scissors models, the pushover
responses were identical. For structures created by assembling subassemblages into a
rectilinear frame, but with real hinges at the midspan of the girders and midheight of the
columns, the pushover responses were again identical. Minor differences in the pushover
responses were obtained when the midspan/midheight hinges were removed. It was

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 127


concluded, therefore, that the Scissors model, when properly used, is generally as effective
for analysis as is the Krawinkler model, given the approximations in the derivations and the
uncertainties involved in the analysis.

Offsets

Figure 5. Kinematics of Krawinkler model (left) and Scissors model (right).

COMPARISON WITH DETAILED FINITE ELEMENT ANALYSIS AND SAC RESEARCH

To evaluate the effectiveness of the elastic modeling techniques developed above, a series
of comparisons was performed using test results provided by Ricles (5) from Phase II of the
SAC Steel Project. The properties used for one of the test specimens are provided in Table
1. Additionally, the SAC subassemblage was modeled using ABAQUS. The ABAQUS
model of the subassemblage is shown in figure 6.

Table 1. Lehigh test C1: Geometric and material properties.


Yield Stress (ksi)
Member Size Length (in.) Grade
Mill Certs. Coupon Test
56.7 flange
Girder W36x150 354 A572 Grade 50 57
62.9 web
53.2 flange
Column W14x398 156 A572 Grade 50 54
52.2 web
Continuity Plate N/A N/A N/A N/A
Doubler Plate (2) @ ¾” A572 Grade 50 57 57.1

Figure 6. ABAQUS model of SAC subassemblage.

128 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


The results of the analysis showed very good agreement between the Krawinkler, Scissors,
and ABAQUS models, and the SAC experimental results. See Downs (1) for a detailed
discussion of modeling techniques used in the ABAQUS analysis.

CONCLUSIONS

Simplified mechanical models such as the Krawinkler model and the Scissors model are
extremely effective in representing both elastic and inelastic panel zone deformations in
steel frame structures. While the Krawinkler model is considerably more complex and has
significant kinematic differences with the Scissors model, the results obtained using the two
models are essentially identical. The results from the simple mechanical models also
correlate well with more advanced finite element analysis, and with experimental results.

The Scissors model, however, is limited to use in frames with equal bays widths and equal
story heights. The Krawinkler model has no such restriction. Very significant errors will
occur in the analysis if the properties derived for the Krawinkler model are used in a
Scissors model. Such inconsistent use of the models was found in several references
reviewed by the authors.

NOTATION

α ratio of effective depth of column to span length


β ratio of effective depth of girder to column height
δY yield displacement of panel
∆ lateral displacement over height H
∆C column contribution to displacement ∆
∆G girder contribution to displacement ∆
∆P panel zone contribution to displacement ∆
∆PA axial contribution to ∆P
∆PF flexural contribution to ∆P
∆PV shear contribution to ∆P
θP,K rotation in Krawinkler spring representing panel resistance
θYP,K yield rotation in Krawinkler spring representing panel resistance
τP shear stress in panel zone due to column shear force VC
τ1 shear stress in panel zone due to unit column shear force
tP panel zone thickness, equal to column web thickness + doubler plate thickness
tCf thickness of column flange
bCf width of column flange
Fy yield stress of steel used in column and doubler plate
G shear modulus of steel
H height of column from center of story above to center of story below
L length of girder from center of bay to center of bay
MF,K moment in Krawinkler spring due to column flange bending resistance
MP,K moment in Krawinkler spring due to panel shear resistance
MYF,K yield moment in Krawinkler spring due to column flange bending resistance
MYP,K yield moment in Krawinkler spring due to panel shear resistance
MYF,S yield moment in Scissors spring due to column flange bending resistance
MYP,S yield moment in Scissors spring due to panel shear resistance
SP,K stiffness of Krawinkler spring due to panel shear resistance
SF,K stiffness of Krawinkler spring due to column flange bending resistance
SP,S stiffness of Scissors spring due to panel shear resistance
SF,S stiffness of Scissors spring due to column flange bending resistance

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 129


VC average shear force in columns above and below the joint
VP horizontal shear force in panel zone
∇P volume of panel zone = αLβHtP

REFERENCES

(1) Downs, William M, (2002). Modeling and Behavior of the Beam/Column Joint Region
of Steel Moment Resisting Frames, M.S.Thesis, Department of Civil and Environmental
Engineering, Virginia Tech, Blacksburg, Virginia.
(2) Krawinkler, H., (1978), “Shear in Beam-Column Joints in Seismic Design of Frames”,
Engineering Journal, v15, n3, American Institute of Steel Construction, Chicago,
Illinois.
(3) Hibbit, Karlson, and Sorensen, (2001). ABAQUS User’s Manual, Verson 6.2.
(4) FEMA (2000). Recommended Seismic Design Criteria for New Steel Moment Frame
Buildings, FEMA-350. Federal Emergency Management Agency, Washington D.C.
(5) Ricles, J. M., (2002). “Inelastic Cyclic Testing of Welded Unreinforced Moment
Connections” Journal of Structural Engineering, ASCE, v128, n4.
(6) Charney, Finley A., (1993). "Economy of Steel Frame Buildings Through Identification
of Structural Behavior", Proceedings of the Spring 1993 AISC Steel Construction
Conference, Orlando, Florida.
(7) Prakesh, V. and Powell, G. H., (1993). DRAIN 2D-X Users Guide, University of
California, Berkeley, California.

130 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


STATISTICAL EVALUATION OF ROTATION CAPACITY OF MOMENT
CONNECTIONS
D. Beg, University of Ljubljana, Slovenia
E. Zupančič, University of Ljubljana, Slovenia

ABSTRACT

The paper describes the reasons that cause rotation capacity of moment
connections strongly dependent on material properties of components of such
connections. Then a statistical evaluation of the rotation capacity of typical
moment connections with respect to the variation of the material properties of
individual components was performed. The Monte Carlo method was used and
10,000 calculations were performed in each analysis. Log-normal distribution
was assumed for yield strength fy and ultimate strength fu with the range of
typical mean values and coefficients of variation from literature. The results
show that considerable variation in rotation capacity is obtained in some cases.

INTRODUCTION

Besides strength and stiffness, rotation capacity is also very important characteristic of
moment connections of steel structures. Until recently there have been no general methods
for the determination of rotation capacity available, except tests and sophisticated FE
analysis. Also EN 1993-1-8 (1), a very detailed standard on connections in steel structures,
only briefly addresses this topic by giving some requirements to achieve sufficient ductility.
For these reasons it is not surprising that several research groups are working on this
problem (2, 3, 4, 5) and the authors of this paper together with I. Vayas have recently
developed an analytical method that enables the assessment of rotation capacity of moment
connections based on a component approach from EN 1993-1-8 (6, 7). From test results and
extensive numerical simulations for each relevant component, deformation capacities and
simplified bilinear force-displacement relations were established. By suitable mechanical
model the components were assembled to model connection behaviour and to determine the
rotation capacity.

Studying the problem and applying the developed method, it became evident that rotation
capacity is very sensitive to the changes of basic material properties, yield stress fy and
tensile strength fu, of individual components. It was observed that nominal values of fy and fu
that are usually known at the design stage do not always give satisfactory results, as the real
values are usually higher and most likely also some differences between individual
components take place.

DESCRIPTION OF THE PROBLEM

The rotation capacity of the joint is governed by its weakest component. When the weakest
component reaches its ultimate resistance, also the resistance of the joint is exhausted. This
is exactly true for a component method, but also in reality (tests, FE simulations) there is very
limited possibility for shedding of forces between different parts of the connection, and the
resistance of the connection usually drops after the resistance of the critical part is reached.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 131


This means that the critical component contributes to the rotation capacity of the connection
its full deformation capacity and other components contribute only deformations reached at
the same loading level (see Figure 1).

Figure 1. Contribution of components to the rotation capacity.

One of the most important issues when determining the rotation capacity is the strength of
components and especially relative difference in strength between components. This is
demonstrated in Figure 2. For the sake of simplicity the connection is composed of only two
components, represented by bilinear inelastic force-displacement diagrams, one in tension
and one in compression (Figure 2a).

a) Mechanical model of the connection.

b) Large rotation capacity c) Small rotation capacity

Figure 2. Influence of strength of components on the rotation capacity.

132 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


In the first case the resistance of both components is similar, with component C2 being
decisive (Figure 2b). As component C1 is in inelastic state when the resistance of C2 is
reached, its contribution to the rotation capacity is large. At slightly higher resistance of
component C1 (Figure 2c) its contribution to the rotation capacity decreases considerably, as
this component remains in the inelastic state.

Such scenario can easily happen in reality. At the design stage only characteristic values of fy
and fu are known and the rotation capacity is determined based on these values (case in
Figure 2b). Real values of fy and fu are higher than the characteristic values, as characteristic
values are more or less lower guaranteed values. For grade S235 the increase of strength
goes up to 30% and for S355 the increase is somewhat smaller. The case in Figure 2c can
be regarded as a situation, where the resistance of component C1 was increased due to
differences between characteristic and real material properties, while the resistance of
component C2 remained almost unchanged. The result is much smaller rotation capacity. In
this case the actual behaviour is on the unsafe side compared to the design values.

Certainly, it is possible to find the opposite situation, where the real behaviour is on the safe
side. The main conclusion is that the rotation capacity is very sensitive to the variability of
material strength parameter. This is not a specific problem of a component method, but is
immanent to the behaviour of joints when ductility is considered.

STATISTICAL PARAMETERS AND METHODS

Statistical analysis of typical moment connections was performed to assess how the variation
of basic material properties fy and fu of individual components influences the rotation capacity
of a connection. For basic random variables yield strength of column web, column flange,
end-plate and tensile strength of bolts was selected. Log-normal distribution was applied to
all random variables with the following four combinations of mean my and coefficient variation
vy:
• A – my = 1.18 fy, vy = 0.08
• B – my = 1.18 fy, vy = 0.10
• C – my = 1.12 fy, vy = 0.05
• D – my = 1.12 fy, vy = 0.08
where fy is a nominal value from EN 1993-1-1.

Different authors (8, 9, 10) also propose log-normal distribution and case A as the most
suitable for structural steel. Case B gives larger scatter and cases C and D smaller scatter of
random variables than case A. For bolts of grade 8.8 in all calculations case C was taken into
account to allow for more favourable statistical distribution of tensile strength in this case. In
addition to the analysis with all four random variables, for case A also the Monte-Carlo
simulation with only one active random variable at a time was performed. Other parameters
were kept constant at their nominal (characteristic values).

DATA ON CONNECTIONS

Five different end-plate moment beam-to-column connections were analysed. They were
designed for the purpose of statistical analysis, each with different decisive component with
the smallest resistance. At the fifth connection all components have approximately equal
resistance. The relevant data on connections are given in Table 1 and Figure 3.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 133


Table 1. Data on connections 1 to 5.
Connections 1 2 3 4 5
Column (S235) HEAA 240 HEA 300 HEB 200 HEA 240 HEA 240
Beam (S235) IPE 300 IPE 300 IPE 300 IPE 300 IPE 300
End plate (S235) 380/170/12 380/190/20 380/190/15 380/190/10 380/150/15
Bolts 8.8 M 16 M 20 M 20 M 16 M 16
w [mm] 110 110 110 110 90
x [mm] 60 60 60 60 60
e1 [mm] 30 30 30 30 30
p1 [mm] 80 80 80 80 80
p2 [mm] 200 200 200 200 200
Critical component CF CWC SWC EP CWC
CWC-Column web in compression, CF-Column flange in bending
SWC-Column web panel in shear, EP- End plate in bending

Figure 3. Geometry of connections.

DETERMINISTIC ANALYSIS OF CONNECTIONS

For all five connections rotation capacity ϕu was determined at nominal values of material
parameters with two methods: FE simulation (ABAQUS (11), solid finite elements) and own
analytical method (6), where the resistance of components was determined in two ways –
according to EN1993-1-8 and by replacing fy with fu when relevant (6) to get more accurate
resistances. Comparison to numerical simulations (case c in Table 2) shows very good
results when more realistic resistance of components is used – case b in Table 2 (very good
agreement in three cases and conservative results in two cases), but some unsafe results
when resistances from EN1993-1-8 were used (case a in Table 2).

Criteria for reaching ultimate rotation ϕu were stated by maximum equivalent plastic strains
reached in different parts of connection: 0.1 for column web in tension, compression on
shear and 0.2 for column flange and end-plate in bending.

RESULTS OF STATISTICAL ANALYSIS

Due to the limited number of pages of this paper it is not possible to present all the results
obtained from the analysis. For connection 1 more results are given and for other
connections only some results for case A of statistical parameters are presented (only for
case a in Table 2).

134 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Table 2. Results for rotation capacity.
a) Analytical – EC3 b) Analytical - fu c) Numerical simulations
Connec- ϕu ϕu ϕu
Failure Failure Failure
tion
mode [rad ] mode [rad ] mode [rad ]
1 CF 0.055 CF 0.065 CF 0.115
2 CWC 0.133 CWC 0.081 CWC 0.087
3 SWC 0.148 CF 0.142 CWC 0.134
4 EP 0.045 EP 0.040 EP 0.103
5 CWC 0.175 CWC 0.104 EP 0.104
CWC-Column web in compression, CF-Column flange in bending
SWC-Column web panel in shear, EP-End plate in bending

250 SWC

200 CWC

150 TWC

EP
100
CF
50

d ,f
2.5 5 7.5 10 12.5 15

Figure 4. Connection 1: M-d diagram for components.

Connection 1

From moment-displacement diagram (Figure 4) it is clear that column flange evidently has
the lowest strength and should influence the rotation capacity to a large extent. In Figures 5
to 11 the calculated values of rotation capacity are plotted in diagrams M - φ (left) and the
obtained distribution function of the rotation capacity is plotted at the right side. From these
results the following observations can be made:

- The results for the ultimate capacity are in the range of 0.05 – 0.138 rad with the most
probable result around 0.057 rad (0.055 rad at nominal values).
- Comparison of results for cases A, B and C shows that the results are very similar, but
clearly reflect the statistical input parameters. Case B (Figure 6) gives the worst results
and case C (Figure 7) gives the most favourable results.
- From Figures 9 – 11 it is evident that only the change of material properties of column
flange strongly influences the rotation capacity, while other components have no
influence. This is expected because their resistance is much larger than the resistance of
column flange, and therefore they do not contribute significantly to the rotation capacity of
the connection.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 135


MRd n

2500
100
80 2000

60 1500

40 1000
Monte Carlo
20 EC3 - 1- 8 500
Abaqus
f Cd f Cd
0.025 0.05 0.075 0.1 0.125 0.15 0.025 0.05 0.075 0.1 0.125 0.15

Figure 5. Connection 1: case A, all random variables active.

MRd n
80 3000
2500
60
2000

40 1500

1000
20 Monte Carlo
EC3 - 1- 8 500
f Cd f Cd
0.025 0.05 0.075 0.1 0.125 0.15 0.025 0.05 0.075 0.1 0.125 0.15

Figure 6. Connection 1: case B, all random variables active.

MRd n
70 4000
60
3000
50
40
2000
30
20 Monte Carlo 1000
10 EC3 - 1- 8

f Cd f Cd
0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.02 0.04 0.06 0.08 0.1 0.12 0.14

Figure 7. Connection 1: case C, all random variables active.

MRd n

7000
50 6000
40 5000
30 4000
3000
20
Monte Carlo 2000
10 EC3 - 1- 8
1000
f Cd f Cd
0.02 0.04 0.06 0.08 0.02 0.04 0.06 0.08

Figure 8. Connection 1: case A, only column web random variable.

136 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


MRd n
700
60 600
50 500
40 400
30 300
20 Monte Carlo 200
10 EC3 - 1- 8
100
f Cd f Cd
0.025 0.05 0.075 0.1 0.125 0.15 0.025 0.05 0.075 0.1 0.125 0.15

Figure 9. Connection 1: case A, only column flange random variable.

MRd n

50
8000
40
6000
30
4000
20
Monte Carlo
10 EC3 - 1- 8 2000

f Cd f Cd
0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.01 0.02 0.03 0.04 0.05 0.06 0.07

Figure 10. Connection 1: case A, only end-plate random variable.

MRd n

50
8000
40
6000
30
4000
20
Monte Carlo
10 EC3 - 1- 8 2000

f Cd f Cd
0.01 0.02 0.03 0.04 0.05 0.06 0.01 0.02 0.03 0.04 0.05 0.06

Figure 11. Connection 1: case A, only bolts random variable.

MRd n
200
1500

150 1250
1000
100
750
Monte Carlo
50 EC3 - 1- 8 500
Abaqus
250
f Cd f Cd
0.025 0.05 0.075 0.1 0.125 0.15 0.025 0.05 0.075 0.1 0.125 0.15 0.175

Figure 12. Connection 2: case A, all random variables active.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 137


MRd n
175
150 1200

125 1000

100 800
75 600
Monte Carlo
50 400
EC3 - 1- 8
25 Abaqus 200
f Cd f Cd
0.05 0.1 0.15 0.2 0.05 0.1 0.15 0.2

Figure 13. Connection 3: case A, all random variables active.

MRd n

2000
100
80 1500

60
1000
40 Monte Carlo
EC3 - 1- 8 500
20 Abaqus
f Cd f Cd
0.05 0.1 0.15 0.2 0.05 0.1 0.15 0.2

Figure 14. Connection 4: case A, all random variables active.

MRd n
2500
120
100 2000

80 1500
60 EC3 - 1- 8
SS tlak 1000
40 PS upogib
ÈP upogib 500
20 Abaqus
f Cd f Cd
0.05 0.1 0.15 0.2 0.05 0.1 0.15 0.2

Figure 15. Connection 5: case A, all random variables active.

Connections 2 to 5

At connection 2 column web in compression is decisive, but the strength of other


components is not much higher. The influence of the decisive component is not so
pronounced as at connection 1. The most probable value of the rotation capacity is around
0.103.

At connection 3 column web in shear is decisive and the results are similar as at connection
1.

Also at connection 4 one component, end-plate, has evidently lower resistance and the
influence of this component is large, while other components have little influence.

138 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Connection 5 was designed such that all components have approximately equal resistance,
only column web in tension is stronger. The scatter of results for the rotation capacity is very
large (0.057 – 0.178) and qualitatively the results are similar as for connection 2 where also
the resistance of its components does not differ much.

CONCLUSIONS

Rotation capacity of moment connections depends on the properties of several components


and on the interaction of these components. Change in strength of one component can
produce large change in the calculated rotation capacity. As at the design stage a designer
does not know the real value of material properties of connections, but only nominal lower
bound values, when fabricated, connections may have considerably different rotation
capacity compared to the calculated value.

Obviously, statistical analysis can give more information about realistic behaviour in such
conditions. For our own analytical method for the calculation of the rotation capacity we
performed statistical evaluation, where material properties of individual components were
random variables.

The following conclusions result from the statistical analysis:

- Scatter of results increases with more unfavourable statistical parameters.


- When one component has a considerably lower strength than other components, then
only this component influences importantly the rotation capacity.
- When one component is evidently decisive, then the distribution function for the rotation
capacity is asymmetric with a tail on the side of larger values and the most frequent value
lies close to the lower bound.
- When more components have approximately equal strength, then the distribution function
is asymmetric to the opposite side or almost symmetric and the most frequent value is
closer to the upper bound.
- Values of rotation capacity, calculated on deterministic basis with nominal values (lower
bound values) of material parameters, have a tendency to be close to the most frequent
value. As the scatter of results is rather large this does not mean that the solution at
nominal parameters is always a good estimate and on the safe side. Depending on the
relations between components, the probability that the rotation capacity lies within (φu
NOMINAL ± 0.2 φu NOMINAL) is assessed to 50% to 80% in the analysed cases.

These results show that the deterministic approach to the calculation of the rotation capacity
based on nominal values of material properties may give unsatisfactory results. A possible
solution could be statistical approach. Based on this approach reduction factors could be
determined to be used in the simple deterministic calculation of the rotation capacity.

REFERENCES

(1) Eurocode 3: Design of steel structures, Part 1.8: Design of joints, prEN 1993-1-8: 2000.
(2) Coelho A., Bijlaard F., da Silva L. On the deformation capacity of beam-to-column
bolted connection. Report AG-XXIII, ECCS-T10, April 2002, Ljubljana. 2002.
(3) Huber G., Tschemmernegg F. Modeling of beam-to-column joints. Journal of
Constructional Steel Research, Vol. 45, 199-21. 1998.
(4) da Silva L. S., Coelho A. M. G. A ductility model for steel connections. Journal of
Constructional Steel Research, Vol. 57, 45-70. 2001.
(5) Kuhlmann U., Sedlacek G., Kuhnemund F., Stangenberg H. Mitteilungen: Vorhandene

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 139


rotationskapazitat wirtschaftlicher anschluskonstruktionen auf der basis des
komponentenverhaltens fur die anwendung plastischer bemessungskonzepte im
stahlbau. Institut fur konstruktion und entwurf stahl, holz- und verbundbau, Universitat
Stuttgart, 2002.
(6) Beg D., Zupančič E., Vayas I. On the rotation capacity of moment connections. Journal
of Constructional Steel Research, Vol. 60, 601-620. 2004.
(7) Zupančič E. PHD Thesis: Rotation capacity of steel connections. University in Ljubljani,
Faculty for Civil and Geodetic engineering. Ljubljana, January 2004 (in Slovene lang.).
(8) Gulvanessian H., Holicky M., Markova J. Calibration of Eurocode reliability elements
considering steel members. Eurosteel conference 2002, Coimbra Portugal, 1511-1520.
2002.
(9) Schleich J. B., Sedlacek G., Kraus O. Realistic safety approach for steel structures.
Eurosteel conference 2002, Coimbra Portugal, 1521-1530. 2002.
(10) Fajkus M., Melcher J., Holicky M., Rozlivka L., Kala D. Design characteristics of
structural steels based on statistical analysis of metallurgical products. Eurosteel
conference 2002, Coimbra Portugal, 1541-1550. 2002.
(11) ABAQUS User's Manual, Version 5.8, 6.1, 6.2 & 6.3, Hibbitt, Karlsson and Sorensen
Inc., USA, 2002.

140 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


ROTATION CAPACITY OF MR BEAM-TO-COLUMN JOINTS
UNDER CYCLIC LOADING

D. Grecea, Politehnica University of Timisoara, Romania


A. Stratan, Politehnica University of Timisoara, Romania
A. Ciutina, Politehnica University of Timisoara, Romania*
D. Dubina, Politehnica University of Timisoara, Romania

ABSTRACT
Actual design codes and recommendations for ductile Moment Resisting Steel
Frames in seismic zones require to beam-to-column joints to sustain a total
rotation of 0.035 rad (in Europe) to 0.040 rad (in United States). The problem is
that in codes, there are no analytical methods to predict the rotation capacity of
MR connections, not even for monotonic loading, which is simpler then
compared with seismic action of cyclic type. Even if last years appeared some
proposals for analytical or numerical based methods to evaluate rotation
procedure for their qualification in terms of ductility, still remain the pre-
qualification tests, both for static and seismic actions.
In the present paper, based on extended experimental studies, comparatively
conducted under monotonic and cyclic loading on different types of
connections, both for steel and composite MR frames, a reduction factor to be
applied to monotonic rotation capacity is proposed in the purpose of evaluation
of cyclic rotation capacity.

INTRODUCTION

Modern design codes characterise the behaviour of a connection by three basic parameters
i.e. strength, stiffness and ductility. In MR connection the ductility is measured in terms of
rotation capacity. From this point of view, connections may be classified, similar to the
classification of sections (EN 1993-1-1 [1]) in terms of their ductility (EN 1993-1-8 [2]) as
follows:
Class 1 joints: Ductile joints.
A ductile joint is able to develop its plastic moment resistance and
to exhibit a sufficiently large rotation capacity.
Class 2 joints: Joints of intermediate ductility.
A joint of intermediate ductility is able to develop its plastic moment
resistance but exhibits only a limited rotation capacity once this
resistance is reached.
Class 3 joints: Non ductile joints.
Premature failure (due to instability or to brittle failure of one of the
joint components) occurs within the joint before the moment
resistance based on a full plastic redistribution of the internal forces
is reached.

The parameters of this classification are shown in Figure 1 (EN 1993-1-8 [2]).
* at present Assoc. Prof. to INSA Rennes France

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 141


Mj

Mj,u Mel,Rd – elastic design moment resistance


Mpl,Rd Mpl,Rd – plastic design moment resistance
Sj,ini – initial rotational stiffness
Mel,Rd Φel – elastic rotation
Φtr – another rotation necessary to get to the
Sj,ini
plastic design moment resistance Mpl,Rd
Φpl – plastic rotation ( Φ pl = Φ Cd − Φ Xd )
Φel ΦXd ΦCd Φ
Φtr Φpl

Figure 1. Characteristic parameters for a beam-to-column connection.

For welded beam-to-column connection, in which the web is stiffened in compression, but
unstiffened in tension, under monotonic load only, EN 1993-1-8 [2] provide the following
formula to estimate the design rotation capacity, ΦCd:
Φ Cd = 0.025 h c h b (1)

where hb is the depth of the beam and hc is the depth of the column.

The deformation capacity of components has been studied by several researchers. Faella at
al. [19] carried out tests on T/stubs and derived analytical expression for the deformation
capacity of this component. Kuhlmann and Kuhnemund [22] performed tests on the column
web subjected to transverse compression at different levels of compression axial force in the
column. Some authors have tried to extract the information of the behaviour of single
components from the tests on a whole joint. Bose et al. [14,15] determined only the strength
of the most important components, while da Silva et al. [23] tried to determine all three
important parameters, stiffness, strength and deformation capacity, at different levels of axial
force in a beam.

Recently, Beg et al. [12] proposed a simple analytical method for calculation of rotation
capacity of MR connection which is compatible with the “component method” used in EN
1993-1-8 to provide the stiffness and strength.

Also Gioncu [20] suggested to use the “local mechanism method” to evaluate the rotation
capacity of top-and-seat cleat bolted MR connections. This method could be used for welded
connections, but probably not for extended-end-plate bolted connections, or other types.

However, even, as shown, there are tentatives to find analytical methods for evaluation of
monotonic rotation capacity, some of these referred in this paper being promising, the basic
procedure provided by design codes on this purpose still remain the prequalification tests.

In case of seismic resistant connection the problem is much more complex, and if the
relevant design codes have specific provisions to request precisely the lower bound values
for the necessary rotation capacity, they do not provide a specific evaluation methodology,
except, again testing.

DUCTILITY OF CONNECTIONS IN SEISMIC DESIGN CODES

EN 1998-1 [3] has introduced three levels of structural ductility class in connection with
design concepts and range of reference values of the behaviour factors as presented in
Table 1.

142 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


For the two structural ductility classes of the dissipative structural behaviour, specific
requirements are introduced concerning structural ductility (behaviour q factor), element
ductility (cross sectional class), material (yield strength and toughness) and joint ductility
(rotation capacity).

Concerning the rotation capacity, according to EN 1998-1 [3], connections in dissipative


zones shall have sufficient overstrength to allow for yielding of the connected parts.

Table 1. Structural ductility classes.


Structural ductility Range of the reference values of the
Design concept
class behaviour factor q
Low dissipative structural
DCL (Low) ≤ 1,5 - 2
behaviour
≤4
Dissipative structural DCM (Medium)
also limited by the values in EN 1998-1
behaviour
DCH (High) only limited by the values in EN 1998-1

Dissipative semi-rigid and/or partial strength connections are permitted, provided that all of
the following conditions are satisfied: a) the connections have a rotation capacity consistent
with the deformations; b) members framing into the connections are demonstrated to be
stable at ultimate limit state (ULS); c) the effect of connections deformation on global drift is
taken into account using non linear static (pushover) global analysis or non linear time history
analysis.

The overstrength condition for connections need not apply if the connections are designed in
a manner enabling them to contribute significantly to the energy dissipation necessary to
achieve the chosen q-factor.

The moment frame connections design should be such that the plastic rotation capacity θp in
the plastic hinge is not less than 35 mrad for structures of ductility class DCH and 25 mrad
for structures of ductility class DCM with q>2.

Even with the limits mentioned above for the rotation capacity of joints, EN 1998-1 does not
specify any formula for this evaluation except testing and design experience.

According to American codes UBC-97 [5] or AISC-2002 [6], the purpose of the earthquake
provisions is primarily to safeguard buildings against major structural failures and loss of life,
not to limit damage or maintain function.

Structures and portions thereof shall, as a minimum, be designed and constructed to resist
the effects of seismic ground motions.

Moment Frame (MF) defined as a building frame system in which seismic shear forces are
resisted by shear and flexure in members and connections of the frame are divided as
follows (AISC-2002): Special Moment Frame (SMF), Intermediate Moment Frame (IMF) and
Ordinary Moment Frame (OMF).

Even if the rotation capacity of the beam-to-column joints is connected with the classification
of frames, the AISC code is not providing any formula for the evaluation of this very important
characteristic.

Anyhow, American FEMA 350 document [8], in Chapter 3: Connection Qualification provides
pre-qualification data and design procedures for alternative types of welded, fully restrained,

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 143


steel moment-frame connections, suitable for use in new construction. This pre-qualification
is extremely important, because a designer is able to choose a type of connection and then
to follow the instructions concerning the cross sectional dimensions, in order to obtain a
certain ductility class of the joint.

Post-Northridge US steel moment connection types, such as welded un-reinforced


flange/welded-web (WUF-W), and reduced beam section (RBS), have been pre-qualified in
an extensive series of SAC Joint Venture tests. These connections, described in FEMA 350
design document, are designed to sustain a total rotation of 0.04 rad before significant
strength degradation and 0.06 rad before complete loss of resistance (Table 3-15 of [3]).

After Stojadinovici [24], failure modes of pre-qualified US connections can be classified


based on observed behaviour during pre-qualification tests, into three categories:
1. beam instability, comprising local beam flange or web buckling and beam lateral-
torsional buckling;
2. lateral-torsional buckling of the column or the entire connection sub-assembly; and
3. low-cycle connection fatigue, when the connection fails by tearing of the base metal
after a small number of large amplitude cycles.

The Japanese code AIJLSD-90 [7] provides specifications concerning only the frame
classification, according to the classification of members and member cross-sections, storey-
drift and overstrength of joints in comparison with the connected members. Any specification
concerning the rotation capacity of beam-to-column joints is not made.

ROTATION CAPACITY: EVALUATION BY TESTS

In this chapter four different experimental studies aiming to evaluate performances of both
steel and composite beam-to-column joints under cyclic loading will be summarised.

Rotation capacity for welded joints (INSA Rennes)

Aribert & Grecea [11] have developed an experimental research program at INSA Rennes.
This research program dealt with 8 beam-to-column welded joints under monotonic and
repeated cyclic loading. The specimens were major axis joints with a symmetrical cruciform
arrangement comprising an H or I column connected to two cantilever beams. No transverse
stiffener was welded in the compression zone of the column web.

The tests were performed according to the Recommended Testing Procedure of ECCS
(1985). Each type of joint was subject first of all to a monotonic loading (CPP11, CPP13,
CPP15 and CPP17, then to cyclic reversal loading (CPP12, CPP14, CPP16 and CPP18).
The cyclic moment-rotation curves for the four tested joints are presented in Figure 2.
Experimental values of ultimate resistance moment and rotation capacity obtained in both
monotonic and cyclic reversal loading can be compared in Table 2.

From the moment-rotation curves, it is observed that the ultimate moment and the initial
stiffness of the joints are not strongly influenced by the repeated cyclic loading, so that in
seismic design the corresponding formulae given in EN 1993-1-8 for the case of static
loading can be used, as reasonable approximations. On the other hand it appears clearly
that the rotation capacity of the joints is systematically reduced by a factor about 2.

144 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Test CPP 12 Test CPP14
300 M [kNm] 200 M [kNm]

200 150

100
100
50
Φ [rad] Φ [rad]
0 0
-0,03 -0,02 -0,01 0 0,01 0,02 0,03 0,04 -0,04 -0,03 -0,02 -0,01 0 0,01 0,02 0,03 0,04
-50
-100
-100
-200
-150

-300 -200

Test CPP16 Test CPP18


400 M [kNm]
600 M [kNm]
300
400
200

100 200
Φ [rad] Φ [rad]
0 0
-0,02 -0,01 0 0,01 0,02 0,03
-100 -0,05 -0,04 -0,03 -0,02 -0,01 0 0,01 0,02
-200
-200
-400
-300

-400 -600

Figure 2. Cyclic behaviour of the joints.

Table 2. Comparison between joint characteristics under cyclic and monotonic loadings.
CPP11 CPP12 CPP13 CPP14 CPP15 CPP16 CPP17 CPP18
Mu [kNm] 230.0 253.0 166.9 180.3 349.2 368.5 467.5 486.2
Φu [rad] 0.064 0.031 0.045 0.023 0.045 0.020 0.052 0.030

Rotation capacity for welded and bolted joints (UP Timisoara)

Investigations on beam-to-column joints, carried out at the laboratory of steel structures at


the Civil Engineering Faculty of Timisoara are presented hereafter.

Three typologies of beam-to-column joints have been tested from a total of 12 specimens. All
the joints are double sided. For all the specimens the design steel grade was S235 (fy=235
N/mm2, fu=360 N/mm2), beams being IPE 360 and columns HEB 300.

Two types of loading were applied: symmetrical and anti-symmetrical and three connection
typologies were tested (Figure 3).

Column HEB 300 Column HEB 300 Column HEB 300

10M20 gr 10.9 Equal strength weld 3M20 gr6.6 Supplementary web plate

Beam IPE360 Beam IPE360


Beam IPE360

Equal Strenght weld Welded cover plate


End Plate t=20

(a) EP (b) W (c) CWP


Figure 3. Connection configurations: (a) bolted, (b) welded and (c) with cover welded plate.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 145


XS-EP specimens – Symmetrical Joints Anti-Symmetrical Joints
400 350
XS-EP2 300
XU-EP1
250

COLUMN FACE [kNm]


COLUMN FACE [kNm]

200

MOMENT AT THE
150
MOMENT AT THE

100
50

0
-0.06 -0.04 -0.02 -50 0 0.02 0.04 0.06
-0.06 -0.04 -0.02 0 0.02 0.04 0.06
-100
-150
-200
-250
-300
-350
-400 TO TAL JOINT RO TATIO N [rad]
TOTAL JOINT ROTATION [Rad]

XS-W specimens – Symmetrical Joints Anti-Symmetrical Joints


5 00 350
X S -W 2 4 00 X U -W 1
250
3 00

COLUMN FACE [kNm]


COLUMN FACE [kNm]

150

MOMENT AT THE
MOMENT AT THE

2 00

1 00 50
0
-0 .0 6 -0 .0 4 -0 .0 2 -5 0 0 0 .0 2 0 .0 4 0 .0 6
-0.0 15 -0 .0 1 -0.0 05 0 0.0 05 0 .0 1 0.0 15 0 .0 2
-1 00
-1 5 0
-2 00

-3 00 -2 5 0
-4 00
-3 5 0
-5 00
T O T A L J O IN T R O T A T IO N [ra d ]
TO T A L J O IN T R O T A TIO N [ra d .]

XS-CWP specimens – Symmetrical Joints Anti-Symmetrical Joints


600 350

X S -C W P 1 X U -C W P 1 250
400
COLUMN FACE [kNm]
COLUMN FACE [kNm]

150
MOMENT AT THE
MOMENT AT THE

200
50
0
-0 .0 4 -0 .0 2 0 0 .0 2 0 .0 4 0 .0 6 -0 .0 7 -0 .0 5 -0 .0 3 -510
-0 .0 0 .0 1 0 .0 3 0 .0 5 0 .0

-2 0 0 -1 5 0

-4 0 0 -2 5 0

-3 5 0
-6 0 0
T O T A L J O IT R O T A T IO N [ra d .] T O T A L J O IN T R O T A T IO N [ra d ]

Figure 4. Cyclic moment-rotation curves for test specimens.

The joints (three different joint configurations and two load types) have been designed
according to EN 1993-1-8. The loading history was made according to the ECCS
Recommendations simplified procedure. The cyclic testing curves are presented in Figure 4.

Table 3. Comparison between computed and experimental joint characteristics.


M (exp)
j , Rd M (j th, Rd) S (exp)
j ,ini S (j th,ini) φ y(exp) φu(exp)
SPECIMEN
[kNm] [kNm] [kNm/rad] [kNm/rad] [rad] [rad]
XS-EP 1 255.6 262.7 69539 142932.2 0.0038 0.033
XS-EP 2 288.9 261.3 44205 140886.8 0.0063 0.038
XS-W 1 305.6 309.3 333953 ∞ 0.0009 0.029
XS-W 2 277.8 317.6 321569 ∞ 0.0009 0.016
XS-CWP 1 316.7 452.1 366309 ∞ 0.0009 0.038
XS-CWP 2 ** 449.0 ** ∞ ** **
XU-EP 1 146.7 169.2 44081 43727.2 0.0033 0.060
XU-EP 2 157.8 169.1 49004 43718.2 0.0028 0.062
XU-W 1 113.3 163.6 63102 68792.1 0.0020 0.052
XU-W 2 131.1 164.1 49681 69062.1 0.0026 0.052
XU-CWP 1 131.1 178.6 60712 75597.2 0.0022 0.064
XU-CWP 2 164.4 177.4 58453 74963.1 0.0026 0.060

146 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Table 3 comprises the results of the experimental tests compared to that of EN 1993-1-8, in
terms of joint bending moments, rotational stiffness and ultimate rotation attained.

Comparing the experimental and computed values of joint moment capacity, it can be
observed that generally, close values are obtained for the XS series. In the case of XU
series, all experimental values are lower than the ones computed by EN 1993-1-8.

In what concerns the initial stiffness of the joints, numerical and experimental results agree
fairly well for the XU series, while significant differences are noticed for XS series. Anyway,
stiffness is much lower for the anti-symmetrical joints both from experimental and computed
stiffness values. This fact is again given by the deformability of the panel zone.

Joint rotations are considerably higher for XU series, generally in the ratio of 1:2. Improved
ductility in the case of XU joints is given by good rotation capacity and stable hysteresis
loops of the web panel in shear. Anti-symmetrical joints have generally increased energy
dissipation capacity with respect to the symmetrical ones.

Rotation capacity for bolted steel and composite joints (UP Timisoara)

Column
Column 8 top & bot. fl.
r.c. slab
8M20 gr.10.9 Beam

120
R20 R20
290

290
Beam
12

12

5 .
5 . 6M20 gr.10.9
8 top & bot. fl.

170
2Ø12 PC52
14
600

3Ø10 PC52

2Ø12 PC52
20 360 20

7Ø12 PC52 2Ø10 PC52 7Ø12 PC52

(a) (b)
Figure 5. Connection configurations for (a) BX-S series and (b) BX-C series of joints.

An alternative to the “standard” European column cross-section (hot-rolled I profiles) is the


use of X-shaped cross-sections, built-up of two hot-rolled profiles welded along the median
axis or built-up sections made out from welded plates, as shown in Figure 5.

The testing program comprised six specimens: three joints under symmetrical loading (BX-
SS and BX-CS), and three joints under anti-symmetrical loading (BX-SU and BX-CU). Tests
were performed in accordance with the ECCS Recommendations complete procedure
(ECCS 1985). The first specimen from each series was tested monotonically, The load was
applied quasi-statically, under displacement control.

Figures 6, and 7 synthetically present the main experimental results expressed in terms of
maximum plastic rotation ϕ max and maximum moment attained M max , for steel and
composite joints. Tables 4 and 5 present the joint characteristics obtained from the tests and
computed analytically in accordance to EN 1993-1-8 and EN1994-1-1 section 8, respectively.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 147


The experimental curves of the monotonic tests, as well as the envelopes of the cyclic tests,
are compared to the analytical results given by the EN 1993-1-8. In the case of symmetrically
loaded joints, it can be observed that the values of the experimental stiffness are close to
those computed by EN 1993-1-8, but the values of the computed yielding bending moment is
generally 10-20% smaller than the experimental ones. For the anti-symmetrically loaded
joints, only the full-shear area approach gives close agreement to the analytical results given
by EN 1993-1-8. The monotonic test is different from the cyclic tests in stiffness, maximum
resistance and maximum rotation, as can be seen in Table 4. As usual, the maximum
rotation in monotonic tests is 1.5-2 times greater than the maximum rotation attained under
cyclic loading. The conventional values of yielding moment and of yielding rotation for the
case of load reversal are closer to the computed values by EN 1993-1-8.
400 400

BX-SS-C1 300 BX-SU-C1 300


COLUMN FACE [kNm]

COLUMN FACE [kNm]


200 200
MOMENT AT THE

MOMENT AT THE
100 100

0 0
-0.1 -0.08 -0.06 -0.04 -0.02 0 0.02 0.04 0.06 0.08 0.1 -0.1 -0.08 -0.06 -0.04 -0.02 0 0.02 0.04 0.06 0.08 0.1
-100 -100

-200 -200

-300 -300

-400 -400
TOTAL JOINT ROTATION [rad] TOTAL JOINT ROTATION [rad]

(a) (b)
Figure 6. Moment - rotation relationships for cyclic specimens of the BX-S series.
350 350

BX-CS-C1 250
BX-CU-C1 250
COLUMN FACE [kNm]

COLUMN FACE [kNm]

150 150
MOMENT AT THE

MOMENT AT THE

50 50

-0.1 -0.08 -0.06 -0.04 -0.02


-50 0 0.02 0.04 0.06 0.08 0.1 -0.1 -0.08 -0.06 -0.04 -0.02
-50 0 0.02 0.04 0.06 0.08 0.1

-150 -150

-250 -250

-350 -350
TOTAL JOINT ROTATION [rad] TOTAL JOINT ROTATION [rad]

(a) (b)
Figure 7. Moment - rotation relationships for cyclic specimens of the BX-C series.

FULL SHEAR
SUPPLEMENTARY
AREA shear area WEB PLATE
shear area

(a) (b)
Figure 8. (a) Full shear approach and (b) EN 1993-1-8 approach.

148 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Table 4. Comparison of test to the analytical results given by EN 1993-1-8 for steel joints.
Total Energy φmax+ φmax- Mmax Mmin Sj,ini+ Sj,ini- φy+ φy- My+ My-
3
Specimen kNm rad mrad kNm x10 kNm/rad mrad KNm
Symmetrically loaded joints
EC3-full As --- --- --- 55.64 2.97 165.40
EC3–red. As --- --- --- 55.64 2.97 165.40
BX-SS-M 9.01 43.20 263.34 48.03 3.26 180.79
BX-SS-C1 41.5 28.0 21.0 271.6 259.1 55.91 59.60 3.26 2.60 197.2 188.0
BX-SS-C2 23.4 17.4 18.1 261.8 259.8 71.24 63.5 2.66 2.39 194.8 206.8
Anti-symmetrically loaded joints
EC3-full As --- --- --- 32.99 4.76 156.91
EC3–red. As --- --- --- 25.19 4.24 106.77
BX-SU-M 24.0 105.5 258.36 51.50 2.28 137.66
BX-SU-C1 135.8 72.5 55.3 269.4 240.6 35.07 29.08 3.77 4.44 153.1 161.2
BX-SU-C2 88.37 39.2 46.8 240.1 236.6 27.82 40.53 5.54 3.37 179.8 161.2
full As - full shear area approach
red As - reduced shear area approach, according to EN 1993-1-8

Table 5. Comparison of test to the analytical results (EN 1994-1-1) for composite joints.
Total Energy φmax+ φmax- Mmax Mmin Sj,ini+ Sj,ini- φy+ φy- My+ My-
3
Specimen kNm rad mrad kNm x10 KNm/rad mrad KNm
Symmetrically loaded joints
EC4 Mom - --- --- --- 57.11 2.34 133.80
EC3* Mom + --- --- --- 99.68 2.30 230
BX-CS-M1 16.60 86.7 244.25 98.47 1.27 155.51
BX-CS-M2 8.4 32.7 316.09 105.27 2.52 231.32
BX-CS-C1 32.0 44.8 14.1 305.5 196.9 102.5 75.05 1.60 1.73 195.2 149.9
Anti-symmetrically loaded joints
EC4-comb.** --- --- --- 41.78 3.84 160.55
BX-CU-M 15.2 90.1 198.24 47.19 2.34 126.23
BX-CU-C1 104.5 60.5 53.8 187.1 193.3 36.87 37.92 3.49 3.38 142.67 137.3
BX-CU-C2 32.2 32.7 30.9 191.3 210.5 -- -- -- -- -- --
EC3* - computed according to EN 1993-1-8, by translation of the centre of compression
EC4-comb.** - mean value between the positive and negative values

The analytical computations according to EN 1994-1-1 section 8 in the case of symmetrical


positive bending leads to safer values in terms of resistance (15% approx.) and smaller
stiffness. Due to slab degradation in cyclic tests, the resistance and stiffness values are
smaller.

The EN 1994-1-1 section 8 does not give the possibility of computing composite joints
subjected to positive moments. For this case, the values of stiffness and moment resistance
presented in Table 5 are computed according to EN 1993-1-8 by a translation of the centre of
compression from the upper beam flange to the middle of the concrete slab (considered
without corrugated sheet). These assumptions lead to comparable values to the tests in
terms of resistance and stiffness. For the case of cyclic loading, there can be observed a
decrease in both resisting moment and stiffness due to rapid slab degradation.

Analytical values of moment resistance and stiffness for anti-symmetrical loading have been
obtained by the mean values for the two connections subjected to anti-symmetrical loading,
taking into account the full shear area approach. This prediction remains only an attempt of
computing the composite joints under anti-symmetrical loading.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 149


Rotation capacity for bolted steel and composite joints (INSA Rennes)

Ciutina [16] has tested some composite connections at INSA Rennes (Figure 9) under
monotonic and cyclic loads. G14 and G15 are composite joints under monotonic and cyclic
loading, and G17 and G18 are composite joints with haunch under monotonic and cyclic
loading.

Column Column
Beam Beam

M M

Dissipative zone
Dissipative zone
(connection)
Haunch (beam)

Figure 9. Tested joints.

Tests were performed in accordance with the ECCS Recommendations complete procedure
(ECCS 1985).

Obtained results concerning the joint characteristics are presented in Figure 10 and Table 6
for test specimens G14-G15 and Figure 11 and Table 7 for G17-G18 respectively.

In this case of tests, it is quite clear again that the rotation capacity of joints subjected to
cyclic loading could be estimated as approximately 0.5 of the rotation capacity of joints
subjected to monotonic loads.

400
Rotation Globale
Global rotation
300 Moment [kNm]

200

100

0
-0.1 -0.08 -0.06 -0.04 -0.02 0 0.02 0.04
-100
Rotation [rad.]
-200

-300 Essai
Test G15
Essai
Test G14
-400
Figure 10. Comparison between tests G14 and G15.

Table 6. Experimental characteristics determined by tests.


+ −
S +j ,ini S −j ,ini M max M max Φ +max Φ −max
Test
[kNm/rad] [kNm/rad] [kNm] [kNm] [mrad] [mrad]
Test G14 ---- 49900 ---- 342 ---- 91,8
Test G15 41098 53630 326 340 24 68

150 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


800

Moment [kNm]
Rotation Globale
Global rotation 600

400

Essai
Test G18 200
Essai
Test G17
0
-180 -160 -140 -120 -100 -80 -60 -40 -20 0 20 40 60 80
-200

-400

-600
Rotation [mrad.]
-800
Figure 11. Comparison between tests G17 and G18.

Table 7. Experimental characteristics determined by tests.


+ −
S +j ,ini S −j ,ini M max M max Φ +max Φ −max
Test
[kNm/rad] [kNm/rad] [kNm] [kNm] [mrad] [mrad]
Test G17 ---- 54734 ---- 639 ---- 88
Test G18 101982 66587 740 665 40 32

A small difference should be remarked for G14 and G15, but this is due to the test conditions
and the fact that the test could not be developed symmetrically. Anyhow, it can be seen that
the sum of the negative and positive rotation is practically identical with the rotation capacity
of the joint subjected to monotonic loads (24+68=92≈91.8 mrad).

In the second case, for G17 and G18, where negative and positive rotations are similar, it
can be seen that 40+32=72≈88 mrad.

CONCLUSIONS

It is clear for all specialists of the field that the rotation capacity of beam-to column joints is
one of the main characteristics which are influencing the seismic behaviour of the steel MR
frames. That is the reason why in the last period, this characteristic has been introduced with
some arbitrary values of reference for different types of frames. Unfortunately, formulae for
evaluating this characteristic are very few. It is evident that the researches have to be
continued in both directions experimental test research and numerical modelling, to establish
new definitions for the evaluation of the rotation capacity of beam-to-column joints.

Rotation capacity of joints is very difficult to establish except of test procedures. From
analytical point of view, the evaluation in static range should be possible, using the method of
components. Contrary, in cyclic range, Zandonini and Bursi [26] have shown that the use of
this method is impossible and have recommended the use of some macro-components,
incorporating some of the interaction effects among elemental components. This method
should be more adequate, because some macrocomponents are ductile ensuring the ductile
behaviour and the rotation capacity, and some others are fragile, needing to ensure them a
certain level of overstrength, in order to ensure the development of the rotation capacity.
Anyhow the ductile components of the joint are the endplate and the column web panel. After

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 151


Ciutina [16], a sort of balance of equilibrium has to be done between the geometrical
dimensions of the two ductile components, otherwise only one of them working for the joint
ductility.

An important observation in our opinion should be that taking into account the complex
behaviour of the moment resisting joints, and the conclusions of Stojadinovici [24] and
Zandonini and Bursi [26], concerning the pre-qualification and the introduction of
macrocomponents, under seismic loads, it is more adequate to refer to rotation capacity of
the joint using the concept of “macrocomponent method”, instead of rotation capacity of the
connection using the concept of “component method” utilised for monotonic loads.

So, what could we do in order to establish the rotation capacity of the joints, without
experimental cyclic tests? Analysing several tested joints under monotonic and cyclic loads,
the authors of this paper have observed that the rotations under cyclic and monotonic loads
are usually in the ratio of 1:2. Using this conclusion, the answer would be very simple.

For the evaluation of the rotation capacity under cyclic loads, we could use an analytical
method for evaluation under monotonic loads, or the well-known computer code DUCTROT
of Gioncu and then affect the results with the correction coefficient for cyclic loads equal to
0.5.

For the same evaluation, the correction coefficient of 0.5 could be used also to affect the
values of rotation capacity of joints determined under monotonic loads. This proposal is
made in order to use the large number of results, existing in literature, thinking also to some
data bank like SERICON.

From the same analytical point of view, we are reminding the importance of the
macrocomponents method to be developed and used after some more numerical
simulations.

It is important also to remind the prequalification connections what FEMA 350 have made, in
order to make a similar prequalification in Europe, with the usual connections utilised by
designers in Europe.

REFERENCES

1. EN 1993-1-1. 2003. Design of steel structures. Part 1-1: General rules and rules for
buildings. European standard.
2. EN 1993-1-8. 2003. Design of steel structures. Part 1-8: Design of joints. European
standard.
3. EN 1998-1. 2003. Design of structures for earthquake resistance. Part 1: General rules,
seismic actions and rules for buildings. European standard
4. EN 1994-1-1. 2003. EUROCODE 4: Part 1.1. General rules and rules for buildings;
Section 8: Composite joints in frames for buildings. European standard
5. Uniform Building Code, Volume 2, Structural Engineering Design Provisions. International
Conference of Building Officials, Whittier, California, USA, 1997.
6. Seismic Provisions for Structural Steel Buildings. American Institute of Steel
Construction, Inc. Chicago, Illinois, USA, 2002.
7. Standard for Limit State Design of Steel Structures. Architectural Institute of Japan, 1990.
8. FEMA 350. 2000. Recommended Seismic Design Criteria for New Steel Moment-Frame
Buildings, SAC Joint Venture.

152 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


9. SAC Joint Venture 1995. Connection Test Summaries. Report No. SAC-96-02,
Sacramento, California, USA
10. COST C1-Recent advances in the field of structural steel joints and their representation
in the building frame analysis and design process, Edited by Jean-Pierre Jaspart,
Brussels-Luxembourg, 1999.
11. Aribert J.M., Grecea D. 1998. Experimental behaviour of partial-resistant beam-to-column
joints and their influence on the q-factor of steel frames. The 11th European Conference
of Earthquake Engineering, Paris, 6-11 September 1998.
12. Beg D., Zupancic E., Vayas I. 2004. On the rotation capacity of moment connections.
Journal of Constructional Steel Research 60 (2004) 601-620.
13. Bertero V. 1997. “General Report on Codification, Design and Applications”, - STESSA
’97 Behaviour of Steel Structures in Seismic Areas, Proceedings of the Second
International Conference 3-8 Aug. 1997, Kyoto, Japan.
14. Bose B., Youngson G.K., Wang Z.M. 1996. An appraisal of the design rules in Eurocode
3 for bolted endplate joints with experimental results. Proc Inst Civil Eng: Struct Build
1996; 116:221-34.
15. Bose B., Sarkar S., Bahrami M. 1996. Extended endplate connections: comparison
between three-dimensional nonlinear finite-element analysis and full-scale destructive
tests. Struct Eng Rev 1996;8:315-28
16. Ciutina A. 2003. Assemblages et comportement sismique de portiques en acier et mixtes
acier-béton : expérimentation et simulation numérique. Thèse de Doctorat. INSA de
Rennes, France.
17. Dubina D., Ciutina A., Stratan A. 2001. "Cyclic Tests of Double-Sided Beam-to-Column
Joints", Journal of Structural Engineering, Vol.127, No.2, Feb.2001, pp.129-136
18. Dubina D., Ciutina A., Stratan A. 2002. "Cyclic tests on bolted steel and composite
double-sided beam-to-column joints", Steel & Composite Structures, Vol.2, No.2, 2002,
pp.147-160
19. Faella C., Piluso V., Rizzano G. 2000. Plastic deformation capacity of bolted T-stubs:
theoretical analysis and testing. Moment resistant connections of steel frames in seismic
area, design and reliability, RECOS. London: E&FN Spoon, 2000.
20. Gioncu V., Mateescu G., Petcu D., Anastasiadis A. 2000. Prediction of available ductility
by means of local plastic mechanism method: DUCTROT computer program. Moment
resistant connections of steel frames in seismic area, design and reliability, RECOS.
London: E&FN Spoon, 2000.
21. Grecea D. 1999. Caractérisation du comportement sismique des ossatures métalliques -
Utilisation d’assemblages à résistance partielle. Thèse de Doctorat. INSA de Rennes,
France.
22. Kuhlmann U., Kuhnemund F. 2000. Rotation capacity of steel joints, NATO Advanced
Research Workshop “The Paramount Role of Joints into the Reliable Response of
Structures, From the Rigid and Pinned Joints to the Notion of Semi-rigidity”,
Ouranoupolis, Greece, 21-23 May 2000.
23. da Silva L., Lima L., Vellasco P., Andrade S. 2002. Experimental behaviour of endplate
beam-to-column joints under bending and axial force. Database reporting and discussion
of results. Report on ECCS-TC10 Meeting in Ljubljana, April 2002.
24. Stojadinovici B. 2003. Stability and low-cycle fatigue limits of moment connection rotation
capacity. Engineering Structures 25 (2003) 691-700.
25. Suita K., Nakashima M., Morisako K. 1998. Tests of welded beam-column
subassemblies. Journal of structural engineering. November, 1236-1252

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 153


26. Zandonini R., Bursi O. 2002. Monotonic and hysteretic behaviour of bolted endplate
beam-to-column joints, Advances in Steel Structures, Vol.1, Edited by Chan, Teng and
Chung, Elsevier Science Ltd., 81-94.
27. ECCS. 1986. Recommended Testing Procedures for Assessing the Behaviour of
Structural Elements under Cyclic Loads, European Convention for Constructional
Steelworks, Technical Committee 1, TWG 1.3 – Seismic Design, No.45

154 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


A PROBABILISTIC EVALUATION OF THE ROTATION CAPACITY OF
END-PLATE BEAM-TO-COLUMN STEEL JOINTS

Luís Simões da Silva, University of Coimbra, Portugal


Luís Borges, University of Coimbra, Portugal
Helena Gervásio, GIPAMB Ltd., Portugal

ABSTRACT
In recent years, the pursuit of increased competitiveness of steel construction
has led to the quest for simpler and more economical joint details. The under-
lying concept of semi-continuity requires that adequate rotation capacity is
available, that is very much dependent on a good balance between ductile and
brittle components. In this paper, a probabilistic approach based on the Monte
Carlo method is implemented to investigate the sensitivity of end-plate beam-
to-column joints to (i) the differences between nominal and actual material
properties of steel and (ii) the difference between actual moment-rotation be-
haviour and the elastic-perfectly plastic response usually assumed in terms of
codes of practice.

INTRODUCTION

The rotation capacity of steel joints constitutes an essential property for the safe response of
steel structures. In non-seismic regions where gravity loading usually controls the behaviour,
the shift to semi-continuous design means that available rotation capacity must safely ex-
ceed the rotation demand. In seismic regions, this becomes an even higher requirement,
coupled with sufficient energy dissipation.

Current design procedures for steel joints do not include explicit evaluation of rotation capac-
ity. In particular, the Eurocode 3 provisions for rotation capacity (1) do not establish any spe-
cific procedures for the evaluation of the rotation capacity of bolted or welded joints. How-
ever, they state the need to ensure adequate rotation capacity either by testing in accor-
dance with EN 1990 (2) or alternatively, using appropriate calculation models based on the
results of tests. Additionally, it provides sufficient conditions that ensure adequate rotation
capacity (3, 4).

For end-plate beam-to-column steel joints, these conditions basically impose that either the
column web panel in shear controls the behaviour of the joint or, alternatively, the end-plate
or the column flange in bending and are reproduced below:
d < 69ε (1)
tw
t < 0.36d fub f y (2)
where d denotes the bolt diameter in the unthreaded part of the shank, tw is the thickness of
the column web, t is the thickness of the end-plate or the column flange, fub is the ultimate
tensile strength of the bolt, fy represents the yield stress of the plate and
ε= 235
fy (3)

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 155


Proper evaluation of the rotation capacity of steel joints within the scope of the component
method (5) requires full characterization of the various components. It is worth recalling that,
within the framework of the component method, a joint is modelled as an assembly of springs
and rigid links that may present a wide range of geometries, with different numbers of bolt
rows and connecting parts. Although the constitutive law of each component is non-linear,
several properties can be usually identified: initial stiffness (Ke), resistance (FY), post-limit
stiffness (Kp) and failure deformation (∆f). Currently, no reliable estimates of post-limit stiff-
ness and failure deformation exist, particularly for the components that usually control the
behaviour, such as the T-stub components or the column web panel (6). Recently, with re-
spect to the equivalent T-stub in tension, Swanson (7), Faella et al. (8) and Girão et al. (9)
addressed the issue of the evaluation of the ductility of this component. Kuhlmann and
Kuhnemund (10) investigated the response of the component column web under transverse
compression. Finally, Beg et al. (11) developed numerical expressions for the deformation
capacity of joint components.

Given that small variations of the post-limit stiffness (in particular for the critical component)
result in large variations of the maximum rotation of the joint (12), it is the aim of this paper to
assess the influence of the various component properties (post-limit stiffness, yield force and
failure deformation) on the available ductility of the joint.

MONTE CARLO METHOD

General description

The Monte Carlo simulation technique (13), applied to steel joints (12) consists of the systematic
solution of the relevant component model, carried out in a deterministic way, to yield a locus of
statistically valid solutions that provide a probabilistic description of the joint response. To
achieve this objective, one of the main tasks in a simulation procedure is the generation of ran-
dom numbers from prescribed probability distributions. This can be achieved using a random
generator available in any computer according to the following procedure: (i) generation of a uni-
formly distributed random number between 0 and 1, and (ii) application of appropriate transfor-
mations in order to obtain the corresponding random number with the specified probability distri-
bution. The Box and Muller Method was adopted in this paper (13), complemented with adequate
checking of both the independence and the randomness of the generated numbers. Finally, it is
noted that many simulations are usually necessary to obtain results with acceptable precision, al-
though resource to sample reduction techniques such as the latin hypercube may be used to limit
the number of simulations.

Computational implementation

The computational implementation of a probabilistic evaluation of the rotation capacity of end-


plate beam-to-column steel joints is summarized in Figure 1 and required the implementation of
two computer programs. The first, NasCON (14), addresses the solution of the component mod-
els, which requires a nonlinear procedure, given the nonlinear behaviour of the springs (compo-
nents). Solution is carried out under a standard Newton-Raphson scheme under load or dis-
placement control. Care is taken to identify and output all component notable points, namely
component yield points and failure of the joint. The second program, MCSimulation (14), imple-
ments the Monte Carlo procedure in the following way:
(i) For a total number of components, n, r ≤ 4n random variables representing the se-
lected random component properties (ke, FY, kp, ∆f) are considered.
(ii) For each random variable r, mr values are generated using the Box and Muller
method (Step 3).

156 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Figure 1. Computational scheme for probabilistic evaluation of the response of steel joints.

(iii) For each simulation, the mr values for each random variable are combined, resulting
in a sample containing m-dimensional values, where
r
m = ∏mj (4)
j=1

(iv) Using the appropriate mechanical model for the chosen joint configuration and the
program NasCON, m moment-rotation curves are obtained for each simulation (Step
4).
(v) Steps (ii) to (iv) are repeated to generate s independent simulations.
(vi) Finally, direct probabilistic assessment is carried out on the s available simulations
for the desired indicators (failure rotation or sequence of yielding of the various
components, for example) – Steps 5 and 6.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 157


STATISTICAL CHARACTERIZATION OF STEEL COMPONENTS

Statistical description of mechanical properties of steel

Although steel is among the most uniform construction materials, and is produced with strin-
gent quality control procedures, the production requirements as specified by the relevant
standards of steel production only demand that minimum values for yield and ultimate stress
are guaranteed. Thus, in practice, all structural steels always present higher yield stress than
the nominal values, usually with a significant scatter between the nominal value and the ac-
tual mean or characteristic yield stress (15). This over-strength has always been regarded as
beneficial because most emphasis in the design of joints was on strength. As ductility be-
comes an issue, excessive over-strength may result in unforeseen behaviours and must be
assessed carefully.

The force-displacement relation for the various components is frequently simplified as a bi-
linear curve, characterized by four independent properties: initial stiffness (Ke), resistance
(FY), post-limit stiffness (Kp) and collapse displacement (∆f). All these properties depend on
the mechanical properties of steel, the major dependencies being highlighted in the following:
Ke= Ke(E), FY= FY(E, fy), Kp= Kp(E, Ep, fy), ∆f= ∆f (E, Ep, fy, fu) (5)
where E, Ep, fy, fu are, respectively, the elastic stiffness, the plastic stiffness, the yield stress
and the ultimate stress of steel. Consequently, being a major source of randomness in the
characterization of the various components, the variability of the yield stress of steel is as-
sessed by considering either a normal distribution or a mixed distribution, obtained by a lin-
ear combination of two normal distributions (16).

Calibration procedure for post-limit stiffness and failure displacement

Because there are no accepted estimates for the post-limit stiffness and collapse displace-
ment of most components, it was necessary to establish representative statistical estimates
for these properties. To achieve this objective, a representative set of experimentally tested
sample joints covering flush and extended end-plate connections taken from the database
SERICON (17) were selected. Subsequently, using the deterministic values of initial stiffness
and resistance obtained according to part 1-8 of EC3 and a genetic algorithm coupled with
NASCon (18), average values of Kp and ∆f were established for each component. The
adopted probability distributions for the component properties are summarized below:
(i) Initial elastic stiffness (Ke): deterministic values from EC3 (1);
(ii) Resistance (FY): given the direct dependence of component resistance on the
yield stress of steel (eq. (5)), the adopted probabilistic distribution is taken as the
product of the deterministic value (obtained according to EC 3 – part 1.8) by the
histogram of the corresponding steel grade divided by the nominal yield strength
value. The two yield stress distributions described above are used: normal or
mixed distribution.
(iii) Post-limit stiffness (Kp): a Gaussian distribution is proposed to simulate its varia-
tion. For components with limited ductility, a coefficient of variation equal to 100%
was selected (so that the generated Kp values could become negative). For com-
ponents with high ductility, a coefficient of variation equal to 50% was chosen (so
that the generated Kp values could approach zero stiffness). Finally, in case of
brittle components the post-limit stiffness, resulting from the linearization of the
behaviour law, is not considered.
(iv) Collapse displacement (∆f): given the scarce experimental evidence, a normal dis-
tribution corresponding to the range suggested in (6) is used.

158 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


FLUSH END-PLATE JOINTS

Reference configuration

In order to explore a typical design condition, a flush end-plate beam-to-column joint tested
by Lima at the University of Coimbra (19) was selected. It presents a critical tension zone,
with compression zone overstrength of 143 % in nominal terms and 164 % in real terms.
Having a critical tension zone, it provides an excellent example to explore the possibility of
brittle components such as the bolts in tension having a significant probability of being
critical. Table 1 describes the flush end-plate connection. It is further noticed that, according
to EC 3, this connection does not meet the requirements for sufficient rotation capacity.

Table 1. Test details for Flush Endplate joint tested in Coimbra - FE1 (19).
Test arrangement Geometry
T-configuration, column with 2 supports, cantilever beam

Joint details Mechanical properties


fy fu
[ N/mm2 ] [ N/mm2 ]
Type of joint : flush endplate with
2 bolts per row Column
Flange 342.9 477.3
Column orientation: strong axis
Web 372.0 448.8
Profiles sizes: IPE 240, HEB 240
Beam
Number, size and grade of Flange 340.1 448.2
bolts: 4 bolts M20 of grade 10.9 Web 363.4 454.3
including washers. Endplate 369.4 503.5
Bolts 900.0 1000.0

Component characterization

Joint M-Rot Curve Joint M-Rot Curve


100 100
80 80
M (kN.m )
M (kNm )

60 60
40 40

20 NASCon - Bilinear Calibration 20 Experimental - FE1


Experimental - FE1 NASCon - M ult ilinear Calibration
0 0
0 50 100 0 50 100
Rotation (m rad) Rotation (m rad)

Figure 2. Comparison between experimental and numerical results for the flush end-plate.
joint for two types of component behaviour: bilinear and smoothed multi-linear.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 159


Table 2 reproduces the component properties both for nominal and measured values of yield
stress. Finally, Figure 2 compares the experimental results with a calculated result based on
a bilinear characterization of the various components and calibrated values of Kp.

Table 2. Component properties for flush end-plate joint.


Fy Fy Kp
Ke ϕ=
Nominal Measured calibrated
[kN/m] ∆f/∆y
[kN] [kN] [kN/m]
3.1 - Column Web in
445.36 602.45 1.48E+06 1.48E+04 10
Transverse Tension
4.1 - Column Flange
375.48 408.32 8.03E+06 8.03E+03 200
in Bending
Tension 5.1 - End-Plate in
305.79 339.36 2.63E+06 1.05E+04 200
Zone Bending
8.1 - Beam Web in
365.68 483.23 ∞ - 200
Tension
10.1 - Bolts in Tension 441.00 441.00 1.63E+06 3.26E+04 3
Critical 305.79 339.36
1 - Column Web Panel in
-474.77 -642.24 1.58E+06 1.83E+04 200
Shear Zone Shear
Critical -474.77 -642.24
2 - Column Web in
-507.06 -634.55 2.13E+06 1.23E+05 12
Transverse Compression
Compression 7 - Beam Flange in
Zone -438.00 -555.57 ∞ - 10
Compression
Critical -438.00 -555.57

Fy Fy
Nominal [kN] Measured [kN]
Ratio C/T 143% 164%
Ratio S/C 108% 116%
Ratio S/T 155% 189%

Simulation schedule

Three cases were tested. Case FA is intended to explore the influence of the post-limit stiff-
ness variability on the joint overall behaviour. Case FA was sub-divided into 3 possibilities: (i)
FA.1 (simulation of the post-limit stiffness of the critical components – column web in tension,
column flange in bending and end-plate in bending - using the measured real values for steel
properties), (ii) FA.2 (same as FA.1 but adding the variability of the post-limit stiffness of the
column web panel in shear) and (iii) FA.3 (same as FA.1 but using the nominal properties of
steel). Case FB adds the variability of the resistance of the components, closely linked to the
distribution of the yield stress of steel. The chosen components now have their resistance FY
simulated following either using a Normal law (FB.1) or a Mixed distribution (FB.2). Finally,
Case FC explores the influence of the collapse displacement variation on the global behav-
iour.

Comparative analysis of results

Starting with an analysis of the variability of the post-limit stiffness (Case FA), it is noted that a
high percentage of the samples reach 100 mrad, failures ranging from 18.2% to 43.5%. Using
measured values of yield stress of steel instead of nominal values leads to an increase of fail-
ures of the bolts in tension (6.7% to 30.7%). The histograms of rotation at failure show the same

160 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


tendency of a small increase of rotation capacity with actual steel properties. The average rota-
tion capacity is quite high (85-90 mrad). Finally, it is interesting to note that, for a rotation of 30
mrad, the average bending moment is 109% higher for actual steel properties than nominal steel
properties. Case FA.2 shows little or no influence of the post-limit stiffness variation of the col-
umn web in shear, hardly surprising since this is the strongest component of the joint. Allowing
the variation of the resistance (case FB) of the most critical components in the tension zone, the
following conclusions can be drawn: (i) a higher number of failures occur (84.5%) when com-
pared to Case FA.1 (measured values for steel: 36.6%) or Case FA.3 (nominal yield stress of
steel: 18%); (ii) there is a clear increase in the number of failures due to the collapse of compo-
nent bolts in tension and a decrease in failures due to the collapse of component column flange
in bending or column web in tension; (iii) most of the failures of the component bolts in tension
take place for small rotations with a 5% percentile of only 8.2 mrad. The consideration of the
variability in the failure deformation of the column web in tension brought no major influence on
the joint rotation capacity (comparing with cases FA and FB).

EXTENDED END-PLATE JOINTS

Reference configuration

Table 3. Test details for extended end-plate joint tested by Humer (17).
Mechanical properties
fy fu
Type of joint: extended endplate [ N/mm2 ] [ N/mm2 ]
with 2 bolts per row Column
Flange 275,9 400,50
Column orientation: strong axis Web 306,60 445,00
Profiles sizes: IPE 450, HE240 B Beam
Number, size and grade of bolts: Flange 284,60 413,00
6 bolts M24 of grade 10.9 including Web 315,60 298,00
washers. Endplate 323,00 n/a
Bolts 900,00 1000,00

ep = 65 bp = 239 ep = 65
w = 109

tbp = 12
ex = 35 lp = 84
185 p = 115

tp = 41
hp = 553
320
HEB 240R

IPE 450R

Test arrangement: T-configuration,


column with 2 supports, cantilever
beam
Geometry tc = 16.4

tb = 14.0

100
= 240 wb = 10.4
bb = 192

wc = 10.4

hc = 242 hb = 454

(mm)

The extended end-plate joint, described in detail in Table 3, was tested by Humer at the Uni-
versity of Innsbruck (17). This joint presents a critical shear zone, with compression zone re-

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 161


sistance over strength of 111% in nominal and real terms, and tension zone resistance over
strength of 142%. Having a critical shear zone (ductile), followed by a critical compression
component, it provides an excellent example to explore the possibility of failure mode shift
from the compressive to the tension zone, even in an extreme design situation.

Component characterization

Table 4 reproduces the component properties both for nominal and measured values of yield
stress. Figure 3 compares the experimental results with a calculated result based on a bilin-
ear characterization of the various components and calibrated values of Kp.

Table 4. Component properties for extended end-plate joint.


Fy Fy Kp
Ke
Nominal Measured calibrated ϕ=∆f/∆y
[kN/m]
[kN] [kN] [kN/m]
3.1 - Column Web in
417.04 544.11 1.50E+06 1.70E+04 20
Tension
4.1 - Column Flange
Row 1

392.23 460.50 3.75 E+06 3.37E+05 200


in Bending
5.1 - End-Plate in
462.29 635.40 2.41 E+07 1.44E+05 200
Bending
10.1 - Bolts in Tension 635.00 635.00 1.31 E+06 1.31E+06 3
3.2 - Column Web in
Tension 175.04 228.37 1.50 E+06 1.50E+06 10
Tension
Zone 4.2 - Column Flange
350.37 411.35 3.75 E+06 3.75E+06 200
in Bending
Row 2

5.2 - End-Plate in
462.29 635.40 2.45 E+07 2.45E+07 200
Bending
8.2 - Beam Web in
677.52 909.90 ∞ - 200
Tension
10.2 - Bolts in Tension 635.00 635.00 1.31 E+06 5.91E+06 3
Critical 567.27 688.87
1 - Column Web Panel in
Shear -416.83 -543.83 6.11 E+05 1.83E+05 200
Shear
Zone
Critical -416.83 -543.83
2 - Column Web in
-461.65 -602.31 2.46 E+06 1.23E+06 12
Transverse Compression
Compres- 7 - Beam Flange in
sion Zone -993.52 -1203.22 ∞ - 10
Compression
Critical -461.65 -602.31

Fy Fy
Nominal [kN] Measured [kN]
Ratio T/C 128% 128%
Ratio T/S 142% 142%
Ratio C/S 111% 111%

Simulation schedule

As for the flush end-plate joint, Case EA is intended to explore the influence of the post-limit
stiffness variability and its implications in the joint overall behaviour. Case EA was subdivided
into 3 possibilities: (i) EA.1 (simulation of the post-limit stiffness of the critical failure compo-
nent – column web in compression - and characterization of components using the nominal

162 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


properties of materials); (ii) EA.2 (same as EA.1 but using actual steel properties); and (iii)
EA.3 (simulation of the post-limit stiffness of the column web in compression and column
web in shear, using actual steel properties). Case EB adds the variability of the resistance of
the components, the chosen critical components now following either a normal law (EB.1 and
EB.3) or a mixed distribution (case EB.2). Finally, Case EC explores the influence of the col-
lapse displacement variation in the global behaviour.

M-Rot Curve Joint


SERICON 109.005
500

400
M (kNm)

300

200

100 NASCon
Experimental
0
0.00 10.00 20.00 30.00 40.00 50.00 60.00 70.00
Rotation (m rad)

Figure 3. Comparison between experimental results for the extended end-plate joint and
numerical results obtained using NASCon and bilinear component behaviour.

Comparative analysis of results

Starting with cases EA.1 and EA.2, that compare the influence of actual vs. nominal steel
properties, 94% of failures were recorded (less than 100 mrad of rotation). Failure of the col-
umn web in compression occurs in 60.9% and 57.5% of the cases (nominal and real, respec-
tively), as expected in a joint where the tension zone is 128% stronger than the compression
zone. However, because of the over strength effect of steel, a shift in failure modes is ob-
served whenever the tension zone controls the behaviour. For nominal steel properties, 33%
of the simulated connections fail in the column web in tension while for actual steel proper-
ties, 36% of failures result from the component bolts in tension. Case EA.3 explores the in-
fluence of the post limit stiffness of the column web panel in shear, the component with the
lowest resistance. A decrease in failures (47%) of the column web in compression together
with an increase of failures of the bolts in tension (44%) is observed, together with some
marginal failures of the column web in tension (4%). Comparison of cases EB.1 and EB.2
shows that the use of the mixed distribution versus a normal distribution for the resistance of
the compression and shear components does not significantly affect the results. A slight in-
crease of column web in compression failures (72% to 79%) is noted, together with a small
decrease of ductility (39 mrad to 29 mrad). The simulation of the component column web in
tension for FY and Kp brought no major changes. The consideration of the variability in the
failure deformation of component column web in compression yields a big influence on the
joint ductility, reducing the average rotation capacity to 25 mrad. Globally, taking 30 mrad as
a ductility limit state for sufficient rotation capacity, the nine cases that were analysed yield
probabilities of not reaching the required rotation ranging from 26% to 62%.

CONCLUSIONS

The two examples of joint configuration that were explored in this paper have revealed some
interesting features. Firstly, the flush end-plate beam-to-column joint designed to be critical in
tension (mode 1: end-plate in bending), exhibits a significant probability of bolt failure in ten-

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 163


sion, especially when actual steel properties are considered, accompanied by a high probabi-
lity of low rotation capacity. Secondly, for an extended endplate beam-to-column joint desig-
ned to be critical in shear and subsequently critical in compression, it was found that there
was a significant probability of failure from the tension zone. The influence of steel over
strength also caused the bolts to become the critical component. Although the results pre-
sented in the work are insufficient to establish any quantitative conclusion about ductility of
steel joints, they are certainly sufficient to reveal the possibility of unexpected brittle failure
whenever the rotation capacity is not explicitly calculated and does not constitute a compul-
sory limit state.

REFERENCES

(1) Eurocode 3, prEN 1993-1-8: 2003, Part 1.8: Design of Joints, Eurocode 3: Design of
Steel Structures, Stage 49 draft., 5 May 2003. CEN, European Committee for Stan-
dardisation, Brussels, 2003.
(2) ENV 1090: 2001. Execution of Steel Structures. CEN, European Committee for Stan-
dardisation, Brussels, 2001.
(3) Zoetemeijer P. Summary of the research on bolted beam-to-column connections. Re-
port 25-6-90-2. Delft University of Technology, Faculty of Civil Engineering, Stevin
Laboratory – Steel Structures, 1990.
(4) Jaspart, JP. Recent Advances in the Field of Steel Joints – Column Bases and Further
Configurations for Beam-to-Column Joints and Beam Splices. Thèse présentée en vue
de l’obtention du grade d’Agrégé de l’Enseignement Supérieur, Année Académique
1996-1997.
(5) Weynand K, Jaspart JP and Steenhuis M. The Stiffness Model of Revised Annex J of
Eurocode 3. In: Connections in Steel Structures III, Proceedings of the 3rd International
Workshop on Connections (eds.: Bjorhovde R, Colson A, Zandonini R), Trento, Italy,
May 8-31, pp 441-452, 1995.
(6) Simões da Silva L, Santiago A and Vila Real P. Post-limit Stiffness and Ductility of End-
Plate Beam-to-Column Steel Joints. Computers and Structures, 80, pp 515-531, 2002.
(7) Swanson JA. Characterization of the strength, stiffness and ductility behavior of T-stub
connections. PhD dissertation, Georgia Institute of Technology, Atlanta, USA, 1999.
(8) Faella, C., Piluso, V., Rizzano, G. Structural Steel Semirigid Connections Theory De-
sign and Software. CRC Press, 2000.
(9) Girão, A., Bijlaard, F., Gresnigt, N. and Simões da Silva, L. Experimental assessment
of the behaviour of bolted T-stub connections made up of welded plates. J. of Con-
structional Steel Research, 60, 269-311, 2004.
(10) Kuhlmann U, Kuhnemund F. Ductility of semi-rigid steel joints. In: Stability and ductility
of steel structures (SDSS 2002); Budapest, Hungary, 2002.
(11) Beg D, Zupančič, Vayas I. On the rotation capacity of moment connections. J. of Con-
structional Steel Research 60(4), 601-620, 2004.
(12) Gervásio, H., Simões da Silva, L., and Borges, L. Reliability assessment of the post-
limit stiffness and ductility of steel joints. J. of Constructional Steel Research 60(4),
635-648, 2004.
(13) Ang, A and Tang, W. Probability concepts in Engineering planning and design. Vols. 1-
2, Wiley, 1975-1984.
(14) Borges, L. Probabilistic Evaluation of the Rotation Capacity of Steel Joints. MSc The-
sis, University of Coimbra, 2003.
(15) Fajkus, M., Melcher, J., Holicky, M., Rozlivka, L., Kala, Z. Design Characteristics of

164 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Structural Steels Based on Statistical Analysis of Metallurgic Products. J. of Construc-
tional Steel Research 60(5), 795-808, 2004.
(16) Weynand K. Sicherheits- und Wirtschaftlichkeitsuntersuchungen zur Anwendung
nachgiebiger Anschlusse im Stahlbau. Shaker Verlag, D 82 (Diss. RWTH Aachen),
Heft 35, Aachen, 1997.
(17) Cruz, P.J.S., Simões da Silva, L., Rodrigues, D.S. and Simões, R.A.D. Database for
the Semi-Rigid Behaviour of Beam-to-Column Connections in Seismic Regions. Jour-
nal of Constructional Steel Research, Vol. 46:1-3, Paper No. 120, 1998.
(18) Borges, L., Lima, L., Simões da Silva, L. and Vellasco, P. An evaluation of the post-
limit stiffness of beam-to-column semi-rigid joints using genetic algorithms. in Topping,
B.H.V. (ed.), Proceedings of 9th International Conference on Civil and Structural Engi-
neering Computing, Civil-Comp Press, Stirling, United Kingdom, paper 69, 2003.
(19) Simões da Silva, L., Lima, L., Vellasco, P., and Andrade, S. Behaviour of flush end-
plate beam-to-column joints subjected to bending and axial force. International Journal
of Steel and Composite Structures, 2004 (in print).

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 165


166 Connections in Steel Structures V - Amsterdam - June 3-4, 2004
PERFORMANCE-BASED SEISMIC DESIGN OF BRACED-FRAME
GUSSET PLATE CONNECTIONS

C.W. Roeder, U. of Washington, Seattle, WA, USA


D.E. Lehman, U. of Washington, Seattle, WA, USA
J. H. Yoo, U. of Washington, Seattle, WA, USA

ABSTRACT
Performance-based seismic design (PBSD) produces structures that meet
multiple performance objectives. Economical application of PBSD produces
structures that:
• have adequate strength and stiffness to remain serviceable during
small, frequent earthquakes, and
• develop cyclic nonlinear deformations while assuring life safety and
collapse prevention during large infrequent earthquakes.
Special concentrically braced frames (SCBFs) and buckling restrained
concentrically braced frames (BRCBFs) can meet these diverse objectives if
the gusset plate connection provides adequate performance. Current
connection design provisions attempt to ensure adequate connection
resistance to avoid premature failure, but the resulting connections may be
massive and uneconomical or provide unacceptable performance. An
analytical and experimental research study to develop improved design
methods for these gusset plate connections is described. A rational, PBSD
procedure is proposed. Future directions of the research study are noted.

INTRODUCTION

Large, infrequent earthquakes induce huge elastic forces in building structures. Therefore,
seismic design of buildings employs relatively small earthquake design forces to assure that
the structure remains serviceable during frequent seismic events, and cyclic, inelastic
ductility is used to prevent loss of life and structural collapse during large seismic events.
This concept is simple and results in economical design, but it is difficult to reliably and
accurately apply in engineering practice. Performance based seismic design (PBSD) is a
recently developed design concept that formalizes this procedure for meeting multiple
design objectives. Concentrically braced frames (CBFs) are stiff, strong steel structures,
which are economical systems for seismic design. The inelastic lateral response of CBFs is
dominated by axial yielding and post buckling deformation of the braces, and special
concentrically braced frames (SCBFs) are designed under guidelines which are intended to
assure good inelastic performance from the brace. These SCBF requirements control the
local and global slenderness of the brace to prevent concentration of local damage during
post buckling deformations. They require that the strength of the connection be greater than
the yield capacity of the brace, and they establish geometric clearances intended to develop
the connection rotations needed to develop brace buckling. Innovative bracing systems,
such as unbonded or buckling restrained braced frames (BRCBFs), have also been
proposed, since they hold promise for improved seismic performance. The seismic
performance of these CBF systems depends on the brace, the connection, and the framing
members. To achieve a superior level of seismic performance, the design of the connection

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 167


must be balanced to meet strength and deformation demands while permitting the brace to
develop the desired elastic and inelastic performance. The connections are normally gusset
plate connections, and the connection seismic performance of this connection is the focus of
this paper.

BRACED FRAME SEISMIC BEHAVIOR

SCBFs (AISC, 2002) provide economical strength and stiffness and are commonly used for
seismic design. The brace provides lateral stiffness to the frame, and so the brace attracts
large axial forces during earthquake loading. As a result, the cyclic inelastic deformation of
an SCBF is achieved through post buckling deformation and tensile yield of the brace as
illustrated in Zones 0-A, A-B, B-C, C-D and D-E in Figs. 1a and 1b. Plastic hinges form
within the brace after buckling, because of the P-δ moments. These hinges cause
permanent plastic deformations and ultimately deterioration of brace resistance. When the
brace is subjected to a tensile force during load reversals (Zones B-C, C-D and D-E)
significant axial deformation is needed to recover the full tensile stiffness and resistance.
This leads to the one-sided axial force-deflection behavior of the brace seen in Fig. 1a, and
SCBFs use braces in opposing pairs to achieve the system inelastic hysteretic behavior
illustrated in Fig. 1c.

Figure 1. Behavior of special concentrically braced frames (Popov et al. 1976).

Braces are normally joined to the beams and columns of the frame through gusset plate
connections as illustrated in Fig. 2. SCBF brace post-buckling behavior places significant
cyclic load and deformation demands on these connections as illustrated by the brace end
rotation shown in Zone A-B of Fig. 1b. These connection rotation demands vary depending
upon whether the brace has in-plane or out-of-plane buckling. Figure 3 is a photo of a
buckled braced from a steel frame damaged during a past earthquake, and it shows the
large end rotations that can occur during inelastic seismic deformations.

168 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Figure 2. Typical gusset plate connections. Figure 3. Connection rotation due to
brace buckling.

Brace buckling may result in deterioration of stiffness and resistance and pinched hysteretic
behavior shown in Fig. 1a and 1c. BRCBFs have been developed to increase ductility and
reduce deterioration in brace resistance. BRCBFs are patented systems where an axially
loaded bar is encased into a stiff tube but is not bonded to the tube. As a result, the bar
yields in tension and compression without brace buckling, and the resulting cyclic inelastic
performance of the brace is symmetric without deterioration as illustrated in Fig. 4. Many
engineers currently prefer BRCBFs for seismic design to SCBFs, because of this inelastic
performance. A significant body of research (Clark et al. 2000, Ando et al 1993, Connor et
al. 1997, and Inoue et al. 2001) has been completed on BRCBF performance. However, the
past research and professional opinions on BRCBF braces are based on the assumption
that braced frames behave as a truss, and the brace has pure axial deformation with no
bending moment. This idealization is frequently not achieved in practice, because of the
gusset plate performance.

Figure 4. Axial force-deformation hysteresis curve for an BRCBF (Clark et al. 2000).

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 169


BRACED FRAME GUSSET PLATE CONNECTION DESIGN REQUIREMENTS

Past research (Kahn and Hanson 1976, Foutch et al. 1987, Astaneh-Asl et al. 1982, Lee and
Goel 1987, and Aslani and Goel 1989) shows that SCBFs can provide good seismic
performance if premature fracture or tearing of the brace and the connection is avoided.
The AISC Seismic Design Requirements for the SCBF system (AISC 2002) provide
guidelines toward meeting these goals. Slenderness limits for the brace and different brace
geometry and configuration requirements are provided to avoid concentration of inelastic
strain that leads to early tearing or fracture. The AISC seismic provisions also require that
the connection be designed to be stronger than the brace, and with out-of-plane buckling,
geometric limits are established to permit the expected end rotation on the connection.
These design rules lead to the common conception that a stronger gusset plate connection
is better, and some very uneconomical and impractical connections such as illustrated in
Fig. 5 have been built. Further, it is difficult to satisfy the out-of-plane buckling geometry
requirements for most practical gusset plate connections. The past performance of braced
frames during earthquakes has been mixed. In some cases, economical and serviceable
performance during earthquake loading has resulted, but in others, the apparent resistance
and ductility were significantly smaller than expected when brace or connection failures such
as illustrated in Fig. 6 occurred.

Figure 5. Photo of excessively large Figure 6. Photograph of brace fracture


gusset plate connection. noted from past earthquake damage.

Figure 7. Deformation mechanisms of BRCBF systems.

170 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


The seismic performance of BRCBFs also depends upon the gusset plate connection
design. The BRCBF connection must support the full tensile and compressive force capacity
of the brace during cyclic inelastic deformation demand, and it must have adequate stability
and lateral restraint to prevent out-of-plane deformation. As a result, out-of-plane rotation or
deformation, buckling, or fracture of the BRCBF connection cannot be tolerated, because
the brace and gusset plate act in series and gusset plate deformation may negate BRCBF
buckling restraint. BRCBF braces have large axial deformation capacity as long as the
beam-column joints respond as pinned joints as illustrated in Fig. 7b. However, BRCBF
braces have large flexural stiffness contributed by the moment of inertia of the surrounding
tube. Further, the gusset plate connections have substantial flexural resistance and in-plane
rotational restraint because of the size and geometry of the connection as illustrated in Fig.
2. As a result, BRCBF deformations may result in significant bending moment in the gusset
plate and the BRCBF brace due to flexural deformation as depicted in Fig. 7c. These
deformation demands may place large stress and strain demands on the gusset plate
connection, and they may have negative impact on the seismic performance of the BRCBF
system. Figure 8 is a photo of a BRCBF brace and gusset plate connection in a test frame.
The gusset plate connection was clearly strong enough by the present seismic design
reasoning. However, the gusset plate buckled at relatively modest inelastic frame
deformations, and the BRCBF brace was damaged because of the resulting connection
deformation.

Figure 8. Photograph of buckled BRCBF brace and gusset plate.

PROPOSED PBSD DESIGN METHOD

Prior discussion demonstrates the need to understand and improve the BRCBF and SCBF
gusset plate connection design to efficiently and economically achieve multiple performance
objectives for seismic design. The stiffness and strength of CBF systems clearly aid in
satisfying the Operational and Immediate Occupancy performance levels, and they provide
economical building design. However, the SCBF and BRCBF connections must
accommodate significant inelastic deformation and strength demands for the Life Safety and
Collapse Prevention performance levels. These later design limit states are less easily
satisfied, because they place both force and deformation demands on the gusset plate

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 171


connection. A design strategy, which considers both connection strength and inelastic
deformation capacity, is needed to meet these later design goals.

A rational hierarchy of yield mechanisms within the frame and the connection are required to
meet these diverse force and deformation demands, because a simple connection
resistance check as presently used in SCBF seismic design leads to variable seismic
performance. The simple resistance checks in the AISC requirements do not assure the
most desirable connection behavior nor do they prevent the least desirable performance.
The proposed design procedure must address this complex braced frame behavior while
retaining a simple design method.

Figure 9. Yield mechanisms and failure modes of CBF components.

The proposed procedure employs concepts developed for PBSD of steel moment-resisting
frames (Roeder 2001 and 2002) during the SAC Steel Project. This design procedure
developed balance conditions to assure desirable yield mechanisms, restrict undesirable
failure modes, and achieve the desired seismic performance. Occurrence of a yield
mechanism changes the stiffness and provides inelastic deformation of the structure without
significant loss in resistance. The occurrence of a failure mode can lead to fracture, loss in
resistance, and reduced inelastic deformation capacity. A single failure mode will produce a
significant reduction in resistance or deformation capacity, but multiple failure modes are
usually required to produce complete connection failure. Similar design procedures are
being developed to achieve the performance objectives of SCBF and BRCBF structures.
The yield mechanisms and failure modes for the CBF systems as illustrated in Fig 9 are
considered in this evaluation. The controlling yield mechanisms for SCBFs are expected to
be inelastic shortening due to post-buckling deformation and tensile yielding of the brace.
The axial load capacity of the BRCBF brace is similar in tension and compression, and the
brace should not buckle. As a result, the controlling yield mechanism for BRCBF systems
will be this axial yield deformation. Secondary yield mechanisms are also possible with
braced frame gusset plate connections. These include local yielding of the gusset plate,
local yielding of the beam and column adjacent to the gusset plate and elongation of
boltholes in brace and gusset plate. These secondary yield mechanisms can contribute
some plastic deformation during large earthquakes, but their deformation capacity is limited
so that they are not capability of providing primary sources of inelastic deformation. CBFs
have many different possible failure modes including tearing or fracture of the brace, net
section fracture of the brace or gusset plate, weld fracture, shear fracture of the bolts, block
shear, excessive bolt bearing deformation, and buckling of the gusset plate.

A rational connection design procedure for seismic design can be based upon the expected
or mean yield resistance of the controlling yield mechanism, and its comparison to the

172 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


expected resistance of the critical failure mode. As noted in prior discussion, favorable yield
mechanisms should have a smaller resistance than less favorable or secondary yield
mechanisms and all failure modes. Greater separation is warranted between the controlling
yield mechanism resistance and failure mode resistance, when the failure mode results in
poor seismic performance than that used for failure modes with better seismic performance
and deformation capacity. As a result, good inelastic performance can be achieved for CFT
gusset plate connections by balancing the controlling yield mechanism resistance with the
resistances associated with all failure modes. This idea can be expressed as Eqs. 1a and 1b
where β is a balance factor.
Ry Fy Ag < β Rfailure (Eq. 1a)
Ry Fcr Ag < β Rfailure (Eq. 1b)

Where Rfailure is the failure mode resistance for an individual failure mode, and Ag is the
gross cross sectional area of the brace. Ry is a factor for adjusting the nominal yield stress,
Fy and Fcr, to the expected or average yield and critical buckling stress, respectively. Two
balance equations may be required for SCBFs because both brace buckling and tensile yield
serve as controlling yield mechanisms for some failure modes. BRCBFs have identical yield
resistance in both tension and compression, and so only Eq. 1a is required.

The β factor used in these balance equations is similar to the φ factor in LRFD design (AISC
2001) in that both are less than 1.0 and both are based upon the performance and variability
of the structural elements. However, φ is based solely upon strength, safety, and statistically
extreme considerations, while β depends upon balancing the expected inelastic seismic
behavior to meet the performance requirements and inelastic deformation capacity.

Table 1. Overview of proposed performance based design strategy.

Different β values will be required for different design limit states and for different failure
mode-yield mechanism combinations, because brace connection design must consider the
full range of seismic behavior at all performance levels. Connection strength and stiffness
must be adequate to insure serviceable seismic performance, but once this adequate

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 173


strength and stiffness is provided, system ductility and deformation capacity become the
dominant concerns. As a result, smaller value of β is required when a given failure mode is
difficult to predict or has undesirable consequences, and larger β values are appropriate for
failure modes that provide larger inelastic deformation capacity or better and more
predictable performance. Further, different seismic excitation demands may be warranted
for different performance based design goals as illustrated in Table 1. The balance
equations are used to ensure the desired progression of yielding, prevent premature failure,
and ensure that undesirable failure modes are prevented. Appropriate β factors must be
established for each failure mode to provide the desired progression of yielding and seismic
inelastic performance. BRCBFs require that the connection satisfy these goals without local
buckling or out-of-plane deformation, while the SCBF connections must tolerate large out-of-
plane rotation while achieving the above resistance goals. Different β factors are likely
needed for these two alternatives.

Some failure modes may be controlled by prescriptive design measures. For example,
fracture or tearing of the SCBF brace is theoretically controlled by global and local
slenderness limits. In addition, geometric constraints are employed with SCBF gusset plate
connections to assure that connection can tolerate brace end rotations due to post-buckling
brace deformation. BRCBFs require great control over the connection deformation to avoid
premature brace damage, and so additional constraints may be needed for out-of-plane
stiffness of these BRCBF connections. This research work focuses on improving the
seismic design of these gusset plate connections through:

• better understanding of these yield mechanisms and failure mode


behaviors,
• development of simple but accurate models for prediction of the
resistance associated with all of these yield mechanisms and failure
modes,
• developing appropriate β factors for assuring adequate seismic
performance of braced frame gusset plate connections with each of these
yield mechanism and failure mode combinations, and
• application of these combined developments through balance conditions
such as provided in Eqs. 1a and 1b.

These efforts are the continuing focus of the ongoing work.

CONTINUING WORK

This research work is in progress and will continue for the next two years. At present, past
research is being compared and evaluated and this database will be used to develop and
verify accurate but simple models for predicting connection behaviors. The inelastic
performance achieved with different yield mechanism and failure mode conditions for SCBF
and BRCBF systems in these past research studies will be assessed. Balance conditions (β
values) necessary to assure the proper combination of behaviors are achieved in practice
will then be developed to establish a proposed gusset plate connection design procedure.
The proposed connection design procedure will then be evaluated experimentally. The
experimental results will be used to modify the design procedure to assure that it provides
the greatest design economy combined with good seismic performance at all performance
levels. Follow-up experiments will be conducted to evaluate the design modifications.

The experiments will be designed to evaluate bolted and welded gusset plate connections.
The test matrix will be developed to consider variation in the type of brace, type of

174 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


connection, and balance conditions. A schematic of the test configuration is depicted in
Figure 10. The goal is to test approximately 25 specimens. Each specimen will consist of a
large-scale braced bay typical of that expected in the bottom story of a 3 to 4 story building.
One or more specimens will evaluate the expected performance with the existing SCBF
design requirements [AISC (2002)] and will serve a reference specimen for that framing
system. Another reference specimen will model a braced frame connection in a BRCBF. The
remaining specimens will evaluate the proposed braced frame connection design criteria for
SCBF and BRCBF systems. In these later tests different failure modes and yield
mechanisms will be evaluated, and improved models for predicting their behavior will be
examined and developed.

The test specimens will be constructed and attached to the laboratory strong floor as
depicted in Figure 10. Each specimen will consist of end gusset plates attached to a brace
and the surrounding beam and column framing. This configuration should insure realistic
boundaries for the test specimens. The imposed displacement history will include cyclic
deformation with multiple cycles of increasing story drift such as employed with the ATC-24
testing protocol. Initial cycles will be at deformations below the initial yield and buckling
loads of the brace to examine Operational and Immediate Occupancy performance limit
states. Subsequently, multiple cycles will be completed at and slightly above the buckling
load and tensile yield load of the brace. Finally multiple cycles will then be completed with
increasing inelastic story drift until ultimate failure of the brace or the connection occurs.
These later cycles will be documented with emphasis upon the Life Safety and Structural
Collapse Prevention performance limit states. Work on these tests will begin in Spring 2004.

Figure 10. Schematic of proposed test assembly.

ACKNOWLEDGEMENT

This research work is funded by the National Science Foundation through Grant CMS-
0301792, Performance-Based Seismic Design of Concentrically Braced Frames. Dr. Steven
L. McCabe is the Program Manage for this research. This financial support is gratefully
acknowledged.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 175


NOTATION

Ag gross cross sectional area of the brace


Rn failure mode resistance for an individual failure mode
Ry factor for adjusting the nominal yield stress
Fcr critical buckling stress
Fy expected or average yield stress
β balance factor
φ resistance

REFERENCES

(1) Ando, N. Takahasi, S. and Yoshida, K., (1993) "Behavior of Unbonded Braces
Restrained by Reinforced Concrete and FRP," ASCE, Composite Construction II, New
York, pgs 869-882.
(2) AISC (2001) "Manual of Steel Construction, Load and Resistance Factor Design," 3rd
Edition, American Institute of Steel Construction, Chicago, IL.
(3) AISC (2002). "Seismic Provisions for Structural Steel Buildings," American Institute of
Steel Construction, Chicago, IL.
(4) Astaneh-Asl, A., Goel, S.C., and Hanson, R.D., (1982) "Cyclic Behavior of Double
Angle Bracing Members with End Gusset Plates," Research Report UMEE 82R7,
Department of Civil Engineering, University of Michigan, Ann Arbor, MI.
(5) Clark, P.W., Kasai, K., Aiken, I.D., and Kimura, I., (2000) "Evaluation of Design
Methodologies for Structures Incorporating Steel Unbonded Braces for Energy
Dissipation," Proceedings 12th WCEE, Auckland, New Zealand.
(6) Connor, J.J., Wada, A., Iwata, M., and Huang, Y.H., (1997) " Damage-Controlled
Structuresl I: Preliminary Design Methodology for Seismically Active Regions," ASCE,
Journal of Structural Engineering, Vol. 123, No. 4, pgs 423-31.
(7) Foutch, D.A., Goel, S.C. and Roeder, C.W., (1987) Seismic testing of a full scale steel
building - Part I, Journal of Structural Division, ASCE, No. ST11, Vol. 113, New York,
pgs 2111-29.
(8) Goel, S.C. (1992). "Earthquake Resistant Design of Ductile Braced Steel Structures,"
Stability and Ductility of Steel Structures Under Cyclic Loading, edited by Y. Fukumoto
and G.C. Lee, CRC Press, Boca Raton, Florida.
(9) Inoue, K., Sawaizumi, S., and Higashibata, Y., "Stiffening Requirements for Unbonded
Braces Encased in Concrete Panels, ASCE, Journal of Structural Engineering, Vol
127, No.6, pgs 712-19.
(10) Kahn, L.F., and Hanson, R.D., (1976). "Inelastic Cycles of Axially Loaded Steel
Members," Journal of Structural Division, ASCE, No. ST5, Vol. 102, pgs 947-59.
(11) Lee, S., and Goel, S.C., (1987). "Seismic Behavior of Hollow and Concrete Filled
Square Tubular Bracing Members," Research Report UMCE 87-11, Department of
Civil Engineering, University of Michigan, Ann Arbor, MI.
(12) Popov, E.P. Takanashi, K., and Roeder, C.W. (1976) Structural Steel Bracing
Systems, EERC Reprot 76-17, University of California, Berkeley, 1976
(13) Roeder, C.W., (2001) “State of Art Report – Connection Performance”, FEMA 355D,
Federal Emergency Management Agency, Washington, D.C.
(14) Roeder, C.W., (2002) "Connection Performance for Seismic Design of Steel Moment
Frames," approved for publication, ASCE, Journal of Structural Engineering.

176 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


EFFECT OF COLUMN STIFFENER DETAILING AND WELD
FRACTURE TOUGHNESS ON THE PERFORMANCE OF WELDED
MOMENT CONNECTIONS

R.J. Dexter, University of Minnesota, United States of America


J.F. Hajjar, University of Minnesota, United States of America
D. Lee, Research Institute of Industrial Science & Technology, South Korea

ABSTRACT
Full-scale cyclically loaded cruciform experiments with weak panel zones and
welded unreinforced flange-welded web (WUF-W) prequalified moment
connections performed well provided that the weld metal has minimum Charpy
V-Notch (CVN) toughness. Welds with low CVN and brittle fracture was
obtained in one specimen despite using weld metal certified to meet minimum
CVN requirements. Unstiffened columns perform well and alternative stiffening
details all performed well, indicating that variations in column stiffening details
did not affect the potential for fracture or low-cycle fatigue. Low-cycle fatigue
performance is compared to strain range-cycles curves extrapolated from high-
cycle S-N curves.

INTRODUCTION

Following the Northridge earthquake of January 17, 1994, extensive damage to steel
moment connections was reported (1-3). This damage most often consisted of brittle
fractures of the bottom girder flange-to-column flange Complete Joint Penetration (CJP)
groove welds. The fractures were caused by the use of low toughness welds combined with
a number of other connection detailing and construction practices that were typical prior to
the earthquake (3-5). Additionally, column stiffening practices have been cited as a possible
contributor to the fractures, largely as a result of observations that many of the connections
fractured during the Northridge earthquake lacked continuity plates and that some had weak
panel zones (6). Finite element analyses (7-10) also have shown an increase in stress and
strain concentrations in the girder flange-to-column flange CJP welds associated with
excessively weak panel zones or insufficient continuity plates. It is presumed that these
stress and strain concentrations increase the potential for fracture. As a result of these
observations, there has subsequently been a tendency to be more conservative than
necessary in designing and detailing of the continuity plates and doubler plates in steel
moment connections.

Recommendations for the seismic design of new steel moment-frame buildings (3) provide
equations for determining whether continuity plates are required, and indicate that any
required continuity plates must be of equal thickness to the girder flange for interior
connections (thinner continuity plates are permitted for exterior connections), unless
connection qualification testing demonstrates that the continuity plates are not required.
Furthermore, the connection of the continuity plates to the column flanges must be made
with CJP welds, and reinforcing fillet welds should be placed under the backing bars.

Design criteria for the limit states related to column stiffening are presented in the AISC

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 177


LRFD Specification (11). The limit states of primary importance for stiffening of connections
include Local Flange Bending (LFB), Local Web Yielding (LWY), and Panel Zone yielding
(PZ). Additional provisions for seismic design of doubler plates and continuity plates were
included in the AISC Seismic Provisions (12). However, AISC (13) removed all continuity
plate design procedures for Intermediate and Special Moment Frames, requiring instead that
they be proportioned based on connection qualification tests.

The tendency towards being more conservative than necessary in column stiffener design
has raised concerns about economy as well as the potential for cracking of the k-area in the
column web near the web-flange junction during fabrication due to high residual stresses
caused by highly restrained CJP welds on the continuity plates or doubler plates (14,15).
Therefore, a combined experimental and computational research study was conducted at the
University of Minnesota to reassess the recent column stiffener design and detailing
provisions and recommendations, and to provide economical alternative stiffener details that
minimize welding along the column k-line while retaining superior performance for non-
seismic and seismic design (16). This paper examines the effect of variations in column
stiffening, including no stiffening as well as more economnical alternative stiffening details,
on the fracture and low-cycle fatigue performance of Welded Unreinforced Flange-Welded
Web (WUF-W) moment connections (3).

The design and results of these tests are reported elsewhere, including an assessment of
LFB and LWY limit states (17) and the cyclic panel zone behavior and design (18). Related
research included nine pull-plate experiments (19-22) that investigated the limit states of LFB
and LWY, primarily for non-seismic design, and further tested the alternative doubler plate
and continuity plate stiffener details. Finite element analyses of all experimental specimens
were also conducted as part of this research as well as parametric studies to extend the
results to member sizes and details not tested (23).

FULL-SCALE CONNECTION TESTS

Six full-scale, girder-to-column cruciform specimens were tested (Table 1). The SAC (24)
loading history was applied including six cycles at each interstory drift level of 0.375%, 0.5%,
and 0.75%, four cycles at 1.0% interstory drift level, and two cycles at each interstory drift
level of 1.5%, 2.0%, 3.0%, and 4.0%.

ASTM A992 wide-flange sections and A572 Grade 50 plate was used. The specimens used
the pre-qualified (3) WUF-W connection detail (Figure 1). The column stiffening was varied
in these specimens including three alternative doubler plate details (i.e., back-beveled fillet-
welded doubler plate, square-cut fillet-welded doubler plate, and groove-welded box doubler
plate; see Figure 2) and a fillet-welded ½ in. thick fillet-welded continuity plate detail.

Table 2 presents the design strength-to-demand ratios using minimum specified material
properties for the LFB and LWY limit states. Note that the design strength for LFB and LWY
is the design strength of the column shape alone and does not include the column
reinforcement, if any. The demand is calculated with various methods:

• Yield strength of the girder flange,


• A value of 1.8 times the yield strength of the flange as in the 1992 AISC Seismic
Provisions (12)
• Equation included in AISC Design Guide No. 13 (25) which amplifies the plastic
moment due to shear and reduces this moment to a force couple of the flanges. This
equation, or slightly modified forms of it, has been widely used for the design of

178 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


column stiffeners in steel moment connections (3).

Table 2 shows that 1992 AISC Seismic Provisions (12) and AISC Design Guide No. 13 (25)
provide similar demand values. For these cases, the demand exceeds the capacity for LFB
for all specimens except CR1, whereas only specimen CR3 has continuity plates (although
the box detail functions as a continuity plate as well as a doubler). The remaining specimens
CR2 and CR5 are underdesigned for LFB for seismic demand.

Table 1. Test matrix of cruciform specimens.


CR4
CR1 CR2 CR3 CR5
and CR4R
Girder W24x94 W24x94 W24x94 W24x94 W24x94
Column W14x283 W14x193 W14x176 W14x176 W14x145
Doubler Plate Detail III
None Detail II Detail II Detail I
(DP) Box (Offset)
DP Thickness NA 0.625 in. 2 @ 0.5 in. 2 @ 0.75 in. 2 @ 0.625 in.
Continuity Fillet-
None None None None
Plate (CP) welded
CP Thickness NA NA 0.5 in. NA NA

Figure 1. Typical welding details used for cruciform specimens (Specimen CR1).

However, the girder flange demand predicted by the latter two methods is very conservative
and can be put in perspective by comparing to the maximum possible uniaxial tensile
strength of A992 steel. The stress in the flange is 1.8 times 50 ksi or 90 ksi, well above the
likely tensile strength of A992 steel. For example, a survey of more than 20,000 mill reports
from (26,27) showed that A992 steel has a mean tensile strength of 73 ksi. The 97.5
percentile tensile strength was 80 ksi, and the maximum value reported was 88 ksi.

Also shown in Table 2 are the panel zone capacity-to-demand ratios (including the strength
of the doubler plate in the capacity) and column-girder moment ratios calculated from the

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 179


2002 AISC Seismic Provisions (28), assuming no axial compression in the column. The
panel zone demand exceeds the capacity for all specimens. Therefore, if column stiffening
were necessary to prevent premature brittle fracture or low-cycle fatigue, these specimens
are a worst-case test since they are underdesigned.

Figure 2. Doubler details: (a) back-beveled fillet-welded doubler (Detail I), (b) square-cut
fillet-welded doubler (Detail II), (c) box (offset) doubler (Detail III).

Table 2. Nominal capacity/demand ratios of PZ yielding, LFB, and LWY limit states.
LFB φRn/Ru LWY φRn/Ru
∑ M *pc PZ
∑ M *pb φvRv/Ru 1.8Yield a 1.8Yield a
a a a
(AISC) Yield 1992 DG13 Yield 1992 DG13 a
(AISC)
Seismic Seismic
3.04 1.69 1.64
CR1 1.50 0.72 2.38 1.32 1.29
(3.38)b (1.88) (1.82)
1.47 0.82 0.80
CR2 0.99 0.66 2.20 1.22 1.19
(1.63) (0.91) (0.89)
1.22 0.68 0.66
CR3 0.89 0.74 2.51 1.39 1.36
(1.36) (0.76) (0.73)
CR4 1.22 0.68 0.66
0.89 0.93 3.19 1.77 1.73
CR4R (1.36) (0.76) (0.73)
0.84 0.47 0.46
CR5 0.73 0.74 2.34 1.30 1.27
(0.93) (0.52) (0.51)
a
Equation used to calculate demand, Ru
b
Values in parentheses reflect use of φ = 1.0

Weld details

E70T-1 (Lincoln Outershield 70) wire with 100% CO2 shielding gas was used for all shop
welding. The girder flange-to-column flange CJP groove welds were made in the flat position
with E70T-6 (Lincoln Innershield NR-305) wire. Welds made with E70T-6 wire are required
by AWS A5.20 (29) and AISC 2002 Seismic Specifications (28) to have notch toughness of
20 ft-lbs at -20°F. FEMA 350 (3) has recommended minimum notch toughness requirements
at two temperatures, 20 ft-lbs at 0°F and 40 ft-lbs at 70°F. According to the Lincoln Electric
Company product family literature, the typical values for NR-305 are 21 to 35 ft-lbs at -20°F
and 21 to 54 ft-lbs at 0°F. As shown in Table 3, Specimens CR1 and CR4 were fabricated
with a 5/64 in. diameter NR-305 wire and the remaining were fabricated with 3/32 in.
diameter NR-305 wire. All CJP welds were ultrasonically tested by a certified inspector in
conformance with Table 6.3 of AWS D1.1-2000 (30) for cyclically loaded joints.

The out-of-position field welds, including the CJP welds connecting the girder web to the
column flange and all reinforcing fillet welds were made with 0.068 in. diameter E71T-8
(Lincoln Innershield NR-203MP) wire for Specimens CR1 and CR4, and 5/64 in. diameter

180 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


E71T-8 (Lincoln Innershield NR-232) wire for the other specimens.

The shear tab was designed to extend approximately 0.25 in. into the top and bottom access
holes and acted as the backing bar for the CJP welds of the girder web to the column flange.
This extension acted as a short runoff tab, allowing the weld to extend the full depth of the
girder web. Ricles et al. (31) recommended that these runoff tabs of the vertical web weld be
ground smooth, which is labor intensive. Since it was felt that this might not be necessary,
these runoff tabs were not ground smooth in the specimens tested in this work.

Table 3. Tested weld material properties (E70T-6 only).


E70T-6a E70T-6
5/64 in. wire 3/32 in. wire
CR1a CR4a CR2 CR3 CR4R CR5
CVN @ 0°F
2.6 2.0 34.3 44.3 33.0 33.0
(ft-lbs)
CVN @ 70°F
19.3 2.3 54.3 73.3 58.7 53.7
(ft-lbs)
Fy (ksi) NA NA 59.5 50.0 56.0 53.5
Fu (ksi) NA NA 79.5 72.5 78.2 75.5
% Elongation NA NA 25.0 23.0 27.5 26.0
a
For Specimens CR1 and CR4, the CVN tests were performed on specimens machined after
the experiment from the welds that did not fracture in the cruciform joints

BRITTLE FRACTURE

Specimen CR4 exhibited premature brittle failure in three of four girder flange-to-column
flange CJP welds in the early stage of the SAC (1997) loading history and was stopped after
one-half cycle at 2.0% interstory drift. It was found that this specimen was unintentionally
prepared with low-toughness weld metal, as shown in Table 3. Note that the AWS
Certificate of Conformance for this wire indicated that the weld metal meets the minimum
toughness requirement of AWS A5.20 (28) of 20 ft-lbs at -20°F (32).

Specimen CR4R was essentially a replicate test of Specimen CR4, except that the weld
metal used for Specimen CR4R met the minimum requirements of FEMA 350 (3). Specimen
CR4R not only performed acceptably according to the SAC (24) requirements, it performed
as well as any of the other specimens successfully tested in this experimental study. The
fact that the box (offset) doubler plate detail performed well in Specimen CR4R indicates that
the detail itself was probably not a factor in the fracture that occurred in Specimen CR4.

Following the premature brittle failure in Specimen CR4, it was found that the previously
tested Specimen CR1 also had relatively low weld-metal notch toughness, with an average
of 2.6 ft-lbs at 0°F and 19.3 ft-lbs at 70°F as presented in Table 3. Specimen CR1, which
was welded using the same wire and the similar welding procedure as Specimen CR4, but
had marginally better notch toughness, performed very well, experiencing 14 cycles of 4.0%
interstory drift before the significant strength degradation. It is important to note that this
Specimen CR1 had no doubler plates or continuity plates, even though doubler plates would
be required as shown in Table 2. Thus this test shows that column stiffeners are not
absolutely required to avoid brittle fractire or low-cycle fatigue, even with this very poor notch
toughness. These two tests have closely bracketed the minimum notch toughness required
for adequate performance of CJP welds.

It is believed that the FEMA 350 requirements (3) for minimum notch toughness are

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 181


adequate, provided they can be consistently met. FEMA 353 (4) requires toughness testing
on each production lot of the specified filler metal. However, lot testing is typically not done
since FEMA 353 also allows this requirement to be waived (upon approval of the Engineer),
instead relying upon the consumable manufacturer’s certification. However, the certification
is not necessarily reliable; for example, this 5/64-in. diameter E70T-6 wire that produced
these brittle welds had been certified by the manufacturer as meeting the minimum 20 ft-lbs
at -20°F required by the AWS certification test (32). Therefore, further evaluation of the CVN
testing requirements for weld metal is warranted.

LOW-CYCLE FATIGUE

The remaining connections exhibited no brittle fractures, but low-cycle fatigue failures occurred
after significant cyclic loading. Figures 3 and 4 show low-cycle fatigue cracks forming at the
beam flange weld. These beam flange weld cracks were the only type of low-cycle fatigue
cracks that actually propagated to cause failure, which is defined here as significant strength
reduction. Low-cycle fatigue cracks did originate at the weld of the beam web to the column
flange and at the weld access hole, as shown in Figure 5. However after propagating for a short
distance they arrested and did not propagate further or lead to failure, so therefore they are not
structurally significant.

(a) (b)
Figure 3. Low cycle fatigue crack developing at the toe of the beam flange weld in a moment-
frame connection after (a) 11 cycles and (b) 17 cycles of 4% drift.

Table 4 shows the cycles at 3% or 4% drift when the first crack was first visible in the CJP
welds, the final cycles at 4% drift when the strength was reduced, and the measured strain
ranges. The pairs of measured strain ranges for each specimen are from the west top flange
and the east bottom flange, respectively (except for CR1 where only the west top flange data
were avaialble). The average of five gages across the width was used to eliminate some of
the scatter and the effect of strain gradients. It is believed that the variation in the measured
strain ranges is random, and that the strain range at 4% drift was relatively consistent among
the specimens, averaging approximately 4.1%.

The performance requirement is that the connections must complete 2 cycles at 4% drift without
a significant reduction in strength in order to be prequalified connections (3). One could
conclude that all these connections (except the original CR4, which experienced brittle fracture)
met this performance requirement and that therefore the performances of the specimens are
equally good. However, there may be some significance to the final number of cycles before

182 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


strength reduction. In particular, it is believed that the differences in number of cycles to strength
reduction between 12 and 16 are due to random variability, and that therefore specimens CR1,
CR2, CR3, and CR4R performed equally well. This means that the variation in column stiffener
detailing and panel zone strength among these specimens had no significant influence on the
low-cycle fatigue performance. For specimen CR5, the number of cycles before strength
reduction is 6. As shown in Table 2, this specimen had the lowest ratio of capacity to demand
for LFB – less than one even for non-seismic (nominal yield strength of flange) and less than 0.5
for either seismic demand. It is likely that the relatively smaller number of cycles in this under-
designed specimen is due to lack of continuity plates.

LCF Crack

Column Column
Web Flange

Slag Inclusion,
LOF

Figure 4. Cross section of beam flange weld showing low cycle fatigue crack
developing at the weld toe.

Figure 5. Low cycle fatigue cracks forming at the end of the beam web to column flange
weld and at the weld access hole.

Table 4. Low-cycle fatigue data.


Specimen CR1 CR2 CR3 CR4R CR5
Column W14x283 W14x193 W14x176 W14x176 W14x145
Doubler Plate None Detail II Detail II Box Detail I
Continuity Plate None None Fillet welded None (box) None
Cycles when Crack
11@4% 2@3% 2@3% 2@4% 1@3%
Visible
Cycles at 4% when
14 16 14 12 6
Strength Reduced
Strain Range in
Girder Flange at 4.2% 4.5%, 3.7% 3.8%, 4.5% 3.8%, 3.8% 3.7%, 4.8%
4% Drift

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 183


As noted above, specimen CR1 had much lower notch toughness than the others.
Therefore, it appears that, as long as the notch toughness is sufficient to preclude brittle
fracture, notch toughness above this minimum level also has little influence on low-cycle
fatigue performance. Finally, for Specimens CR1 through CR4, there is little correlation
betweeen the cycles when the crack was first detected in the CJP welds, which was highly
variable, and the final number of cycles when strength was reduced. Therefore, it is believed
that the number of cycles at which the crack is first detected is not significant.

Most past research on low-cycle fatigue has involved pressure vessels and some other types
of mechanical engineering structures. Since low-cycle fatigue is an inelastic phenomenon,
the strain range is the key parameter rather than the stress range as in high-cycle fatigue.
The Coffin-Manson rule (33) has been used to relate the strain range in smooth tensile
specimens to the fatigue life. Manson suggested a conservative lower-bound simplification,
called Manson’s universal slopes equation (34):
σu
∆ε = 3.5 N −0.12 + ε 0f .6 N −0.6 (1)
E
where: ∆ε is the total strain range, σu is the tensile strength, and εf is the elongation at fracture.

Note that the first term in Equation 1 is the elastic part of the total strain range (which is relatively
insignificant when there are fewer than 100 cycles) and the second term is the plastic part of the
total strain range. Figure 6 shows a plot of Manson’s universal slopes equation where σu is 450
MPa and εf is 25%, typical minimum properties for Grade 50 structural steel. Many studies have
shown that Manson’s universal slopes equation is conservative compared to experimental data
from smooth specimens (34,35). However, because of buckling at greater strain ranges, most of
the experimental data are for strain ranges less than 1%, i.e., for cycles greater than 1000.
Limited data exist at higher strain ranges – some are shown in Figure 6 for A36 steel smooth
specimens machined from the flanges of wide-flange sections (35).

At this time, very little is understood about low-cycle fatigue in welded or bolted structural details.
For example, it is a very difficult task to predict accurately the local strain range at a location of
cyclic local flange buckling. However, Krawinkler and Zohrei (36) and Ballio and Castiglioni
(37,38) showed that the number of cycles to failure by low-cycle fatigue of welded connections
could be predicted by the local strain range in a power law that is analogous to an S-N curve.
Ballio and Castiglioni (37,38) showed that the power law would have and exponent of 3, just like
the elastic S-N curves. Krawinkler and Zohrei (36) also showed that Miner’s rule (39) could be
used to predict the number of variable-amplitude cycles to failure based on constant amplitude
test data.

Therefore, it may be possible to predict and design against low-cycle fatigue using strain-range
vs. number-of-cycles curves that are extrapolated from the high-cycle fatigue design S-N curves.
Figure 6 shows the AISC S-N curves (11) for Categories A and C, converted from stress range
to strain range by dividing the stress ranges by the elastic modulus, and extrapolated up to one
cycle.

There are only limited data to support this approach. Figure 6 shows the strain range-number of
cycles data from these WUF-W beam-to-column connection tests. The number of cycles plotted
in Figure 6 is the number of cycles at 4% drift. If the effect of all the previous cycles is included
using Miner’s rule (39), they add up to an equivalent of about one additional cycle at 4%. Since
flange strains right at the weld toe were used rather than nominal values, this is analogous to a
hot-spot approach for high-cycle fatigue. For high-cycle fatigue, the Category C S-N curve is a
suitable baseline S-N curve for the hot-spot approach (40,41). It appears that the Category C S-
N curve is also a good lower bound to these low-cycle fatigue data. The scatter in the data is
substantial, as is also true in high-cycle fatigue.

184 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Also shown in Figure 6 are previously unpublished data for smaller coupon-type specimens with
transverse butt welds, which would be expected to be Category C details. These are some of
the only available data with fewer than 5,000 cycles. These coupon data are also in reasonable
agreement with the extrapolated Category C curve as a lower bound.
100
welded coupon
Manson's Equation smooth specimen
WUF-W connection tests
Strain Range [%]

Category A
10
Category C

0.1
1 10 100 1000 10000
Cycles
Figure 6. Comparison of standard S-N curves presented in terms of strain range and Manson’s
universal slopes equation for Grade 50 (350 MPa yield strength) steel to low-cycle fatigue test
data and the connection test data.

CONCLUSIONS AND RECOMMENDATIONS

1. When properly detailed and welded with notch-tough filler metal, the WUF-W steel
moment connections can perform adequately even though relatively weak panel zones and
low local flange bending strengths were chosen.
2. The failure mode of the specimens other than the original CR4 was low-cycle fatigue
(LCF) crack growth and eventual rupture of one or more girder flange-to-column flange
complete joint penetration (CJP) groove welds. Low cycle fatigue may be conservatively
predicted using strain-range vs. cycles curves derived from the stress based S-N curves for
high-cycle fatigue.
3. Specimens CR1 and CR4 were unintentionally prepared with weld metal that had CVN
values that were much lower than the minimum requirements. The premature brittle failure of
specimen CR4 reconfirmed that achieving the required minimum CVN toughness in the
girder flange-to-column flange CJP welds is critical. These low toughness welds occurred
despite the certification of the filler metal; the certification is only required annually, unlike the
way that each heat of steel is tested. A study should be conducted to fully characterize the
typical variability in the CVN and other properties of the weld. An evaluation of the need for
lot testing should be performed. Consideration should also be given to use of filler metals
with a distribution of CVN such that there is a sufficiently small probability of not meeting the
minimum required values, and therefore lot testing may not be required.
4. Application of the alternative column stiffener details (i.e., back-beveled fillet-welded
doubler plate detail; square-cut fillet-welded doubler plate detail; groove-welded box doubler
plate detail; fillet-welded 1/2 in. thick continuity plates) in the WUF-W steel moment
connections was successfully verified. No cracks or distortions were observed in the welds
connecting these stiffeners before the rupturing of the girder flange-to-column flange CJP

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 185


welds.
5. Specimens CR1, CR2, CR4R, and CR5, none of which had continuity plates (although
Specimen CR4R included the offset doubler plate detail), met the requirements for two
cycles at 4.0% interstory drift, athough only Specimen CR1 met the seismic requirements of
AISC and FEMA with respect to continuity plates. Continuity plates may thus not be
necessary in many interior columns in steel moment connections, and design provisions
permitting the design, or lack of inclusion, of continuity plates are recommended for
reintroduction into the AISC Seismic Provisions.
6. For a wide range of column sections and doubler plate detailing, strain gradients and
strain magnitudes well above the yield strain in the girder flanges did not prohibit the
specimens from achieving the connection prequalification requirement of completing two
cycles at 4.0% interstory drift without significant strength degradation. This was even the
case for specimen CR1, which had notch toughness in the weld metal that was significantly
below the requirements. These results indicate that the column reinforcement detailing may
not have a significant effect on the potential for brittle fracture at the girder flange-to-column
flange weld. Note that this is contrary to previous finite-element analyses reported in the
literature using theoretical fracture criteria that have predicted a significant effect of using or
omitting continuity plates.
7. If continuity plates are required, fillet-welded continuity plates that were approximately
half of the girder flange thickness performed well. The results showed that only minor local
yielding occurred in these continuity plates in part of the most stressed cross sections at
peak drift level and that these strains were not sufficient to cause cracking or distortion in the
continuity plate or to change the strain gradients in the girder flange substantially. Since
continuity plates do not significantly yield, it may not be necessary to size the welds large
enough to develop the continuity plate. Rather the weld and the plate may only need to be
designed for the difference between the demand and the capacity of the column shape
without the continuity plate. Continuity plates with undersized fillet welds should be tested to
confirm that the weld need not develop the full continuity plate strength.
8. In all the tests except the original CR4, the seismic performance of the relatively weak
panel zones was stable and ductile, and the panel zones exhibited good energy dissipation.
Lee et al. (17-23) provide recommended changes to the AISC panel zone strength
equations, as well as detailed evaluations of current AISC provisions for local flange bending,
local web yielding, and panel zone shear.

ACKNOWLEDGEMENT

This research was sponsored by the American Institute of Steel Construction, Inc. and by the
University of Minnesota. In-kind funding and materials were provided by LeJeune Steel
Company, Minneapolis, Minnesota; Danny’s Construction Company, Minneapolis,
Minnesota; Braun Intertec, Minneapolis, Minnesota; Nucor-Yamato Steel Company,
Blytheville, Arkansas; Lincoln Electric Company, Cleveland, Ohio; and Edison Welding
Institute, Cleveland, Ohio. Supercomputing resources were provided by the Minnesota
Supercomputing Institute. The authors wish to thank T. V. Galambos and P. M. Bergson,
University of Minnesota, L. A. Kloiber, LeJeune Steel Company, and the members of the
technical advisory group on this project for their valuable assistance.

REFERENCES

1. Youssef, N. F. G, Bonowitz, D., and Gross, J. H. (1995). A Survey of Steel Moment-


Resisting Frame Buildings Affected by the 1994 Northridge Earthquake, Report No.
NISTIR 5625, National Institute of Standards and Technology, Gaithersburg, Maryland.

186 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


2. Northridge Reconnaissance Team (1996). Northridge Earthquake of January 17, 1994,
Reconnaissance Report (Supplement C-2 to Volume 11), EERI, Oakland, California.
3. Federal Emergency Management Agency (FEMA) (2000). Recommended Seismic
Design Criteria for New Steel Moment-Frame Buildings, Report No. FEMA 350, FEMA,
Washington, D.C.
4. FEMA (2000). Recommended Specifications and Quality Assurance Guidelines for
Steel Moment-Frame Construction for Seismic Applications, Report No. FEMA 353,
FEMA, Washington, D.C.
5. Fisher, J.W., R.J. Dexter, and E.J. Kaufmann, “Fracture Mechanics of Welded Structural
Steel Connections,” Report No. SAC 95-09, FEMA-288, March 1997
6. Tremblay, R., Timler, P., Bruneau, M., and Filiatrault, A. (1995). “Performance of Steel
Structures during the 1994 Northridge Earthquake,” Canadian Journal of Civil
Engineering, Vol. 22, pp. 338-360.
7. El-Tawil, S., Mikesell, T., Vidarsson, E., and Kunnath, S. (1998). “Strength and Ductility
of FR Welded-Bolted Connections,” Report No. SAC/BD-98/01, SAC Joint Venture,
Sacramento, California.
8. El-Tawil, S., Vidarsson, E., Mikesell, T., and Kunnath, S. K. (1999). “Inelastic Behavior
and Design of Steel Panel Zones,” Journal of Structural Engineering, ASCE, Vol. 125,
No. 2, pp. 183-193.
9. El-Tawil, S. (2000). “Panel Zone Yielding in Steel Moment Connections,” Engineering
Journal, AISC, Vol. 37, No. 1, pp. 120-131.
10. Mao, C., Ricles, J. M., Lu, L., and Fisher, J. W. (2001). “Effect of Local Details on
Ductility of Welded Moment Connections,” Journal of Structural Engineering, ASCE, Vol.
127, No. 9, pp. 1036-1044.
11. American Institute of Steel Construction (AISC) (1999). Load and Resistance Factor
Design Specification for Structural Steel Buildings, AISC, Chicago, Illinois.
12. AISC (1992). Seismic Provisions for Structural Steel Buildings, AISC, Chicago, Illinois.
13. AISC (1997). Seismic Provisions for Structural Steel Buildings, AISC, Chicago, Illinois.
14. AISC (1997). “AISC Advisory on Mechanical Properties Near the Fillet of Wide Flange
Shapes and Interim Recommendations, January 10, 1997,” Modern Steel Construction,
AISC, Chicago, Illinois, February, p. 18.
15. Tide, R. H. R. (2000). “Evaluation of Steel Properties and Cracking in k–area of W
Shapes,” Engineering Structures, Vol. 22, No. 2, pp. 128-134.
16. Lee, D., Cotton, S., Dexter, R. J., Hajjar, J. F., Ye, Y., and Ojard, S. D. (2002). “Column
Stiffener Detailing and Panel Zone Behavior of Steel Moment Frame Connections, ”
Report No. ST-01-3.2, Department of Civil Engineering, University of Minnesota,
Minneapolis, Minnesota.
17. Lee, D., Cotton, S. C., Hajjar, J. F., Dexter, R. J., Ye, Y., and Ojard, S. D. (2004). “Cyclic
Behavior of Steel Moment-Resisting Connections Reinforced by Alternative Column
Stiffener Details. I. Connection Performance and Continuity Plate Detailing,”
Engineering Journal, AISC, submitted for publication.
18. Lee, D., Cotton, S. C., Hajjar, J. F., Dexter, R. J., Ye, Y., Ojard, S. D. (2004). “Cyclic
Behavior of Steel Moment-Resisting Connections Reinforced by Alternative Column
Stiffener Details. II. Panel Zone Behavior and Doubler Plate Detailing,” Engineering
Journal, AISC, submitted for publication.
19. Prochnow, S. D., Dexter, R. J., Hajjar, J. F., Ye, Y., and Cotton, S. C. (2000). “Local
Flange Bending and Local Web Yielding Limit States in Steel Moment-Resisting
Connections,” Report No. ST-00-4, Department of Civil Engineering, University of
Minnesota, Minneapolis, Minnesota.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 187


20. Prochnow, S. D., Ye., Y., Dexter, R. J., Hajjar, J. F., and Cotton, S. C. (2002). “Local
Flange Bending and Local Web Yielding Limit States in Steel Moment-Resisting
Connections,” Connections in Steel Structures IV, Roanoke, Virginia, October 22-24,
2000, American Institute of Steel Construction, Chicago, Illinois, pp. 318-328.
21. Dexter, R. J., Hajjar, J. F., Prochnow, S. D., Graeser, M. D., Galambos, T. V., and
Cotton, S. C. (2001). “Evaluation of the Design Requirements for Column Stiffeners and
Doublers and the Variation in Properties of A992 Shapes,” Proceedings of the North
American Steel Construction Conference, AISC, Chicago, Illinois.
22. Hajjar, J. F., Dexter, R. J., Ojard, S. D., Ye, Y., and Cotton, S. C. (2003). “Continuity
Plate Detailing for Steel Moment-Resisting Connections,” Engineering Journal, AISC,
Vol. 40, No, 4, Fourth Quarter, pp. 189-211.
23. Ye, Y., Hajjar, J. F., Dexter, R. D., Prochnow, S. C., and Cotton, S. C. (2000).
“Nonlinear Analysis of Continuity Plate and Doubler Plate Details in Steel Moment
Frame Connections,” Report No. ST-00-3, Department of Civil Engineering, University of
Minnesota, Minneapolis, Minnesota.
24. SAC (1997). “Protocol for Fabrication, Inspection, Testing, and Documentation of Beam-
Column Connection Tests and Other Experimental Specimens,” Report No. SAC/BD-
97/02, SAC Joint Venture, Sacramento, California.
25. AISC (1999). “Stiffening of Wide-Flange Columns at Moment Connections: Wind and
Seismic Applications,” AISC Design Guide No. 13, AISC, Chicago, Illinois.
26. Dexter, R. J. (2000). “Structural Shape Material Property Survey, ” Final Report to
Structural Shape Producer's Council, University of Minnesota, Minneapolis, Minnesota.
27. Bartlett, F.M., R.J. Dexter, M.D. Graeser, J.J. Jelinek, B.J. Schmidt, and T.V. Galambos
(2003). “Updating Standard Shape Material Properties Database for Design and
Reliability”, Engineering Journal, AISC, Vol. 40, No. 1, 1st Quarter, pp 2-14.
28. AISC (2002). Seismic Provisions for Structural Steel Buildings, AISC, Chicago, Illinois.
29. American Welding Society (AWS) (1995). Specification for Carbon Steel Electrodes for
Flux Cored Arc Welding, AWS A5.20-95, AWS, Miami, Florida.
30. AWS (2000). Structural Welding Code – Steel, AWS D1.1-2000, AWS, Miami, Florida.
31. Ricles, J. M., Mao, C., Lu, L., and Fisher, J. W. (2002). “Inelastic Cyclic Testing of
Welded Unreinforced Moment Connections,” Journal of Structural Engineering, ASCE,
Vol. 128, No. 4, pp. 429-440.
32. Lincoln Electric Co. (1999). Certificate of Conformance for Innershield NR-305
(E70T-6), Lincoln Electric Co., Cleveland, Ohio, August 6, 1999.
33. Coffin, L.F. Jr., “A Note on Low Cycle Fatigue Laws”, Journal of Materials, Vol. 6, No. 2,
pp.388-402, 1971.
34. Itoh, Y.Z., Kashiwaya, H., “Low-cycle fatigue properties of steel s and their weld metals”,
Journal of Engineering Materials and Technology, Vol. 111, No. 4, pp. 431-437, 1989.
35. Howdyshell, P., J.C. Trovillion, and J.L. Wetterich, “Low-Cycle Fatigue of Structural
Materials,” Materials and Construction: Proceedings of MatCong 5, the 5th ASCE
Materials Engineering Congress, Bank, L. (ed.), 10-12 May 1999, Cincinnati, OH,
American Society of Civil Eng., Reston, VA., pp. 148-155.
36. Krawinkler, H., and Zohrei, M. (1983). “Cumulative Damage in Steel Structures
Subjected to Earthquake Ground Motion,” Computers and Structures, Vol. 16, No. 1-4,
pp.531-541, 1983.
37. Ballio, G., and Castiglioni, C.A. (1995). “A Unified Approach for the Design of Steel
Structures Under Low and/or High Cycle Fatigue,” Journal of Constructional Steel
Research, Vol. 34, No. 1, pp. 75-101
38. Castiglioni, C.A. (1995). “Cumulative Damage Assessment in Structural Steel Details,”

188 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


IABSE Symposium San Francisco, Extending the Lifespan of Structures, pp. 1061-1066.
39. Miner, M., “Cumulative Damage in Fatigue,” Transactions of the American Society of
Mechanical Engineers, Vol. 67.
40. Dexter, R.J., Tarquinio, J.E., and Fisher, J.W., "Application of the hot-spot stress fatigue
analysis to attachments on flexible plate," Proceedings, 13th International Conference on
Offshore Mechanics and Arctic Engineering, ASME, Vol. III, Materials Engineering, 85-
92.
41. Dexter, R. J., “Fatigue and fracture", The Structural Engineering Handbook, 2nd Edition,
Lui, B. M. ed., CRC Press, 2004.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 189


190 Connections in Steel Structures V - Amsterdam - June 3-4, 2004
ANALYSIS OF BOLTED END-PLATE JOINTS: CYCLIC TEST AND
STANDARD APPROACH

L. Dunai, Budapest University of Technology and Economics, Hungary


N. Kovács, Budapest University of Technology and Economics, Hungary
L. Calado, Instituto Superior Tecnico, Lisbon, Portugal

ABSTRACT
Experimental program is performed on bolted end-plate type joints of 19 steel
and steel-concrete composite specimens under cyclic loading. The test
specimens are designed to study and characterise the typical cyclic behaviour
modes of this type of joints. The paper summarises the details of the
experimental program, the cyclic joint behaviour and the cyclic characteristics.
The joint design characteristics (moment resistance and initial stiffness) are
determined by the Eurocode standard and compared to the experimental joint
parameters. After the verification of the standard procedure a parametric study
is completed to study the influence of the structural details on the design
parameters.

INTRODUCTION

The paper presents an experimental and Eurocode standard based analyses of end-plate
type joints. The experimental program is performed in a co-operation project between the
Budapest University of Technology and Economics and the Instituto Superior Tecnico in
Lisbon. The subject of this experimental program is to test bolted end-plate type steel and
composite joints (19 specimens) under cyclic reversal loading. The experimental program is
published in (1), in this paper only the short summary of the specimen characteristics, the
test set-up and procedure and the behaviour modes are presented.

The aim of the research is to observe the typical cyclic behaviour modes of the studied
connection type, characterize it quantitatively and find structural solution for the improved
cyclic behaviour. On the basis of the experimental results a calculation method is under
development for the determination of the monotonic and cyclic design parameters of the
joints. In the first phase of it Eurocode 3 and 4 based numerical study is performed on the
monotonic joint characteristics (the moment resistance and the initial stiffness). The paper
presents the standard procedure with the comparison of the design values and the
experimental results. Since cyclic tests are performed the envelope curve of the hysteretic
moment–rotation diagram is applied to characterize the monotonic behaviour. The
experimentally verified Eurocode procedure is to be extended to describe cyclic behaviour.
The developed, Eurocode based, calculation method is also presented in the paper with the
comparison of the experimental results.

After the verification of the proposed method a parametric study is performed to study the
influence of structural details on the design moment resistance and initial stiffness. In the
further step of the research the cyclic characteristics (degradation characteristics, absorbed
energy, ductility, etc.) of the joint will be evaluated to gain the improved structural solution for
cyclic behaviour of this type of joints.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 191


TESTING PROGRAM

The test programs are performed on end-plate type bolted joints with different parameters of
each specimen. In case of the first test program the aim is to have information about the
cyclic behaviour modes of the joint (2). During the second experimental program the main
parameters of the specimens are identical with those of the first case; the only difference is
the composite member section (3). In case of the third program the focus is on the cyclic
local buckling of the slender composite section (4).

Specimens, test arrangement

Altogether three test sets are performed on end-plate type joints. In case of the first test set
the specimens are designed with H-shaped steel element (hot-rolled or welded) as it is
illustrated in figure 1 a). In the second and third test set composite member – steel section
with concrete filling and reinforcement between the flanges – is applied as figure 1 b) shows.
Table 1 contains the main characteristics of each specimen.

a) Steel joints b) Composite joint


Figure 1. Specimens’ cross-section.

Table1. Details of the specimen

Steel member Composite member I. Composite member II.


End- End- End-
Bolt Bolt Bolt
Test Member plate Test Member plate Test Member plate
M16 M16 M24
[mm] [mm] [mm]
CB1 HEA-200 25 8.8 CCB1 HEA-200 25 8.8 CCF1 weld. (6mm) 30 10.9
CB2 HEA-200 16 8.8 CCB2 HEA-200 16 10.9 CCF2 weld. (6mm) 20 10.9
CB3 weld. (6mm) 25 8.8 CCB3 weld. (6mm) 25 8.8 CCF3 weld. (4mm) 20 10.9
CB5 HEA-200 12 8.8 CCB4 HEA-200 19 8.8 CCF4 weld. (4mm) 20 4.8

The test setup is developed at the Instituto Superior Tecnico to test beam-to-column joints of
steel frames in cantilever type arrangement (1), as it can be seen in figure 2. The loading
history is defined in accordance with the ECCS Recommendations (5). It is decided to use
the same loading history in all test cases to be able to compare the results of similar
specimens. Inductive displacement transducers are used to measure the displacements
during the testing procedure.

Figure 2. Test arrangement.

192 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Experimental results

On the basis of the measured data the cyclic moment-rotation curves are determined. The
calculation of the moment-rotation diagrams are performed in the so-called ‘joint reference
section’. In this way the disturbing effect of the local buckling of the flanges is eliminated. The
moment-rotation diagrams illustrate the various hysteretic behaviours and characterise
qualitatively the joints by reflecting the tendencies of the rotational stiffness, the moment
carrying capacity, the rotational capacity and the absorbed energy. The behaviour modes are
detailed in (1) and summarised as the follows.

‘Pure’ bolt-failure:
The ‘pure’ bolt-failure occurred when the end-plate is thick enough to resist the failure without
significant deformation (CCB1 and CCB3). The governing phenomenon is the plastic
deformation and bolt fracture after pinching of the bolt shanks. The most representative
phenomenon is the rigid-body type rotation of the specimen, which is caused by the plastic
elongation of the bolt shanks and appears as large deformations on near-zero moment level
as it is shown in figure 3 a). This type of joint has small deformation and moment carrying
capacity limited by the ductility and grade of the bolts. While the bolts govern the behaviour
of the specimen the concrete filling has no significant effect on the cyclic behaviour since the
load carrying and deformation capacities are limited by the characteristics of the bolts. The
specimen after failure is presented in figure 3 b).
60 rigid-body
moment [kNm]

40 type rotation
20
0
-20
-40
-60
-50 -40 -30 -20 -10 0 10 20 30 40 50
rotation [mrad]

a) M-Θ diagram b) Specimen after failure


Figure 3. Bolt-failure.

‘Pure’ plate-failure:
The ‘pure’ plate-failure type behaviour mode occurs when the end-plate is thin comparing to
the other elements (CCB2). The governing phenomenon is the elastic/plastic deformation
and plastic hinge of the end-plate. In the followings subsequent cycles cracks occurred and
propagated in the plate near to the welding of the flanges, which caused resistance
degradation. Plastic deformation of the bolts is not significant. Figure 4 a) shows the
representative moment-rotation diagram of the ‘pure’ plate-failure. The deformation capacity
of this joint is reached higher value than in case of the bolt-failure. The absorbed energy is
also higher. The crack of the end-plate is shown in figure 4 b).
60
40
moment [kNm]

20
0
-20
-40
-60
-60 -45 -30 -15 0 15 30 45 60
rotation[mrad]
a) M-Θ diagram b) Specimen after failure
Figure 4. Plate-failure.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 193


Combined plate+bolt-failure:
Combined behaviour of the connecting elements is developed when both the end-plate and
the bolts have significant plastic deformation (CCB4). The governing phenomenon is the
plastic flow in the end-plate and in the bolt shanks. The final failure is caused by the cracking
and breaking of the end-plate with remarkable plastic flow of the bolts. In figure 5 a) the
moment-rotation diagram of this combined-failure mode is presented. The moment-rotation
diagram has full hysteretic curves, which means large energy absorption of the joint. Small
rigid-body type rotation is observed due to the plastic elongation of the bolts and the bending
of the bolts heads (see figure 5 b). This type of behaviour is the most favourable from the
connecting elements type failure, since besides the significant moment carrying capacity the
joint has remarkable deformation capacity. This phenomenon can be explained by the fact
that due to the bolt elastic/plastic elongation effectively reduces the plate deformations,
consequently, prevents the early development of the end-plate cracks, and avoid the early
failure.
60
moment [kNm]

40
20
0
-20
-40
-60
-60 -45 -30 -15 0 15 30 45 60
rotation [mrad]
a) M-Θ diagram b) Specimen after failure
Figure 5. Combined plate+bolt-failure.

Plate buckling-failure:
In case of slender welded composite member the plate buckling-failure of the flanges is
observed (the sections belong to the Class 4 in Eurocode 3). The phenomenon of the local
plate buckling is the following: the elastic buckling of the compressed flanges is appeared in
early cycles. After this the phenomenon became plastic plate buckling and later plastic lines
are developed and the phenomenon turned into a yield mechanism. The final collapse
caused by the cracking and fracture of the tension flanges due to low cycle fatigue. The
buckling behaviour is the symmetric buckling of the flanges since the typical asymmetric
buckling pattern of the steel sections cannot be developed due to the supporting effect of the
concrete. The final failure is the cracking and fracture of the tension flange and also the
cracking of the web. The representative moment-rotation diagrams of these behaviour
modes are plotted in figure 6 a). The moment carrying and rotational capacity of these joints
are favourable and besides this small degradation of the absorbed energy is observed. The
significant resistance decreasing can be seen after the cracking of the flange and falling out
of the crushed concrete (figure 6 b).
75
50
moment [kNm]

25

0
-25
-50
-75
-100 -75 -50 -25 0 25 50 75 100
rotation [mrad]

a) M-Θ diagram b) Specimen after failure


Figure 6. Local plate buckling.

194 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


STANDARD APPROACH

As it is discussed before only monotonic behaviour is concerned in accordance of the


Eurocode 3, 1.8 (6) (hereinafter EC3). The EC3 defines the moment-rotation relationship of
the joint as figure 7 a) shows with the design moment resistance (MjRd), the rotational
capacity (φcd) and initial rotational stiffness (Sjini). The standard procedure study monotonic
behaviour, but during the tests the specimens are imposed by cyclic loading. Since
monotonic test are not prepared the joints are described by the envelope curves that belong
to the cyclic moment-rotation diagrams (see figure 7 b).

50
S jini
25

moment [kNm]
M jRd
0

-25

-50
-60 -40 -20 0 20 40 60
xd cd rotation[mrad]

a) Standard M-Θ diagram b) Cyclic and envelope curves


Figure 7. Moment-rotation diagram.

Review of the EC3 design method

The EC3 gives the following formula to calculate the moment resistance:
M j,Rd = Σh r Ftr,Rd (1)
In our case we have only one connecting member, which connected to a rigid support, so the
half of the joint is studied. For this reason the tension resistance of the bolt-row (Ftr,Rd) is
defined as the resistance of end-plate in bending (Ft,ep,Rd), which formulae depends on the
failure mode. The failure modes are clearly defined from experimental tests as previous
sections discuss (Plate-failure – Mode1; Combined plate+bolt-failure – Mode2; Bolt-failure –
Mode3). The Eurocode 4, 1.1 (7) gives the moment resistance of the composite section, as it
is illustrated in figure 8.

tg fck fs fy
tf

Fc
xc
dt

Fs Fy

M pl,Rd
h1
h

F'y
zc

zy
zs

F's
z'y
z's

bp

Figure 8. Moment resistance of composite section.

The rotational stiffness of the joint is calculated from the following formula:
2
hr
Sj = E ⋅ (2)
1
µ⋅Σ
ki
Two stiffness coefficients of EC3 are taken into account: end-plate in bending (k5) and bolts
in tension (k10).

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 195


Comparison of design and experimental results

The moment resistance and initial stiffness are calculated for each specimen for both positive
(Mj,Rd,exp1) and negative (Mj,Rd,exp2) hemicycles. The following notifications are drawn for the
calculation method:
- The yielding stress (fy) of the steel material is from material tests and it is increased by
10% to consider the hardening of the material.
- The experimentally determined rotational stiffness (Sj,ini,exp1 and Sj,ini,exp2) is evaluated from
the unloading part of the hysteretic curve, since initial part of the loading path is unsteady
caused by the uncertainties of the specimen and the test equipment.

The experimental and numerical moment resistances are shown in figure 9. In case of the
local buckling type failure the standard values reach 82-105% of the experimental moment
resistances, which are in good coincidence (see figure 9 a). When the failure is occurred in
the connecting elements (Mode1-Mode3) significant differences between the experimental
and the standard values (56-87%) are found, as figure 9 b) shows.

100 100

80 x 80
moment [kNm]

moment [kNm]

60 60

40 40

20 20
0 1 2 3 4 5 6 7 8 9 10 10 12 14 16 18 20 22 24 26 28 30
flange/web thickness [mm] end-plate thickness [mm]
exp.results EC results exp.results EC results mod EC results

a) CCF Specimens b) CCB Specimens


Figure 9. Moment resistances.

On the bases of the experimental results the modification of the design method is proposed
to apply the standard EC3 design procedure for cyclic behaviour: During cyclic loading the
bolts have residual plastic deformation/elongation. Due to this effect the lever arms are
modified as follows (see figure 10).

F F F
hr1

hr2

hr3

a) Mode1 b) Mode2 c) Mode3


Figure 10. Consideration on the lever arms.

Mode1: 'pure' plate-failure. No significant bolt elongation, but the bolt heads have
deformation; hr1 is the distance between two bolt rows (see figure 10 a).
Mode2: combined failure. Plastic bolt elongation and end-plate deformation occurred; hr2
lever arm is measured between the tensioned bolt row and the half distance of the
compressed bolt row and edge of the end-plate (see figure 10 b).
Mode3: 'pure' bolt-failure. The bolts have plastic elongation without significant deformation of
the end-plate; hr3 is the distance between the tensioned bolt row and the lower edge
of the end-plate.

196 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


As the initial stiffness is concerned the design and experimental results are presented in
figure 11 a) and b) in case of tests CCF and CCB, respectively. As the diagrams show the
EC3 procedure overestimates the initial stiffness of the joint. In case of the presented
experimental program, the overestimation is in a wide range (109-188%). This result could
be caused by the fact that the experimental results are from cyclic test, note, however that in
(8) similar results are found in case of monotonic tests, too.

70 70
60 60

S jini [kNm/mrad]
S jini [kNm/mrad]

50 50
40 40
30 30
20 20
10 10
10 20 30 40 10 20 30 40
end-plate thickness [mm] end-plate thickness [mm]
exp. results EC results mod. EC results exp. results EC results mod. EC results

a) CCF Specimens b) CCB Specimens


Figure 11. Rotational stiffness.

In case of the presented experimental program, the initial stiffness Sjini depends on two
stiffness parameters (k5,k10). The modification of these parameters is as the follows:
t 4p
k 5 = 0,9 ⋅ Σl eff,1 ⋅ (3)
m 4x
As
k 10 = 1,1 ⋅ (4)
Lb
The above modification of the stiffness coefficients gives better agreement of design and
experimental values in case of specimen CCB. But in case of specimens CCF with the
formulae (3) and (4) the stiffness is also overestimated. In this case the application of larger
M24 bolts increases extremely the initial stiffness also in case of the thin end-plates, where
the k5 coefficient should have influence. For this reason the reduction of the multiplication
factor of the k10 coefficient in accordance with the bolt diameter gives better coincidence with
the experimental and design values.

In figure 12 the envelope moment-rotation diagrams are shown with the results of the EC3
calculation and the developed modified design method. In case of CCF the initial stiffness is
overestimated as shown in figure 12 a). Good agreement is found in case of CCB1 as figure
12 b) shows.

100 100
75 75
50 50
moment [kNm]

EC3 results
moment [kNm]

25 25 EC3 results
mod. EC3 results
0 0 mod. EC3 results

-25 Envelop curve -25 Envelop curve


-50 -50
-75 -75
-100 -100
-75 -50 -25 0 25 50 75 -25 -20 -15 -10 -5 0 5 10 15 20 25
rotation [mrad] rotation [mrad]

a) CCF1 b) CCB1
Figure 12. Experimental and design values.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 197


PARAMETRIC STUDIES

After the EC3 procedure is studied and modified, the evaluation of the monotonic joint
parameters (moment resistance and initial stiffness) is extended to apply them to
characterize cyclic joint behaviour. The original and modified standard procedures are used
to prepare parametric study with the following program:

In figure 13 the moment resistance is presented via the end-plate thickness and shows the
results of the standard EC3 procedure with the proposed modification. Figure 13 indicates
the region of the certain failure modes (Mode1, Mode2 and Mode3). Due to the modification
of the EC3 procedure the limits of the failure modes have a small alteration. The failure
modes of the modified EC3 procedure are always in accordance of the tests results.

100
Mode1 Mode2 Mode3
80
moment [kNm]

60
40
20
0
10 12 14 16 18 20 22 24 26 28 30
end-plate thickness [mm]
EC3 results mod. EC3results

Figure 13. Moment resistance by EC3 and modified EC3 method.

The effect of the bolt grade and the steel material is shown in figure 14 a) and b),
respectively. Figure 14 a) shows the application of the higher bolt grade modifies the limits of
the failure modes by moving the curve right and up. This means that the failure occurred in
thicker end-plate on higher moment level. The application of altering bolt grades does not
change the shape and the tangent of the diagram. If we increase the grade of the steel
material the diagram is moved to left, the failure occurred in thinner plate, but approximately
on the same moment level as figure 14 b) shows.

100 60
80 50
moment [kNm]
moment [kNm]

40
60
30
40
20
20 10
0 0
10 12 14 16 18 20 22 24 26 28 30 10 12 14 16 18 20 22 24 26 28 30
end-plate thickness [mm] end-plate thickness [mm]
M16 8.8 M16 10.9 M16 12.9 S235 S275 S355

a) Bolt grade effect b) Steel material effect


Figure 14. Moment resistance vs. end-plate thickness.

In case of the composite section the studied structural details are the flange and the web
thicknesses as figure 15 shows. The application of higher grade of steel material causes
the increase of the tangent of the curve, which means that it has significant effect mainly in
case of larger flange thickness, as figure 15 a) shows. In case of the web the increasing of
the steel grade move the curve up without significant changing of the tangent, so it has
remarkable effect also in case of smaller web thickness (see figure 15 b).

198 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


300 150
250 125
moment [kNm]

moment [kNm]
200 100
150 75
100 50
50 25
0 0
0 2 4 6 8 10 12 14 16 18 20 0 2 4 6 8 10 12 14 16 18 20
flange thickness [mm] web thickness [mm]
S235 S275 S355 S235 S275 S355

a) Effect of flange thickness b) Effect of web thickness


Figure 15. Moment resistance vs. flange/web thickness.

The parametric study includes the investigation of the initial stiffness. Figure 16 a) shows the
effect of the modified k5 and k10 parameters. Coefficient k5 has the influence on the initial
stiffness in case of relatively thin end-plate and k10 in case of thicker plate region. If formula
(3) is used to calculate k5 it decreases the initial stiffness in case of relatively thin plates, but
it has no significant effect on thicker end-plates. Formula (4) of k10 coefficient decreases the
curve initial tangent and move down the original curve. For these reasons the modification of
both coefficients is required to have better agreement between the experimental and design
values.

50 100 M30
x
Sjini [kNm/mrad]

40 M27
75
Sjini [kNm/mrad]

30 M24
M22
20 50
M20
10 M16
25
0 M12
M10
10 15 20 25 30 35 40 45 50 M8
0
end-plate thickness [mm] 10 15 20 25 30 35 40 45 50
EC3 results
mod. k10
mod. k5
mod. EC3 results
end-plate thickness [mm]

a) Effect of stiffness coefficients b) Effect of bolt diameter


Figure 16. Initial stiffness vs. end-plate thickness.

The effect of the bolt diameter is presented in figure 16 b). In case of the studied joint the
larger bolt diameters (over M16) have significant effect on the initial stiffness, since it
remarkable increases the stiffness. When the bolt diameter is relatively small the behaviour
in governed by the end-plate.

SUMMARY

The paper presents the comparison of the design and the tested characteristics of bolted
end-plate type joint. The performed test program and the behaviour modes are summarized.
The design methods of Eurocode 3 and 4 are followed to evaluate the moment resistance
and initial stiffness of the end-plate type joint and the composite section. The EC3 gives
method to evaluate monotonic joint behaviour. Since the performed tests are cyclic tests the
envelope curves of the hysteretic diagrams are assumed to cover the monotonic behaviour.

From the comparison it is found that in case of the studied joint the EC3 procedure does not
give good agreement between the design and experimental results. The moment resistance

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 199


is under- and the initial stiffness is overestimated. For this reason on the bases of the design
and experimental results a modification is proposed to extend the EC3 procedure and to
apply it to evaluate joint parameters of the cyclic behaviour in the next phase of the research.
The method is verified by the experimental results. A parametric study is performed by the
EC3 and the modified EC3 methods to characterize the effect of structural details on the
design moment resistance and initial stiffness. The conclusions are drawn on the influence of
the end-plate thickness, the flange/web thickness and bolt grade.

ACKNOWLEDGEMENT

The research work was conducted under the financial support of the Portuguese – Hungarian
Intergovernmental Science and Technology Cooperation Program ICCTI/OM TÉT P-4/99,
TÉT P-11/01 and the National Scientific Research Foundation OTKA –F 037869.

NOTATION

As : the tensile stress area of the bolt or of the anchor bolt,


E : modulus of elasticity,
Ftr,Rd : effective tension resistance of bolt-row r,
hr : the distance from bolt-row r to the centre of compression,
ki : stiffness coefficient,
Σl eff,1 : effective length,
mx : location of the bolts
tp : end-plate thickness,
µ : the stiffness ratio.

REFERENCES

(1) Kovács N, Calado L. and Dunai L., (2004). Behaviour of bolted composite joints;
Experimental study, Journal of Constructional Steel Research, Volume 60, Issues 3- 5
March-May 2004. (Special issue: Eurosteel 2002 Third European Conference on Steel
Structures), pp 725-738.
(2) Ádány S., (2001). Numerical and experimental analysis of bolted end-plate joint under
monotonic and cyclic loading, Ph.D. Dissertation, Budapest University of Technology
and Economics, Department of Structural Engineering.
(3) Kovács N., Ádány S., Calado L. and Dunai L., (2001). Experimental Program on Bolted
End-plate Joints of Composite Members, Report ICIST, DTC No. 15/01, Lisbon,
Portugal.
(4) Kovács N, Ádány S, Calado L and Dunai L., (2002). Experimental program on the
bolted end-plate type joints of slender section columns, Report ICIST No. 10/02,
Lisbon, Portugal.
(5) ECCS (1986). Recommended Testing Procedure for Assessing the Behaviour of
Structural Steel Elements under Cyclic Loads, TWG 1.3, No. 45.
(6) prEN 1993-1-8:2003, Eurocode 3: Design of steel structures, Part 1.8: Design of joints
(7) prEN 1994-1-1:1992, Eurocode 4: Design of composite steel and concrete structures,
Part 1.1: General rules and rules for buildings
(8) Ciutina A. L. and Dubina D., (2003). Influence of column web stiffeners on the seismic
behaviour of beam-to-column joints, Proc. of Conference on Behaviour of Steel
Structures in Seismic Areas, STESSA, pp 269-275. Naples, Italy.

200 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


ROTATIONAL CAPACITY AND DEMAND IN TOP-AND-SEAT ANGLE
CONNECTIONS SUBJECTED TO SEISMIC LOADING

Roberto T. Leon*, Jong Wan Hu**, and Corey Schrauben****

* Professor, Georgia Institute of Technology Atlanta, Georgia, USA


** PhD candidate, Georgia Institute of Technology Atlanta, Georgia, USA
**** Project Engineer, W. P. Moore and Associates, Atlanta, Georgia, USA

ABSTRACT

This paper explores strength, stiffness and rotational capacity rotational in thick
top-and-seat (cleated) angle connections subjected to monotonic and cyclic
loads. The results of test on two full-scale connections are described first, and
are then compared to published curve-fitting models for these types of
connections. The data indicate that the curve-fitting constants of some existing
mathematical models cannot be extrapolated to thick angles. The
experimental results also show that these connections are capable of providing
very ductile behavior and constitute an ideal back-up structural system in steel
frames.

INTRODUCTION

Two major earthquakes in 1994 and 1995, the Northridge and Kobe earthquakes, highlighted
the vulnerability of welded connections in steel moment frames and drew attention to the role
that partial strength, partial restraint connections played in the post-fracture behavior of those
structures. As a result, several research projects were carried out to investigate the
behavior of weak to moderately stiff bolted connections as part of the SAC Program (1). As
part of SAC (2), the behavior of bolted T-stub and top-and-seat angle was studied both at the
component and full-scale connection level. This paper discusses the results of the two full-
scale tests (3,4); the companion component tests have been described elsewhere (5). The
results of these tests, along with other modern tests by Azizinamini et al. (6), are then
compared with the predictions from several curve-fitting models available in the literature
(7,8). The latter include the Fry and Morris (1975) model based on odd-power polynomial
formulas (9), the Richard and Abbott (1975) four-parameter model (10), the Chen and Lui
(1985) exponential model (11), and the Kishi and Chen (1986) linear model (12,13). None of
these models limits the rotational capacity, as they are based on tests with flexible angles in
which a full plastic mechanism develops in the angles. This paper investigates the
extension of these models to situations where thick angles may lead to tension bolt failures.

EXPERIMENTAL PROGRAM

Two full-scale top-and-seat angle connections were tested as part of the SAC project (4).
As there was extensive data available on the performance of connections with thin angles (t
< 16 mm) and shallow beams (d < 360 mm) subjected to cyclic loads (3), the emphasis was
on testing connections with thicker angles (t = 25 mm) and deeper beams (d = 460 mm). The
specimens consisted of FE 350 (Grade 50) W460x60 (W18x40) beams bolted to W360x216

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 201


(W14x145) columns. The connections to the flanges were made up of L203x152x25 (L8x6x1)
angles 203 mm (8 in.) wide with yield and ultimate strengths measured at 360 and 529 MPa
(52.3 and 76.7 ksi), respectively. The connections were made with four M10.8 22 mm (A490
7/8 in.) bolts transferring the shear to the beam flanges and two similar bolts transferring the
tension to the column flange. The difference between the two connections tested was in the
gage distance in the column bolts, i.e., in the 152 mm (6 in.) leg. For specimen FS-01, the
distance was 63 mm (2.5 in.), while it was 101 mm (4 in.) for specimen FS-02. These
intended to represent two extremes of behavior with respect to prying forces in the tension
bolts. Both specimens had a 228x80x8 mm (9 by 3-1/8 by 5/16) shear tab welded to the
column and bolted to the beam with five M10.8 22 mm (A490, 7/8 in.) bolts

4000
400

2000
200

Moment (kN-m)
Moment (kip-in)

0 0

-200
-2000

-400
-4000
-0.04 -0.02 0.00 0.02 0.04

Total Rotation (rad)


Figure 1. Moment-rotation curve for FS-01.

4000
400
3000

2000
200
Moment (kN-m)
Moment (kip-in)

1000

0 0

-1000

-200
-2000

-3000
-400
-4000
-0.06 -0.04 -0.02 0.00 0.02 0.04 0.06

Total Rotation (rad)


Figure 2. Moment-rotation curve for FS-02.

202 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Specimen FS-01 (Figure 1) failed by tension bolt fracture at slightly less than 0.04 radian of
total rotation. The slip plateau is evident in Figure 1, with some deterioration of the frictional
capacity with cycling. For FS-01, the location of the tension bolt line was less than one inch
from the k-line of the angles, allowing only minimal prying action as the angle was unable to
deform. The test produced localized yielding in the beam flange only, with no evidence of the
formation of plastic hinges in the angles. It should be noted that the load history used is very
severe, corresponding to the demands in a high seismic zone.

Specimen FS-02 (Figure 2) produced a tension bolt failure as in FS-01. However, higher
displacements were reached due to the movement of the tension bolts away from the k-line
of the angles. Much more visible prying action was observed and less damage to the beam
was incurred. While plastic hinges did form above the k-line of the angle leg to the beam,
very little yielding occurred near the tension bolt line in the column. The angles in these tests
were extremely thick relative to the size of the beam and the tension bolts were clearly not
strong enough to allow complete formation of the plastic hinges in either the beam or angles.

CURVE-FITTING MODELS

In the USA, when top-and-seat connection capacities are required, it is common to resort to
published equations for moment-rotation curves. The latter are derived from curve-fitting to
test data. A good curve-fitting model requires a simple formulation that represents, to some
degree, the mechanics of the problem, but uses a minimum number of parameters to achieve
good results. The models proposed by Ang and Morris (1974), Frye and Morris (1975),
Richard and Abbott (1975), Chen and Lui (1985), and Chen and Kishi(1986) cover a wide
range of connection types (9-13). The data used, however, are typically from tests on small
specimens and some data relate to very old tests in which the degree of pretension and
actual material properties is uncertain. Figure 3 shows the standardized parameters
typically used for the mathematical models described in Table 1.

f (Fastener Dia.) = w lt = l
tt = t Flange Angle

Top angle
ga
kt gt
Web Angle

la d

ks p gs
Seat angle

ts (thickness of seat angle) ta (thickness of web angle) ls

Figure 3. Definition of geometric parameters.

COMPARISONS TO EXPERIMENTAL DATA

Table 2 gives details of eight modern specimens tested under monotonic and cyclic loads. Six
specimens come from Azizinamini (6), with companion specimens at two thickness (10 mm
and 13 mm) tested under both static and cyclic loads (14S1/C1 and 12S2/C2); the other two
specimens represent different web detail connections (14S3 and 14S4). The remaining two

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 203


tests used in the comparisons are the thick-angle tests described earlier in this paper (4). It
should be noted that the Azizinamini tests were run with Fe 250 (A36) steel as opposed to Fe
350 (A572 Grade 50) as in the tests reported herein.

Table 1. Curve-fitting models for connection behavior (7-13).


Power Model
Model Polynomial Model
Three Parameter
Developer Fry and Morris (1975) Ang and Morris (1984)
Curved Fitting C1=8.46 x 10-4, C2=1.01 x 10-4 φo=5.17 x 10-3 , (KM)o=745.94
Parameter C3=1.24 x 10-4 n=4.32
Geometric Parameter d, t, f, l d, t, f, l
-1.5 -0.5 -1.1 -0.7 -1.06 -0.54 -1.28 -0.85
Standardized Constant K=d t f l K=d t f l

Standardized Ramberg-Osgood
General Form of the
φc = C1(KM) +C2(KM) +C3(KM)
1 3 5
function
model
[φc/φo] = [KM/(KM)o]×[1+(KM/(KM)o)n-1]
Calibration Tests Rathbun(1936), Mains(1944) Hetchman and Johnston(1947)
Power Model
Model
Three Parameter Four Parameter
Developer Chen and Kishi (1987) Richard and Abbott (1975)
Rki, Mu,
Parameters n=0.827((φo<1.9x10-3rad) kp, k, Mo, n=2
n =1.398log10φo+4.93(φo>1.9x10-3rad)
lt, tt, kt, gt, w, d, ls, ts, ks, gs, da, ta, ka, lt, tt, kt, gt, w, d, ls, ts, ks, gs, da, ta,
Geometric Parameter
ga, la (Including web angle) ka, ga, la (Including web angle)
General Form of the M=[(k-kp) φc] / [(1+|(k-kp) φc / Mo|n )n-1]
φc = (M/Rki)⋅(1/(1-(M/Mu)1/n))
model + kpφc
Calibration Tests Azizinamini et al.(1985)

Table 2. Specimen details (3,4).

Specimen Type of Beam Top and Bottom Flange Angle Web Angle

No. Test Selection tt lt (ls) g p ta Lc

*14S1 Static W14x38 3/8’’ 8’’ 2.5’’ 5.5’’ 1/4’’ 8.5’’


*14S2 Static W14x38 1/2’’ 8’’ 2.5’’ 5.5’’ 1/4’’ 8.5’’
*14S3 Static W14x38 3/8’’ 8’’ 2.5’’ 5.5’’ 1/4’’ 5.5’’
*14S4 Static W14x38 3/8’’ 8’’ 2.5’’ 5.5’’ 3/8’’ 8.5’’
*14C1 Cyclic W14x38 3/8’’ 8’’ 2.5’’ 5.5’’ 1/4’’ 8.5’’
*14C2 Cyclic W14x38 1/2’’ 8’’ 2.5’’ 5.5’’ 1/4’’ 8.5’’
**FS-01 Cyclic W18x40 1’’ 8’’ 2.5’’ 5.5’’ 5/16’’ 9’’
**FS-02 Cyclic W18x40 1’’ 8’’ 2.5’’ 5.5’’ 5/16’’ 9’’

204 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


800 1200
1100
700
1000
600 900
800

Moment(K-in)
.

500
Moment(K-in)

700
400 600
500
300 14S1 14S2
400 F- M M o.
F- M M o.
200 300 A- M M o.
A- M M o.
200 C- K M o.
C- K M o.
100 100 R- A M o.
R- A M o.
0 0
0 0.01 0.02 0.03 0.04 0 0.01 0.02 0.03 0.04
Rot at i on( r ad) Rot at i on( r ad)

800 1000

700 900
800
600
700
.
.

Moment(K-in)
Moment(K-in)

500 600
400 500
14S3 400 14S4
300 F- M M o.
F- M M o.
300
200 A- M M o. A- M M o.
200 C- K M o.
C- K M o.
100 100 R- A M o.
R- A M o.
0 0
0 0.01 0.02 0.03 0.04 0 0.01 0.02 0.03 0.04
Rot at i on( r ad) Rot at i on( r ad)

Figure 4. Comparison for monotonic tests.

COMPARISON OF MOMENT ROTATION CURVES

The four specimens subjected to static load shown in Table 2 will be first compared using four
models (Figure 4). For the Ang and Morris model, regardless of the dimension of each
specimen, an identical reference moment and rotation angle are used. This model is
capable of providing a good, if somewhat larger, estimate of initial stiffness in spite of
neglecting the effect of the double web angles. On the other hand, the Fry and Morris model
consistently underestimates both strength and stiffness of the connections. The power
models have parameters based on the specimen’s dimensions and show good agreement
with the experimental moment-rotation curves even in the nonlinear range. When n, the
shape parameter, is increased, the behavior of the connection becomes closer to an
idealized elasto-plastic model. The connection has no further moment resistance when the
moment approaches ultimate moment (Mu). The Richard and Abbott model needs four
parameters: the initial stiffness (k), the strain hardening stiffness (kp), a reference moment
(Mo) and a parameter n defining the shape of the curve. These values are shown in Table 3
based on the results of the Azizinamini tests.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 205


All cyclic tests show an increased strength with cycling due to cyclic hardening, and show
clearly the results of the Bauschinger effect. Comparisons for the four cyclic tests shown in
Table 2 are given in Figure 5.

Table 3. Summary of static test results (N-mm/rad, N-mm).


Initial Stiffness of secant line Moment at Slope of M - φc Moment at
Specimen -3 -3 -3
stiffness to 4.0 x 10 rad. 4.0 x 10 rad. curves at 24x10 rad. 24x10-3 rad.
14S1 215 x 108 123 x 108 491 x 105 655 x 106 755 x 105
14S2 333 x 108 172 x 108 686 x 105 142 x 107 107 x 106
14S3 131 x 108 100 x 108 401 x 105 813 x 106 737 x 105
14S4 251 x 108 140 x 108 560 x 105 938 x 106 929 x 105

14C1 14C2

600 800

600
400
400
.

.
Moment(K-in)

Moment(K-in)

200
200

0 0
-0.015 -0.01 -0.005 0 0.005 0.01 0.015 -0.015 -0.01 -0.005 0 0.005 0.01 0.015
-200
-200
F- M M o. F- M M o.
-400
-400 A- M M o. A- M M o.
C- K M o. -600
C- K M o.
-600 -800
Rot at i on( r ad) Rot at i on( r ad)

FS01 CA01 FS01 CA02


4000 4000

3000 3000

2000 2000
.
.

Moment(K-in)
Moment(K-in)

1000 1000

0 0
-0.04 -0.03 -0.02 -0.01 0 0.01 0.02 0.03 0.04 -0.04 -0.02 0 0.02 0.04 0.06
-1000 -1000

-2000 F- M M o. F- M M o.
-2000
A- M M o. A- M M o.
-3000 -3000
C- K M o. C- K M o.
-4000 -4000
Rot at i on( r ad) Rot at i on( r ad)

Figure 5. Comparison for cyclic tests.

The use of models developed for monotonic loading to predict cyclic behavior is based on the
observation that the envelopes of cyclic tests match closely the envelope for static tests; i.e.,
the behavior of specimens 14S1 and 14S2 match closely the envelope from 14C1 and 14C2
respectively. For the thick clip angle specimens (FA-01 and FA-02), the Chen and Kishi

206 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


model substantially overpredicts both the stiffness and the strain hardening behavior.
Surprisingly, the Ang and Morris model prediction agrees rather well with the test envelope.
On the other hand, the Fry and Morris model underpredicts the strength and stiffness by an
appreciable margin.

Tables 4 through 7 show a comparison between the experimental results and the
mathematical models in terms of both the stiffness and moment. Four parameters are used
for comparisons in these tables. First, *K-secan is the elastic stiffness of a secant line to 4.0 x
10-3 rad resulting from a polynomial model or a power model. Similarly, **M-secant is moment
at 4.0 x 10-3 rad. After reaching the yield moment, the slope of the moment rotation curve
remains constant. A slope parameter (***Kp) was measured at 24x10-3 rad and used in the
comparisons of the monotonic tests. To compare with experimental curves after yield, the
static moment at 24x10-3 rad was used. For the cyclic tests with thick angles, the moment at
20x10-3 rad was used as this is still in the hardening range of the envelope. The Richard-
Abbott model requires parameters obtained from experimental results, and thus its results
can also be included in the comparisons.

Table 4. Comparison of experimental model and Fry-Morris model (N-mm/rad, N-mm).


Comparison *K-secant %Diff **M-secant %Diff ***Kp ****Mp
%Diff %Diff
Specimen (F-M Mo.) (F-M Mo.) (F-M Mo.) (F-M Mo.)

14S1 746 x107 39.3% 298 x105 39.3% 111 x107 -68.9% 657 x105 13%
7 5 7 5
14S2 862 x10 49.7% 345 x10 49.7% 121 x10 15% 759 x10 28.8%
7 5 7 5
14S3 746 x10 25.6% 298 x10 25.6% 111 x10 -36.1% 657 x10 10.9%
7 5 7 5
14S4 746 x10 46.7% 298 x10 46.7% 111 x10 15.3% 657 x10 29.3%
7 5
14C1 746 x10 49.2% 298 x10 49.2% - - - -
7 5
14C2 862 x10 38.5% 345 x10 38.5% - - - -
8 5 6
FS01CA01 207 x10 61.5% 827 x10 61.5% - - 170 x10 58.3%
8 5 6
FS01CA02 207 x10 54.8% 827 x10 54.7% - - 170 x10 50%
Ave(%) 45.65% 45.65% -18.7% 31.72%
Var(%) 10.42% 10.42% 35.75% 17.49%

Table 5. Comparison of experimental model and Ang-Morris model (N-mm/rad, N-mm).


Comparison K-secant M-secant Kp Mp
%Diff %Diff %Diff %Diff
Specimen (A-M Mo.) (A-M Mo.) (A-M Mo.) (A-M Mo.)

14S1 162 x108 -31.7% 647 x105 -31.7% 120 x107 -82% 115 x106 -53%
8 5 7 6
14S2 190 x10 -10.4% 757 x10 -10.4% 119 x10 15.8% 135 x10 -26.2%
8 5 7 6
14S3 162 x10 -61.4% 647 x10 -61.4% 120 x10 -47.2% 115 x10 -56.7%
8 5 7 6
14S4 162 x10 -15.5% 647 x10 -15.5% 120 x10 27.7% 115 x10 -24.3%
8 5
14C1 162 x10 -10.2% 647 x10 -10.2%
8 5
14C2 190 x10 -9.8% 757 x10 -9.8%
8 6
CA01 432x10 24.3% 173 x10 24.3% 307 x106 24.4%
8 6 6
CA01 432x10 6% 173 x10 6% 307 x10 9.3%
Ave(%) -13.6% -13.6% 21.42% -21.08%
Var(%) 23.8% 23.79% 60.71% 29.76%

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 207


Table 6. Comparison of experimental model and Chen-Kishi model (N-mm/rad, N-mm).
Comparison K-secant M-secant Kp Mp
%Diff %Diff %Diff %Diff
Specimen (C-K Mo.) (C-K Mo.) (C-K Mo.) (C-K Mo.)

14S1 111 x108 9.2% 443 x105 9.2% 231 x106 64.7% 759 x105 -0.005%
8 5 6 6
14S2 194 x10 -13.3% 775 x10 -13.3% 446 x10 68.6% 118 x10 -10%
8 5 6 5
14S3 162 x10 -62.5% 399 x10 -62.5% 257 x10 68.3% 655 x10 11%
7 5 6 5
14S4 997 x10 28.8% 574 x10 28.8% 378 x10 59.7% 958 x10 -3.2%
8 5
14C1 111 x10 24.6% 443 x10 24.6%
8 5
14C2 194 x10 11.3% 775 x10 11.3%
9 6
CA01 124 x10 -132% 497 x10 -132% 629 x106 -54.5%
9 5 6
CA01 127 x10 -176% 506 x10 -176% 643 x10 -89.6%
Ave(%) 38.73% 38.73% 65.5% -24.3%
Var(%) 72.58% 72.58% 3.6% 33.2%

Table 7. Comparison of experimental model and Richard-Abbott model (N-mm/rad, N-mm).


Comparison Specimen M-secant (R-A Mo.) %Diff Mp (R-A Mo.) %Diff
5 5
14S1 515 x10 -6.8% 750 x10 0.6%
5 6
14S2 689 x10 -0.5% 107 x10 0.3%
14S3 395 x105 1.43% 728 x105 1.2%
5 5
14S4 606 x10 -8.06% 923 x10 0.6%
Ave(%) -3.428% 0.675%
Var(%) 4.03% 0.82%

CONCLUSIONS

Except for the Richard – Abbott model, which has to show good agreement by definition
since its constants were based on the data used for the comparisons, large discrepancies
between the experimental results and the models are obvious. In general, the results for
thick clip angles evidence a larger percent difference than those for thin angles. As the table
and figure indicate, the Chen-Kishi model based on Azizinamini’s experimental data shows
good agreement for the thin to medium clip angle cases. However, it does not provide
satisfactory results for the behavior of the thick clip angle connections. Most important, none
of the curve-fitting models is capable of predicting the rotational capacity when the tension
capacity of the bolts governs the behavior. Ongoing work will compare these conclusions with
results from both Eurocode and another polynomial models developed by Citipitioglu et al.
(14).

REFERENCES

1 Liu, J., and Astaneh-Asl, A. (2000). Cyclic Tests of Simple Connections Including the
Effects of Slabs, J. of Structural Engineering, ASCE, v. 126, n. 1, pp. 32-39.
2 Swanson, J.A., and Leon, R.T. (2002). Bolted Steel Connections: Tests on T-stub
Subcomponents, J. of Structural Engineering, ASCE, v. 126, n. 1, pp. 50-56.
3 Leon, R.T., Swanson, J.A., Schrauben, C., and Smallidge, J. (2000). Experimental Test

208 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


and Analytical Models for Bolted Connections, Final Report MAE Center Project ST-7,
Georgia institute of Technology, Atlanta GA.
4 Schrauben, C.S. (1999). Behavior of Full-Scale Bolted Beam to Column T-stub and Clip
Angle Connections Under Cyclic Loading, MS Thesis, Georgia Institute of Technology
5 Swanson, J.A., and Cao, X. (2000). Strength Determination of Heavy Clip-Angle
Connection Components, in Connections in Steel Structures 4 (R. Leon and S. Easterling,
eds.), AISC, pp. 234-243.
6 Azizinamini, J.A., Altman, W.G., Radziminski, J.B. et al. (1982). Moment-Rotation
Characteristics of Semi-Rigid Steel Beam-Column Connections, Earthquake Hazard
Mitigation Program, University of South Carolina, Columbia.
7 Chan, S.L.,and Chui, P.P.P. (2000). Nonlinear Static and Cyclic Analysis of Steel Frames
with Semi-Rigid Connections, Elsevier, Oxford, England.
8 Chen, W.F., and Kim, S.E. (2000). LRFD Steel Design using Advanced Analysis, CRC
Press, Boca Raton.
9 Frye, M.J., and Morris, G.A. (1975). Analysis of Flexibility Connected Steel Frames, Can.
J. Civil Eng. v. 2, n. 3, pp. 280-291
10 Ang, K.M., and Morris, G.A. (1984). Analysis of Three-Dimensional Frames with Flexible
Beam-Column Connections, Can. J. Civil Eng. v.11, pp.245-254
11 Richard, R.M., and Abbott, B.J. (1975). Versatile Elastic Plastic Stress-Strain Formula, J.
of Engineering Mechanics, ASCE, v.101, n. 4, pp. 511-515
12 Chen, W.F., and Kishi, N. (1989). Semi-rigid Steel Beam-to-Column Connections:
Database and Modeling, J. of Structure, ASCE, v.115, n.1, pp.105-119
13 Chen, W.F., and Kishi, N., Matsuoka, K.G. (1993) Design Aid of Semi-Rigid Connections
for Frame Analysis, J. of Engineering., AISC, 4th quarter, pp.90-107
14 Citipitioglu, A.M., Haj-Ali, R.M., and White, D.D. (2002). Refined 3D Finite Element
Modeling of Partially Restrained Connections Including Slip, Journal of Constructional
Steel Research, v. 58, nos. 5-8, pp. 995-1014

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 209


210 Connections in Steel Structures V - Amsterdam - June 3-4, 2004
SEISMIC PERFORMANCE OF DEEP COLUMN-TO-BEAM WELDED
REDUCED BEAM SECTION MOMENT CONNECTIONS

J.M. Ricles, Lehigh University, U.S.A.


X. Zhang, Lehigh University, U.S.A.
J.W. Fisher, Lehigh University, U.S.A
L.W. Lu, Lehigh University, U.S.A

ABSTRACT
An experimental study was conducted to investigate the seismic behavior of
reduced beam section (RBS) moment connections to a deep wide flange
column. The test matrix for the experimental program consisted of six full-scale
interior RBS connections, where the column for the specimens ranged in depth
from a W24 to a W36 wide flange section. All but one of the specimens had a
composite floor slab. The results from the study show that a composite floor
slab provides restraint to the top flange of the beams; reducing the magnitude
of beam top and bottom flange lateral movement in the RBS, column twist, and
strength degradation due to beam instability in the RBS. The performance of
each of the test specimens was found to meet the seismic connection
qualification criteria in Appendix S of the AISC Seismic Provisions, and thereby
have sufficient ductility for seismic resistant design. The results of the
experimental study, along with a nonlinear finite element study were used to
develop seismic design recommendations for RBS connections to deep
columns.

INTRODUCTION

RBS beam-to-column moment connections are often utilized in the design of special steel
moment resistant frames (SMRFs). The details of a typical RBS connection are shown in
Figure 1(a), where the flanges of the beam are reduced in width, away from the column face.
Complete joint penetration (CJP) groove welds attach the beam flanges to the column. The
beam web often is welded to the column flange with a CJP groove weld. By design, the RBS
connection develops inelastic deformations primarily in the region where the beam flange
width has been reduced (referred to herein as the RBS), limiting the inelastic strain developed
in the beam flange-to-column CJP groove welds. With the reduction of the beam flange width,
an RBS connection is more prone to inelastic local buckling of the beam web and flanges in
the RBS. For economical reasons, design engineers in the U.S. prefer to use deep columns
in SMRFs (as large as 914 mm in depth corresponding to a W36 wide flange section) in order
to control seismic drift. Previous tests on RBS connections have been performed primarily on
columns with depths corresponding to a W12 and W14 wide flange section (Roeder (1)),
where the depth was about 305 mm to 356 mm. Some tests using W27 wide flange column
sections (686 mm depth) were conducted by Chi and Uang (2), where the connection was an
exterior connection (i.e., only one beam was connected to the column). It was observed in
these tests, that as a result of inelastic beam web and flange local buckling in the RBS, a
lateral displacement of the beam compression flanges occurs. Shown in Figure 1(b) is the
movement of the compression flanges (the top and bottom flanges of the right and left-hand
beams, respectively), where F1 and F2 represent the beam flange compression forces of the

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 211


two beams. Due to an eccentricity created by the lateral movement of the compression
flanges, a torque is applied to the column. Deep columns tend to have thinner flanges and a
web than a shallower column, resulting in a reduced torsional resistance. Consequently, there
have been concerns that the use of an
(a) RBS connection to a deep column in a
SMRF can lead to inferior seismic
Erection bolt
performance because of the connection
being susceptibility to torsional loading
from the beams.

The lack of knowledge of the performance


of RBS beam-to-deep column
R
connections under seismic loading led to
Doubler Plate a study on this topic at Lehigh University
(3). The study involved both finite element
analysis and experimental tests. The
effects of the column depth, a composite
floor slab, panel zone strength, beam web
slenderness, and supplemental lateral
bracing at the end of the RBS section
(b) were examined. Six full-scale specimens
FF1
1
were subsequently tested involving
e1
different column and beam sizes, a
e2
composite floor slab and supplemental
F
F2 2
lateral bracing. Results and conclusions
from the experimental study, along with
Figure 1. (a) RBS connection details, and some design recommendations are
(b) RBS local buckling and lateral beam presented in this paper.
flange movement.

TEST MATRIX

The test matrix for the experimental program is given below in Table 1, where some of the
details of the six full-scale RBS beam-to-deep column connection specimens are
summarized. All specimens represented an interior RBS connection in a perimeter SMRF
with a composite floor slab, with the exception of SPEC-6 which did not have a composite
floor slab. The parameters investigated in the experimental program included: (1) column
size; (2) beam size; (3) the floor slab; and (4) supplemental lateral brace at the end of the
RBS.

The beam and column section sizes for each specimen were selected on the basis of
introducing different degrees of torsional effects, predicted by the recommended design
procedure of Chi and Uang (2), while also satisfying the weak beam-strong column criteria in
the ASIC Seismic Provisions (4). The design procedure by Chi and Uang considers the total
normal stress in the column at 4% story drift due to axial load, flexure load, and torsion. The
predicted total normal stress in the column flange is shown plotted in Figure 2 for various
column sections, including those of the test specimens. Figure 2 indicates that SPEC-2,
SPEC-4, and SPEC-5 are predicted to develop column flange yielding. The columns for all
specimens and the beams for SPEC-3 through SPEC-6 were fabricated from A992 steel. The
beams for SPEC-1 and SPEC-2 were fabricated from A572 Gr. 50 steel. Both A992 and A572
Gr. 50 have a nominal yield strength of 345 MPa.

212 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Table 1. Test matrix.
Supp.
Yield Stress Tensile Stress
Column Beam Doubler Floor Lat.
SPEC Flange/Web Flange/Web
size Size Plate Slab Brace @
(MPa) (MPa)
RBS
Beam Col Beam Col
6x800x
1 W36x230 W36x150 Yes No 343/378 356/393 478/492 496/514
1067
13x610x
2 W27x194 W36x150 Yes No 343/378 372/392 478/492 520/502
1067
13x610x
3 W27x194 W36x150 Yes Yes 365/396 356/403 508/506 497/521
1067
10x8160x
4 W36x150 W36x150 Yes No 365/396 365/396 508/506 508/506
1067
10x610x
5 W27x146 W30x108 Yes No 344/353 363/399 471/469 499/513
914
13x533x
6 W24x131 W30x108 No Yes 344/353 334/359 471/469 499/493
914
Shear Plate
Continuity plate
700 Doubler plate
SPEC-4 E71T-8
SPEC-5 and 6 had E70T-1 E70T-6
600 W36x150 (TYP)
E70T-1 CJP(TC-U4a-GF)
W30x108 beams; all
SPEC-5 SPEC-2 others had W36x150
Total Stress (MPa)

500 W27x194 beams


W27x146 erection bolts
10"
400 SPEC-1 E71T-8 RBS
W36x230 Shear tab
CJP Flange Cut

6"
No Run- Beam
300 6" 6" off Tabs

6"
SPEC-6 W24x131 E70T-1
(supplemental bracing) E71T-8
200

100 SPEC-3 W27x194


27" 9"
(supplemental bracing) E70T-6
E70T-1 R CJP(TC-U4a-GF)
0
E71T-8
0 200 400 600 800 1000 1200 1400 Column
E70T-1
Column Section Weight (kg/m)

Figure 2. Column total stress per Chi and Uang Figure 3. Specimen typical connection
(2) versus column section weight. details (Note: 1 inch = 25.4 mm).

CONNECTION AND COMPOSITE FLOOR SLAB DETAILS

The elevation of a typical connection detail is shown in Figure 3. Each specimen was
designed in accordance with the criteria recommended by Engelhardt (5) for RBS
connections, where the design moment in the beam at the column face is limited to Mpn,
where Mpn is the nominal plastic capacity of the beam. For the six specimens in the test
matrix, the average value of the beam design moment at the column face was equal to
0.973Mpn. The reduction in flange width at the center of the RBS for each specimen was 50%
of the original flange width, which complied with the design criteria of Engelhardt. The RBS
was flame cut, with the burned surface ground to a surface roughness of 500 micro-inches,
as recommended by FEMA 353 (6). Each specimen had continuity plates the same thickness
as the beam flanges and designed for a balanced panel zone condition. Complete details are
given in Ricles et al. (3). The weld procedure specifications used in the fabrication of the
connections were prequalified in accordance with AWS D1.1/D1.1M:2002 (7). All welds were
done using the flux core arc welding procedure, and conformed to the AWS 5.20-95
Specification (8). The beam flange-to-column flange CJP groove field welds and beam web-
to-column CJP groove field welds utilized E70T-6 and E71T-8 electrodes, respectively. All
shop welds (e.g., shear tab to the column, doubler and continuity plates) were performed
using E70T-1. The run off tabs for the beam flanges were removed following the placement of

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 213


the CJP groove welds, and the weld at the edges of the beam flanges ground to a smooth
transition. The backing bar of the top flange weld was left in place and a reinforcement fillet
weld was provided between the bottom surface of the backing bar and the column flange
using the E71T-8 electrode. The beam bottom flange backing bar was removed using the air-
arc process, back gouged, and reinforced with a fillet weld using an E71T-8 electrode. No run
off tabs were used for the vertical beam web CJP groove welds. All CJP groove welds were
inspected using the ultrasonic test procedure in order to evaluate whether they complied with
the criteria in AWS D1.1 (7) for weld quality.

The specimen composite floor slab had a total thickness of 133 mm, and consisted of 27.6
MPa nominal compressive strength concrete cast on a 20-gage zinc coated metal deck. A
W4x4 welded wire mesh with wire 152 mm on center was placed in the floor slab prior to
pouring the concrete. The width of the floor slab was 1220 mm to one side, with a 305 mm
overhang on the other side to simulate the conditions for a perimeter SMRF. The ribs of the
decking ran parallel to the main beam (i.e., the beams with the RBS connections) of each test
specimen. To develop the composite action, 19 mm diameter shear studs were placed
outside the RBS region at 305 mm spacing along the beams to attach the deck to the main
beams as well as transverse W14x22 floor beams. These transverse beams were placed at a
spacing of 3048 mm to provide lateral bracing to the main beams and column, where the
distance of 3048 mm satisfied the AISC Seismic Provisions (4).

SPEC-6, which had no composite floor slab, had a supplemental lateral brace at the end of
the RBS in addition to the other lateral bracing noted above for the beams. The lateral
bracing was attached to a W36x150 section that was placed parallel to the beams of the test
specimen to simulate a parallel beam in the prototype building. This parallel beam in the test
setup was allowed to move horizontally with the test specimen, but restrained from out-of-
plane movement. The corresponding stiffness of the lateral bracing setup satisfied the AISC
LRFD Specification (9). SPEC-3 also had supplemental lateral braces, but these were
anchored in the floor slab.

TEST SETUP, LOADING PROTOCOL, INSTRUMENTATION

The test setup is shown in Figure 4 (a), with the lateral bracing detail given in Figure 4(b) for
the main beams. The ends of the members in the test setup had pin-connected boundary
conditions, using cylindrical bearings to simulate inflection points at the beam midspan and
column midheight in the prototype frame. The ends of each beam away from the column were
supported by instrumented rigid links, which simulated a roller boundary condition and
enabled horizontal movement of the end of each beam. The lateral bracing detail shown in
Figure 4(b) was used to prevent out-of-plane movement of the beams and column (the
diagonal double angles were not used at the column), and designed for strength and stiffness
in accordance with the AISC LRFD Specification (9). The top of the column was braced
against torsion, while at the base of the column a clevis was used to create the pin boundary
condition. The beams were also braced at the rigid links in order to stabilize the test setup.
The torsional bracing provided at both ends of the column in the test setup was evaluated
using a nonlinear finite element model (3) to examine whether the stiffness would be
representative of the torsional restraint at the column inflection points in the prototype
structure. It was found to be satisfactory and not influence the test results by over-restraining
the ends of the column from twisting.

The specimens were tested by imposing a cyclic story drift history based on the loading
sequence defined in Appendix S of the AISC Seismic Provisions (4). The loading protocol
consisted of initial elastic cycles of story drift, followed by cycles of increased amplitude to
cause inelastic response. A test was terminated when either a fracture occurred, resulting in

214 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


a significant loss of specimen capacity, or after reaching a story drift of 6%. Each specimen
was instrumented to enable measurement of the applied loads, reactions at the rigid links,
specimen story drift; strains in the beam, column, panel zone, and continuity plates; in
addition to panel zone deformation, plastic beam rotation, twisting of the column, and lateral
displacement of the beam at the center of the RBS.
SYM
(a) CL
10' 10' (b)

21"
Load Cell
Actuator
Column Shear Stud
Floor Beam 12" 48" Floor Slab
(North Side Only) Floor Beam(North Side Only)
6'-6"

1312" Floor Slab No Diagonal Bracing 131 "

51 4"
2

A325
3
4"
diam.
Beam (East) Double W14x22
Beam (West) (TYP)
6'-6"

Beam Web Stiffener with W36x150 Angle


Load Cell Diagonal Brace to Floor Load Cell 2 L2x2x5 16
1312"

Beam (North Side Only)

14'-9" 14'-9"
29'-6"
Setup Lateral Bracing

Figure 4. (a) Test setup and (b) beam lateral bracing detail for specimens with a
composite floor slab (Note: 1 inch = 25.4 mm).

TEST RESULTS

A summary of test results for each specimen is given in Table 2, where Rv/Vpz, θmax, Mf/Mpn,
Kφ,col, φ, δflg, and bf are equal to the ratio of panel zone shear capacity-to-panel zone shear
force corresponding to the plastic flexural moment developing in the RBS, specimen drift from
the last cycle prior to any fracture or strength deterioration to below 80% of the specimen
nominal capacity, ratio of maximum measured beam moment developed at the column face-
to-nominal beam flexural capacity, column elastic torsional stiffness, specimen column twist
at 4% story drift, lateral displacement of the beam bottom flange at the RBS at 4% story drift,
and beam flange width, respectively. Typical observed behavior during the testing of a
specimen consisted of yielding in the RBS and the panel zone, followed by cyclic local web
and flange buckling in the RBS. Following the development of local bucking in the RBS,
lateral movement of the bottom beam flange began to occur in the RBS of specimens with a
composite floor slab at 2% to 3% story drift. The combined effect of cyclic local buckling and
lateral flange displacement resulted in a gradual deterioration in specimen capacity to occur
during subsequent cycles where the story drift amplitude was increased. This is evident in the
lateral load-story drift hysteretic response of SPEC-4 shown in Figure 5. The lateral
displacement of the bottom beam flange occurred when it was in compression, and caused
some column twist to develop. Figure 7(a) and (b) shows photographs of SPEC-4 at 4% and
6% story drift, where the yielding in the members and panel zone in the connection region
and lateral beam flange movement in the RBS are visible. The maximum column twist among
the specimens with a floor slab at 4% story drift was 0.037 rads. (SPEC-4). 4% story drift is
the drift at which connections are judged for qualification for seismic use by the AISC Seismic
Provisions (4). SPEC-4, like the other specimens, developed a flange fracture in the RBS
where extensive local flange buckling had occurred (see Figure 7(c)). This occurred at a story
drift of 6%, and was caused by local buckling in the beam flange that led to large cyclic
strains, resulting in a low cycle fatigue failure. SPEC-6, which had a supplemental brace and
lateral bracing attached to the beam that is parallel to the test beam, had minimal
deterioration in capacity as well as column twist (0.004 rads. at 4% story drift), see Figure 6.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 215


2000 2000

1500 1500

1000 1000

Lateral Load (kN)


Lateral Load (kN)

500 500

0 0

-500 -500

-1000 -1000

-1500 -1500

-2000 -2000
-6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6
Story Drift (% rad.) Story Drift (% rad.)

Figure 5. Lateral load-story drift Figure 6. Lateral load-story drift


hysteretic response of SPEC-4. hysteretic response of SPEC-6.

(a) Yielding and


local buckling in
connection
region, 4% drift

fracture

(b) Beam bottom flange lateral movement (c) Beam bottom flange fracture
at RBS, 4% story drift at RBS, 6% story drift
Figure 7. Photographs of SPEC-4 during testing.

The reduced amount of deterioration in the capacity of SPEC-6 was due to the specimen
having a weaker panel zone than the other specimens. As noted in Table 2, for SPEC-6 the
ratio of Rv/Vpz is equal to 1.03. Rv is based on the ASIC Seismic Provisions (4). All other
specimens have a value of 1.14 or greater for the ratio of Rv/Vpz. Consequently, these other
specimens developed a larger amount of yielding and local buckling in the RBS than SPEC-6,
leading to local buckling and deterioration in specimen capacity.

In Table 2, all specimens are shown to have a value for θmax that exceeds 0.04 rads., which is
the current criteria in Appendix S of the AISC Seismic Provisions (4) for qualifying a
connection for seismic use. A summary of the ratio of Mf/Mpn in Table 2 indicates that the
maximum beam moment developed at the column face in the specimens exceeded the
design value of Mpn for which the specimens were designed, with SPEC-5 having the
maximum value of 1.2. The increase in the moment Mf is attributed to the composite floor
slab increasing the moment capacity in the RBS. SPEC-6 had a valve of Mf equal to 1.0, and
had no composite floor slab.

216 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Table 2. Test results.

Rv θmax Mf Kφ,col φ(1) δflg (1) 0.2bf


SPEC
Vpz (% rad) M pn (kN-m/rad) (rad) (mm) (mm)

1 1.26 4.0 1.03 3190 0.016 53 61


2 1.14 4.0 1.13 1404 0.025 34 61
3 1.28 5.0 1.15 1404 0.006 35 -(2)
4 1.24 4.0 1.11 947 0.037 38 61
5 1.21 5.0 1.20 900 0.007 26 53
6 1.03 4.0 1.00 577 0.004 5 -(2)
Note: (1) Corresponding to 4% story drift
(2) Chi and Uang criteria (2) for transverse beam flange movement does not apply to
cases with supplemental braces.

6000
SPEC-4
SPEC-6
4000
Strain (microstrain)

2000
εy=1765 µε
0

-2000

-4000

-6000
-150 -100 -50 0 50 100 150
Distance across column flange (mm)

Figure 8. Longitudinal strain profile across column flange, just below


RBS connection; 4% story drift.

An examination of the results for column twist φ in Table 2 reveals that column twist tends to
increase when the elastic torsional stiffness of the column Kφ,col is reduced. However, for a
smaller beam section size φ is reduced, although the column torsional stiffness is smaller
(e.g., SPEC-5). This phenomenon is associated with a smaller demand on the column when
a smaller beam is used. The column twist is reduced significantly in specimens with a
supplemental brace (SPEC-3 and SPEC-6). The reduction in column twisting in SPEC-6 is
also attributed to a weaker panel zone, which reduced the amount of yielding and local
buckling in the RBS, and subsequently less lateral movement in the RBS. An examination of
the measured specimen beam flange lateral displacement δflg in Table 2 shows these results
to be less than the value of 0.2bf, which is the value recommended by Chi and Uang (2) for
determining the design torque T applied to the column. Consequently, the use of the value of
0.2bf for determining the design torsional loading on the column from the RBS will result in a
larger column design torque. This is evident by comparing the column total normal stress at
the connection based on Chi and Uang’s recommendation with the measured specimen
response (see Table 3). The criterion by Chi and Uang anticipates column flange yielding
occurring in SPEC-2, 4, and 5; see Figure 2, where the nominal yield stress is 345 MPa. The
measured column flange longitudinal strains in these specimens indicated no yielding in
SPEC-2 and 5, with some minor yielding occurring in SPEC-4 (a maximum strain of 2 to 4
times the yield strain developed). The measured longitudinal strains across the column flange
just below the connection are shown in Figure 8 for SPEC-4 and SPEC-6. These results are

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 217


representative of typical specimen behavior, and show little evidence of a strain gradient
across the flange that would result from the effects of warping normal stresses due to column
torsion.

DESIGN RECOMMENDATIONS

The strains in the beam bottom flange near the column face were examined to evaluate the
stress distribution across the beam flange that leads to a torque T applied to the column.
Shown below in Figure 9(a) is the distribution of longitudinal stress across the beam bottom
flange at 4% story drift. These stresses are based on measured longitudinal strains in the
specimens. These results correspond to a negative beam moment at the column face (i.e.,
when the bottom flange of the beam is in “compression”). Similar results for longitudinal
stress across the beam compression flange were obtained from finite element studies (see
Figure 9(b)). The results in Figure 9 show a trend where the stress distribution across the
beam flange has a reduction in stress, which is due to a moment in the plane of the beam
flange caused by the lateral movement of the beam flange at the RBS. This moment is
equivalent to the torque T that is applied by the beam flange to the column. Shown in Figure
10(a) is an idealized uniform longitudinal stress distribution prior to lateral movement of the
beam flange in the RBS (at 2% story drift). The idealized longitudinal stress distribution at 4%
story drift based on the measured and finite element analysis results is given in Figure 10(b).
At 4% drift local buckling and lateral beam flange movement has occurred in the RBS.
Elastic-perfectly stress-strain behavior is assumed in Figure 10, where Fye is the yield stress.
400 100
SPEC-1
300
SPEC-2
SPEC-3 0
(b)
SPEC-4
Longitudinal stress (MPa)

200 SPEC-5
-100
Stress (MPa)

100

0 -200

-100
-300
-200
(a)
-400
-300

-400 -500
-150 -100 -50 0 50 100 150 -150 -100 -50 0 50 100 150
Distance across beam flange (mm) Distance across beam flange (mm)

Figure 9. Longitudinal stress distribution across beam flange for (a) all test
specimens, and (b) finite element analysis of SPEC-2.

Fye Fye

(a) (b)

Figure 10. Idealized longitudinal stress distribution across beam bottom flange at
(a) 2% story drift and (b) 4% story drift.

218 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


For the longitudinal stress distribution shown in Figure 10(b), T can be shown (3) to be equal
to
11
T= Fyeb 2f t f (1)
150

where Fye, bf, and tf are equal to the expected beam flange yield stress (1.1Fy), the beam
flange width, and beam flange thickness, respectively.

A design procedure was thus developed in order to determine the total design longitudinal
stress ftotal in the column flange that is attached to an RBS connection. The procedure
involves determining the elastic warping normal stresses fw that develop in the column flange
due to the torque T (10) and superimposing them with the column flange normal stresses
due to bending (fb) and axial loading (fa) to obtain the total normal stress ftotal, where

f w = EWnOθ " (2)

In Equation (2) E, WnO, and θ” are equal to the Young’s modulus, normalized warping
function at the column flange tip (10), and the second derivative of the angle of twist in the
column (10), respectively, where θ” is a function of the torque T.

The total stress ftotal is compared to the criteria in the AISC LRFD Specification (9), Equation
(H2-1), where

f total = φFy (3)

in which φ and Fy are the resistance factor (0.9) and nominal yield stress of the column
flange, respectively. The above design procedure is similar to that developed by Chi and
Uang (2), except for the method in which the torque is determined.

Table 3. Comparison of column normal flange compression stresses with design procedure.
Experimental
Warping
Axial Total normal results, total
stress fw
load Bending stress ftotal (MPa) stress & strain,
(MPa)
SPEC Column Beam stress stress 4% story drift
fa fb (MPa) Chi Chi
Pro- Pro- Strain Stress
(MPa) and and
posed posed (µε) (MPa)
Uang Uang
1 W36x230 0 190 128 66 318 256 1277 255
2 W27x194 0 299 182 101 481 400 2151 372(1)
W36x150
3 W27x194 0 332 0 0 332 332 1797 356(1)
4 W36x150 0 337 321 163 658 500 3296 365(1)
5 W27x146 0 252 180 95 432 347 1598 319
W30x108
6 W24x131 0 347 0 0 347 347 2525 334(1)
Note: (1) Yield stress of the column flange.

The total normal column flange stress based on the above procedure is compared in Table 3
to the measured stress of the test specimens, as well as the stress predicted using the
procedure by Chi and Uang (2). The comparisons in Table 3 indicate that a more accurate
prediction of the total normal stress in the column flange is made using the above procedure
compared to the procedure developed by Chi and Uang (2). The difference between the two

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 219


methods is the normal warping stress fw predicted by the above procedure is based on a
more accurate value of the torque T applied to the column. For specimens with a
supplemental lateral brace it was assumed that the restraint of the supplemental brace
resulted in no torque applied to the column (i.e., the normal warping stress fw is equal to
zero). This results in a lower predicted stress than the measured response.

SUMMARY AND CONCLUSIONS

An experimental program was conducted in order to evaluate the seismic performance of


RBS connections to deep wide flange columns. The study involved testing six full-scale
specimens to evaluate the effects of column depth, beam size, composite floor slab, and a
supplemental lateral brace.

Based on the experimental study, the following main conclusions are noted:

1. A composite floor slab can significantly reduce the lateral displacement of the beam
bottom flange in the RBS and the amount of twist developed in the column. The slab
appears to be effective in reducing the twist in deeper columns attached to an RBS
connection, and enables the cyclic strength of the beam with an RBS connection to
be better sustained.
2. All of the specimens were able to satisfy the criteria in the AISC Seismic Provisions
(4) for qualifying the connection for seismic use.
3. A weaker panel zone in a deep column RBS connection will not develop as much
column twist and strength degradation as a connection with a stronger panel zone.
However, a weaker panel zone can significantly increase the potential for ductile
fracture of the connection (3). It is recommended that connections be designed with a
balanced panel zone strength condition.
4. A supplemental brace at the end of the RBS significantly reduced the transverse
movement of the beam flanges in the RBS and column twist that leads to cyclic
degradation in specimen capacity.
5. Basing the column torque on a transverse movement of the beam flange in the RBS
of 0.2bf for calculating column flange warping stresses appears to be conservative. A
new procedure for estimating the torsional load applied to the column due to the local
and lateral buckling in the RBS shows improvement in predicting the correct column
flange normal stress.

ACKNOWLEDGEMENTS

The research reported herein was supported by a grant from the American Institute of Steel
Construction (Mr. Tom Schlafly program manager) and from the Pennsylvania Department of
Community and Economic Development through the Pennsylvania Infrastructure Technology
Alliance (PITA) program. The following companies donated materials for the experimental
testing conducted in this research project: Arcelor International America of New York, NY
(steel sections); Nucor Vulcraft Group of Chemung, NY (metal decking); and the Lincoln
Electric Company of Cleveland, OH (welding wire). The support provided by the funding
agencies and companies is greatly appreciated.

220 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


REFERENCES

(1) Roeder, C. W. (2000). “Connection Performance State of Art Report,” Report No.
FEMA-355D, FEMA, Washington, D.C.
(2) Chi, B. and Uang, C.-M. (2002). “Cyclic Response and Design Recommendations of
Reduced Beam Section Moment Connections with Deep Columns,” Journal of
Structural Engineering, ASCE, 128(4): 464-473.
(3) Ricles, J., Zhang, X., Lu, L.W., and J. Fisher, (2004). “Development of Seismic
Guidelines for Deep-Column Steel Moment Connections,” ATLSS Report No. 04-13,
ATLSS Engineering Research Center, Lehigh University, Bethlehem, PA.
(4) “Seismic Provisions for Structural Steel Buildings,” (2002). American Institute of Steel
Construction, Chicago, Illinois.
(5) Engelhardt, M. D. (1999). “The 1999 T. R. Higgins Lecture: Design of Reduced Beam
Section Moment Connections,” Proceedings: 1999 North American Steel Construction
Conference, American Institute of Steel Construction, Toronto, Canada, pp. 1-1 to 1-29.
(6) “Recommended Specifications and Quality Assurance Guidelines for Steel Moment-
Frame Construction for Seismic Applications,” (2000). Report No. FEMA 353, Federal
Emergency Management Agency, Washington D. C.
(7) “Structural Welding Code – Steel,” (2002). AWS D1.1/D1.1M:2002, American Welding
Society, Miami, Florida.
(8) “Specification for Carbon Steel Electrodes for Flux Cored Arc Welding,” (1995).
ANSI/AWS A5.20-95, American Welding Society, Miami, Florida.
(9) “Manual of Steel Construction-Load and Resistance Factor Design,” (2001). Third Ed.,
AISC, Chicago, Illinois.
(10) Seaburg, P., and C. Carter, (1997). “Torsional Analysis of Structural Steel Members,”
American Institute of Steel Construction Steel Design Guide Series, ASIC, Chicago,
Illinois.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 221


222 Connections in Steel Structures V - Amsterdam - June 3-4, 2004
BOLTED LINKS FOR ECCENTRICALLY BRACED STEEL FRAMES
A. Stratan, the Politehnica University of Timisoara, Romania
D. Dubina, the Politehnica University of Timisoara, Romania

ABSTRACT
Eccentrically braced steel frames represent a suitable solution for multi-storey
buildings located in seismic areas. A bolted connection between the link and
the beam is suggested to facilitate replacement of damaged dissipative zones
(links) after a moderate to strong earthquake, which reduces repair costs. A full
scale testing program was carried out in order to demonstrate the feasibility of
this concept and to evaluate the performance of bolted links. The paper
summarises the results of the testing program.

INTRODUCTION

Design of multi-storey structures in high-seismicity areas is usually based on dissipative


structural response, which accepts significant structural damage under the design
earthquake. It is believed however, that design criteria specified in modern seismic codes will
prevent structural collapse, ensuring life safe. The earthquakes of Loma Prieta (1989),
Northridge (1994) and Hyogoken-Nanbu (1995) showed that generally, modern structures
behaved as expected. However, the unexpectedly high economic losses following these
earthquakes urged for a limitation of damage to structures in future earthquakes, leading to
the development of Performance-Based Design (PBD), Hamburger, 1996 (1). Its objectives
include minimizing structural and non-structural damage under low and moderate earthquake
intensities, which is equivalent to reduction of the total cost (initial and repair).

ed
e

Figure 1. Bolted link concept.

On the other hand, capacity-based design, applied in most of the current seismic design
codes allows design of structures that promote plastic deformation in predefined areas only,
called dissipative zones. In the case of a bolted connection between dissipative zones and
the rest of the structure it is possible to replace the dissipative elements damaged as a result
of a moderate to strong earthquake, reducing the repair costs. Application of this philosophy
to eccentrically braced frames, where link elements serve as dissipative zones, is presented
in figure 1. The connection of the link to the beams is realized by a flush end-plate and high-
strength bolts. Bolted connection allows he link element to be fabricated from a lower-yield
steel grade, assuring an elastic response of the elements outside removable link element.
This system may be applied to both homogeneous structures (eccentrically braced frames
alone) and dual ones (eccentrically braced spans combined with moment-resisting spans).
The latter system has the advantage of more uniform transient and smaller permanent lateral
displacements, which is beneficial for replacing damaged links, as well as for the building
function, Stratan and Dubina, 2002 (2). Extended end-plate bolted connections for

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 223


eccentrically braced frames with link-column connection configuration were previously
suggested and investigated experimentally by Ghobarah and Ramadan, 1994 (3). Their
inelastic performance was found to be similar to fully-welded connections.

EXPERIMENTAL PROGRAM

Specimens and experimental set-up

An experimental program was carried out to determine cyclic performance of bolted links and
to check the feasibility of the suggested solution. The removable link was fabricated from
IPE240 profile of S235 grade steel, while the rest of the structure – from S355 grade steel.
Four link lengths were considered (e=400, 500, 600 and 700 mm, see figure 1), to study the
influence of moment to shear force ratio. All links are classified as short ones according to
AISC, 1997 (4). Another parameter considered was the spacing of web stiffeners, provided to
prevent web buckling and to improve rotation capacity of the link. Two limit values of stiffener
spacing were considered to AISC, 1997 (4): "close" spacing - 30tw-h/5, for a rotation capacity
0.08 rad, and "rare" spacing - 52tw-h/5, for 0.02 rad rotation capacity.

support

actuator
link

Figure 2. Experimental set-up for removable bolted links.

For combination of link length and stiffener spacing, three specimens were tested: one
monotonically and two cyclically. A total of 24 specimens were thus obtained, each being
denoted as L[x][n]-[t], where: [x] – L for "rare" spacing of stiffeners, H for "close" spacing of
stiffeners; [n] – 7, 6, 5, 4 for link length; [t] – m for monotonic, c1 for cyclic 1 and c2 for cyclic
2 specimens. Thus, LL7-m specimen is one with rare spacing of stiffeners (L) of length 700
mm (7), monotonically loaded (m). The complete ECCS 1985 (5) loading procedure was
applied, consisting of one monotonic and two cyclic tests. The monotonic test was used to
determine the yield force Fy and displacement Dy, at the intersection of the initial stiffness
and the tangent to the F-D curve having 10% of the initial stiffness. Yield displacement was
determined for each monotonic test with rare stiffeners, and used to apply cyclic loading to
the specimens of the same length. The cyclic tests consisted of four cycles in the elastic
range (±0.25Dy, ±0.5Dy, ±0.75Dy and ±1.0Dy), followed by groups of three cycles at
amplitudes multiple of 2Dy (3x±2Dy, 3x±4Dy, 3x±6Dy, etc.) The loading was applied quasi-
statically, in displacement control. Bolts were preloaded to 100% of the full preload value for
friction-grip bolts in the case of the monotonically loaded (m) and the first of the cyclically
loaded (c1) specimens, and 50% for the second cyclically loaded specimen (c2).

Previous experimental research on beam-column joints with end plates, Dubina et al., 2000
(6) showed a series of problems that undermined their cyclic performance: (1) fillet welds are
inappropriate in the case of cyclic loading; (2) full-penetration 1/2V weld with the root at the
exterior part of the beam cross-section promotes fragile ruptures, due to cracks initiated at
weld root; (3) weld-access hole acts as a stress concentrator, causing brittle ruptures of the

224 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


beam flange. Welding details used for the link to end plate connection were chosen so as to
prevent the causes of poor performance mentioned above. Thus, link flange was welded to
the end plate with a full-penetration 1/2V weld, realised from the exterior part of the cross-
section (weld root at the interior); the weld access hole was eliminated completely, and
reinforcing fillet weld was applied at the interior part of the flanges and on the web.

Table 1. Characteristics of the materials used for fabrication of removable link specimens.
component fy (Reh), N/mm2 fu, N/mm2 fu/fy A, %
IPE240 flange 268.0 401.9 1.50 29.2
IPE240 web 337.8 426.7 1.26 30.8
t=25 250.8 413.1 1.65 36.3

Standard tensile tests were performed on coupons extracted the materials used to fabricate
the link specimens. Results presented in table 1 revealed a higher yield strength of the web
in comparison with flanges of the link.

Design of connections

Bolted connection between the link element and the beam is located in a zone of maximum
stresses. There are two possible strategies for connection design. The first one is to provide
a sufficient overstrength of the connection over the link shear resistance. The second one is
to assure a ductile behaviour of the bolted connection itself. The former strategy was
followed in this case, as it facilitates replacement of damaged link elements.

Capacity design of the connection involves two steps: determination of the yield strength of
the dissipative element (link plastic shear resistance), and of the overstrength to allow for
strain hardening. Two design provisions available at the date of the experimental program
set-up were considered: Eurocode 8, 1994 (7) and AISC, 1997 (4). Though plastic shear
resistance is determined using similar formulations in the two codes, the European seismic
design provisions, referring to Eurocode 3 (8), consider the contribution of the fillet radius to
the shear area, resulting in a capacity 40% higher than the one of the American code, which
considers only the web area. The overstrength required for elements outside links also differ
substantially. Previous experimental research, Kasai and Popov, 1986 (9), indicated link
ultimate shear resistance about 1.5 times the plastic shear resistance. Eurocode 8 requires
an overstrength of only 1.2, while AISC 1997 results in overstrength factors between 1.38
and 1.88. Reduced overstrength factor in European codes is counterbalanced by higher
plastic shear resistance, the maximum shear force estimated to the two codes having similar
values. A relatively conservative estimation of maximum shear force was adopted in this
study (1.75 factor, applied to the web area, corresponding to a 1.25 factor applied to the
Eurocode 3 shear area):
Vmax = 1.75 ⋅ Vy = ( h − 2 ⋅ tf ) ⋅ tw ⋅ fy / 3 (1)

Maximum moment for connection design was determined as:


Mmax = Vmax ⋅ ed / 2 (2)

Design of connections to the forces determined to equations (1) and (2) was based on
Eurocode 3, 1997 and its Annex J (8). M20 gr.10.9 high-strength bolts were used. End plate
thickness (25 mm) was chosen so as to provide a mode 3 (bolts in tension) failure mode of
the equivalent T-stub, preventing excessive deformation of the end plate. A linear distribution
of bolt forces was then assumed, and the bolts checked for tension, shear, combined tension
and shear resistance, assuming a partial safety factor γMb=1.25. Demand to capacity ratio for
combined tension and shear ranged from 0.7 for the LH4 and LL4 specimens to 0.98 for the
LH7 and LL7 specimens. Additionally, bolt slip resistance was checked.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 225


Data processing

The instrumentation consisted of the actuator load cell, and a series of displacement
transducers used to measure both absolute and relative displacements. The basic force-
displacement relationship used to characterize the monotonic and cyclic response of bolted
links was actuator force (equal to link shear force) and total displacement DT, which includes
slip in connection and endplate deformations. Response of link elements is characterised by
the shear distortion angle γ - shear force F relation. For classical links, the distortion γ is
determined as the difference of end displacements divided to the link length, Engelhardt and
Popov, 1992 (10). With the notations from figure 3, γ is expressed as:
γ = DT / b (3)

Assuming that the edges of the panel bounding the link remain straight after deformation, the
same angle γ may be determined from the deformations of the diagonals (DD1 and DD2):

a 2 + b 2 ⋅ ( DD2 − DD1)
γ = (4)
2⋅a⋅b

d+
d

DD
π D1
d+D

2
γ


d a 2
-γ1

a
π
DT b
b

(a) (b) (c)


Figure 3. Deformation of a classical link (a), idealisation of the panel zone (b) and its
deformation (c).

θ
DALS

γL
γΤ

θ γ
DALJ

γ
θ L
γ
DT

(a) (b)
Figure 4. Deformation of a bolted link (a) and its idealisation (b).

Values of angle γ determined according to equations (3) and (4) have close values in the
case of classical links. However, in the case of removable bolted links, the behaviour of the
link is more complex, and angle γ determined from equations (3) and (4) will be different.
Total link deformation is given by the sum of: (1) shear distortion of the link panel - γ, (2)
rotation in the two connections θM=θS+θj, and (3) slip in the connections, characterised by the
equivalent rotation γAL=(DALS+DALJ)/ed, and can be expressed as:
γ T = γ + θM + γ AL (5)

226 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


It can be directly obtained from the total displacement DT:
γ T = DT ed (6)

Instrumentation permitted both direct (6) determination of characteristic deformations, and


indirect one (5), using the component deformations. A satisfactory correlation was observed
between the two methods.

BEHAVIOUR OF SPECIMENS

Strength characteristics obtained from nominal and measured geometry and strength are
presented in Table 2. Account was taken of the different flange and web yield strength in
determining the link plastic moment: M y = Wpl ,w ⋅ fy ,w + Wpl * ⋅fy ,f .

Measured characteristics of steel showed higher increase of plastic shear force in


comparison with plastic moment, which caused a decrease of the 1.6My/Vy limit. Even so, the
links are classified as short. At the same time, maximum shear force and moment used for
connection design are considerably higher than the initial estimates based on nominal
characteristics. Connection strength was checked using estimates of maximum forces
determined from measured geometrical and mechanical characteristics, considering a partial
safety factor γMb=1.0 for the connection. Results indicated that the connection should have
responded in the elastic range, though with little reserve for the longer LL7 and LH7
specimens. However, at large displacements, both bolt failures and end-plate deformations
were observed during the tests. Two types of bolt failures were observed: (1) by thread
stripping, which results in a ductile response (dominant in this experimental program), and
(2) by fracture in bolt shank, which results in a brittle response.

Table 2. Yield and maximum forces evaluated from nominal and measured characteristics.
Wpl, Wplw, Wpl*, My, 1.6My/Vy, Vmax, Mmax,
specimen Vy, kN
cm3 cm3 cm3 kNm mm kN kNm
LH7, LL7 366.6 75.29 291.31 185.4 86.2 743 278.1 83.4
LH6, LL6 366.6 75.29 291.31 185.4 86.2 743 278.1 69.5
nominal
LH5, LL5 366.6 75.29 291.31 185.4 86.2 743 278.1 55.6
LH4, LL4 366.6 75.29 291.31 185.4 86.2 743 278.1 41.7
LH7, LL7 366.6 75.43 291.2 266.7 103.5 621 400.1 120.0
LH6, LL6 366.6 75.43 291.2 266.7 103.5 621 400.1 100.0
measured
LH5, LL5 366.6 75.43 291.2 266.7 103.5 621 400.1 80.0
LH4, LL4 366.6 75.43 291.2 266.7 103.5 621 400.1 60.0
Note: Mmax determined per equation (2)

400 400
LL7−c1 LH7−c1
200 200
F, kN

F, kN

0 0

−200 −200

−400 −400
−0.2 −0.1 0 0.1 0.2 −0.2 −0.1 0 0.1 0.2
γT, rad γT, rad

Figure 5. Force-total deformation relationship F-γT for specimens LL7-c1 and LH7-c1.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 227


Bolted connections had important contributions to the overall link response and in general did
not showed an elastic response. Connection suffered important degradations at the Lx7
specimens, and caused a pronounced pinching effect with a reduced energy dissipation
capacity (see figure 5). Element degradation started by bending of the end plate and bolt
thread stripping, followed by local buckling of link flanges and web. Closer stiffener spacing
had as main effect isolation of local flange and web buckling in outer web panels. Failure was
attained by complete degradation of bolt threads.

400 400
LL4−c1 LH4−c1
200 200
F, kN

F, kN
0 0
web
−200 breathing −200

−400 −400
−0.2 −0.1 0 0.1 0.2 −0.2 −0.1 0 0.1 0.2
γT, rad γT, rad

Figure 6. Force-total deformation relationship F-γT for specimens LL4-c1 and LH4-c1.

(a) (b) (c)


Figure 7. Failure by connection degradation at the LH6-c2 specimen (a); plastic web buckling
at the LL4-c1 specimen (b), and strengthening of the brace to beam welded connection (c).

Smaller length of Lx6 reduced the damage to connections and the pinching behaviour.
Failure was attained by complete damage to bolts (see figure 7a), but also by web cracking
after repeated plastic web buckling in the case of LL6-c2 specimen, with rare stiffeners.

Starting with Lx5 specimens, connections were characterised by a more stable response,
plastic web buckling being more important and preceding the one of the flanges. Failure of
LL5-c1 and LL5-c2 specimens, with rare stiffeners, was attained by tearing of the web on
three edges, at the cracks initiated in the base metal at the web-stiffener and web-end plate
welds. Closer stiffener spacing in the case of LH5-c1 and LH5-c2 specimens reduced web

228 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


tearing due to severe and repeated buckling (but did not eliminate it completely), failure
being attained by damage of the connection.

Response of specimens from the Lx4 series was dominated by web shear. Connection had a
quasi-elastic response. Flange buckling was observed only after important web buckling.
Hysteretic response was characterised by "full" cycles with high energy dissipation capacity
(see figure 6). Due to higher web slenderness of the LL4-c1 and LL4-c2 specimens, web
buckling was pronounced, and plastic web "breathing" was observed, as web buckling wave
was changing direction at reversals of load direction (figure 6). Repeated buckling lead to
web tearing along the diagonals (see figure 7b). Close spacing of stiffeners at the LH4-c1
and LH4-c2 specimens prevented this phenomenon, failure initiating through web tearing
along the stiffener weld, which extended on three edges of the web.

High stresses are present at the beam to brace welded connection, next to the beam to link
bolted connection. Higher grade steel of the elements outside removable link did not provide
sufficient overstrength in this zone. Due to repeated cyclic loading, the lower beam to brace
welded connection fractured during the test of LL7-c1 specimen. Removing the weld and
applying a new weld did not help, and the lower beam-brace assembly was completely
replaced to due to extensive damage in the zone between the lower connection and the
brace. To mitigate this problem after similar failure of the new subassembly during testing of
the LL6-c1 specimen, stiffeners were added in the affected zone to increase the shear area
and provide a smooth transfer of stresses from the bolted link element to the brace and the
beam (see figure 7c). The performance of the subassembly modified in this way was
satisfactory for the rest of tests.

COMPARATIVE ANALYSIS OF RESULTS

Elastic response of links was characterised by the total initial stiffness KγT, determined from
V-γT relationship, as well as shear stiffness of the web Kγ, stiffness of connections KθJ and
KθS, determined from M-θJ, and M-θS relationships. Initial shear stiffness of the link (Kγ) was in
good correlation with the theoretical one (Kγth=G⋅As), and not influenced much by the different
considered test parameters. There was an important scatter in experimental values of
connection rotational stiffness. Full preloading increased the stiffness of connection by
approximately 50%. Upper connection resulted more flexible in comparison with the lower
connection. Unsymmetrical distribution of moments and lack of fit at the upper connection
may be attributed to this behaviour. Reduction of total initial stiffness of the bolted link in
comparison with the classical solution is important, as a result of both the semi-rigid end-
plate, and slip in the connection. Therefore, either explicit modelling of the semi-rigid
connection behaviour, or consideration of an equivalent link stiffness is necessary for global
analysis of frames with bolted links.

Table 3. Yield Vy and maximum Vmax shear forces.


parameter specimen LL7 LL6 LL5 LL4 LH7 LH6 LH5 LH4
Vyth, kN 266.7*
m 228.0 209.0 189.5 191.0 201.6 217.8 198.2 201.5
Vy, kN c1 234.8 218.3 245.0 174.4 227.4 212.9 229.6 236.6
c2 216.5 216.9 175.4 223.7 211.1 222.2 231.5 249.0
Vmaxth, kN 400.1**
m 304.9 333.3 348.1 388.3 270.1 307.5 352.5 420.6
Vmax, kN c1 296.9 308.4 343.3 360.9 305.2 318.5 364.1 400.6
c2 289.6 313.9 355.7 362.5 301.6 324.4 364.0 402.9
Note: average of positive and negative values presented for specimens c1 and c2
* plastic shear resistance based on measured geometry and yield strength
** Vmaxth=1.5Vyth

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 229


Connection slip was defined when relative displacement between the end plates of a
connection exceeded 0.15 mm, according to C133/82 (11). The only specimen that did not
slip was the first tested LL7-m. Slip resistance of the connection was reduced by cyclic
loading and partial preload of bolts, rendering ineffective limitation of slip deformations.

Yield force determined from V-DT relationship was not influenced by the test parameters and
was controlled by shear response of the web. Lower experimental values (see table 3) are
partially explained by the procedure used to determine yield force, which underestimates it
for high initial stiffness. On the other hand, experimental maximum force presents an
increase from the longer to the shorter links (effect of connection strength) and is higher for
closer stiffeners (prevention of web plastic buckling).

The maximum moment determined from equation (2) was lower than the theoretical one
used to design the connections. Poor performance of connections could be explained by the
fact that vertical displacement in the experimental set-up was constrained, which generated
supplementary tension in the connections at large displacements. Further research is
needed to validate this assumption and to check its application to real structures. Following
the experimental observations in this study, in order to reduce damage in bolted connections,
it is recommended to limit the length of bolted links to ed ≤ 0.8 ⋅ My Vy , which corresponds to
links LL4 and LH4.

Table 4. Ultimate displacement DTu and corresponding deformationγTu.


specimen LL7 LL6 LL5 LL4 LH7 LH6 LH5 LH4
m 93.0 136.6 144.2 118.4 140.8 138.8 137.9 125.9
DTu,
c1 58.1 64.4 42.3 30.4 68.4 71.7 58.6 37.8
mm
c2 55.4 66.3 62.5 33.5 65.5 68.2 72.8 37.6
m 0.155 0.273 0.360 0.395 0.235 0.278 0.345 0.420
γTu c1 0.097 0.129 0.106 0.101 0.114 0.143 0.147 0.126
c2 0.092 0.133 0.156 0.112 0.109 0.136 0.182 0.125
Note: minimum of positive and negative values presented for c1 and c2 specimens

Ultimate link displacement DTu, representing the stable hysteretic response is presented in
table 4. Cyclic loading reduced by 40% to 70% rotation capacity, with the maximum reduction
for short links. A slight reduction of ultimate displacements was observed for short links. In
terms of deformations (γTu), rotation capacity increases slightly for shorter links, with the
exception of LL4 and LH4 specimens. With the exception of longer links with rare stiffeners
(LL7), specimens showed a stable deformation capacity of at least 0.1 rad. Ductilities larger
than 10 were observed, with a number of 16 to 22 cycles in the plastic range. Bolt preloading
did not affect rotation capacity, as oppose to closer spacing of stiffeners, which improved link
deformation capacity.

Behaviour of long specimens was much influenced by the response of the bolted connection,
characterised by a gradual reduction of strength due to bolt thread stripping, and a pinching
cyclic response. The latter effect reduced the energy dissipated in the group of cycles of
constant amplitude. Full bolt preloading reduced partially this effect. Response of short
specimens was controlled by the shear of the link web, characterised by important hardening
and energy dissipation capacity, but a more rapid degradation of strength after web tearing.
Stiffener spacing had maximum importance for short links. Their effect was to limit plastic
local buckling of the web, increasing the maximum force and deformation capacity, and
providing a more stable cyclic response. However, after the attainment of ultimate
deformation, failure of LH4 specimens was more rapid in comparison with LL4 specimens.

Distribution of ductility demands between end pate and link web resulted in improved overall
deformation capacity in comparison with "pure" failure modes, determined by concentration
of plastic deformations in connection or web alone. This effect is characteristic of

230 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


intermediate length specimens LL6-LL5 and LH6-LH5. However, it is difficult to achieve this
response in practice, due to variability of mechanical characteristics of structural steels.

CONCLUSIONS

Experimental investigations on removable bolted links demonstrated the technological


feasibility of the solution. Performance of short removable links and possibility to be easily
replaced makes them attractive for dual eccentrically braced frames. Very short links, that
assure an elastic behaviour of the connection are preferred, due to much easier replacement
of damaged links. Concentration of damage in the removable link (performing like passive
energy dissipation devices) may be accomplished by the capacity design principles, including
fabrication of the link from a steel with lower yield strength in comparison with rest of the
structure. The beam zone between the link end plate and brace is subject to high stresses,
therefore its reinforcement by stiffeners is recommended. Welding details between the link
and end plate showed a very good performance, which is attributed to: (1) elimination of weld
access hole; (2) full penetration weld in 1/2V between the link flange and end plate, realised
from the exterior of the profile; (3) a fillet weld on the interior contour (web and flanges) of the
cross-section. Lack of weld access hole has the advantage of reduced fabrication cost in
addition to higher connection performance.

Longer links and closer stiffener spacing imposed higher demands on the connection. Cyclic
response elements for which connection represented the weaker element were characterised
by: (1) a reduction of maximum force in comparison with elements dominate by web shear;
(2) a pinching behaviour with stiffness and strength degradation in cycles of constant
amplitude; (3) failure by gradual strength degradation due to bolt thread stripping. Post
elastic connection response was ductile, due to thread stripping. This failure mode is not
generally characteristic for bolts. Bolt failure by shank rupture would have caused a more
brittle response of long links.

Response of short links was governed by web shear, stiffener spacing being important for
their performance. In the case of rare spacing of stiffeners, inelastic response of short links
was determined by plastic web buckling, which lead to strength degradation by alternative
buckling in the direction of the two diagonals. Closer stiffener spacing limited plastic web
buckling, leading to: (1) attainment of the maximum possible shear strength; (2) a stable
hysteretic response; (3) a larger rotation capacity, but also (4) a more rapid failure by web
tearing on the panel edges.

With the exception of very short links, connections were partial-strength. On the basis of
present experimental program, in order to prevent excessive connection damage, it is
recommended to limit link length ed to 0.8⋅My/Vy. Design strength of short removable links
limited to this length may be computed as for classical short links. Full bolt preloading
resulted in higher initial stiffness, a more stable hysteretic response and a larger deformation
capacity, and therefore is recommended for removable short links. Semi-rigid connections
with flush end plate reduce substantially initial stiffness of removable short links in
comparison with classical solution. Global analysis of eccentrically braced frames with
removable links requires either explicit modelling semi-rigid connections, or consideration of
an equivalent shear stiffness of the removable link.

ACKNOWLEDGEMENT

Support of the Romanian National Education Ministry (MEC-CNCSIS) and World Bank
through the C16 Grant “Reliability of Buildings Located in Strong Seismic Areas in Romania"
and MEC-CNCSIS grant AT10/218 "Seismic response of dual eccentrically braced frames
with removable links" is gratefully acknowledged.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 231


NOTATION

e, ed – clear length of the link between braces, and length of the bolted link
tf, tw, h – flange thickness, web thickness, and cross-section height
Dy, Fy – yield displacement, yield force
fy (Reh), fu, A – (upper) yield stress, tensile strength, elongation at rupture
Vy , My – plastic shear resistance, plastic moment
Vmax, Mmax – maximum shear force, maximum moment
γMb – partial safety factor for bolt resistance
DT – total link displacement
a, b – link panel dimensions
γ – link shear distortion angle
γAL – equivalent link rotation angle due to connection slip
θM, θS, θj – average, bottom, and top connection rotation
γT – total link distortion angle
DD1, DD2 – measurements of link diagonal displacement transducers
DALJ, DALS – measurements of link slip displacement transducers
fy,w, fy,f – web and flange yield stress
Wpl,w, W*pl – plastic modulus of the web and flanges (W*pl = Wpl - Wpl,w, Wpl)
KγT, Kγ, KθJ and KθS – total initial, web shear, and connection stiffness
DTu, γTu – ultimate displacement, ultimate deformation

REFERENCES

(1) Hamburger, R.O. (1996). "Implementing performance-based seismic design in


structural engineering practice". In: Proceedings of 11th World Conference on
Earthquake Engineering, Acapulco, Mexico. Paper no. 2121. Oxford: Pergamon.
(2) Stratan, A., and Dubina, D. (2002). "Control of performance of dual frames with
eccentric bracing", Proc. Stability and Ductility of Steel Structures – SDSS 2002,
Budapest, Hungary, 26-28 september 2002.
(3) Ghobarah, A. and Ramadan, T. (1994). "Bolted link-column joints in eccentrically
braced frames". Engineering Structures, Vol.16 No.1: 33-41.
(4) AISC-97, (1997). "Seismic Provisions for Structural Steel Buildings". American
Institute of Steel Construction, Inc. Chicago, Illinois, USA.
(5) ECCS (1985). "Recommended Testing Procedures for Assessing the Behaviour of
Structural Elements under Cyclic Loads", European Convention for Constructional
Steelwork, Technical Committee 1, TWG 1.3 – Seismic Design, No.45
(6) Dubina, D., Ciutina, A., Stratan, A., (2000). "Cyclic Tests on Bolted Steel Double
Sided Beam-to-Column Joints". The International Workshop Connections in Steel
Structures IV: Steel Connections in the New Millenium. October 22-25, 2000,
Roanoke, Virginia, USA.
(7) Eurocode 8 (1994). "Design provisions for earthquake resistance of structures". CEN
European Committee for Standardisation.
(8) Eurocode 3 (1997). "Design of steel structures. Part 1-1: General Rules and Rules for
Buildings ", CEN. European Committee for Standardisation.
(9) Kasai, K., and Popov, E.P., (1986). "General Behaviour of WF Steel Shear Link
Beams", ASCE, Journal of Structural Engineering, Vol.112, No.2: 362-381.
(10) Engelhardt, M.D. and Popov, E.P. (1992). "Experimental performance of long links in
eccentrically braced frames". Journal of Structural Engineering, Vol.188, No.11:3067-
3088.
(11) C133/82 (1982). "Technical guide for connections id steel structures with high
strength friction grip bolts ". ICB, INCERC (in Romanian).

232 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


STRUCTURAL PROPERTIES OF CONNECTIONS
FOR RACK STRUCTURES

Carlos Aguirre A., Federico Santa Maria University, Valparaiso, Chile

ABSTRACT
Rack Structures are constructed easy to be assembled, that is
the connections must be temporary elements in order to allow
the owner to change the lay out according to the needs, which
makes that bolted and welded connections do not qualify for
these purposes. This paper presents experimental findings about
the beam-column connection under static and cyclic loads. The
connection is extremely flexible, so the rack structures should be
classified as Partially Restrained Construction, according to
AISC design code. The similarity of the static and cyclic failure
modes indicate that the failure is controlled entirely by the
connecting elements.

INTRODUCTION

The design of rack supporting structures has become increasingly complex in recent years;
the properties of the connections determine the behaviour of the structure and the
performance of the structural system in the event of a destructive earthquake occurrence.

There is not much information available about these systems and there are some evidence
showing a poor performance of the racks supporting systems during the past earthquakes.
Rack structures in seismic zones are requested to comply with the local building codes and
they must be engineered to meet the code requirements of the building structures. Even
though rack structures are quite different to the buildings they use to be placed inside a
building, so it is necessary to control the lateral deflections in order to avoid the hammering
of the structures, the collapse of the racks and eventually the collapse of the surrounding
structure. The connections designed to be easily unlocked as well as the user needs makes
difficult to keep control on the amount of beams to be removed, which can produce at last
low redundant structures. The rack supporting structure studied in this paper has been used
successfully in Chile for several years and they have survived the March 3, 1985 Chilean
earthquake. The beam is connected to the column by using hooks that are fabricated with
the beam; these hooks are inserted into columns slots, so they can be easily disconnected
from the column. Details and connecting elements are shown in Fig. 1, the thickness of the
elements is usually 2 mm, however, in some columns this thickness is increased up to 3 mm.

LABORATORY TESTS

In order to study the behavior of this type of rack structures, there were conducted several
static and cyclic tests on these connections. Moment-rotation curves were determined as a
result of the tests.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 233


SPEEDLOCK JOINT CONNECTING ELEMENTS

HOOK

OPENING FOR
BEAM
SAFETY CLIP

301

BEAM BEAM 48 64

80

Figure 1. Connection details (dimensions in mm).

Mechanical properties of the steel

The mechanical properties of the connecting elements were obtained from the testing of
several coupons. The set up is shown on Fig. 2, the axial displacement measurements were
obtained from high resolution photography. The yielding and rupture values are shown on
Table 1.

Figure 2. Tension test for beam properties determination.

Geometric properties of the connections

The properties of the steel shapes are presented on Table 2, according to the notation
indicated in Fig. 3.

234 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Table 1. Mechanical properties of the elements.
Element Fy (MPa) Fr (MPa)
Beam 327 435
Column 280 382
Connection zone 304 353

Table 2. Dimensions of the elements.


Column Beam Connection

Test A B C D E F G H I J K L M N O
1 80 22 46 28.5 74.5 15.5 50 2 53 121 2 48 64 301 2
2 80 22 46 28.5 74.5 15.5 50 2 53 121 2 48 64 301 2
3 80 22 46 28.5 74.5 15.5 50 2 53 121 2 48 64 301 2
4 80 23 48.7 27.6 76.3 17.3 50 3 53 121 2 48 64 301 2
5 80 23 48.7 27.6 76.3 17.3 50 3 53 121 2 48 64 301 2
Cíclico 80 22 46 28.5 74.5 15.5 50 2 53 121 2 48 64 301 2

COLUMN SECTION BEAM SECTION SPEEDLOCK CONNECTION


A I

O
B
HOOKS
C
E BEAM OPENING FOR
SAFETY CLIP
H D
K J

F G F N

L M

Figure 3. Notation for geometry dimensions on Table 2.

Fig. 4 shows the set up of the tests, the testing frame is placed on the floor, supported on
wheels in a self reacting fashion, that is, there is no need of fixed points to take the reactions.
The testing frame allows the testing of two beams at the same time, one of them in a normal
position and the other one either in the normal position or upside down. The inverted position
was included in some of the tests to take into account the possibility of a sign change of the
seismic loads, able to produce the unlocking of the connection.

LVDT transducers were installed on the beams and strain gauges to measure the axial
deformations on the columns; they were discarded afterwards because the column axial
deformation measurements turn out to be negligible. The load was measured by a load cell
placed between the jack and the beam. There were conducted five static tests and three
cyclic ones. Fig. 5 shows a sketch of the geometry of the specimen, the table on the right
indicates the load and the beam position.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 235


Figure 4. Test set up.

Normal Beam T
Right Side Connection x P
TEST x(m) R.S.B. L.S.B.
Static Loads
Stat. 1.4 Normal Inverted
Tests 1 and 4: x= 1.40
Test 2: x= 0.70 Stat. 0.7 Normal Inverted
Tests 3 and 5: x= 0.20
Stat. 0.2 Normal Inverted
Cyclic Loads
Stat. 1.4 Normal Inverted
x=0.70
Stat. 0.2 Normal Normal
Cyc. 0.7 Normal Normal
Left Side Connection
P Tests 1, 2, 3 and 4: Inverted Beam
x
Test 5 y cyclic test: Normal Beam

Figure 5. Sketch of the specimen geometry.

Three positions of the load were considered to take into account the influence of the shear
force in the moment-rotation curve. In a first arrangement, both beams were connected in
the same position they work. In a second arrangement, one of the beams was connected in
an inverted position. Fig 6 shows a detail of the beam to column connection, the transducers
are fixed to the beam and there is a reference point, fixed to the column, in front of every
transducer. Under the left transducer it is placed a safety clip fitted in a special hole. In some
cases the builder does not include the clip or this is replaced by a bolt.

Monotonic loading tests

There were conducted five tests to determine the behavior and the collapse mechanism of the
connection. The tests were conducted as far as the collapse of any of both connections or up to
the maximum displacement of the loading device. An Electrack 2000 jack was used the
maximum displacement capacity of this device is 300 mm.

In the first test a bolt was placed instead of the safety clip at the inverted beam to avoid
unlocking. In order to detect potential yielding of the column, deformations in three points of the
column sections were measured. The failure occurred at the beam in the normal position; the

236 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


load was applied until the outermost hooks became broken. The bolt remained undamaged at
the end of the test; figure 7a shows the specimen after the test.

The second test was basically the same than the first one but the eccentricity of the load was
smaller (0.70 m). The failure was again after the outermost hooks of the connection of the
beam placed in the normal way, were broken, the failure pattern is shown in figure 7d. It can
be observed that the first hook (on the top) is missing and the second and fourth (on the
bottom) failed. The column deformations seems to be unimportant, except for the local dents
that can be observed at the hooks holes after the failure.

Figure 6. Beam to column connection.

Figure 7. Failure modes pattern.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 237


In the third test, the load was applied with an eccentricity of 0.20 m and the safety clip was
used on both beams. The failure occurred at the inverted beam connection as it is shown at
figure 7c; the connection with the beam placed in a normal position did not fail. It can be
seen the safety clip, placed under the transducer, on the upper part of the same figure.

Test N° 4 intended to confirm the results of the first one, the column selected was thicker
(3mm instead 2mm) and the specimen included the safety clip. The failure occurred at the
inverted beam, all the hooks yielded. Figure 7d shows that the load was applied until three
hooks became torn, and the fourth one, close to the safety clip, finished absolutely deformed.
The clip did not suffer any damage, just like in the case of the normal beam connection.

Specimen five was arranged with both beams on the normal position. The failure mode was
similar and the hooks became broken at the end of the test.

All of the tests exhibit similar curves, a comparison of the moment-rotation curves obtained
from connections with beams in a normal position are shown on figure 8. A more detailed
comparison for connections with the beam in an inverted position is presented by Irisarri (1).

2500

2000
Moment [N-m]

1500 Test 1
Test 2
1000 Test 3
Test 4
500 Test 5A
Test 5B
0
0 0.01 0.02 0.03 0.04 0.05

Rotation [rad]

Figure 8. Moment-rotation (beams connected in a normal position).

Cyclic loading tests

The set up arrangement was modified to perform the cyclic tests; a double action Sheffer
jack, with a 500 mm displacement stroke was used. The transducers were also changed to
allow a larger deformation range. Additional transducers were included to measure the gap
of the connection and the displacement of the application point of the load. The loading
history included series of 3 cycles of equal maximum displacement (± 30, ±60 and ± 90 mm),
the maximum displacement of the series was increased until the connection failed. The
behavior of the samples was essentially the same of the behavior observed in the static
tests. A first failure occurred at the outermost hooks, and the connection kept a residual
strength given by the central hook, the safety clip, and the end plate. Figure 9 shows a
typical moment-rotation curve.

238 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Figure 9. Typical moment – rotation curve.

The similarity of the static and cyclic failure modes indicate that the failure is controlled
entirely by the hooks. Such a result could be expected, because the bearing strength of the
plate is normally larger than the shear strength of the hook (when both components are
made with the same steel). The following zones can be identified:

1. A first horizontal line (zero moment) represents a gap; it reaches about 0.02 rad
after several cycles. The gap increases as long as the loading cycles increases;
this effect is a result of the yielding of the hooks and the residual deformations.
2. There is a second zone with a large stiffness, it occurs after the gap when all the
hooks are in the elastic range.
3. The third zone occurs after the yielding of the outermost hooks. It can be
appreciated a reduction of the stiffness, this smaller stiffness is provided by the
hooks who still remain in the elastic range.
4. The stiffness of the unloading path is the same than the elastic one. The yielding
process increases the gap of the connection.
5. When all the hooks have yielded, the moment keeps constant.

The slopes of the curves are easy to be identified, thus, it is possible to describe the behavior
of the Moment-Rotation Curve in terms of several straight lines. The hysteretic model is
shown in red on Fig. 10. The aforementioned behavior is a result of the progressive failure of
the hooks, starting from the outermost of them.

The gap is represented by the horizontal zero line, which is about 0.002 radians in the first
loading step. R1 is the elastic stiffness before the occurrence of the failure of the hooks. A
reduced stiffness follows the elastic loading process as it is indicated in the third loading
path; this new reduced stiffness was called R2. The model assumed that R2 represents the
stiffness from the failure of the outermost hooks up to the maximum strength.

All the tested connections show a similar strength range, from 2000 N-m up to 2500 N-m, the
only exception was the Specimen N° 3, whose strength was 1484 N-m. It was found no
difference on the behavior when the connection is normal compared to the inverted
connection. The hysteretic curve is stable, as long as the number of cycles increases the
initial 0.002 radians gap increases. Even though the connection hooks yield, the deformed
shape was the same during the rest of the loading process. Table 3, shows the maximum
value of the moment, rotation and load, for the static loading case.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 239


2500

2000

1500
R2
1000
Moment [N-m]
R1
500 R 1

0
-0,03 -0,02 -0,01 R1 0 0,01 0,02 0,03 0,04 0,05 0,06 0,07
R1 -500

-1000

R2
-1500

-2000
Rotation [radians]

Figure 10. Hysteretic model.

The connection of the inverted beam with bolt (IBB) has more resistance than the normal
beam without either bolt or clip (NBW) that was observed during the first two tests where the
failure occurred at the normal beam. The rest of the tests were performed with the safety clip
in both connections, in such a case, the inverted beam had a smaller resistance than the
normal beam so the failure occurred at the inverted beam, except in the last test where both
beams were connected in a normal position.

Table 3. Maximum values.

Number Mmax θmax Pmax Comments


(N-m) (rad (N)
Test 1 2192 0.03 1566 NBW-IBB ; Failure at normal beam connection
Test 2 2403 0.04 3434 NBW-IBB ; Failure at normal beam connection
Test 3 1484 0.04 7420 NBC-IBC ; Failure at inverted beam connection
Test 4 2430 0.02 1735 NBC-IBC ; Failure at inverted beam connection
Test 5 2034 0.02 10168 NBC-NBC ; Failure at normal beam connection
NBW : normal beam connection without safety device
NBC : normal beam connection with safety clip
IBW : inverted beam connection without safety device
IBB : inverted beam connection with bolt
IBC : inverted beam connection with clip

CONNECTION CLASIFICATION

The connection can be classified as semi-rigid, but it exhibits a different behavior than other
type of connections. Fig. 11 is a non-dimensional M-θ curve comparison with some other
typical cases, the vertical axis is the moment of the connection divided by the plastic moment
of the beam, and the horizontal axis is the rotation of the connection divided by the plastic
rotation of the beam.

240 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


The rack connection, either bolted or clipped, is different to other semi-rigid or flexible
connections. It reaches about three times the moment capacity of a typical semi-rigid
connection, but this happens at much higher level of deformations, which is a result of the
small slope of the Moment-Rotation curve.

Figure 11. Moment - rotation curves for semi-rigid connections.

Bjorhovde, Colson, and Brozzetti (2), based on a series of 55 set of connection data tests
collected by Kishi and Chen (3), proposed a criteria to determine whether a connection is
rigid, semi-rigid or flexible. They defined three zones assuming that the connection is a beam
segment whose length defines the connection type, for semi-rigid connections the segment
should be between two and seven times the beam height (2d~7d). Fig. 12 shows the non-
dimensional M-θ curve for the connection. It is based on the aforementioned definition, the
axis have the same meaning than the axis of figure 11.

The boundary between the semi-rigid and the flexible zone in terms of moment ratio is 0.2,
the rigid zone falls out of the limits of the graphic.

According to the classification, the equivalent length for this connection is 170d, that is, ten
times larger than the length of a double angle connection, which is extremely flexible. The
curve shows a continuing hardening, so it starts flexible and it moves to the semi-rigid zone
at larger deformation levels.

0.6

0.5

0.4 le=10d
M/Mp

0.3 Clipped
Bolted
0.2

0.1

0
0 20 40 60 80

θ/θp
Figure 12. Moment – rotation curve classification.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 241


CONCLUSIONS

1. Since the rack connections presented are essentially flexible, the rigid approach for the
analysis is not correct; the proper way to analyze the structure is taking into account the
non linear properties of the connection.
2. Even though several types of connections have been tested, and there are several
predictive equations, there is not too much information about these kinds of connections.
It is necessary to perform more tests in order to have a deeper knowledge and a
calibrated model of these rack structures.
3. The connection requires a minimum locking which can be obtained by using the clip, a
bolt or another safety device. The clip seems to be a good solution, because is easier for
the purpose of locking-unlocking the rack structure. However, the easy unlocking of
beams can produce less redundant structures, which makes the structure more
vulnerable to lateral loads.
4. Once a hook has yielded that portion of the connection stop working, the connection
become less redundant and a redistribution of the bending moments towards the center
of the beam span takes place. As a consequence an earlier failure of the beam is
probably to occur.
5. The failure mode is controlled entirely by the hooks, so the failure takes place at the
beam. The beam is a component easy to be replaced and the hooks acting as fuses
prevent the columns failure.
6. The hysteretic loops are stables with a gap near the zero moment of the curve. The
connection unloading path follows a line parallel to the initial loading part of the curve,
and during the next loading steps, it follows the same initial line. The gap increases with
the loops and the rotation increases too.
7. The shear forces seem to have no effect on the moment-rotation curve.
8. The ability of the connection to support vertical loads, the loss of redundancy, and the
type of the failure in a brittle fashion makes them useful for vertical loads but they should
not be used for seismic loads. The connection can reach more than 50% of the beam
capacity before the failure, just like a semi-rigid connection. However the rotation level is
just like a typical flexible connection which is about 0.02 radians when the strength is
about 10% of the full beam capacity.

REFERENCES

1. Irisarri, A. (1998) “Conexiones Semi-rígidas en Estructuras de Acero y su influencia en


el Análisis”. Memoria para optar al título de Ingeniero Civil, Universidad Técnica
Federico Santa María, Valparaíso, Chile.
2. Bjorhovde R., Colson A. and Brozzetti J. (1990) “Classification System for Beam-to-
Column Connections” Journal of Structural Engineering ASCE Vol 116, Nº11, pp 3059-
3076.
3. Kishi, N. and Chen, W.F. (1986), “Data Base of Steel Beam-to-Column Connections”,
Volumes I and II, Structural Engineering Report N° CE-STR-86-26, Purdue University,
West Lafayette, Indiana.

242 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


EXPERIMENTAL BEHAVIOUR MODES OF
COLD-FORMED FRAME-CORNERS

L. Dunai, Budapest University of Technology and Economics, Hungary


P. Fóti, Budapest University of Technology and Economics, Hungary

ABSTRACT
The paper is related to a frame-type cold-formed structural system that is
developed as a load bearing structure of smaller industrial-type and residential
buildings. In the structure C-section structural members are connected to each
other by simple site connections using self-drilling screws. The design is based
on test, where the local (screw) and global (joint) behaviour modes are
investigated. In the paper the experimental programs are emphasized. The test
results of the two-level test programs are shown and discussed as an
interacting phenomena: the joint behaviour is derived and explained by the
local phenomenon. The behaviour modes are classified and the effects of
different joint characteristics are determined.

INTRODUCTION

In the framework of a joint project of Lindab Hungary Ltd. and Budapest University of
Technology and Economics a new cold-formed structural system is under development. The
aim of the research and development activities to give alternative, cold-formed steel
structural solutions for different functional purposes, such as smaller industrial-type buildings,
residential buildings, floor sub-systems, flat roof covering sub-systems. The current phase of
the development is based on the available Lindab C/Z and sheeting profiles and self-drilling
screw type connections. In the applied structural solutions there is a priority on the traditional
frame-type of arrangements due to the traditional Lindab fabrication and erection technology
of the region. In the second step of the R&D project product development is planned to
increase the efficiency of the structural systems.

The load bearing structure of the building system is cold-formed frame, built-up from C-
profiles using self-drilling screws. The configurations of the frames are as follows (figure 1):
• two-hinged portal frames,
• two-hinged portal frames with tie bar,
• two-storey frames with floor beam.

Figure 1. Frame configurations.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 243


The frames are developed with the geometrical characteristics, as follows:
• span: 3 – 10 meters,
• frame spacing: 1 – 1.5 meters,
• column height: 2.5 – 4.5 meters,
• roof slope: 20 – 45 degrees,
• C-profiles: 150 – 300 mm height and 1.5 – 3.0 mm thickness.

The beams are single and the columns are single or double C-profiles. The double C-profiles
are used in box arrangement. The joints are solved by back-to-back self-drilling screw
connections between the webs of the connecting C-profiles, resulting semi-rigid and partial
strength characteristics. A piece of C-profile as an additional element is used in the eve joint.
The design of the cold-formed frame is based on the pertinent parts of Eurocode 3 Part 1.1
(1) and 1.3 (2) standards as presented by Fóti and Dunai (3). The design is supported by
experimental tests as follows:
• beam-to-column joint tests,
• self-drilling screw shear tests,
• compression member tests,
• composite floor beam tests.

The paper deals with the first two – interacting – test programs. Bellow the experimental
studies are summarized. Then the evaluation strategy of the test results is discussed. The
joint experienced behaviour modes are shown with a focus on the interaction of the failure
phenomena of the local connecting element and the global joint. Finally the conclusions and
the further steps of the ongoing research and development activities are drawn up.

FRAME CORNER TESTS

Test program

Full-scale tests are performed on beam-to-column frame corner joints as a part of the
development of the first prototype of cold-formed Lindab frames; see Dunai et al (4). The
results of the studies can be extended for the behaviour of the eve – beam-to-beam – joint,
too. Altogether 38 experiments are performed and the investigated parameters are as
follows:
• size of the fasteners (diameter: 5.5 and 6.3 mm),
• arrangement – number and placing – of fasteners (5-22 pieces of self-drilling
connectors),
• thickness of connected C-profiles (1.0 – 3.0 mm),
• height of the connected C-profiles (150 – 300 mm).

Test arrangement

Figure 2 shows the two types of test setups which are used in the experiment, including
typical joints. In test setup A the lengths of the column and beam elements are 1.0 and 1.5 –
2.0 meters, respectively. The concentrated load is applied on 1.0 – 1.5 meters arms for
different profile heights. Lateral bracings are applied on the top flange by hat profiles (purlins)
and a C-profile (bracing) on the compression side of the joint.

The test setup B is designed for residential frame system, using a 1.0 m long boxed column
element and a 2.5 m long beam section. The C-profiles have different upper and lower flange
widths; the boxed column is build up from two C-profiles slipped into one another and joined
partly on the flanges by self-drillers. Lateral bracing is solved by two pairs of hot rolled
columns keeping the beam in the plane of the frame by fork supports.

244 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


0
20
1500

F
19
20

25
00

58
0

1050
1000

165

200

Figure 2. Frame corner test arrangements A and B.

Measurement system

The main purpose of the experiments is to obtain the moment-rotation relationships of the
joints. The rotations are measured in three points of the joint region by rotation measurement
devices: on the column under the connection and on the beam both sides of the connection
(see figure 2, numbered boxes). The displacements are measured in four places: vertically
under the force and in the middle of the beam; horizontally, transverse direction in the joint;
and horizontally, lateral direction in the joint. Strains are also measured; strain gauges are
placed on the flanges of the beam near to the joint.

SELF-DRILLING SCREW TESTS

Isolated self-drilling screws of the frame corner joints are tested under shear, in single
overlapped condition in the “local” experiments; see Dunai et al (5). The screw types and
sizes are the same as applied in the “global” tests. The plates of the specimens are partly cut
out from the global specimens and partly obtained from base material of the C-profiles. Cold-
formed steel strips nominal basic yield strength of 350 MPa and ultimate tensile strength of
420 MPa, with thickness from 1.0 to 3.0 mm are used. In the program one, two and three
screwed specimens are tested, as it is shown in figure 3. In the tests the force-displacement
relationship of the specimens are measured.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 245


Altogether 38 specimens are tested with the following parameters:
• placement of screws,
• types of screws (SFS, LD),
• plate thickness (1.0, 1.5, 2.0, 2.5 and 3.0 mm),
• combined plate thickness,
• number of screws (1, 2 and 3).
Note, that the properties of the base material are also determined by tensile tests.

Figure 3. Self-drilling screw specimens.

EVALUATION OF TEST RESULTS

The test results of the “global” joint and “local” screw tests are evaluated and presented by
Fóti and Dunai (6), and Fóti (7), respectively. In the papers the typical behaviour modes are
identified and illustrated. In the current paper the results of the two types of tests are
combined and presented with the purpose to obtain correlation between the component and
the joint behaviour. This background is required to develop a component based design
method for the joint as it is presented by Fóti and Dunai (8).

In the traditional approach a bolted connection under shear is designed by separating it into
connector and base material, and checked on the bases of the “independent” failure by the
application rules of the given standard. The results are connector resistances, from which the
number of connectors can be determined. In the investigated details the joint is assumed as
a multi-component structure. The interaction in the element and connector behaviour
requires another level of design methodology. The connector and its region are considered
as a unit. The difference between the separations is similar to the difference between the
meanings of connection and joint. This approach follows the philosophy that is used in the
moment resisting end-plate joint design of Eurocode 3.

If the screw is separated with its region, the interactive failure modes that are affected by
both screw and material properties can be followed. These possible interactive failure modes
are obtained from the tests, as follows:
• tilting and pull-out,
• tilting and bearing,
• tilting and shear and
• tilting and tension failure of fastener.

It means, that only two failure modes: pure bearing failure (not experienced) and pure screw
shear failure are non-interactive – independent modes. In all other cases the properties of
the screws and the base material affect the behaviour of each other.

246 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


To connect the joint and reference local tests, the similarities and the differences of the
conditions should be analyzed first. Main similarities between the conditions of built-in and
separated screw testing are the followings:
• same steel properties,
• same screw types,
• same steel plate thickness,
• mainly pure shear acts on the screws,
• single overlapped arrangements.
Main differences between the conditions of built-in and separated screw testing can be
summarized as follows:
• different number and arrangement of screws,
• stiffening effect of the group of screws: less chance for pull-out in built-in screws,
• more in-plane bending due to the eccentricities: more chance for pull-out of built-in
screws,
• local buckling and distortional buckling creates tension force in built-in screws,
• combined plastic deformations.

In the following sections the combined (local and global) test results are presented and
discussed for typical joint failure modes.

JOINT FAILURE MODE #1: SHEARING

This failure mode is occurred when relatively small number of screws is applied in relatively
thick profiles. In this case the tilting and pull-out of fasteners cannot take place.

Shearing of screws appeared in two different ways, as follows:

The first mode of shearing failure is shown on the moment-rotation relationship of figure 4.a.
The assumed resistance of the profile is nearly reached in the first mode when the dominant
screw is sheared suddenly. The downward steps over the top equals to the shearing
resistance of the screws, one by one. Certain load bearing capacity is experienced on a
lower load level since the remaining screws are still active.
15
8
„a” „b”
Moment [kNm]

6 10
Moment [kNm]

4
5
2

0 0
0 0,03 0,06 0,09 0,12 0,15 0,00 0,01 0,02 0,03 0,04 0,05
Rotation [rad] Rotation [rad]

Figure 4. Joint shearing failure.

At the maximum level of the moment the shearing failure is occurred in more screws at the
same time in the second mode and it resulted in a sudden collapse of the whole joint, as
shown in figure 4.b.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 247


Figure 5 shows the force-displacement relationship of the reference local screw test using
the same thickness of plates, failed by tilting and shearing mode. It is a two-screwed
specimen and the plateau refers to tilting of the screws. After significant tilting one screw is
sheared and the other is worked under reduced load bearing capacity; this screw is sheared
after another plastic-like state, caused by further tilting. The phenomenon is illustrated by the
photos in figure 6.
15

Force [kN] 10

0
0 10 20 30
D is p la c e m e n t [m m ]

Figure 5. Screw tilting and shearing failure.

Figure 6. Illustration of shearing failures.

JOINT FAILURE MODE #2: PULL-OUT

This failure mode is experienced when relatively small number of screws is applied in a
relatively thin profile. The typical moment–rotation curve can be seen in figure 7.
2

1 ,5
Moment [kNm]

0 ,5

0
0 ,0 0 0 ,0 4 0 ,0 8 0 ,1 2 0 ,1 6 0 ,2 0
R o ta tio n [ra d ]

Figure 7. Joint pull-out failure.

248 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


At the ultimate level having large rotations of the screws’ axes (over 60 degrees), the screws
started to be pulled out from the section. On the curve it can be seen as the sudden
reduction of the moment when one thread of the screw is jumped out. When it is stopped on
the next thread the moment capacity is reached again. Then the phenomenon is continued
and a series of jumping are appeared on the curve. This phenomenon results in a ductile
behaviour. Plastic damages of the profiles are localized under the screws. In the case of
specimen with more screws the beam finally is failed by local buckling as an interactive
behaviour of element and connection, and classified as buckling type failure.

Figure 8 shows the force-displacement relationship of the reference local screw tests using
the same thickness of plates, failed by tilting and pull-out mode. The failure mode is
illustrated by the photo. As it can be seen the jumping phenomenon of the screw behaviour is
inherited in the global joint behaviour.
9

6
Force [kN]

0
0 15 30 45
Displacement [mm]

Figure 8. Screw tilting and pulling-out failure.

JOINT FAILURE MODE #3: BUCKLING

General

This phenomenon may appear in the column or in the beam, or in both of them. In all of
these cases, the connectors are strong enough to avoid the failure but due to tilting and/or
bearing the behaviour is combined. The place where local buckling appears is determined
basically by the arrangement of the connectors. Using spread screws the profile’s failure is
experienced in three places: next to the joint in the beam (figure 9), just bellow the fasteners
in the column (figure 10), and near to the joint in the column.

Figure 9. Buckling in the beam. Figure 10. Buckling in the column.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 249


Interaction of tilting/bearing and buckling

This failure mode is obtained when strong enough connection is applied in thinner plates and
the dominant behaviour is appeared in the element by plate buckling. Typically obtained in
case of four groups of self-drilling screws. Figure 11 shows that the initial part of the
moment–rotation curve is nearly linear then the stiffness is gradually decreasing due to
plastic deformation of the components and tilting of screws. In the ultimate behaviour under
the screw group in the compression side of the joint local buckling is experienced, with a
stronger interaction of the screw deformation and local buckling under the group of screws
on the compression side.
6,0
Moment [kNm]

4,0

2,0

0,0
0 20 40 60 80 100 120
Displacement [0,001 rad]

Figure 11. Interaction of tilting/bearing and buckling.

The reference screw local behaviour is illustrated in figure 12. The force-displacement
diagram starts with a clear linear part, followed by plastic-like behaviour, caused by the
interaction of tilting and bearing. A jump on the diagram shows a pull-out effect, directly
followed by bearing failure.
12

8
Force [kN]

0
0 5 10 15
Displacement [mm]

Figure 12. Screw tilting and bearing failure.

Interaction of tilting/pull-out and buckling

This failure mode is experienced when the tilting/pull-out resistance of the connection and
the buckling resistance of the element are close to each other. It is typical when spread of
self-drilling screws is used and in the behaviour the most effected screw (farthest from the
centroid of the screw arrangement) is dominant. Due to the large deformation capacity of the
screws by tilting, a long plateau of the moment-rotation curve at the ultimate level is
obtained, as it is shown in figure 13. After significant plastic deformation, local buckling
appeared in the beam due to the concentrated bolt force effect on the compression side of
the profile. Due to the increased rotations of the joint in the dominant screws pull-out
phenomenon is experienced.

250 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


The reference local screw experiment is the same as it is shown in the case of failure mode
#2 of the joint (see figure 8).
4

Moment [kNm]
2

0
0,00 0,05 0,10 0,15 0,20
Rotation [rad]
Figure 13. Interaction of tilting/pull-out and buckling.

JOINT FAILURE MODE #4: DISTORTION

Distortion typically appeared in specimens of test setup B when the connectors are strong
enough to avoid failure, as it is shown in figure 14. This failure mode is in certain correlation
with the buckling resistance, but also greatly effected by the eccentricity of the setup. The
phenomenon is the following: first the web of the beam moves out from the initial plane due
to the eccentricity, and the column web acts as a bended plate. It causes decrease in the
stiffness of the whole joint, as it is shown in figure 15. In combination with it tilting of the
screws appears, and the jumps on the curve refer to pull-out of screws. The reference screw
local behaviour is illustrated in figure 8.

Figure 14. Illustration of distortion of column.

20

16
Force [kN]

12

0
0 100 200 300 400
Displacement of the beam end [mm]

Figure 15. Distortion failure.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 251


CONCLUSIONS

In the paper the results of an ongoing research and development work on light-gauge cold-
formed frame joints are presented. On the bases of the results of the R&D activities the
following conclusions can be drawn.

The studied web-to-web screwed connections of cold-formed C-profiles result in semi-rigid


and partial strength joints, which can be applied to create frame rigidity. By the analysis of
the test results on global joints and local sheared screws the typical experimental behaviour
modes are determined and explained. From the test results of the joints and sheared screws
the design resistances and stiffness are derived and applied in practical design. The
combined screw and joint test results are capable to derive design model for this type of
joints for more general application, as proposed by Fóti and Dunai (8).

The research and development work is continued in two directions: further joint tests are in
progress on glued and bolted specimens with the purpose to extend the system to bigger
spans, and test-based design model is under development for more general application.

AKNOWLEDGEMENT

The research work is conducted under the financial support of the OTKA T035147 project
and OM ALK00074/2000 R&D project with the cooperation of BUTE and Lindab Ltd.

REFERENCES

(1) Eurocode 3, (1992). Design of Steel Structures, Part 1.1 General Rules and Rules for
Buildings.
(2) Eurocode 3, (1996). Design of Steel Structures, Part 1-3. General rules –
Supplementary rules for cold-formed thin gauge members and sheeting.
(3) Fóti, P. and Dunai, L. (2001). Design aspects of cold-formed portal frames, 3rd
International Conference on Thin-Walled Structures, Eds. Zaras, J., Kowal-Michalska,
K., Rhodes, J., Elsevier Publisher, pp. 203-208, Krakow, Poland.
(4) Dunai, L., Fóti, P., Kaltenbach, L. and Kálló, M. (2000). Experimental study on frame
corner joints built-up from cold-formed C-profiles, Department Reports (1-3) (in
Hungarian), Technical University of Budapest, Department of Steel Structures,
Hungary
(5) Dunai, L., Fóti, P. and Kálló, M. (2001). Experimental study on screwed connections
under shear, Department Reports (1-2) (in Hungarian), Budapest University of
Technology and Economics, Department of Structural Engineering, Hungary
(6) Fóti, P. and Dunai, L. (2000). Interaction phenomena in the cold-formed frame corner
behaviour, 3rd International Conference on Coupled Instabilities in Metal Structures,
CIMS 2000, Eds. Camotim, D., Dubina, D., Rondal, J., Imperial College Press, pp. 459-
466, Lisbon, Portugal,.
(7) Fóti, P. (2002). Behaviour modes of screwed connections under shear – experimental
research, (in Hungarian), Scientific Publications of the BUTE Department of Structural
Engineering, Műegyetem Kiadó, pp. 29-36, Budapest, Hungary.
(8) Fóti, P. and Dunai, L. (2002). Test based design method of moment resisting joints in
cold-formed structures, International Colloquium on Stability and Ductility of Steel
Structures, Prof. O. Halász Memorial Session, Ed. Iványi, M., Akadémia Kiadó, pp.
211-218, Budapest, Hungary.

252 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


PRELIMINARY COMPONENT METHOD MODEL OF STORAGE
RACK JOINT

A. Kozłowski, University of Beira Interior, Portugal


and Rzeszów University of Technology, Poland
L. Ślęczka, Rzeszów University of Technology, Poland

ABSTRACT
Steel storage pallet racks are three-dimensional framed structures, similar to
multi-storey building structures. For practical reason (possibility of easy access
to stored products) pallet racks are not braced in down-aisle direction, so the
only source of the stiffness required for down-aisle stability is the stiffness of the
connections between columns and beams, and the stiffness of the column
bases.
The paper presents selected test results of beam-to-column joints of one of the
commercially available pallet racks system in Poland. The component method
was used to assess main joint properties. Results obtained using proposed
model were compared with test results.

INTRODUCTION

Steel pallet racks are regular, 3D multi-storey, multi-bay structures, used in industry and
warehouses for the storage of palletised goods. Design analysis of such structures is carried
out on 2D sub-assemblies, separately in the cross- and down-aisle directions. Stability in the
cross-aisle direction is provided by bracings. Because of lack of bracings in the down-aisle
direction, structure analysis is carried out by adopting a semi-continuous sway frame model,
i.e. unbraced frame with semi-rigid joints.

Most of the recent design codes and papers, e.g. Baldassino and Zandonini (1), Baldassino
and Bernuzzi (2), Markazi et al. (3) recommend an experimental-theoretical approach and
require experimental tests of beam-to-column connections to obtain semi-rigid joint
characteristics that can be applied in the global analysis. Such experimental tests were
conducted for one of the commercially available steel racks system in Poland. Observation of
joint behaviour during tests suggests possibility of using the component method to evaluate
the main join properties: moment resistance and initial stiffness. The aim of the paper is to
present preliminary model of the component method for investigated rack joints.

BEAM-TO-COLUMN CONNECTION TEST RESULTS

A sketch of tested beam-to-column connection with its dimensions is shown in figure 1. The
joint consists of a beam-end connector with tabs, made of a 4-mm thick cold-formed angle,
welded to each end of the beam. The leg of the angle with tabs is in contact with column web
after assembly, while there is a 2-mm gap between the second leg, perpendicular to the
beam axis and the column web (figure 1).

Tests were carried out on five joint groups. Every tested joint group had the same connector and

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 253


column section, while the size of the beam was different in each group. Group B had a beam of
RHS 100 x50 x3 section, group C from RHS 80 x 40 x3 and group E had a beam of RHS
120 x60 x3 section.
A A
Column
Beam-end-connector Beam-end-connector

Rectangular hollow section Slots RHS


(Beam) Beam

250
19
50
37

A-A Cantilevered tabs


2 mm gap
2
75

90

Figure 1. Dimensions of tested joint.

Testing procedures, instrumentation and detailed test results of analysed joints can be found
in Kozłowski and Ślęczka (4); below the main test results are summarized.

Table 1. Material properties.


Yield stress Ultimate stress Elongation
Part of the joint MPa MPa %
Column 384.2 431.4 30.7
Connectors 346.6 403.3 37.2
Beam 404.7 433.3 36.9

Sj,ini 0.1S j,ini


4.0
1
M [kNm]

M pl,exp
1
M u,exp
3.0

2.0
experimental curve

1.0

0.0 φ [rad]
0 0.02 0.04 0.06 0.08 0.1

Figure 2. Experimental M − φ curve, specimen B5.

254 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Table 2. Main test results of joints.
Group B Group C Group E
Initial rotational stiffness S j ,ini [kNm / rad ] 139,40 116,70 123,06
Plastic flexural resistance M pl ,exp [kNm] 3,19 2,36 3,20
Ultimate flexural resistance M u ,exp [kNm] 3,90 3,27 3,73

The plastic flexural resistance M pl ,exp was obtained as a point of intersection of a straight
line representing initial stiffness S j ,ini , with the line of slope 0.1 S j , ini , which is tangent to the
non-linear part of the curve obtained in test (figure 2).

IDENTIFICATION OF THE COMPONENTS

The component method is now one of the most effective methods to analyse and predict the
rotational behaviour of different types and different configurations of connections. It is mainly
used in case of steel joints made from hot rolled sections, prEN 1993-1-8 (5), and composite
joints, prEN 1994-1-1 (6), but recently has found also other applications e.g. Fink et al (7).

The application of the component methods is usually performed in three stages. The first
stage is the identification of components in the analysed joint, where the complex joint is
subdivided into parts. The second stage is predicting for each component its individual initial
stiffness, strength and deformation capacity. The behaviour of each component is described
by a bilinear relationship between displacement and force. Components that do not affect the
stiffness of the joint are modelled as rigid-plastic, while other components are modelled as
elasto-plastic elements. The third stage is the evaluation of flexural strength and rotational
stiffness of the whole joint. In this stage the lever arms hi should also be predicted for every
group of components.

In the case of analysed storage rack joint, the following components can be identified.

Column web in tearing

A part of the column web situated between the slot and the flange of the column is subjected
to a distributed load transmitted by the tabs in the connector. Fixed-ended beam of a span
equal to the height of the slot, loaded by contact stresses can be used to model the
behaviour of this component (figure 3). The resistance of the component depends mainly on
shear stresses. This component is active only in a tension part of the joint.

V
0.06 F

A 19 A
5

F 0.94 F

A-A AV 2.0

12

Figure 3. Model to determine the resistance and stiffness of column web in tearing.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 255


According to figure 3, the resistance Fcw, tear can be obtained considering the relationship:
1
0.94 Fcw, tear = Av f u.cw → Fcw, tear = 6.36 kN (1)
3
The initial stiffness of spring element modelling this component is given by:
F
k cw, tear = (2)
δ
where deformation δ under load F is the deflection of the beam caused by bending and
shear. Using deflection equations, it can be found that kcw,tear = 5599 kN / cm .

Column web in bearing

This component is active both in tension and compression zones of the joint (figure 4).
Tab

19
4

In tension
In compression zone zone

Figure 4. Local stresses due to the pressure acting on the hole walls.

The resistance of the column web in bearing can be evaluated using analogy to bolted
connections, according to EC3:
Fcw, b = 2.5α f u , cw d t cw (3)

where bolt diameter should be replaced by the thickness of the tab d = t tab = 4.0 mm .
In the case of joint with tabs α = f u , co / f u , cw = 0.935 . Finally, the resistance of the column
web in bearing can be predicted as Fcw, b = 8.07 kN .
The stiffness of the component can be also estimated, as suggested by EC3, for snug
tightened bolts.
k cw, b = 24 k b k t d f u , cw (4)

where d = t tab = 4.0 mm , k b = 1.25 and k t = 1.5 t cw / d M 16 ≤ 2.5 . Using the above
equations it can be found that k cw, b = 90.7 kN / cm .

Column web in tension (compression)

A part of the column web is subjected to tension or compression. The buckling resistance of
the column web in compressions can be computed according to Chen and Newlin (8) as:
3
t cw
Fcw, c = 10766.1 f y, cw = 42.2 kN (5)
d wc
The resistance of the tensioned part of the column web can be determined according to EC3
as:

256 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Fcw, t = ω beff t cw f y , cw (6)

Assuming the tangent of the spreading angle 1:2.5 (figure 5), beff = 25 mm and
Fcw, t = 19.2 kN .

Tension zone Compression zone

1:2.5
tcw=2.0 tcw=2.0

1:1
5 5

be‘ ff
beff

dwc=172 mm dwc=40

Figure 5. Model to determine the resistance and stiffness of column web in tension
and compression.

The axial stiffness of the column web in tension and compression can be evaluated by the
relationship
'
k cw = E beff t cw / d wc (7)
'
Effective width of column web for stiffness calculation beff in compression zone is equal the
length of zone where distributed load from bearing is acting ( 5.0 mm ). The effective width in
tension zone is predicted considering an 45D angle of the load spreading (figure 5). So, the
axial stiffness in the compression zone is k cw, c = 525 kN / cm , and in the tension zone is
k cw, t = 488 kN / cm .

Tabs in shear

Another component affecting the rotational behaviour of the rack joint is the tab of the
connector, subjected to bending and shear from local bearing stresses. To predict the
resistance and stiffness of this component, the model of a cantilever beam, subjected to a
concentrated load can be used (figure 6). The distance between the load and fixed end of
beam, measured by the developed length of tabs is equal l = 7.0 mm . The resistance of the
component can be evaluated as shear resistance:
Ft , s = Av ( f u , co / 3 ) (8)
Area of bearing

l
F

13 ttab

Figure 6. Tab in shear and bending.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 257


Because of the existence of undercuts in tabs, made during punching in time of production,
effective thickness of tabs is equal t tab, eff = 3.0 mm , instead the nominal value
t tab, nom = 4.0 mm . The resistance is based on the ultimate tensile strength of the connector
f u , co rather than the yield stress, so resistance of the component can be calculated as
Ft , s = 9.1 kN .
The initial stiffness k t , s of this component can be calculated using equation (2), where δ is
the deflection of the beam (in point where the concentrated load is acting) from bending and
shear:
l3 1 .2 l
δ = + (9)
3 E I G AV
Both moment of inertia I and shear area AV should be calculated taking into account the
effective thickness of the tabs. Using above equations it can be predicted that
k t , s = 2714 kN / cm . This component is active both in tension and in compression zone of
the joint.

Connector in bending and shear

Behaviour of the connector can be modelled taking into account a cantilevered part of the
connector, protruding over the flange surface of the beam, as shown in figure 7.
δ
X

F 4
Ext.

A-A 70

Int.
A A
6 13 x
37

Figure 7. Connector in bending and shear.

Resistance of the component can be found as plastic resistance of element subjected to


bending and shear:
Vco, s = AV f y , co / 3 = 19.2 kN (10)

M co, b = W pl f y , co = 42.22 kNcm (11)

The initial stiffness k co of this component is calculated as the ratio of the force to the
deflection of the component under this force.

Because of the openings (slots) in connector, the deflection of the component was evaluated
as for shear-wall structures under lateral loads acc. to Benjamin (9), including the effect of
bending and shear. Every tested group of joints have different height of the beam, so the
initial stiffness of the component is different for each group. In every group two values of
initial stiffness has been calculated: for external slot (force is acting in tab more distant to the
beam), and for internal slot (force is acting in tab closer to the beam). Predicted values of the
initial stiffness for each group are listed below.

258 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Group B: k co, ext = 540 kN cm , k co, int = 5664 kN cm ;
Group C: k co, ext = 368 kN cm , k co, int = 2689 kN cm ;
Group E: k co, ext = 662 kN cm , k co, int = ∞ ;
This component is active both in tension and in compression zone of the joint.

Connector web in tension (compression)

Regarding the tension and compression zone of the connector web, the resistance can be
estimated according to EC3 as:

Fcow = ω beff t co f y , co (12)

where the effective width beff of the connector web is found accounting for the spreading of
the stresses transmitted by the beam flange (figure 8). Assuming the tangent of the
spreading angle 1 : 2.5 , it can be adopted that beff = 27 mm and Fcow, t = Fcow, c = 37.4 kN .

dwc=33

beff

tco=4.0

Figure 8. Connector web in tension or in compression.

The axial stiffness of the connector web in tension and compression can be evaluated from
the relationship:
'
k cow = E beff t co / d wc (13)

Assuming an angle of 45D for the load spreading, the value of the initial stiffness
k cow, t = k cow, c = 2800 kN / cm can be calculated.

Beam flange in tension (compression)

Weld toe line

beff

Figure 9. Stress distribution in the beam flange.

The last component, beam flange in tension and compression, has to be considered only
in the evaluation of the joint resistance. Its resistance can be computed as:

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 259


Fbf , c = Fbf , t = beff t bf f y , bf (14)
where the spreading of stresses to calculate effective width of the beam is depicted in
figure 9. The value of resistance can be predicted as Fbf , c = 29.9 kN .

a) kcw,tear kcw,t kt,s kco,ext b) c)


kco,int
kcow,t d) e)
50 50
50 50 ht

hb-tfb

hb-tfb
hb-tfb
50 50 hc
S1 S2 Sj,ini
50 kcow,c 50

kcw,b kcw,c

Figure 10. Mechanical model and procedure for evaluating the rotational stiffness.

EVALUATION OF THE ROTATIONAL STIFFNESS

The mechanical model adopted to predict the initial rotational stiffness is shown in figure 10a,
and the procedure to evaluate the rotational stiffness is depicted in figures 10b-10e. The first
step is the computation of effective stiffness of each row (figure 10b), next predicting the
equivalent overall stiffness in compression and in tension zone and lever arms (figure 10c).
Third step is change of two springs with axial stiffness into one spring with rotational stiffness
(figure 10d) and finally predicting the initial stiffness of the whole model, (figure 10e).
Adequate relationships to make conversion of mechanical model from figure 10a to model
depicted in figure 10e, can be found in mechanics handbooks or in Faella et al (10). In the
calculation it is assumed, that contribution of springs situated in the centre part of the joint is
small, so influence of this row was neglected. Comparison of predicted values of initial
rotational stiffness with values from tests is shown in table 3.

Table 3. Prediction of the initial stiffness and comparison with experimental data.
Group C Group B Group E
Test results [kNm / rad ] 116.7 139.4 123.1
Component method [kNm / rad ] 133.0 148.4 156.4
Difference: (component–test)/test [%] 13.9 6.4 27.0

EVALUATION OF THE FLEXURAL RESISTANCE

Behaviour of such joints should be considered in two stages: before direct contact of
connector with column flange (elastic stage) and after this contact occurred (plastic stage).
Mechanical model adopted to predict the elastic flexural resistance and lever arms for each
group of components are presented in figure 11a. Because the weakest component governs
the resistance of each row, model depicted in figure 11a can be simplified into model
presented in figure 11b. The resistance of components marked “1” in figure 11b is equal
F = Fcw, tear = 6.36 kN when component is in tension (column web in tearing) or
F = Fcw, b = 8.07 kN when component is in compression (column web in bearing). The
resistance of the connector in bending (modelled as rotational spring) is equal to
M co,b = 42.22 kNcm , and the resistance of components marked “2” (beam flange) is equal
F = Fcof , b = 29.9 kN . Model presented in figure 11b, represents only the initial behaviour of

260 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


joint, when the gap between the column flange and connector (fig. 1) is still existing. When the
rotation of the connector increases, the gap disappears and contact stresses starts acting
between column flange and connector directly, (figure. 11c), what change the lever arm. This
model should be used to assess ultimate resistance of the joints. This requires “step by step”
calculation to predict the phase of transition between these two stages of behaviour.
a) Fcw,tear Fcw,t Ft,s b) “1” c) “1”
Mco,b Mco,b
Fcow,t Mco,b
50 Fbf,t 50 50
50 50 50

hb-tfb
hb-tfb

hb-tfb
50 50 hi
50 Fbf,c 50 “2”
“2”
Fcow,c
Fcw,b Fcw,c

Figure 11. Mechanical models for the elastic flexural resistance evaluation
Group E Group B Group C
Fcw,tear Fcw,tear ψ Fcw,tear

Fcw,tear
170

150
160
110

120 100
120

80

100
60
60
70

Mco,b

28
Mco,b Mco,b
65

25

85
75
22

Figure 12. Plastic resistance model of each tested joint group

Plastic flexural resistance of joints in group E and B is governed by resistance of column web
in tearing Fcw, tear in tension zone and connector in bending and shear M co, b in
compression zone (figure 12). Bearing stresses between connector and column (when the
gap disappeared) are non-uniformly distributed – it is assumed triangle distribution. Plastic
flexural resistance can be predicted using equation:
3
M pl = Fcw, tear ∑ hi + M co, b (15)
i =1
where hi is the distance of i − th component from the centre of compression. Predicted
values are presented in the table 4. In case of group C moment resistance can also be
calculated using formula (15), but now the resistance of the first row is reduced to ψ Fcw, tear
due to essential influence of the component M co, b also in the tension zone.

Table 4. Prediction of the flexural resistance and comparison with experimental data.
Group C Group B Group E
Test results [kNm] 2.36 3.19 3.20
Component method [kNm] 2.48 3.0 3.12
Difference: (component–test)/test [%] 5.1 -5.9 -2.5

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 261


CONCLUSIONS

Comparison of predicted by means of the component method results with those obtained
from tests shows the high level of accuracy, both for the stiffness and for the flexural
resistance, in case of analysed joints.

One of the main advantages of component method is the possibility to predict mechanical
characteristic of joints without expensive experimental work. Traditionally, up till now, behaviour
of such complex joints as steel pallet rack joints is predicted by testing. Presented in this paper
component approach can be used as complementary method.

The second important advantage of the component method is the possibility of clear and precise
specification of the influence of each component on the resistance and stiffness of the joint. The
weakest component can be easy identified and improved. It makes easier to conduct the
optimisation of joints, especially in such elements, which are produced in long series, as steel
storage pallet racks.

NOTATION

AV shear area
beff effective width
d wc depth of the column web
f y , cw , f u , cw yield and ultimate stress of column section
f y, co , f u , co yield and ultimate stress of connector
t co , t cw , t bf , t tab thickness of the connectors, column web, beam flange and tab

REFERENCES

(1) N. Baldassino, R. Zandonini. (2002). Design by testing of steel storage pallet racks.
Proceedings of the 3rd European Conference on Steel Structures, Coimbra 2002,
Volume I, pp. 689-698.
(2) N. Baldassino, C. Bernuzzi. (2000). Analysis and behaviour of steel storage pallet
racks. Thin-Walled Structures 37 (2000) 277–304.
(3) F. D. Markazi, R. G. Beale, M.H.R. Godley. (1997). Experimental Analysis of Semi-
Rigid Boltless Connectors. Thin-Walled Structures, Vol.28, No.1, pp.57-87.
(4) A. Kozłowski, L. Ślęczka. (2002). Experimental analysis of beam-to-column joints in
steel storage pallet racks. Proceedings of the 3rd European Conference on Steel
Structures, Coimbra 2002, Volume II, pp. 897-906.
(5) prEN 1993-1-8 (2003). Eurocode 3: Design of Steel structures. Part 1.8: Design of
joints. (Stage 49 draft). CEN, Brussels.
(6) prEN 1994-1-1 (2002). Eurocode 4. Design of composite steel and concrete structures.
Part 1.1: General rules and rules for buildings. CEN, Brussels.
(7) J. Fink, D. Rubin, K. Hollmann. (2003) Anwendung des Innsbrucker
Komponentenmodells bei der Optimierung eines modernen Deckenschaltischs.
Stahlbau 72, Heft 1.
(8) W.F. Chen, D.E. Newlin. (1973) Column Web Strength of Beam-to-Column
Connections. Journal of the Structural Division. ASCE. Vol.99, No ST9.
(9) J. R. Benjamin. (1959). Statically Indeterminate Structures. McGRAW-HILL.
(10) C. Faella, V. Piluso, G. Rizzano. (2000). Structural steel semirigid connections. Theory,
Design and Software. CRC Press.

262 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


STRENGTH, STIFFNESS AND DUCTILITY OF COLD-FORMED STEEL
BOLTED CONNECTIONS
D. Dubina, the Politehnica University of Timisoara, Romania
A. Stratan, the Politehnica University of Timisoara, Romania
A. Ciutina, the Politehnica University of Timisoara, Romania
L. Fulop, the Politehnica University of Timisoara, Romania
Z. Nagy, Lindab Ltd., Bucharest, Romania

ABSTRACT
The paper summarises the results of an experimental program carried out at
the Politehnica University of Timisoara in order to evaluate the performance of
eaves (knee) and ridge (apex) joints of pitched roof cold formed steel portal
frames under monotonic and cyclic loading. Three different configurations of
ridge and knee joints have been tested. The behaviour and failure mechanisms
of joints have been observed in order to evaluate their rigidity, strength and
ductility.

INTRODUCTION

Previous studies by Lim and Nethercot, 2004 (1) and Chung and Lau, 1999 (2) showed that
bolted joints in cold formed steel portal frames have a semi-rigid behaviour. Also, this type of
joints are partially resistant (Lim and Nethercot 2003 (3), Wong and Chung 2002, (4)). An
important contribution to the global flexibility of the joints, besides the bearing effect (e.g. bolt
hole elongation), is due to the deformation induced by the local buckling or distortion of the
thin walled profiles. In an unwisely configured joint premature local buckling can cause the
failure of the joint itself well below the expected load bearing capacity.

Previous studies focused on monotonic tests only. Present paper summarises the testing
programme developed at the "Politehnica" University of Timisoara, on a series of knee and
apex (ridge) joints used for pitched roof cold-formed steel frames. The joints were tested
under both monotonic and cyclic loading aiming to observe the effect of loading type on the
response parameters. In this paper authors extend the results already reported, Dubina et.
al. (5), with the moment-rotation curves and ductility factors characterising the monotonic and
cyclic behaviour of the joints.

TESTING PROGRAM

Specimens

In order to be able to define realistic specimen configurations a simple pitched roof portal
frame was first designed with the following configuration: span 12 m; bay 5m; height 5m and
roof angle 10°. This frame was subjected to loads common in the Romanian design practice
as follows: self weight 0.35 kN/m2 (γULS=1.1), technological loads 0.15 (γULS=1.1) and
symmetric snow load 0.72 (γULS=2.0). These loads were totalling an approximately 10 kN/m
uniformly distributed load on the frame. The frame was analysed and designed according to
the current EN 1993-1-3 (6) rules. The size of the knee and the ridge specimens, and the

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 263


testing setup were chosen to obtain in the connected members a similar bending moment as
observed in the structure.

From the design the elements of the portal frame resulted back-to-back built up sections
made by Lindab C350/3.5 (SUB350 - fy=350N/mm2). In accordance to these cross sectional
dimensions three alternative joint configurations, using welded connecting gusset elements
(S235 - fy=235N/mm2), were proposed (see figure 1 and figure 2).

-RIS- 1-1 -RIP- 3-3

1 M20 gr.6.6 3 M20 gr.6.6

1025 65
220
65
83 -RSG- 1025 65
220
45
M20 gr 6.6 (FB-only) 2 M20 gr.6.6 M20 gr 6.6 (FB-only)
1458 1420

2-2

220
45
1025 65
1420

Figure 1. Main dimensions of ridge connections.

M20 M20 M20

-KIS- -KSG- -KIP-


1445
1535

1535
2354

2354

2354
M20 gr.6.6
M20 gr.6.6 M20 gr.6.6
1 2 3
65
270 65

270 65

M20 gr 6.6 (FB-only) M20 gr 6.6 (FB-only)


360
65

65

65

M20 M20 M20


419

419

419

420 420 413


65 65 360
65 1133 65 1133 65
65 65 65 833
1747 1747 1737

Figure 2. Main dimensions of knee connections.

(a) (b) (c)


Figure 3. Bolt configuration in the cross section.

The connecting bolts are subjected to shear and their design was carried out assuming the
rotation of the joint around the center of the bolt group and a linear distribution of the arising
forces in each bolt, depending on the distance from the rotational center. In the design of the

264 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


joints the bending moment reduced in the rotation center of the joint was used, not the
theoretical one at the corner of the frame.

One group of specimens (KSG and RSG) used spaced gussets, (figure 3.c). In this case,
bolts were provided only on the web of the C350 profile. In the other cases, where two
different details were used for the connecting bracket – e.g. welded I sections only (KIS and
RIS), and welded I section with plate bisector (KIP and RIP), respectively - bolts were
provided on the web only (figure 3.a), or both on the web and the flanges (figure 3.b). The
case where bolts were also on the flanges had in their name the distinctive FB (see table 1).

Table 1. Tested specimens.


Number of
Element type Code Loading type
specimens
RIS-FB-M Monotonic 1
RIS (Ridge connection
RIS-FB-C1* Cyclic - Modified ECCS procedure 1
with I Simple profile)
RIS-FB-C2* Cyclic - Low cycle fatigue 1
RSG-M Monotonic 1
RSG (Ridge connection
RSG-C1 Cyclic - ECCS procedure 1
with Spaced Gusset)
RSG-C2 Cyclic - Modified ECCS procedure 1
RIP (Ridge connection RIP-M Monotonic 1
with I profile and end RIP-M Monotonic 1
Plate) RIP-C1 Cyclic - ECCS procedure 1
KSG-M Monotonic 1
KSG (Knee connection
KSG-C1 Cyclic - Modified ECCS procedure 1
with Spaced Gusset)
KSG-C2 Cyclic - Low cycle fatigue 1
KIS-M Monotonic 1
KIS (Knee connection
KIS-FB-M* Monotonic 1
with I Simple profile)
KIS-FB-C* Cyclic - Modified ECCS procedure 1
KIP (Knee connection KIP-M Monotonic 1
with I profile and end KIP-FB-M* Monotonic 1
Plate) KIP-FB-C* Cyclic - Modified ECCS procedure 1
Total number of specimens 18
*FB Specimens (RIS, RIP, KIS, KIP) with supplementary bolts on the flange (figure 3.b)

Test setup

Monotonic and cyclic experiments were made for each specimen typology, all specimens
being tested statically. Figure 4 shows the test setup and specimen instrumentation. For
monotonically loaded specimens the loading velocity was approximately 3.33mm/min, and
the ‘yield’ displacement was determined according to the ECCS procedure (figure 5).

Figure 4. Loading scheme and instrumentation.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 265


P αy /10
Pmax
Py

αy vu
vy vmax v

Figure 5. ECCS procedure for determining the yield displacement.

For the cyclic tests several alternative loading procedures were used: (1) the standard ECCS
cyclic procedure, ECCS, 1985 (7), (2) a modified cyclic procedure, suggested by the authors,
which is based on the ECCS proposal and (3) a cyclic procedure for low cycle fatigue. Some
inadequacies were found as far as the ECCS procedure is concerned, and will be discussed
later. As seen in figure 4 several parameters of the response were monitored, but in this
paper only the basic force versus displacement are discussed. All graphs and numeric
values refer to Fact as force and the average of Dglft and Dglsp as displacement.

TEST RESULTS

Monotonic tests

The monotonic tests identified failure modes of the different joint configurations. All
specimens had a failure due to local buckling of the cold formed profiles; however two
distinctive modes were identified for specimens with flange bolts and those without (figure 6
and figure 7).

(a) (b)
Figure 6. Failure of ridge specimens RIS-M (a) and RIP-FB-M (b).

If no bolts are provided on the flange of profile, initially minor bearing elongation of the bolt
holes were observed, the failure being due to stress concentration in the vicinity of first bolt
row. The resulting concentration of compressive stress in the C profile causes the local
buckling of the web (figure 8.a) followed by web-induced flange buckling. This phenomenon
occurred in a similar way in the case of RSG and KSG specimens. No important differences
were observed between specimens where no bolts were provided on the flanges. In the case
of the specimens with flange bolts, the stresses concentrated in the vicinity of the firs bolt row
on the flange. In this case no initial elongation of the bolt holes were observed; the buckling
was firstly initiated in the flange, and only later was extended into the web (Figure 8.b).

266 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


(a) (b)
Figure 7. Failure of knee specimens (a) KIS-M and (b) KIS-FB-M.

(a) (b)
Figure 8. Failure mode of specimens with and without bolts on the flanges (FB).

Qualitative FEM simulation (Figure 9) shows locations of stress concentration in case of


specimens with (FB) and without flange bolts. The FEM simulation also demonstrates that
the load distribution in the bolts is not linear. The rotation centre of the joint is shifted towards
the first bolt rows and the force in the first bolts is an order of magnitude higher than the force
in the last two bolts. Therefore the initial assumption yields unsafe results concerning the
design of joint.

Figure 9. Stress concentration in case of different bolt arrangements.

Comparative experimental curves for ridge and knee connections are presented in figure 10.
There are no significant differences among the specimens without flange bolts (RSG-M, RIP-
M, and KSG-M, KIS-M). The explanation for this is because the connecting bolts had higher
rigidity and capacity compared to the other components of the joint. On the other hand, there
is important gain in load bearing capacity when bolts are installed also on the flanges,
although this solution is more difficult to fabricate (RIS-FB-M and KIS-FB-M).

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 267


Comparison for Ridge Connections Comparison for Knee Connections

200 80
RSG-M KSG-M
70
RIP-M KIS-M
160
RIS-FB-M 60 KIS-FB-M

50
120

P (kN)
P (kN)

40
80 30

20
40
10

0 0
0 20 40 60 80 0 50 100 150
v (mm) v (mm)

(a) (b)
Figure 10. Comparative results from monotonic tests for ridge (a) and knee (b) joints.

Based on notations in figure 5 the yield displacement (vy), and corresponding force (Py), were
determined for all monotonically tested specimens. These values are summarised together
with the values of maximum force (Pmax), and the corresponding displacement (vmax) in table
2. The load bearing capacity of the specimens with bolts on the flanges was higher, and only
in this case the capacity of the joint reached the capacity of the connected C profile. It can
also be observed that all specimens have limited ductility. In table 2, PRdth represents the
calculated capacity of the member, while PRd,bth represents the bearing capacity of the bolted
connection. All strength values were calculated using measured characteristics of the cold-
formed steel (fy=452N/mm2 , fu=520N/mm2).

Table 2. Monotonic results: forces and displacements.


Pyexp vy Pmax vmax vu PRdth PRd,bth Peexp / Peexp /
Element
(kN) (mm) (kN) (mm) (mm) (kN) (kN) PRdth PRd,bth
RIP-FB-M 190.4 34.1 193.1 39.1 40.0 164 162 1.16 1.18
RIP-M 130.2 31.2 133.0 35.9 36.6 164 121 0.79 1.08
RIS-M 127.3 26.7 130.5 31.5 32.0 164 121 0.78 1.05
RSG-M 134.0 37.5 137.9 42.0 42.1 164 121 0.82 1.11
KIS-M 56.4 62.6 58.1 74.2 83.5 70.0 65.4 0.81 0.86
KIS-FB-M 75.6 65.8 76.2 71.5 72.4 70.0 59.9 1.08 1.26
KIP-M 70.6 54.0 71.5 59.0 59.9 74.4 66.7 0.95 1.06
KIP-FB-M 91.8 61.1 92.6 65.7 66.7 74.4 73.6 1.23 1.25
KSG-M 55.3 52.5 56.6 60.4 60.4 70.0 65.4 0.79 0.85

Obviously, the specimens with unbolted flanges that failed prematurely by web buckling due
to stress concentration around the first row of bolts, would be the weakest part of such a
frame. Consequently, this joint typology is not recommended to be used in practice. Table 3,
associated with figure 11 show the characteristic values of moment-rotation curves for
flange-bolted specimens (RIP-FB and KIP-FB).

Table 3. Monotonic results: parameters of moment-rotation curves.


Specimen Kini, kNm/rad φy, rad φu, rad µ Mmax, kNm
RIP-FB-C1-pos 3486.9 0.045 0.054 1.19 154.3
RIP-FB-C1-neg 4249.4 -0.040 -0.045 1.14 -158.1
KIP-FB-C_pos 5587.5 0.035 0.050 1.43 167.6
KIP-FB-C_neg 4762.3 -0.032 -0.043 1.33 -159.3

268 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


250 250

200 200

150 150
M, kNm

M, kNm
100 100

50 RIP-FB-M 50 KIP-FB-M

0 0
0.000 0.020 0.040 0.060 0.080 0.100 0.120 0.000 0.020 0.040 0.060 0.080 0.100 0.120
φ TOT, rad φ TOT, rad

Figure 11. Monotonic moment-rotation curves.

Comparing figure 10 and figure 11 one observes that the shapes of the F-d and M-φ curves
are similar, as expected. The ductility µ = φu φ y is 1.19 for knee joint KIP-FB-M, and 1.16 for
apex joint – RIP-FB-M. The reduction of the maximum moment (Mmax) and of ultimate rotation
(φu) in the case of RIP-FB-M specimen is due to the effect of axial compression, which is
significant in this case.

Cyclic tests

In case of the cyclic loading, the degradation of the specimens initiated with elongations in
the bolt holes caused by bearing. Compared to monotonic loading, in this case the
phenomenon was, amplified due to the repeated and reversal loading. However, the failure
occurred also by local buckling, as in case of monotonic tests, but at the repeated reversals,
the buckling occurred alternatively on one and the other side of the profile. This repeated
loading caused the initiation of a crack at the corner of the C profile, in 2-3 cycles following
the buckling, closed to the point where the first buckling wave was observed in the flange
(figure 12). The crack gradually opened in the flange and web causing an important decrease
of the load bearing capacity in each consecutive cycle.

Figure 12. Propagation of cracks due to repeated buckling.

An important observation during the cyclic testing was that the recommended ECCS
procedure for cyclic testing proved to be unsuitable for the limited ductility subassemblies.
The increase of displacement from 1vy directly to 2vy is too sudden making it impossible to
asses the characteristics of the cyclic behaviour (figure 13.a - RIP-C1). As it can be
observed, before the attainment of the displacement limit vy there was no significant damage
in the specimen, while at the displacement level 2vy it already failed. Therefore the failure
mode was not different from the one observed during the monotonic tests. For the other
cyclically tested specimens a modified ECCS procedure was used. In this procedure half of
the difference between vmax and vy was used as increment after the reaching of the

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 269


displacement level of vy. The consequence of the use of this much smaller increment is that
the assessment of the cyclic characteristics became possible (figure 13.b).
Comparison for Ridge Connections Comparison for Knee Connections

200 90

150
60
100
30
50
P (kN)

P (kN)
0 RIP-C1 0
-60 -30 0 30 60 -105 -35 35 105
-50
-30
-100
-60
-150 RIP-C1 KIS-FB-C
RIS-FB-C1 KIP-FB-C
-200 -90
v (mm) v (mm)

(a) (b)
Figure 13. Comparative results from cyclic tests.

The hysteretic curves show a stable behaviour up to the yield limit (vy) with a sudden
decrease of the load bearing capacity afterwards. Therefore the low ductility of the
specimens must be underlined again. Further, the cycles show the effect of slippage in the
joint (i.e. pinching) and strength degradation in repeated cycles. The strength degradation is
stronger in the first repetition, while in the consequent cycles the behaviour is more stable.
Based on the unstabilised envelope of the cyclic curves the strength, capacity and ductility
characteristics of the joints have been determined, and are reported in Table 4. Again, joints
without flange bolts were weaker.

Table 4. Cyclic results: forces and displacements.


Py+exp vy+ Py-exp vy- Pmax Pmin vmax vmin vu+ vu-
Element
(kN) (mm) (kN) (mm) (kN) (kN) (mm) (mm) (mm) (mm)
RIP-FB-C1 191.1 36.8 -191.2 -37.0 191.2 -194.4 40.1 38.6 40 -39
RIP-FB-C2 170.9 33.8 -202.2 -40.4 179.2 -199.5 36.8 -39.7 - -
RIP-C1 122.0 29.3 -112.2 -21.1 132.1 -133.6 35.2 -27.5 33 -38
RSG-C1 140.0 29.8 -134.7 -31.2 140.9 -137.4 33.9 -35.4 34 -40
RSG-C2 130.0 34.2 -138.1 -47.6 135.9 -139.6 -41.6 48.6 42 -53
KIP-FB-C 79.9 58.7 -80.0 -50.0 81.1 -83.4 61.3 -57.95 62 -63
KIS-FB-C 74.9 61.3 -83.3 -57.5 75.8 -84.7 67.9 -59.82 68 -67
KSG-C1 58.9 47.1 -66.1 -56.9 60.9 -67.1 64.4 -59.48 64 -70
KSG-C2 55.6 50.5 -59.8 -52.5 56.7 -62.4 59.3 -52.94 - -

For this reason, as previously, the moment-rotation curves parameters are shown in table 5
and figure 14 for RIP-FB and KIP-FB joints only. Apparently, the cyclic ductility values are
greater than the monotonic ones, at least for positive cycles in case of knee joints. In fact,
this is not true, because the ultimate rotation capacity for KIP-FB-C (positive) is similar with
that of KIP-FB-M, but the yield rotation is reached earlier in cyclic loading. Practically, cyclic
loading makes weaker the knee joint, and the maximum moments clearly show that.

Table 5. Cyclic results: parameters of moment-rotation curves.


Specimen Kini, kNm/rad φy, rad φu, rad µ Mmax, kNm
RIP-FB-M 3836.9 0.041 0.048 1.16 155.9
KIP-FB-M 4256.2 0.043 0.051 1.19 193.2

270 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


200 200
150 150
100 100

M, kNm
50 50
M, kNm

0 0
-0.100 -0.050 -500.000 0.050 0.100 -0.100 -0.050 -500.000 0.050 0.100
-100 RIFB-C1-env+ KIP-FB-C1-env+
-100
RIFB-C1-env- KIP-FB-C1-env-
-150 -150
RIFB-C1 KIP-FB-C
-200 -200
φ TOT, rad φ TOT, rad

Figure 14. Cyclic moment-rotation curves.

CONCLUDING REMARKS

The experiments described in this paper have been carried out recently and the
interpretation of the results is underway, but already some conclusions can be drawn as
follows.

The calculation model for the connection, based on the linear distribution of the force on
each bolt is not correct. The force distribution is unequal due to the flexibility of the
connected member. In fact, the force is an order of magnitude bigger in the outer bolt rows
compared to most inner one. There are two main components, namely the bearing of the
bolts and the local buckling of the connected profile, which interact and determine both the
rigidity and the load bearing capacity of the joint. A correct model for the behaviour must
include both these components.

A connection with bolts only on the web of the profiles is always partial strength. If the load
bearing capacity of the connected beam is to be matched by the connection strength, bolts
on the flanges become necessary.

The ductility of the connection is limited both under monotonic and cyclic loads and the
design, including the design for earthquake loads, should take into account only the
conventional elastic capacity corrected with safety factors. Because there is no significant
post-elastic strength, there are no significant differences in ductility and capacity of cyclically
tested specimens compared with the monotonic ones. However, if the joints are loaded
under the limit of their maximum capacity, even cyclically, their strength is not too much
affected. Consequently, if the joint detailing and connection components sizing may provide
at least 20% overstrength, the cold-formed steel pitched-roof frames could be classified as
class L of ductility (low) according to EN 1998-1 (8).

NOTATION

γULS partial safety factor for resistance


fy yield stress
fu tensile strength
Fact actuator force
Dglft, Dglsp global displacements
F force
d displacement
Py, vy yield force and displacement

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 271


Pmax, vmax maximum force and corresponding displacement
vu ultimate displacement
PRdth calculated capacity of the member
PRd,bth bearing capacity of the bolted connection
Kini initial joint stiffness
φy, φu yield and ultimate joint rotation, respectively
µ ductility
Mmax maximum moment

REFERENCES

(1) Lim, J.B.P. and Nethercot, D.A. (2004). "Stiffness prediction for bolted moment-
connections between cold-formed steel members", Journal of Constructional Steel
Research, Vol.60, No.1: 85-107
(2) Chung, K.F. and Lau, L. (1999). "Experimental investigation on bolted moment
connections among cold formed steel members", Engineering Structures, Vol.21,
No.10: 898-911
(3) Lim, J.B.P. and Nethercot, D.A. (2003). "Ultimate strength of bolted moment-
connections between cold-formed members", Thin-Walled Structures, Vol.41, No.11:
1019-1039
(4) Wong, M.F. and Chung, K.F. (2002). "Structural behaviour of bolted moment
connections in cold-formed steel beam-column sub-frames", Journal of Constructional
Steel Research, Vol.58, No.2: 253-274
(5) Dubina, D., Stratan, A, Ciutina, A., Fulop, L., Zsolt, N. (2004). "Monotonic and cyclic
performance of joints of cold formed steel portal frames". 4th International Conference
on Thin-walled Structures, ICTWS'2004, Loughborough, UK, 23-24 June 2004 (in
print).
(6) EN 1993-1-3 (2001). "Eurocode 3: Design of steel structures. Part 1-3: General
Rules. Supplementary rules for cold-formed thin gauge members and sheeting". CEN
- European Committee for Standardization.
(7) ECCS (1985). "Recommended Testing Procedure for Assessing the Behaviour of
Structural Steel Elements under Cyclic Loads", European Convention for
Constructional Steelwork, TWG 13 Seismic Design, Report No. 45, 1985
(8) EN 1998-1 (2003). "Eurocode 8: Design of structures for earthquake resistance. Part
1: General rules, seismic actions and rules for buildings". CEN - European Committee
for Standardization.

272 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


AN ALTERNATIVE APPROACH TO DESIGN
OF ECCENTRICALLY LOADED BOLT GROUPS

L. S. Muir, P.E., Cives Steel Company, The United States


W. A. Thornton, Ph.D., P.E., Cives Steel Company, The United States

ABSTRACT

The instantaneous center method used to calculate the ultimate capacity of


eccentrically loaded bolt groups assumes that the bolts are the controlling
element in the connection in order to obtain the maximum permissible load
based on bolt shear alone. In other words , a weak bolt, strong plate model is
assumed. Prior to the AISC LRFD 3rd Edition Manual (AISC 2001), this
approach was satisfactory since the designer did not have to be concerned
that the bolts would tear out through the material as long as an edge distance
of 1.5 times the bolt diameter was maintained. The increased likelihood that
bolt tear-out will be the limiting failure mode under the 3rd Edition Specification,
requires that a new model be developed to better approximate the true
maximum permissible load for the connection. Using the Lower Bound
Theorem of Limit Analysis, the authors have developed an iterative weak plate,
strong bolt model which maximizes connection capacity for cases where bolt
tear-out, and not bolt shear, is the limiting factor.

INTRODUCTION

The usual approach to eccentrically loaded bolt groups assumes that the bolts have less
strength than the plates or members they connect. This is called the “weak bolt/strong plate”
model.

When thin plates or members are used, and bolts are close to the edges of plates or
members, the bearing strength of the plate or member at these bolts can be less than the
shear strength of the bolts.

In this case, the “weak bolt/strong plate” model is not correct, because the force distribution
based on bolt shear strength, assumed with this model cannot be achieved.

This paper presents a “weak plate/strong bolt” model which can be used to supplement the
“weak bolt/strong plate” model. The capacity of the connection will be the greater of the
values that result from the two models, but will not exceed the capacity obtained using the
“weak bolt/strong plate” considering only bolt shear.

THE WEAK BOLT/STRONG PLATE MODEL

This is the model that assumes that an eccentrically-loaded bolt group rotates about an
“instantaneous center of rotation” (AISC, 2001). This instantaneous center (ic) determines
the forces and direction of these forces on each bolt. The method can use an elastic or an
inelastic constitutive equation, and the compatibility equation usually assumes that bolt
deformation is linearly proportional to distance from the ic. This method is well established,

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 273


and computer programs and design charts are available to simplify its use.

However, this model does not allow for the situation that exists when bolts close to an edge
cannot develop the force dictated by the ic location. This is a strong bolt/weak plate situation
where the bolt force is dictated by the bearing strength of the plate, rather than the shear
strength of the bolt.

If the force induced in a bolt close to an edge exceeds the edge distance bearing strength, all
bolts in the group must have their forces reduced in the ratio of the edge distance strength to
the bolt shear strength. This must be done to guarantee that the calculated location of the ic
does not change. If the above ratio is 50%, the capacity of the connection is reduced to 50%
of what the weak bolt/strong plate model would otherwise predict. It can be seen that the
strength of one bolt in a group can degrade the strength of the entire group. The weak
plate/strong bolt model is introduced to mitigate this strength degradation.

THE WEAK PLATE/STRONG BOLT MODEL

Figure 1. Typical bolt group and plate – weak plate/strong bolt model.

Consider Fig. 1. This shows a typical plate and bolt group. In the following, bold face
symbols are used to represent vectors. Point O is any arbitrary point used as an origin. The
centroid of the bolt group will usually be used. Fi is the force on the i-th bolt, ri is the position
of the i-th bolt with respect to the origin, and ei is the edge distance for the i-th bolt,
measured from the edge of the bolt hole to the edge of the plate, along as the line of action
of the force Fi. P is the applied load, ex is the eccentricity with respect to the origin, and i and
j are unit vectors in the x and y directions, respectively.

Note that in Fig. 1, the bolt forces Fi are not necessarily perpendicular to the location radii ri.
As noted above, the point O, used as an origin for the ri vectors, is completely arbitrary, but
the bolt group centroid is a convenient point to use.

There is no ic involved in this method. Instead, the bolt forces are each allowed to achieve
any magnitude and direction that will maximize the design strength of the connection, subject

274 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


to the plate bearing, bolt shear, and equilibrium constraints.

In the notation of Fig. 1, the problem of the weak plate/strong bolt can be formulated as:
n
Find Fi such that P = ∑ Fi • j → max subject to the constraints:
i =1
n
∑ Fi • i = 0
i =1

n
∑ Vi xFi − e x ixPj = 0
i =1

Fi • Fi ≤ (φ1.2Fu t )2 e 2i i = 1, 2, ---, n

1.2Fu tei ≤ 2.4Fu dt i = 1, 2, ---, n


1.5Fu tei ≤ 2.4Fu dt i = 1, 2, ---, n

Fi • Fi ≤ φR bs 2 i = 1, 2, ---, n

In the above, t is the plate thickness, Fu is the plate strength, d is the bolt diameter,
e i 2 = e i • e i , and n is the number of bolts. The bolt shear constraint is added to insure that
bolt shear limit state Rbs is not exceeded.

Because of the Lower Bound Theorem of Limit Analysis, any solution to the above problem
will be less than the collapse solution. The problem is formulated as a non-linear
programming problem. The solution, which is the greatest lower bound, approximates the
actual collapse solution and will be less than or at most equal to the collapse solution.

Let the solution to the above problem, the weak plate/strong bolt problem, be denoted by Pp.
If the conventional weak bolt/strong plate problem solution is denoted by Pb, then the
{
proposed capacity is P = max Pp , Pb . }

SOLUTIONS TO SOME SPECIFIC EXAMPLES

Figure 2. Specific solved cases.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 275


To date, the authors have developed ad hoc solutions to the above general problem for the
examples shown in Fig. 2.

The simplest of these cases, the two-bolt group subjected to a vertical eccentric load, Case
(a) in Fig. 2, provides a good example of the procedures which can be used to optimise the
connection. First the equations limiting the connection capacity are derived which are
independent of the direction of the forces. These limit states are bolt shear and bearing, and
are defined by the following equations, where η represents the percentage of the vertical
force resisted by the first bolt, s is the bolt spacing, and ex is the eccentricity:

min (φR bs ,2.4Fu dt ) min (φR bs ,2.4Fu dt )


Capacitybs1 = Capacitybs 2 =
2 and 2
⎛e ⎞
η2 + ⎜ x ⎟ (1 − η)2 + ⎛⎜ e x ⎞⎟
⎝ s ⎠ ⎝ s ⎠

Since the equation for the bolt shear limit state represents the strong-plate/weak-bolt model,
and the bolt capacity is optimised when the bolts vertical load is distributed evenly between
the bolts (η = 0.5) its validity can be proven by re-writing it as a solution for the C-value given
in the Table 7-17 (AISC, 2001) as follows:

1
C=
2
⎛e⎞
0.25 + ⎜ ⎟
⎝s⎠

It should be noted that the values obtained using this equation are slightly higher than those
shown in Table 7-17 because of the empirical nature of the load deformation equation used
by AISC, which results in a maximum normalized force of 0.982 instead of 1.

Next equations for the connection capacity based on bolt tear-out are derived. Under
downward vertical loading it is clear that the top bolt may tear-out through the top of the
plate, while the bottom bolt may tear-out through the side of the plate. This results in the
following capacities:

⎡ ⎤
⎢ ⎥
⎢L φh ⎥
Capacity to − top = φ(1.5)(Fu )(t )⎢ cv − ⎥
⎢ η 2 ⎥
⎛e⎞
⎢ 2 ⎜ ⎟ + η2 ⎥
⎢⎣ ⎝s⎠ ⎥⎦

⎡ ⎤
⎢ ⎥
⎢L φh ⎥
Capacity to −side = φ(1.5)(Fu )(t )⎢ ch − ⎥
⎢ 1− η 2 ⎥
⎛e⎞
⎢ 2 ⎜ ⎟ + (1 − η)2 ⎥
⎢⎣ ⎝s⎠ ⎥⎦

In the above φh is the hole diameter and Lcv, Lch are the vertical and horizontal clear edge
distances.

Having derived the equations for each the of the connection capacities based on each of the

276 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


limit states, we can now obtain the connection capacity, as follows:

Assume plate thickness = 1/4”


Bolt shear strength = 44.2 kips
Horizontal edge distance = Vertical edge distance = 1.5”
Eccentricity = 3”
Bolt spacing = 3”
Fu = 58 ksi

The maximum capacity is found when η = 0. The individual limits are as follows:

26.1
Capacitybs = = 26.1kips
2
⎛ 3⎞
02 + ⎜ ⎟
⎝ 3⎠

26.1
Capacitybs = = 18.5kips
2
(1 − 0)2 + ⎛⎜ 3 ⎞⎟
⎝ 3⎠

⎡ ⎤
⎢ ⎥
⎢ 1.0625 ⎥
Capacity to − top = 0.75(1.5)(58)(0.25)⎢1.5 − ⎥=∞
⎢ 0 2 ⎥
⎛3⎞
⎢ 2 ⎜ ⎟ + 02 ⎥
⎢⎣ ⎝3⎠ ⎥⎦

⎡ ⎤
⎢ ⎥
⎢ 1.0625 ⎥
Capacity to −side = 0.75(1.5)(58)(0.25)⎢ 1.5 − ⎥ = 18.3
⎢1−0 2 ⎥
⎛3⎞
⎢ 2 ⎜ ⎟ + (1 − 0 )2 ⎥
⎢⎣ ⎝3⎠ ⎥⎦

Therefore the capacity of the connection is the minimum value of 18.3 kips. AISC in Table
10-9 reports the capacity of this connection as 23.0 kips. Using η = 0.5 which maximizes the
bolts results in a capacity of 16.7 kips.

It should be noted that when checking bearing and tearout a hybrid of equations J3-2a and
J3-2b in the form of R n = 1.5Lc tFu ≤ 2.4dtFu is employed. This approach has been used for
several reasons, First, the weak plate/ strong bolt model is predicated on the Lower Bound
Theorem of Limit Analysis which guarantees that the applied external forces in equilibrium
with the internal force field are less than or, at most, equal to the applied external force that
would cause failure, provided that all the limit states are satisfied. This theorem is, strictly
speaking, only valid as long as the connection is sufficiently ductile to allow redistribution of
the forces. The use of the 1.5 factor is consistent with this approach since considerable
deformations will occur as the forces in the bolts are distributed. Also since this is a weak
plate/strong bolt model, the use of the 1.5 factor helps to minimize the thickness of the plate,
thereby reducing the likelihood of bolt fracture and increasing the ductility of the connection.
Further, the instantaneous center method itself is based on an assumption that rather large

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 277


deformations will occur in the connection. The load deformation curve which is used to
determine the force at each bolt is based on a maximum, and relatively large, deformation of
0.34” for a 3/4" diameter A325 bolt, which tends to uphold the use of the 1.5 factor for bolt
tearout. The 2.4dtFu portion of the hybrid equation is included in an attempt to allow some
deformation and redistribution of loads without cause excessive deformations. Even though a
few bolts are allowed a capacity based on the larger deformation, the remaining bolts
designed to the 2.4 bolt bearing equation, will tend to limit the overall deformation of the
connection. Finally, bolt plowing is assumed to occur in a standard single plate shear
connection, which is also consistent with the use of the 1.5 factor.

Table 1. Connection capacities using various models.


Bolts: 1” A490-X Plate: 1/4" A36 Eccentricity = 3” Edge Distance = 1 1/2"
Instantaneous
Instantaneous Instantaneous
AISC Table Center Method w/ Proposed
Center Method Center Method
10-9 Hybrid Eq. J3-2a Method
w/ Eq. J3-2a Bolt Capacity
& J3-2b
(a) 23.0 13.4 16.7 18.3 38.8
(b) 36.7 26.2 32.8 33.5 78.5
(c)* 48.9 41.3 51.6 51.6 127
* Net shear is the controlling limit state for this connection, so the capacity is 48.9 kips.

It appears from Table 1 that the greatest benefit is achieved for cases with fewer bolts, this is
because the horizontal component is larger and the bolt tearout becomes more critical. Of
course as the eccentricity increases, the weak plate/strong bolt model becomes more
advantageous to larger number of bolts as well. The model is also suited to analyze
connections subjected to simultaneous axial and shear loads. A 1/4” plate was chosen for
this example because it can be easily compared to the existing values presented in AISC
Table 10-9. However it is not just the connection plate which can be the critical element, and
some beam webs are thinner than 1/4”. In such cases the weak plate/strong bolt model can
be used to optimize the connection to even greater advantage.

One concern that arises is that the weak plate/strong bolt model does not take into account
the load-deformation behavior of the bolts. It is assumed that the load can redistribute freely
to optimize the connection. This is not a major concern for the simple cases shown here,
because when the bolt shear and not plate bearing or tearout governs the capacity, the
capacity is consistent with the instantaneous center method, which incorporates the load-
deformation behavior of the bolts. However for more complicated geometries the capacity
should be limited to that obtained using the instantaneous center method.

CONCLUSION AND FURTHER WORK REQUIRED

As has been shown, the weak plate/strong bolt model can be used to mitigate the effects of
the new AISC bearing and bolt tear-out requirements. However because of the nature of the
problem, the number of equations to be solved increases exponentially as the number of
bolts involved increases. The authors have developed ad hoc solutions to the three cases
shown in Figure 2, as well as two further cases involving both shear and axial loads.
However a non-linear programming solution still needs to be developed to increase the
usefullness of the model for increased number of bolts with larger eccentricities.

278 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


NOTATION

t = the plate thickness


Fu = the plate strength
d = the bolt diameter
ei 2 = ei • ei = clear edge distance
n = the number of bolts
Rbs = bolt shear constraint
η = percentage of the vertical force resisted by the first bolt
s = bolt spacing
ex = the eccentricity
φh = hole diameter
Lcv = vertical clear edge distance
Lch = horizontal clear edge distance

REFERENCES

1. AISC, 2001, Load and Resistance Factor Design, 3rd ed., American Institute of Steel
Construction, Chicago, Illinois, pages 7-7,8 & 7-38 to 7-85

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 279


280 Connections in Steel Structures V - Amsterdam - June 3-4, 2004
EXPLORING THE TRUE GEOMETRY OF THE INELASTIC
INSTANTANEOUS CENTER METHOD FOR ECCENTRICALLY
LOADED BOLT GROUPS

L.S. Muir, P.E., Cives Steel Company, The United States


W.A. Thornton, P.E., PhD, Cives Steel Company, The United States

ABSTRACT

Since 1971, presentations of the Instantaneous Center of Rotation Method for


determining the capacity of eccentrically loaded bolt groups have included
figures which indicate that the instantaneous center is located along a line
perpendicular to the applied load and passing through the center of gravity of
the bolt group. It is the purpose of the paper to show that this geometry is
incorrect for all but a few very specific instances. The primary implication of
this conclusion is that the search for the instantaneous center can not be
limited to the one-dimensional line, but instead must include all points in a two-
dimensional plane.

INTRODUCTION

Figure 1. Figure 2.

The AISC LRFD, 3rd Edition Manual (1) presents Figure 7-2, reproduced here as Figure 1, to
demonstrate the geometric relationship assumed when using the Instantaneous Center of
Rotation Method to calculate the ultimate capacity of eccentrically loaded bolt groups. This
figure indicates that the instantaneous center is located along a line perpendicular to the
applied load and passing through the center of gravity of the bolt group. Crawford and Kulak
(2) also indicate that this is the intended geometry in their paper. It will be shown however
that it is impossible to satisfy equilibrium and the requisite nonlinear load-deformation
relationship of the bolts with this geometry, except for the case where the load is parallel to
one of the symmetry axes of a doubly symmetric bolt group, or a linear elastic constitutive

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 281


equation is used. Therefore, the search for the instantaneous center of rotation cannot be
limited to the one-dimensional line shown in Figure 1 but instead must include all points in a
two-dimensional plane.

INVESTIGATION

In order to prove that equilibrium cannot be satisfied while maintaining both the geometric
constraints of Figure 1 and the load-deformation constraints represented by Figure 2, we will
look at the simple case of a two-bolt group, illustrated in Figure 3 (see section Notation).

From Figure 2 the required load-deformation relationship, based on empirical data, is:

0.55
R i = R ult ⎛⎜1 − e −10∆ ⎞⎟
⎝ ⎠

Since the total deformation at each bolt is assumed to vary linearly with its distance from the
instantaneous center, and the bolt furthest from the instantaneous center is assumed to
reach its ultimate stress when ∆=0.34 inches, the force on each bolt can be calculated as
follows:

0.55
⎛ ⎛ r ⎞⎞
⎜ −10(0.34)⎜⎜ 1 ⎟⎟ ⎟
⎜ ⎝ r2 ⎠ ⎟
R1 = R ult ⎜1 − e ⎟
⎜ ⎟
⎜ ⎟
⎝ ⎠

0.55
R 2 = R ult ⎛⎜1 − e −10(0.34) ⎞⎟ = 0.9815R ult
⎝ ⎠

For simplicity assume that R 2 = 1 . Note that theoretically R 2 should be unity but R 2 =0.982
because of the empirical nature of the load-deformation equation.

0.55
⎛ ⎛ r ⎞⎞
⎜ −10(0.34)⎜⎜ 1 ⎟⎟ ⎟
⎜ ⎝ r2 ⎠ ⎟
Also for simplicity introduce η = ⎜1 − e ⎟ .
⎜ ⎟
⎜ ⎟
⎝ ⎠

Rx
The tangent of the applied load from Figure 3 is tan θ =
Ry

From the geometry the component forces on the bolts can be calculated as:

r2 y r
R 2x = R ult R 2 y = R ult 2 x
r2 r2
r1y r
R 1x = R ult η R 1y = R ult η 1x
r1 r1

282 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


The tangent of the resultant resisting force on the bolts can then be found to be:

r1y r2 y
η +
tan θ R =
∑ Rx
=
r1 r2
∑Ry ⎡η 1 ⎤
r0 x ⎢ + ⎥
⎣ r1 r2 ⎦

Since tan θ R must equal tan θ for equilibrium to be satisfied it can be shown that:

r1y r2 y ⎡η 1 ⎤
η + = r0 y ⎢ + ⎥ (1)
r1 r2 ⎣ r1 r2 ⎦

It can also be shown that this equation can only be satisfied when r1y = r2 y = r0 y which only
occurs when θ = 0 with a doubly symmetric bolt group or when η = r1 / r2 which is the linearly
elastic case.

Having proven that the prescribed geometry cannot be satisfied for the two bolt case, it can
also be demonstrated for other bolt groups by moving the instantaneous center along the line
perpendicular to the applied load and passing through the centroid of the bolt group and then
plotting the resultant theta angle, θ R , as shown for a three bolt group in Figure 4. As can be
seen from the graph θ R is asymptotic to the value of θ , but will never equal θ for a finite r0.
Therefore, equilibrium cannot be achieved.

Figure 4.

THE AISC COEFFICIENTS C FOR ECCENTRICALLY LOADED BOLT GROUPS

Finding errors in the presentation of the theory underlying the calculation of the
instantaneous center calls into question the values in Tables 7-17 through 7-24 presented in
the AISC Manual (1). Fortunately these values were produced by a program based on
Brandt’s work (Brandt (3, 4)). Brandt’s procedure never restricts its search for the

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 283


instantaneous center to the line perpendicular to the applied load and passing through the
centroid of the connection, though from a programming standpoint this would seem to be the
most efficient approach. Instead both the location of the instantaneous center and the angle
of the resultant are allowed to drift off of the values implied by Crawford and Kulak’s theory
until equilibrium is satisfied. Being based solely on the load-deformation relationship and
equilibrium, the C-values presented by AISC are correct. The authors have checked
numerous cases and are satisfied that both the load-deformation relationship and equilibrium
are satisfied using Brandt’s approach.

NOTATION (Also see Figure 3)

Figure 3.

R ult = ultimate shear strength of one bolt, kips


Ri = nominal shear strength of one bolt at a deformation ∆ i , kips
∆i = total deformation in the bolt and the connection elements, in.
∆= maximum deformation in the bolt and the connection elements = 0.34 in.
θ= angle of the applied load measured from the vertical
θR = angle of the calculated resisting force measured from the vertical
ri = distance from the center of the bolt to the instantaneous center
rix = horizontal distance from the center of the bolt to the instantaneous center
riy = vertical distance from the center of the bolt to the instantaneous center
r0 = distance from the instantaneous center to the applied load
r0 x = horizontal distance from the instantaneous center to the centroid of the bolt group
r0 y = vertical distance from the instantaneous center to the centroid of the bolt group
0.55
⎛ ⎛ r1 ⎞ ⎞
⎜ ⎜
−10(0.34)⎜ ⎟ ⎟ ⎟
⎜ ⎝ r2 ⎠ ⎟
η = ⎜1 − e ⎟
⎜ ⎟
⎜ ⎟
⎝ ⎠

284 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


REFERENCES

1. AISC, 2001, Load and Resistance Factor Design, 3rd ed., American Institute of Steel
Construction, Chicago, Illinois, pages 7-7,8 & 7-38 to 7-85
2. Crawford, S. F. and G. L. Kulak Eccentrically Loaded Bolted Connections Journal of
the Structural Division, ASCE, Vol. 97, ST3, March 1971, pages 765-783.
3. Brandt, G. Donald: Rapid Determination of Ultimate Strength of Eccentrically Loaded
Bolt Groups. Engineering Journal, American Institute of Steel Construction, Vol. 19, No.
2, 2nd Quarter 1982.
4. Brandt, G. Donald: A General Solution for Eccentric Loads on Weld Groups.
Engineering Journal, American Institute of Steel Construction, Vol. 19, No. 3, 3rd
Quarter 1982.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 285


286 Connections in Steel Structures V - Amsterdam - June 3-4, 2004
FOUR-PLATE HEB-100 BEAM SPLICE BOLTED CONNECTIONS:
TESTS AND COMMENTS
M.D. Zygomalas and C.C. Baniotopoulos
Institute of Steel Structures, Aristotle University of Thessaloniki, Greece

ABSTRACT
The present research work concerns the laboratory testing of a full strength
bolted splice connection for a HEB100 profile. Taking as reference results
those obtained from a continuous intact beam subjected to a concentrated
force at the middle of its span, eight simply supported beams with full strength
bolted splice connections have been tested. The obtained test results
significantly diverge from the reference ones obtained from the intact beam.

INTRODUCTION

The use of non-pretensioned bolts in full strength splice connections is a common practice in
several countries (1)(2). However, in such splices first due to the difference between the
diameter of bolts and the respective holes and second due to a possible preliminary
imperfection (rotation of the connected parts of the beam), an additional significant deflection
is caused as soon as the loading is applied on the beam and before its value reaches the
calculated value.

Scope of the present research work was to experimentally define the relationship between
the moment at the middle of the span and the deflection at the same point. As a matter of
fact, the latter is tightly connected to the Serviceability Limit State of the beam, even before
the total loading at the middle of the span has been applied (3).

For the laboratory tests, HEB100-beams have been selected because the chosen profile is
very often used as a structural member in a plethora of steel structures (cf. e.g. the purlins of
steel roofs). The material used was Fe 360 and the non-pretensioned bolts were M12-8.8
and M14-8.8. Four groups of different connections were used, once with bolts M12 and once
with bolts M14. Two steel plates having cross section 6x80 mm were used to connect the
webs, whereas another two with cross section 12x100 mm to connect the flanges.
The full strength bolted splices have been designed and constructed so that they exhibit at
least the same load-bearing capacity as the intact beam (4).

CALCULATION OF THE CONNECTIONS

The maximum moment and shear force that the selected profile HEB 100 can carry under
the plastic limits of strength are equal to:

Mpl.y.Rd = Wpl.y * fy/γM0 = 104,2 * 23,5/1,1 = 2221,818 kNcm and

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 287


Vpl.y.Rd = 1,04 * h * tw * f y/(3 ½ * γM0) = 1,04*10*0,6*23,5 /( 3½ * 1,1) = 76,966 kN

For the connection of the two separated parts of the beam, two horizontal steel plates with
cross section 10*1,2 cm were used for the flanges and two vertical ones with a cross section
0,6*5,5 cm for the web (cf. e.g. Figure 1).

CALCULATION OF THE ADDITIONAL PARTS OF THE WEBS

Figure 1. The splice connection (left) and the vertical additional plate (right).

The additional horizontal plates on the flanges were used to transfer the moment in the
connection area, whereas the additional vertical ones were used to carry the shear forces.
Because of the symmetry of beam and loading, only half of a vertical plate was drawn and
calculated. In Figure 1 a half of one of the two additional vertical plates is shown where the
geometrical position of the center of the holes is the point of the load application. Note that in
such splices the shear forces are not the principal problem. The same position and the same
kind of bolts were used for all specimens. The experimental results showed that these
vertical additional parts of the connection did not take or carry any force during the
experiment.

In the following equations, calculations and symbols from EC3 were used (4). A 6 bolt
connection was used for M14-8.8 bolts for the half of the connection of a flange. The position
of the bolts is shown in Figure 1. The horizontal distance between the two connected pieces
of the beam was equal to 1,5 cm for all the specimens.

Py=79,966/2=38,48 kN M=9*Py=348,347 kNcm

Py: (Vx,.p=0 Vy,.p=Py/3=12,828) Ip=Σ(xi2+yi2)=2*52=50

M (Vy,M=M*xi/Ip=346,347*5/50=34,635 kN Vx,M=Myi/Ip=0 kN)

Vx=0 Vy=12,828+34,635=47,463 kN

t=min(tΛκ=0,6 s/2=0,6/2=0,3) t=0,3

α=min(e1/(3*1,5)=4/4,5=0,889 p1/(3*1,5)-1/4=0,861 80/36=2,2 1) α=0,861

Fv.Rd=0,6*fub*A/γMb=0,6*80*1,54/1,25=59,136 KN > 47,463 kN

Fb,Rd=2,5*α*fu*d*t/γMb=2,5*0,861*80*1,4*0,3/1,254=57,859 > 47,463

288 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


The cross section of 0,5*5,5 cm was not available in the market and for this reason, a cross-
section 0,6*8 cm was used for the splice under investigation.

CALCULATION OF THE ADDITIONAL PARTS OF THE FLANGES

Two groups of four different positioning of the bolts were used for the specimens. For the first
group, bolts type M12-8.8 were used calculated for the connections on the flanges. For the
second group, bolts type M14-8.8 were used in an effort to optimize the deformation of the
connected areas. Calculations were not repeated because the change in the diameter of the
bolts was in the safe side.

SPECIMENS B1 AND B2

The suggested position of the bolts is shown in Figure 2 (diameters for the holes do =13 and
do =15 mm) corresponding to specimens B1 and B2.

Figure 2. The position of the holes in the additional horizontal splice plates and flanges
(specimens B1 and B2).

Calculation of the strength of the connection

do = 13 mm A=11,3 mm fub= 80 kN/cm2 fyb=64 kN/cm2

t = min ( t=1 tΛπ=1,2) t = 1 cm

α = min ( 4/(3*1,3)=1,026 8/(3*1,3)-1/4=1,8 80/36=2,22 1) α=1

Fsd = Msd/z = 22,218/(10+0,6+0,6)*10-2= 198,375 kN

Fv.Rd = 0,6*fub*A/γMb= 0,6*80*1,13/1,25=43,392 > 198,375/5=39,675 kN

Fb.Rd=2,5*α*fu*d*t/γMb=2,5*1*80*1,2*1/1,25=192 > 39,675 kN

SPECIMENS B3 AND B4

In an effort to add more stiffness to the area of the connection in order to obtain better results
during deformation, it was decided that the distance between the holes to be greater in the
specimens B3 and B4 than the distance in the specimens B1 and B2. The positions of the
bolts in specimens B3 and B4 is shown in Figure 3.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 289


Figure 3. The position of the holes in the additional splice plates and the flanges
(specimens B3, B4).

Calculation of the strength of the connection

Similarly to the calculation for specimens B1, B2, the present one leads to safe results.

SPECIMENS B5 AND B6

Figure 4. The position of the holes in the additional splice plates and the flanges
(specimens B5 and B6).

Calculation of the strength of the connection

do = 13 mm A=11,3 mm fub= 80 kN/cm2 fyb=64 kN/cm2

t = min ( t=1 tΛπ=1,2) t = 1 cm

α = min ( 4/(3*1,3)=1,026 4/(3*1,3)-1/4=0,776 80/36=2,22 1) α=0,776

Fsd = Msd/z = 22,218/(10+0,6+0,6)*10-2= 198,375 kN

Fv.Rd = 0,6*fub*A/γMb= 0,6*80*1,13/1,25=43,392 > 198,375/6=33,067 kN

Fb.Rd=2,5*α*fu*d*t/γMb=2,5*0,776*80*1,2*1/1,25=148,992 > 33,067 kN

290 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


SPECIMENS B7 AND B8

Figure 5. The position of the holes in the additional splice plates and the flanges for
specimens B7, B8.

Calculation of the strength of the connection

do = 13 mm A=11,3 mm fub= 80 kN/cm2 fyb=64 kN/cm2

t = min ( t=1 tΛπ=1,2) t = 1 cm

α = min ( 3/(3*1,3)=0,769 3/(3*1,3)-1/4=0,776 80/36=2,22 1) α=0,519

Fsd = Msd/z = 22,218/(10+0,6+0,6)*10-2= 198,375 kN

Fv.Rd = 0,6*fub*A/γMb= 0,6*80*1,13/1,25=43,392 > 198,375/6=33,067 kN

Fb.Rd=2,5*α*fu*d*t/γMb=2,5*0,519*80*1,2*1/1,25=99,648 > 33,067 kN

SPECIMEN B9 (REFERENCE)

This specimen was used as an intact beam without any kind of splice connection. The
respective test results were used as reference for the rest experimental results of the
specimens B1 - B8.

EXPERIMENTS AND TEST RESULTS

All the experiments were performed at the Laboratory of the Institute of Steel Structures,
Department of Civil Engineering, Aristotle University of Thessaloniki, Greece. All the
specimens used were simply supported beams with a span of 1 meter and a concentrated
vertical force in the middle of their span.

A composite hydraulic machine was used with an upper limit of force equal to 500 kN. A
computer was used to record and store 1000 pairs of values corresponding to the force and
the vertical displacement for each one of the specimens. In the sequel, the diagrams
corresponding to each one of the specimens from B1 to B8 in parallel with the diagram of the
reference beam B9 in Figures 6 and 7 are depicted. Figure 8 shows the diagrams for all the 9
specimens including the reference one in order to give a general comparison of the splices
test results with those of the intact beam.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 291


300 300

250 250

200 200
B9
kN
kN

150 150

100 100
B9
50
B1 50
B2
0 0
0 10 20 30 40 50 0 10 20 30 40 50 60 70

mm mm

300 300

250 250

200 200
B9
kN
kN

150 150

100
B9 100

50
B3 50
B4
0 0
0 10 20 30 40 50 0 10 20 30 40 50 60 70

mm mm

Figure 6. Figures of the specimens and the diagrams for B1, B2, B3, B4 and B9.

292 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


300 300

250 250

200 200
B9
kN

kN
150 150

100 100 B9
50
B5 B6
50

0 0
0 10 20 30 40 50 60 70 0 10 20 30 40 50 60 70
mm mm

300 300

250 250

200 200
kN

kN

150 150

100 B9 100
B9
50
B7 50
B8
0 0
0 10 20 30 40 50 60 0 10 20 30 40 50 60 70

mm mm

Figure 7. Figures of the specimens and the diagrams for B5, B6, B7, B8 and B9.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 293


300

250

200
kN

150

100

50

0
0 10 20 30 40 50 60 70

mm
B1 B2 B3 B4 B5 B6 B7 B8 B9 (REF.)

Figure 8. Force-deflection diagrams for the connected beams from B1 to B8 and the
reference beam B9.

COMMENTS AND CONCLUSIONS

The diagrams of the test results (Figures 6, 7, 8) for the connected beams from B1 to B8
gave a significant deflection that corresponds to a much lower force than this that
corresponds to the reference curve of beam B9. In particular, only the 1/10 of the load used
for reference beam B9 was used to give the same deflection to the connected beams.

After the end of the testing program, it was observed that all the vertical additional pieces of
the web remained undeformed (intact) due to the fact that they didn’t carry any loads. The
latter certifies the initial assumption taken into account in the calculation of the connection.
Having in mind that it is always critical to check the serviceability limit state of such beams,
the failure of the specimens due to excessive deflection around the area of the splice
connection is obvious. The maximum acceptable percentage according to EC3 § 4.3.2 is
equal to 1/200=0.05% for total loading (4). The experimental results gave a deflection of
approximately 15/1000=1.5% below the 1/10 of the total load.

Two were the reasons for the observed significant deflections around the area of the splice
connections: The first is the difference of 1 mm between the diameters of the holes and
those of the bolts. The existence of the clearance gives a kind of freedom of sliding as soon
as a small (1/10) load is applied on the beam. At this first step, the force loaded only the two
additional horizontal plates on the flanges which reacted as autonomous simply supported
beams instead of the full cross-section of the beam. This is the reason for the existence of
linear part at the beginning of the diagrams for the connected beams from 0 to near 15 mm.
The second reason of the observed significant deflection is due to the insertion of the thread
of the bolts into the mass of the steel in the vicinity of the contact area of the holes, as shown
in Figure 9.

294 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


As a conclusive remark it is noteworthy that the choice of the position of the splice
connection must be very carefully chosen: It must be in a position along the length of the
beam where the maximum moment is less than 1/10 of the maximum moment capacity of the
profile HEB 100.

Figure 9. The holes (with the traces of the insertion of the thread into the steel mass) after
the experiment in an additional horizontal splice plate and in the beam flange.

ACKNOWLEDGEMENTS

The work reported here has been partially supported by the European Union Research and
Training Network (RTN) “Smart Systems. New Materials, Adaptive Systems and their
Nonlinearities. Modeling, Control and Numerical Simulation”, with contract number HPRN-
CT-2002-00284.

NOTATION

d diameter
fu ultimate stress
fy yield stress
t thickness
α distance

REFERENCES

(1) Ivanyi, M. & Baniotopoulos, C.C. (2000) (eds), Semi-rigid Joints in Structural Steelwork,
Springer Wien, New York, p. 350.
(2) Baniotopoulos, C. C. & Wald, F. (2000) (eds), The Paramount Role of Joints into the
Reliable Response of Structures. From the Classic Pinned and Rigid Joints to the
Notion of Semi-rigidity, Kluwer, Dordrecht, p. 480.
(3) Kontoleon, M. J., Kaziolas, D. N., Zygomalas M.D. & Baniotopoulos, C. C. (2003),
Analysis of Steel Bolted Connections by Means of a Nonsmooth Optimization
Procedure, COMPUTERS & STRUCTURES 81, 2455-2465.
(4) ENV1993-1-1 (1993). Eurocode 3, Design of Steel Structures. CEN, Brussels.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 295


296 Connections in Steel Structures V - Amsterdam - June 3-4, 2004
BOLTED CONNECTIONS WITH
HOT DIP GALVANIZED STEEL MEMBERS WITH
PUNCHED HOLES
G. Valtinat, Technical University Hamburg-Harburg, Germany
H. Huhn, Technical University Hamburg-Harburg, Germany

ABSTRACT
Bolted connections of hot dip galvanized steel members with punched holes
are usually taken for mast and tower constructions. These structures are often
loaded by wind, so it makes good sense to examine the fatigue behaviour of
the connections. One of the findings described in this paper is that the
punching of the holes and the galvanizing process have a negative influence
on fatigue behaviour. This poses the question: how can the fatigue resistance
of bolted steel connections of galvanized steel members with punched holes
be improved? The main idea is to use preloaded bolts. A preload perpendicular
to the surface of the members protects the area around the hole by reducing
the notch stresses. This means the structures can be strengthened and their
lifetime will be prolonged.

INTRODUCTION

In a large of test programme we investigated the complete behaviour of bolted connections


of steel members in various applications, notably on masts and towers for electric
transmission lines, antenna towers and lattice towers for wind turbines. These connections
are usually made of hot dip galvanized steel members. Galvanizing is the best corrosion
protection for this type of construction because the cleaning of heavily corroded masts later
on, such as by shot blasting and repainting, is extremely expensive and time consuming. So
the hot dip galvanizing of all steel members is advantageous. But considering the load
carrying behaviour of the bolted connections, the sensitivity of the load displacements and
the fatigue resistance, there are also some disadvantages brought about by hot dip
galvanizing.

Another important influence on these properties is the production of the bolt holes
themselves. For the profiles used, such as angle profiles, it is much cheaper to punch the
holes than to drill them. Punching produces a very heavy impact on the steel until a severe
material hardening occurs around the holes. The ductility is reduced tremendously, the
ultimate tensile strength is usually affected and cracks may be initiated. The area affected
extends about 2 to 3 mm around the holes. This is the reason why reaming of the holes is
often required with thicker plate material. Hot dip galvanizing (with its high temperature) that
takes place after punching may also promote the aging of the material. All these influences
have a negative effect the fatigue behaviour of the connections.

STATIC TESTS

Before we investigated this special problem it was necessary to gain knowledge on the
influence of punching and hot dip galvanizing on the load carrying behaviour of bolted

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 297


connections under static loads. We had several research programmes (1, 2) on the relations
of load displacement compared with those of bolted connections with both galvanized and
non galvanized members with drilled holes. Figure 1 shows the load displacement lines of
two typical double sided lap connections, which are designed to break by ovalisation and
bearing of the holes, which were the weakest part of the connection. We see that both
connections reach nearly the same ultimate capacity (about 300 kN), but the connection with
drilled holes (dashed line) has a much higher plastic displacement than that one with
punched holes. Nevertheless, a connection which reaches 6 mm displacement due to hole
ovalisation has, in our opinion, enough ductility for static loading.

Figure 1. Load-displacement curves of bolted double sided lap connections


with drilled and punched holes – rupture by ovalisation and bearing.

Figures 2 and 3 show the difference of the rupture modes of two connections which were
designed to break by net section rupture. In figure 2 we see the net section of a member with
a drilled hole; only one crack runs through the section with a high reduction in plastic
thickness reduction due to Poisson's ratio. By contrast, figure 3 shows the net section of a
member with a punched hole. Many initial cracks start from the hole wall, and one of them
finally runs through the net section but without any reduction in plastic thickness. This
confirms the previous assumption that the ductility is very much reduced.

Figure 2. Net section rupture of a member Figure 3. Net section rupture of a member
with a drilled hole in a static with a punched hole in a static
short term test. short term test.

298 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


MATERIALS TESTING

The structure of the material at the edge of a punched hole was examined (3). Microscopic
photos were taken at different levels of the thickness of the plate. They show how much the
material along the hole wall changes its grain structure when the punch is driven through it.
Just after the entrance of the punch a special zone at the upper part of the plate is formed,
figure 4a. In this first zone the material flows in the cutting direction of the punch from the
edge into the hole. The plastic flow passes over to the second zone in the middle of the plate
thickness, figure 4b. Then shear cutting takes place, and is followed by the third zone where
the punch leaves the plate, shown in figure 4c. Here a rougher surface is built which is
widened conically.

Figure 4. Micro-section of the material structure at the edge of a punched hole


(calibration line 0.1 mm).
a) Zone at the upper side when punch enters the plate
b) Section in middle of the member
c) Zone at the lower side when punch leaves the plate

In addition, the distribution of the hardness in different sections parallel to the surface at
intervals of 0.5 mm has been worked out (4). In an adequate distance from the hole edge the
average value of the Vickers hardness is 150 HV 0.2 for the uninfluenced area. From 2.5 mm
distance up to the hole edge we found an increase of the hardness up to a value of about
330 HV 0.2. Figure 5 shows the effect of cold-work hardening due to punching.

Figure 5. Distribution of Vickers hardness HV 0.2 in the nearer area around a


punched hole.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 299


With the cold-work hardening the tensile strength will also increase. To check this, we made
tension tests with micro tensile specimen. The cross section of the test pieces is only
2.0 x 0.5 mm. So it is possible to produce five test specimen out of the close-up range of a
punched hole. The position of the specimen is tangential to the edge of the hole. Figure 6
shows the results of the test in form of stress-strain curves. Here you can see the curve M5
which is typical for steel S 235 JR G2. This specimen has a distance of 2.7 mm measured
from the hole edge and lies in the uninfluenced area. If you now look at the curves of the
specimen which were nearer to the hole edge you can see that the tensile strength of the
material increases and that the elongation at fracture decreases rapidly. This means a loss of
ductility and an embrittlement due to cold-working during the punching process. This material
data explains why there are so many small cracks on the surface around a punched hole,
while a member with drilled holes has only one final crack in failure mode.

Figure 6. Stress-stain curves of the micro tensile test specimen


(x distance of the test piece from the edge of the hole).

SLIP AND CREEP TESTS

Many slip tests under static load have been carried out over the years. In the following some
actual friction coefficients µ of double lap joints with different surface treatments are
represented

µ = 0.2 for mill scale,


µ = 0.2 for hot dip galvanized surfaces with pure zinc on it,
µ = 0.35 to 0.38 for hot dip galvanized surfaces without pure zinc,
µ = 0.6 for hot dip galvanized surfaces without the removed pure zinc, but with
additional alkali-zinc-silicate friction paint.

Specimen with pure zinc on its surfaces were dipped into the zinc bath without removing the
pure zinc layer afterwards, they had a glossy appearance. To separate and remove the pure
zinc layer the members were put in a centrifuge after galvanizing. The appearance of these
specimen was matte and lacklustre.

Friction coefficients for different surface treatments which result from former slip tests can be
find in the literature.

As a result it can be seen that the friction paint produces a remarkable increase of the friction
capacity. But it is necessary to check under which percentage of that value µ the connection
does not slip-creep and maintains its rigidity under static long term loads. Therefore, we
examined the creep behaviour of such connections (5).

300 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Figure 7 presents three creep-slip tests of a high strength friction grip connection with hot dip
galvanized contact surfaces with alkali-zinc-silicate friction paint. In theses cases the long
term test loads were 90%, 68% and 50% of the friction capacity taken from short term tests.
The 90%-test did not fulfil the requirement for displacement (< 300 µm within 30 years), while
the 68%- and the 50%-tests did succeed. Of course, all curves in the Figure, which resulted
from one year or two year tests, have been extrapolated in a log-linear diagram according to
EC 3. This extrapolation is shown by the crosses.

Figure 7. Creep-slip curves of long term tests with alkali-silicate-zinc-paint on


the hot dip galvanized contact surfaces.

After these tests the question came up as to whether the friction capacity µ may be
influenced by a long term loading, such as included by the creep tests. Therefore, the
behaviour of the slip connection with hot dip galvanized contact surfaces having alkali-zinc-
silicate friction paint has been tested after creep tests, while in the creep tests the actual
friction has been exploited only up to 60% till 70%. The subsequent slip tests were performed
until slip occurred and they should give the remaining friction capacity. Figure 8 shows in the
left diagram the results of the initial short term slip tests with HVM16 and HVM20 bolts. The
series II and III in the middle of the diagram show the ultimate slip capacities after long term
tests (no reduction!). Series IV at right hand in the diagram show the slip capacities after
fatigue tests with two million load cycles (no reduction!).

Figure 8. Behaviour of the slip coefficient of hot dip galvanized surfaces with alkali-
silicate-zinc-paint during long term testing and fatigue loading.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 301


Besides the tests described above, another treatment of the contact surfaces has been
investigated, namely high strength friction grip connection with hot dip galvanized contact
surfaces with a slight shot blasting (sweeping) instead of alkali-zinc-silicate friction paint. The
loading procedure was the same as before. The result of the tests was analogous. After long
term tests (series II and III) and after fatigue tests with two million load cycles (series IV)
there was no reduction of the ultimate slip capacities in either case.

A last question: Is there constancy or an increase of the displacement during fatigue loading
up to more than one million (M) cycles? The solution is illustrated by figure 9. Under high
exploitation of about 70% of the static slip capacity and after 1.1M to 2M cycles the
displacement ended in a maximum 0.12 mm. We see a kind of hysteresis which converges
on a value much below the limit of 0.300 mm. Nine more tests show the same, or even
better, results.

Further investigation have been performed on hot dip galvanized steel members with
punched slotted holes, friction paint and high strength preloaded bolts. Results can be taken
from the literature (6).

Figure 9. Creep displacement for hot dip galvanized connections with


preloaded high strength bolts during fatigue loading.

A frequently asked question is: Can we rely on a certain constancy of the bolt preload FV or
do we need to take into account a considerable reduction? This takes place since in a steel
package there are at least four galvanized zinc layers on the members, another four
galvanized zinc layers on the washers and two layers on the bolt head and nut when
clamped together. The total amount of zinc layer may add up to 0.5 mm, which surely creeps
under the high preload of the bolt. Therefore, a second tightening procedure should go over
the connections after two hours or an overtightening of the bolts should be envisaged.

But what is the amount of the preload reduction? Since in our tests the preload is
permanently measured, we can give relevant information on this topic. Apart from that there
is the possibility of overtightening the bolts by a well defined percentage to cover the creep of
the bolt force. Rotation-preload diagrams of six tightening tests with HV bolts M 20x100 -
10.9 show that an additional rotation of 20° makes an increase of about 10% of the required
preload according to DIN 18800-7, EC 3 and EN 1090, which usually covers the creep
influence.

302 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


FATIGUE TESTS

Masts and towers are, in addition to their dead load, mainly loaded by wind. The wind may
come from different directions, leading to oscillating movements and stresses in the structure
and in the members as well as in the joints. The resulting number of cycles can be very high
and is in the range of the fatigue strength. The use of bearing-type connections in hot dip
galvanized members with punched holes leads to questions about the load carrying
behaviour of these joints under cyclic loading. The knowledge of the fatigue behaviour of the
joints is necessary to predict the lifetime of the structures.

In earlier research projects at the TUHH, low-cycle-fatigue and high-cycle fatigue tests were
carried out. Members with holes and bearing-type connections with both punched and drilled
holes, but without any preload of the fasteners were examined (7, 8). The test specimens
consisted of S 235 JR G2 (formerly: RSt 37-2) and the loading was of simple sinus wave
form, while the ratio κ between the lower and upper tension in the net section was +0.1. The
test specimen for the shear connection (type 1) and the member with a hole (type 2) are
shown in figure 10. The bolts were hand tightened after they had contact with the wall of the
hole in order to exclude the influence of friction due to preloading of the bolts. The
dimensions of the test specimens were selected primarily to produce a net section fatigue
failure of the middle member, and not a failure in the bolts or cover plates.

Figure 10. Test specimen of bearing-type connections (type 1) and members


with a hole (type 2) for fatigue testing.

After the fatigue failure of the test specimens we examined the crack surfaces to get
information about the crack initiation. Figure 11 shows two representative net section failures
of members with a hole. It was remarkable that the starting point of the crack front for
punched holes lies in the first zone on the upper corner of the hole edge (figure 11a) and
never in the third zone where the punch leaves the plate. From here the fatigue crack front
runs through the material until the net section is so much weaken that finally a static crack
causes the rupture. In members with drilled holes a surface crack at the wall of hole is
predominant (figure 11b).

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 303


Figure 11. Fatigue rupture mode with marked crack fronts after fatigue loading.

INFLUENCE OF PUNCHING AND GALVANIZING

The analysis of the fatigue tests happens grafically as S-N curve (stress S over load
cycles N) in log-log diagrams. The higher the curve in the diagram the higher is the fatigue
resistance of the test specimen.

The experimental S-N curves of different hot dip galvanized test specimens are shown in
figure 12. Members with hole and bearing-type connections are compared. As expected, the
members with a hole were able to withstand a higher stress range ∆σ at the same number of
cycles N up to failure than the joints. A comparison between the test specimen with punched
holes and the test specimen with drilled holes shows the negative influence of punching. The
S-N curve for both different structural members with punched holes lies under the
corresponding S-N curve for drilled holes.

Figure 12. S-N-curves of hot dip galvanized members with hole and shear-
bearing connections with drilled and punched holes, FV = 0
(L = member with a hole, V = bearing-type connection s = punched hole,
b = drilled hole, f = hot dip galvanized, κ = stress relation).

Fatigue tests of hot dip galvanized and non-galvanized bolted connections with punched and
drilled holes were carried out to determine the influence of galvanization performed after
punching or drilling. The results are compared in figure 13. It can be seen clearly that the
fatigue life decreases owing to galvanizing and punching. This is the case for members with

304 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


both punched and drilled holes. The figure also shows that the influence of punching on
fatigue behaviour is roughly equal to influence of galvanizing. This means that a non-
galvanized structural member with a drilled hole has the highest fatigue resistance, for
example 2M cycles at a constant stress range ∆σ = 80 N/mm2. If the member has a punched
hole or is galvanized, the influence is nearly the same; the fatigue life decreases with a ratio
of 2.0. Now the fatigue failure for a stress range ∆σ of 80 N/mm2 is at 1M cycles. If the
member is both punched and galvanized there is an additional effect and the number of load
cycles decreases to 500,000.

Figure 13. S-N-curves of non-galvanized and hot dip galvanized shear-bearing


connections with drilled and punched holes, FV = 0 in comparison with
EC 3 (nf = non galvanized).

COMPARISON WITH EUROCODE 3

Figure 13 shows a comparison of experimental results of tested bearing-type connections.


They are classified as detail category 112 of EUROCODE 3. From N = 10,000 to a constant
amplitude fatigue limit at N = 5 x 106 the slope of the curve is m = 3.0. The slope then
changes to m = 5.0 up to the cut-off limit (N = 1 x 108). All results of the tests are clear below
the S-N curve of EC 3. The S-N curve of EC 3 shows the material resistance with P = 95%
confidence interval and 5% probability of failure, while the mean fatigue strength line of the
tests has a P = 50%. Therefore, the statistical analysis confirms that the experimental results
are even lower than the S-N curve of EC 3. The experimental tests for members with a hole
have the same results. This leads to the conclusion that dimensioning using EC 3 is not safe
for these structures. This is valid for hot dip galvanized members as well as for non-
galvanized members with drilled or punched holes.

The relation of these results to the corresponding S-N curves of the EC 3 is of great interest
with regard to the position of the points of failure and the slope of the S-N curve relative to
the S-N curves of EC 3. The slope of the fatigue strength curves of bolted joints lies between
m = 5.9 and m = 6.7. The results show that the slope of the S-N curves with m = 3.0 is too
high for these experimental curves.

In literature (9) we found other fatigue tests of connections in comparison with EC 3. The
result of the experimental tests is the same as before. Most of the fatigue life data lies below
the S-N curve of EC 3. This leads to the question: Is the EC 3 is suitable for use? Because
the detail category 112 is specifically permitted to non-preloaded high strength bolts.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 305


THEORETICAL INVESTIGATIONS

Nevertheless, independent from the comparison with EC 3 it is necessary to strengthen the


bolted connections in hot dip galvanized steel constructions under fatigue loading. The
reason for the low fatigue resistance of bearing-type connections is the unfavourable stress
distribution in the net section of the member. Near to the hole there is a high peak of the
notch stress. In the 1950s and 1960s high strength friction grip connections with preloaded
bolts were developed (10). Tests have shown that the fatigue behaviour of members with
preloaded bolts is much better than that with non-preloaded bolts. The increase in capacity,
which means the increase in stress range ∆σ or load cycles N was so great that sometimes
the fatigue behaviour of plain bars could be reached. This is due to the high pressure under
the washers of the bolts. This high pressure gives a certain protection of the area around the
hole, so that the stress distribution in the net section became much more favourable, even
after slip of the connection.

FATIGUE TESTS WITH PRELOADED FASTENERS

The idea of covering the area around the hole by a high strength preloaded bolt can be
transmitted to the problem of fatigue behaviour of hot dip galvanized steel members with
punched holes. In a 2003 finished research project on this topic many fatigue tests were
carried out at the TUHH. The aim was the strengthening of the fatigue resistance of the
constructions. For this the negative influence of punching and galvanizing on material data
near to the hole must be reduced. To ensure the comparison between the fatigue tests of the
earlier research project of the members without any preload, we use the same geometry of
the test specimen for this project (see figure 10).

In the diagram of figure 14 the results of various fatigue tests can be seen. Hot dip
galvanized bearing-type connections with punched holes were compared with friction-type
connections having 50% and 100% preloaded high strength bolts and with an unnotched
specimen. The open circles and the lowest S-N-curve show the load cycles of connections
with non preloaded bolts. The full grey circles represent the results of equivalent connections
with preloaded bolts tightened up to 50% of the required preload for a high strength bold
M16. Even a preload of only 50% increases the fatigue behaviour enormously. The increase
can be better expressed by the stress range ∆σnet than by the number of cycles. We found a
step from 72 N/mm2 to about 180 N/mm2 at 1 million cycles. There was a further increase of
fatigue strength up to 233 N/mm2 at a preload of 100% of the required preload (full black
circles). This value is near to the fatigue failure of unnotched bars. They reach a stress range
∆σnet of 280 N/mm2 also regarded at 1 million cycles.

These regarded S-N curves for the friction-type connections have a friction coefficient µ of
0.38 between the middle member and the cover plates. This value is valid for hot dip
galvanized surfaces without any pre-treatment. As a consequence of the influence of friction
identified previously, we painted the surface of the members with alkali-silicate coating to
increase the friction coefficient. With this treatment we achieve a value of µ = 0.6. The results
of the fatigue tests are also shown in figure 14. The open grey circles represent the results of
tests with 50% preloaded bolts and the open black circles the results with 100% preload. It
can be seen that it is possible to achieve a further increase of the fatigue resistance in
comparison with the corresponding S-N curves for friction-type connections without any
treatment of the hot dip galvanized surface.

306 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Figure 14. Comparison between S-N curves for hot dip galvanized bearing-type
connections with punched holes, friction-type connections with 50%
and 100% preloaded bolts and unnotched specimen (RS = reference
tests with unnotched bars, asz. = alkali-silicate-zinc-coating).

In the next diagram (figure 15) we compared the S-N curves again with the corresponding S-
N curves for non-galvanized connections with drilled holes and with the S-N curve of
EUROCODE 3 (detail category 112). As seen before, the preload increases the fatigue
behaviour of the connections. One interesting scientific finding is that the negative effect of
punching in combination with galvanizing is neutralized because the experimental S-N
curves for 50% and 100% are nearly identical. And these S-N-curves now lie above the S-N
curve of the detail category 112 of EC 3 for high number of load cycles.

Figure 15. S-N-curves of non-galvanized and hot dip galvanized connections


with drilled and punched holes with preloaded high strength bolts,
FV = 0, 50% and 100% of 0,63⋅fu,b⋅ASp in comparison with EC 3 (detail
category 112).

The problem which is still exists is the evaluation of the slope with m = 3.0. The real
behaviour of the tests shows that the slope is smaller than m = 3.0. In particular, the preload
causes a smaller slope, which can be explained by the crack growth. The crack can only be
observed in the protected area under the washers. If the crack leaves this area the rupture of

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 307


the member begins directly. When the slope of the EC 3 is changed from m = 3.0 to m = 5.0
there is a better fitting for the whole S-N curve (see figure 15) and especially for the low-
cycle-fatigue behaviour. The slope of m = 5.0 is chosen with reference to DS 804 which is
valid for railway structures. In DIN 15018 and DIN 4132 the slope is once more flatter.

Besides the fatigue resistance the crack growth is of great interest. The literature gives many
information about the stress distribution of members with cracked and uncracked net
sections but we found nothing about the influence of the pressure under the washers of a
preloaded high strength bolt relating to this problem. Based on experimental tests and finite
element parameter studies we did further investigations to clear the effects of the preload on
the notch stresses near to the hole, the crack initiation and the crack growth. Finally a
simplified calculation method for the lifetime prediction of the steel members was developed
(11, 12).

With that a large test programme with static, low-cycle- and high-cycle-fatigue tests is
finished which had the aim to give answer about the bearing capacity and the fatigue
behaviour of hot dip galvanized steel members with punched holes.

CONCLUSION

The idea of protecting the net section area around a punched hole of a hot dip galvanized
member by the use of preloaded high strength bolts with two washers has shown that a
remarkable influence on the fatigue life can be achieved. The advantage of this method is the
ease of handling with maximum of efficiency. Even the negative influence of punching in
combination with hot dip galvanizing can be neutralized.

ACKNOWLEDGEMENTS

We greatly acknowledge that the institution AiF (Arbeitsgemeinschaft industrieller


Forschungsvereinigungen in Bonn, Germany), the Minister of Economy of the Federal
Republic of Germany, Berlin, and the GAV (Gemeinschaftsausschuss Verzinken e.V.,
Düsseldorf) gave financial supported this research (Research project AiF no. 11097/N1 and
AiF no. 12547/N1).

REFERENCES

(1) Valtinat, G., Dangelmaier, P. (1993). Schraubenverbindungen mit gestanzten Löchern


in zugbeanspruchten, feuerverzinkten Bauteilen. Bericht Nr. 119 des
Gemeinschaftsausschuß Verzinken e.V., Forschungvorhaben GAV-Nr. FD 18, AiF-Nr.
7448, Düsseldorf.
(2) Valtinat, G., Wilhelm, M. (1995). Vergleich des Last-Verschiebungs-Verhaltens und der
Traglast von Schraubenverbindungen mit gestanzten und gebohrten Löchern in
zugbeanspruchten, feuerverzinkten Bauteilen. Bericht Nr. 129 des
Gemeinschaftsausschuß Verzinken e.V., Forschungsvorhaben GAV-Nr. FD 20, AiF-Nr.
9305, Düsseldorf.
(3) Valtinat, G., Grycz, J., Wilhelm, M. (1991). Feuerverzinkte Stahlbau-Verbindungen mit
hochfesten Schrauben und gestanzten Löchern. Vortrags- und
Diskussionsveranstaltung 1990 des GAV, Düsseldorf, Germany, ISSN 0344-3582, 115-
132.
(4) Valtinat, G., Huhn, H. (2003). Betriebsfestigkeit von stählernen gleitfesten
Verbindungen von feuerverzinkten Bauteilen mit gestanzten Löchern und hochfesten

308 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


vorgespannten Schrauben. Gemeinschaftsausschuß Verzinken e.V.,
Forschungsvorhaben GAV-Nr. FD 23/II, AiF-Nr. 12547/N1, Düsseldorf.
(5) Valtinat, G., Albrecht, F., Dangelmaier, P. (1993). Gleitfeste Verbindungen mit
feuerverzinkten Stahlteilen und reibfesten Beschichtungen oder anderen
reibbeiwerterhöhenden Maßnahmen, Teil I und Teil II. Bericht Nr. 122 des
Gemeinschaftsausschuß Verzinken e.V., Forschungsvorhaben GAV-Nr. FG 23, AiF-Nr.
7571, Düsseldorf.
(6) Valtinat, G. (1996). Gleitfeste vorgespannte Verbindungen mit Langlöchern bei
feuerverzinkten Stahlbauteilen für Fassaden-Unterkonstruktionen. Bericht Nr. 132 des
Gemeinschaftsausschuß Verzinken e.V., Forschungsbericht AiF-Nr. 9266, GAV-Nr. FG
25, Düsseldorf.
(7) Valtinat, G. (1996). Low-cycle-fatigue-Verhalten und Schwingfestigkeitsunter-
suchungen an Schraubenverbindungen mit feuerverzinkten Bauteilen und gestanzten
Löchern. Bericht Nr. 135 des Gemeinschaftsausschuß Verzinken e.V.,
Forschungsvorhaben GAV-Nr. FD 21, AiF-Nr. 9864, Düsseldorf.
(8) Valtinat, G., Huhn, H. (2000). Betriebsfestigkeit von stählernen Lochstäben und
Schraubenverbindungen mit feuerverzinkten Bauteilen und gestanzten Löchern.
Gemeinschaftsausschuß Verzinken e.V., Forschungsvorhaben GAV-Nr. FD 23, AiF-Nr.
11097/N1, Düsseldorf.
(9) DiBattista, J. D., Adamson, D. E. J., Kulak G. L. (1998). Fatigue Strength of Riveted
Connections. Journal of Structural Engineering, 124 (7), 792-797.
(10) Steinhardt, O., Möhler, K. (1954). Versuche zur Anwendung vorgespannter Schrauben
im Stahlbau, I. Teil. Berichte des Deutschen Ausschusses für Stahlbau, Heft 18.
(11) Hadrych, I. (2000). Wachstum von Ermüdungsrissen an Niet- und Schraubenlöchern
unter Berücksichtigung von Vorspannkräften der Verbindungsmittel. Dissertation,
Technische Universität Hamburg-Harburg.
(12) Huhn, H. (probably published 2004). Ermüdungsfestigkeit von Schraubenverbindungen
aus feuerverzinkten Stahlbauteilen mit gestanzten Löchern. Dissertation, Technische
Universität Hamburg-Harburg.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 309


310 Connections in Steel Structures V - Amsterdam - June 3-4, 2004
HIGH STRENGTH HALF ROUND HEAD AND NUT HV-BOLTS
FOR ANCIENT STEEL CONSTRUCTIONS
G. Valtinat, Technical University Hamburg Harburg, Germany

ABSTRACT
More and more ancient and historic steel structures need repair and restora-
tion. Sometimes the authorities are on the way to decide to take the complete
building down and replace it by a new one. In many cases the “Authority for the
Preservation of Ancient Monuments” (Denkmalschutz-Behörde) insists on
equal looking and identical constructions such as trusses, arches and so on.
Especially the connections have to look similar to riveted connections. To seek
workers on the market who are skilled for the job of hot driven riveting is hope-
less. Therefore the rivets have to be replaced by specially designed high
strength bolts with round heads and similar looking nuts. One German com-
pany specialised in this field and created the so called “Rundkopf-HV-
Schrauben AF” (Round head HV-bolt assembly AF) which looks like a rivet
when installed, because the hexagon for the tightening process has been re-
duced to a very small height. From a distance of 1 to 2 m the hexagon nearly
cannot be identified. These bolts have been accepted in a recent greater re-
construction work at the Frankfurt central station. In tightening tests these bolts
behave different from normal HV-bolts as you will see in the document.

INTRODUCTION

More and more ancient steel constructions come under control because of their further ser-
viceability for the origin purpose. I name some of them here such as the big four bay hall of
the Frankfurt main station (Figure 1), the equivalent buildings in Cologne and in Hamburg-
Dammtor as well as in Liverpool many years ago. Very often the experts come to the conclu-
sion that the building, especially the steel construction, needs a complete restoration since
corrosion has done its work over many years and nobody took any action to start earlier with
the restoration. Not seldom the discussion ends with the result, that at least a replacing of
more or less parts of the construction is necessary. Before such actions start usually the au-
thority for the preservation of historic monuments (in Germany the Denkmalschutzbehörde)
play an important role about the aspect of such buildings after the repair. And they put their
recommendations and requirements on the table, which generally say:

After the reconstruction the steel construction must look the same way as before.

This means that the design engineers cannot work as in the case of a new building but they
must deal with historic methods of calculation, fabrication, erection and material. Often old
books from former times help, but sometimes special skills are gone and nobody has any
idea how to recover them.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 311


Figure 1. Frankfurt/Main Central station, restoration of the 4 big halls under full traffic.

Such an ancient skill p. e. is the hot riveting of steel constructions on site. It needs special
tools and a clever group of craftsmen on one hand and it is much to expensive today as
every handwork is. If the main construction parts of these halls such as the main girders are
riveted spatial lattice truss girders, arches or multi bay arches the construction firm cannot
but produce similar looking members either in the workshop for transport or on the site. The
individual members have to be connected on site with simple and cheap rivet similar looking
fasteners. The authority for the preservation of historic monuments (Denkmalschutzbehörde)
will not at all give any permission to change something and therefore the bolt manufacturers
have to have new ideas to make a bolt looking like a rivet, make it high strength, make every
wanted diameter and not only the bolt but also the nut, even if one of them, bolt head or nut,
is positioned inside a spatial truss girder which cannot easily be seen.

DEVELOPMENT OF SPECIAL BOLTS, NUTS AND WASHERS

The well known German bolt manufacturer August Friedberg in Gelsenkirchen developed
such a type of bolt, which should be installed in the steel construction building in Frank-
furt/Main central railway station. The bolts have a half round head with a very small but high
enough hexagon part on it next to the washer. The nut was formed similarly. The length had
to be adjusted so that when tightened the bolt end fits into the half round of the nut. All bolts,
nuts and washers should be hot dip galvanized.

Since the dimensions of those bolts and nuts differ in a certain way from the normal dimen-
sions of high strength bolts the building authority required an independent expertise about
their installation and performance. This expertise was not possible without a test series on
M16 and M20 bolts. We have been asked to perform such test series and evaluate all nec-
essary tightening and working data.

The many many discussions with other expert colleagues in the international group ECCS
TC 10 Connections and in the CEN/TC 185/WG 6 - most of these colleagues I see in the
auditorium - and a lot of similar tests in cooperation with friends of the „bolt experts family“
and literature (1, 2, 3, 4) gave us the capability to do the tests also in a very short time.

312 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


We proposed tightening tests with tightening from the nut side and from the bolts head side
because both procedures were envisaged for erection.

The Figures 2, 3 and 4 show drawings of the bolts, nuts and special distance washers.

Figure 2. Half round head HV bolt.

Figure 3. Half round HV nut. Figure 4. Special Distance washer to adjust


the bolt length, h = 2, 3 and 4 mm.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 313


TEST PROGRAMME

We performed the following 4 test programmes tightening either the nuts or the bolt head

Test series no. 1:


15 AF-Half round head HV-bolts M20x100, tzn, grade 10.9
similar to DIN 6914, tightening by the nut.
Half round HV-nut M20 similar to DIN 6915, tzn, grade 10
HV-washer 21 according to DIN 6916, tzn, grade 10.9, under the bolt head
HV-washer 21 flat, similar to DIN 6916, tzn, grade 10.9,
but only 2 and 3 mm thick, without chamfer under the nut.

Test series no. 2:


6 AF-Half round head HV-bolts M20x100, tzn, grade 10.9
similar to DIN 6914, tightening by the bolt head.
Half round HV-nut M20 similar to DIN 6915, tzn, grade 10
HV-washer 21 according to DIN 6916, tzn, grade 10.9, under the bolt head
HV-washer 21 flat, similar to DIN 6916, tzn, grade 10.9,
but only 2 and 3 mm thick, without chamfer under the nut.

Test series no. 3:


15 AF-Half round head HV-bolts M16x100, tzn, grade 10.9
similar to DIN 6914, tightening by the nut.
Half round HV-nut M16 similar to DIN 6915, tzn, grade 10
HV-washer 17 according to DIN 6916, tzn, grade 10.9, under the bolt head
HV-washer 17 flat, similar to DIN 6916, tzn, grade 10.9,
but only 2 and 3 mm thick, without chamfer under the nut.

Test series no. 4:


6 AF-Half round head HV-bolts M16x100, tzn, grade 10.9
similar to DIN 6914, tightening by the bolt head.
Half round HV-nut M16 similar to DIN 6915, tzn, grade 10
HV-washer 17 according to DIN 6916, tzn, grade 10.9, under the bolt head
HV-washer 17 flat, similar to DIN 6916, tzn, grade 10.9,
but only 2 and 3 mm thick, without chamfer under the nut.

The continuous tightening was performed with motor and gearbox, so that the speed was
about 6 rounds per minute. While tightening the following measurement were taken stepwise
all 15° of rotation:

1. Preload (FV) and


2. Torque (MA).

The distance between the two measurements was always 15° rotation. They have been reg-
istered by the computer and immediately after the individual test presented on the screen.
Figure 5 shows such a type of test diagram for the rotation(ϕ)-preload(FV)-behaviour.

TEST DATA EVALUATION

We have collected all three important test data as rotation(ϕ), torque(MA) and preload(FV).
They can be drawn one against the other in diagrams. We produced three diagrams of each
test data set such as

314 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


1. rotation (ϕ) versus preload(FV),
2. torque (MA) versus preload(FV) and
3. rotation (ϕ) versus torque(MA).

Type of quick test diagram from measurements

150

100
preload [kN]

50

0
0 60 120 180 240 300 360 420 480 540 600
rotation [°]

Figure 5. Rotation (ϕ)-preload (FV)-diagram.

From these diagrams all required proof data, which are specified in the prEN 14399-1, and
prEN 14399-2 (Fitness for purpose documents) can be read.

Results of the test series no. 1: 15 M20 assemblies, nut tightening

15 AF-Half round head HV-bolts M20x100, tzn, grade 10.9


similar to DIN 6914, tightening by the nut

250

HV-Schrauben
200 nach DIN 18800 T7:
FV = 160 kN

150
preload [kN]

100 10,9-Halbrundkopf-HV-
Schrauben
red.FV = 120 kN

50

0
0 60 120 180 240 300 360 420 480 540 600

rotation [°]

Figure 6. Rotation (ϕ)-preload (Fv)-diagram for 15 bolts M20, nut tightening.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 315


15 AF-Half round head HV-bolts M20x100, tzn, grade 10.9
similar to DIN 6914, tightening by the nut

250
MA = 450 Nm = required torque HV M 20
HV-Schrauben
200 acc. DIN 18800 T7:
FV = 160 kN
preload [kN]

150

100 10.9-HRH-HV-Schrauben Scatter ∆MA = 116 Nm


redFV = 120 kN for reduced preload red.FV = 120 kN

50

min MA = 346 Nm max MA = 462 Nm

0
0 100 200 300 400 500 600 700 800 900 1000
torque [Nm]

Figure 7. Torque (MA)-preload (FV)-diagram for 15 bolts M20, nut tightening.

15 AF-Half round head HV-bolts M 20x100, tzn, grade 10.9


similar to DIN 6914, tightening by the nut
1000

900

800

700
10,9-Half-round-head HV bolt
600
torque [Nm]

M A = 500 Nm
500

400
HV-bolts
acc. DIN 18800 T7:
300
M A = 450 Nm
200

100

0
0 60 120 180 240 300 360 420 480 540 600
rotation [°]

Figure 8. Rotation (ϕ) versus torque (MA)-diagram for 15 bolts M20, nut tightening.

316 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Results of the test series no. 2: 6 M20 assemblies, bolt head tightening

6 AF-Half round head HV-bolts M20x100, tzn, grade 10.9


similar to DIN 6914, tightening by the bolt head
250

HV-bolts
200 acc. DIN 18800 T7:
FV = 160 kN

150
preload [kN]

100
10.9-Half round head HV bolts
red.FV = 100 kN

50

0
0 60 120 180 240 300 360 420 480 540 600

rotation [°]

Figure 9. Rotation (ϕ)-preload (Fv)-diagram for 6 bolts M20, head tightening.

6 AF-Half round head HV-bolts M20x100, tzn, grade 10.9


similar to DIN 6914, tightening by the bolt head

250
M A = 450 Nm = required torque for HV M 20
HV-bolts
200 acc. DIN 18800 T7:
F V = 160 kN
preload [kN]

150

100
10.9-Half round head HV Scatter ∆M A = 71 Nm
bolts, redF V = 100 kN for reduced preload red.F V = 100 kN

50

min M A = 594 Nm max M A = 665 Nm

0
0 200 400 600 800 1000 1200
torque [Nm]

Figure 10. Torque (MA)-preload (FV)-diagram for 6 bolts M20, head tightening.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 317


6 AF-Half round head HV-bolts M20x100, tzn, grade 10.9
similar to DIN 6914, tightening by the bolt head

1300
1200
1100
1000
900 10.9-Half round head HV bolt
800 MA = 700 Nm
torque [Nm]

700
600
500
400
HV-bolts
300 acc. DIN 18800 T7:
200 MA = 450 Nm
100
0
0 60 120 180 240 300 360 420 480 540 600
rotation [°]

Figure 11. Rotation (ϕ) – torque (MA)-diagram for 6 bolts M20, head tightening.

Results of the test series no. 3: 15 M16 assemblies, nut tightening

15 AF-Half round head HV-bolts M16x100, tzn, grade 10.9


similar to DIN 6914, tightening by the nut

150
HV-bolts
acc. DIN 18800 T 7:
FV = 100 kN

100
preload [kN]

AF-Half round head HV bolts


M16x100:
50
red.FV = 75 kN

0
0 60 120 180 240 300 360 420 480 540 600
rotation [°]

Figure 12. Rotation(ϕ)-preload(Fv)-diagram for 15 bolts M16, nut tightening.

318 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


15 AF-Half round head HV-bolts M16x100, tzn, grade 10.9
similar to DIN 6914, tightening by the nut

150

HV-bolts
acc. DIN 18800 T 7:
FV = 100 kN
AF-Half round head HV
100 bolts M16x100:
red.FV = 75 kN
preload [kN]

Scatter ∆MA = 75 Nm
for red. preload red.FV = 75 kN
50

min MA = 180 Nm max MA = 255 Nm

0
0 50 100 150 200 250 300 350 400 450 500 550 600
torque [Nm]

Figure 13. Torque(MA)-preload(Fv)-diagram for 15 bolts M16, nut tightening.

15 AF-Half round head HV-bolts M16x100, tzn, grade 10.9


similar to DIN 6914, tightening by the nut

600
550
500
450
400 AF-Half round head HV
bolts, M16x100
torque [Nm]

350
MA = 275 Nm
300
250
200 HV-bolts,
acc. DIN 18800 T7:
150
MA = 250 Nm
100
50
0
0 60 120 180 240 300 360 420 480 540 600
rotation [°]

Figure 14. Rotation(ϕ)-torque(MA)-diagram for 15 bolts M16, nut tightening.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 319


Results of the test series no. 4: 6 M16 assemblies, bolt head tightening

6 AF-Half round head HV-bolts M16x100, tzn, grade 10.9


similar to DIN 6914, tightening by the bolt head
150

HV-bolts
acc. DIN 18800 T 7:
FV = 100 kN

100
preload [kN]

50 10,9-Half round head HV-bolts


red.FV = 60 kN

0
0 60 120 180 240 300 360 420 480 540 600
rotation [°]

Figure 15. Rotation(ϕ)-preload(Fv)-diagram for 6 bolts M16, head tightening.

6 AF-Half round head HV-bolts M16x100, tzn, grade 10.9


similar to DIN 6914, tightening by the bolt head

150 MA = 250 Nm

HV-bolts
acc. DIN 18800 T 7:
FV = 100 kN

100
preload [kN]

10,9-Half round head HV-bolts


red.FV = 60 kN

50 Scatter ∆MA = 112 Nm


for a reduced preload red.FV = 60 kN

min MA = 223 Nm max MA = 335 Nm

0
0 50 100 150 200 250 300 350 400 450 500 550 600 650 700
torque [Nm]

Figure 16. Torque(MA)-preload(Fv)-diagram for 6 bolts M16, head tightening.

320 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


6 AF-Half round head HV-bolts M16x100, tzn, grade 10.9
similar to DIN 6914, tightening by the bolt head

700

600

500
torque [Nm]

400

HV-bolts, acc. DIN 18800 T7:


300 M A = 250 Nm

200

10,9-Half round head HV-bolts


100 red.F V = 60 kN

0
0 60 120 180 240 300 360 420 480 540 600

rotation [°]

Figure 17. Rotation(ϕ)-torque(MA)-diagram for 6 bolts M16, head tightening.

Evaluation of the test results

In the diagrams Figures 6 to 17 horizontal and vertical full lines mark the required data for the
preloads and torques as they are mentioned in the German standard, while the dashed lines
show the results for half round head HV-bolts and nuts.

The bolts were always put from the outside into the hole, the nut was inside the spatial truss
girder. In those cases, where the full preload was necessary tightening from the inside by the
nut has been proposed with the result: 120 kN for M20 bolts and 75 kN for M16 bolts. In
those cases where only a partial preload was necessary such as in bearing type connections
the tightening could be managed by bolt head rotating from the outside. Since in those cases
no further lubrication has been used the scatter in preload was much greater than usual, the
recommended torque had to be higher and the reliable preload was smaller: 90 kN for M20
bolts and 60 kN for M16 bolts.

The reason why the full preload of normal high strength bolts could not be reaches, was the
fact that the size of the nut is much different, the nut has less material to withstand the tan-
gential tension stress with very small deformation (see Figure 3).

CONCLUSION

Half round head HV-bolts and nuts can be used instead of rivets for the restoration of old his-
toric steel constructions to preserve them as an ancient building monument. The carrying ca-
pacity is somewhat smaller than known from usual high strength bolts. I would like to rec-
ommend to perform more tests to get a good knowledge about the special behaviour of such
bolts in erection and in service.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 321


REFERENCES

(1) Joint Program: Fitness for Purpose Tests of High Strength Bolt Assemblies with High
Strength Nuts according to prEN 783. Report des Arbeitsbereichs Stahlbau und Holz-
bau der Technischen Universität Hamburg-Harburg, Hamburg 1994.
(2) Report no. 1: International Tightening Tests with High Strength Preloaded Bolts of the
Quality 10.9 according to the System HR. Common Report of the Research Centre of
the Belgian Metalworking Industries, Liège, and the Department of Stahlbau und Holz-
bau der Technischen Universität Hamburg-Harburg, Hamburg 1997.
(3) Report no. 2: International Tightening Tests with High Strength Preloaded Bolts of the
Quality 10.9 according to DIN 6914 – 6916. Common Report of the Research Centre of
the Belgian Metalworking Industries, Liège, and the Department of Stahlbau und Holz-
bau der Technischen Universität Hamburg-Harburg, Hamburg 1996.
(4) Valtinat, G., E. Piraprez and E. Greff: International Tightening Tests with High Strength
Bolts M20x100 of the Systems HR and HV. Heft 4 der Schriftenreihe Stahlbau und
Holzbau der Technischen Universität Hamburg-Harburg, Hamburg 1998.

322 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


A UNIFIED APPROACH TO DESIGN FOR BLOCK SHEAR

R.G. Driver, G.Y. Grondin, and G.L. Kulak


Dept. of Civil & Environmental Engineering, University of Alberta, Canada

ABSTRACT

Design equations in Canadian, American, and European standards for


determining the block shear capacity of structural members vary significantly.
A detailed survey of test results published in the literature, combined with the
experimental results of two recent research projects at the University of
Alberta, demonstrates that these equations do not always provide an
acceptable level of safety. For coped beams, in particular, it was found that
the North American standards do not provide an acceptable level of safety
and that predictions of the strength of such connections with two lines of bolts
are often on the unsafe side by a considerable margin.
Reliability methods are used to provide a simple, unified approach to design
for block shear failure of gusset plates, angles, tees, and coped beams. This
method provides both an adequate and a consistent level of safety and
predicts capacities that compare well with test results over a broad array of
connection configurations. Of particular importance, the proposed approach
also reflects the failure modes that have been observed in tests, which is an
aspect that is lacking in some current design equations.

INTRODUCTION

Block shear failure is characterized by a combination of rupture on the tension plane and
yielding on the shear plane(s) of a block of material. Although this mode of failure can occur
in either welded or bolted connections, it is more common in the latter because of the
reduced area that results from the bolt holes. This paper deals only with bolted connections.

Design rules in various codes and standards base block shear capacity on a combination of
yield and rupture strength of the net or gross areas in shear and tension on the failure
planes. North American design provisions are examined in order to assess the level of safety
provided by each design standard for gusset plates, angles, tees, and coped beams. The
Eurocode 3 design equations are also examined as one alternative to current North
American practice. Finally, a unified approach that provides a uniform level of safety for all of
these cases is proposed.

DESIGN PROVISIONS

CSA–S16–01

In the current Canadian design standard, CSA–S16–01 (1), the block shear capacity of
gusset plates and angles is taken as the lesser of the following two equations:
Pu = A nt Fu + 0.6 A gv Fy (1)

Pu = A nt Fu + 0.6 A nv Fu (2)

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 323


The capacity of coped beams is provided by similar equations except that the contribution of
the tension area to the connection capacity is reduced by one-half to reflect the non-uniform
stress distribution on the tension plane. Thus, for coped beams:
Pu = 0.5 A nt Fu + 0.6 A gv Fy (3)

Pu = 0.5 A nt Fu + 0.6 A nv Fu (4)

where Fy and Fu are the yield and the tensile strength of the material, respectively, A nt and
A nv are the net tension and shear areas, respectively, and A gt and A gv are the gross
tension and shear areas, respectively. Equations (1) and (3) are based on the observation
that rupture on the tension plane occurs before rupture on the shear planes. This is
supported by several test programs, including the University of Alberta test observations
presented below. Equations (2) and (4) are provided in order to limit the capacity of the shear
planes to the rupture strength of the net shear area.

AISC 1999

The block shear provisions in the AISC LRFD 1999 Specification (2) make use of two
equations that depend on the relative strength of the net tension and shear areas of the
connection:
when Fu A nt ≥ 0.6 Fu A nv , Pu = Fu A nt + 0.6 Fy A gv ≤ Fu A nt + 0.6 Fu A nv (5)

when Fu A nt < 0.6 Fu A nv , Pu = Fy A gt + 0.6 Fu A nv ≤ Fu A nt + 0.6 Fu A nv (6)

where all the variables are as defined above. The block shear capacity combines either
rupture on the net tension area, A nt , with yielding on the gross shear area, A gv , or yielding
on the gross tension area, A gt , with rupture on the net shear area, A nv . Both equations
have an upper bound corresponding to the combination of rupture on both the net tension
and net shear areas. Although equation (5) represents a plausible failure mode, the validity
of equation (6) is questionable (3). Although not yet adopted, the current proposal for the
2005 AISC Specification is to use the S16-01 equations (equations (1) through (4)), except
that the 0.5 coefficient for the tension plane is used for angles as well as for coped beams.

Eurocode

Eurocode 3 (4) combines shear yield acting over the gross shear area with rupture on the net
tension plane. Thus, the block shear capacity of a gusset plate or tension member can be
obtained as follows:

Pu = Fu A nt + Fy ( )
3 A gv (7)

(Eurocode uses the von Mises shear yield stress equal to Fy 3 , whereas the North
American standards simply take this as 0.60 Fy .)

Although equation (7) is not given specifically in Eurocode, it can be deduced from the
description of the mode of failure. A more elaborate procedure is presented for block shear
failure in beam webs, where a series of equations, when combined, gives the following
result:
1 1
Pu = w (L gt − k dh ) Fu + A gv F (8)
3 3 y

where w is the web thickness, L gt is the gross tension length, k is a tension area coefficient
(0.5 for one line of bolts or 2.5 for two), and dh is the bolt hole diameter.

324 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


GUSSET PLATE TESTS

There are a large number of gusset plate tests reported in the literature for which block shear
is the failure mode. Table 1 presents a summary of the results for 133 gusset plate tests from
eight different sources.
All gusset plate tests show that the ultimate load is reached when the tensile ductility of the
gusset plate material at the first (i.e., inner) transverse row of bolts is exhausted. This is true
even in cases where oversize holes were used and in cases where the connection was short
(i.e., little shear area available). Recent tests conducted at the University of Alberta (5) show
clearly that fracture on the net tension plane takes place before shear fracture on the shear
planes, as illustrated in Figure 1 on the left. The displacement of a block of material is seen
only when the test is continued until the parts separate (Figure 1, right), and this occurs well
past the ultimate load capacity of the connection. Figure 1 also illustrates that shear rupture
does not take place at the net area, as assumed in some design equations.

Tension fracture

Figure 1. Block shear failure in gusset plates (5).

Although many tests have been conducted on gusset plates, the range of geometries tested
in the past is fairly limited. Most tests have been conducted on gusset plates with two lines of
bolts with from two to four bolts in each line. Huns et al. (5) have extended the number of
bolts in a line to eight using finite element analysis. The 10 tests reported in Table 1 from
reference (5) include five results from a finite element model that has been validated against
existing test results.

The test-to-predicted ratios for all 133 gusset plates presented in Table 1 average 1.18 for
CSA–S16–01, 1.19 for AISC LRFD, and 1.10 for Eurocode 3. Thus, all design standards
provide a conservative estimate of the block shear capacity of gusset plates.

ANGLE AND TEE TESTS

Experience and test results show that block shear is potentially a failure mode for angles and
tees, particularly when the connection is short. Tests of 41 such members (including nine
structural tees from Orbison et al. (16) connected at the stem that are considered to give the
same failure mode as angles) are shown in Table 1. Although Epstein (14) reported the
results of 144 tests, only the three specimens that failed in block shear with the tension face
perpendicular to the load direction are included in Table 1. Other test programs have also

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 325


investigated block shear in single angle connections and structural tees (14, 15, 16) with one
line of bolts. Block shear failure of angles and tees can be affected by both out-of-plane and
in-plane eccentricity. Tests on structural tees have been used to assess the effect of out-of-
plane eccentricity (16). Although Orbison et al. (16) found that out-of-plane eccentricity was
not a significant factor, some tests (15, 16) have shown that the block shear capacity
decreases with an increase in eccentricity. In-plane eccentricity is present in all of the angle
and tee tests reported in Table 1. Table 1 indicates that the block shear capacity is predicted
reasonably well by the North American and European design standards.

Table 1. Block shear test results.


Test Test Test Test
Source No. Tests
S16-01 AISC EURO Eq. (13)

Gusset Plates
1.20 1.22 1.19 1.05
Hardash and Bjorhovde (6) 28
(0.07) (0.08) (0.06) (0.06)
1.22 1.22 1.05 0.96
Rabinovitch and Cheng (7) 5
(0.06) (0.06) (0.05) (0.05)
1.18 1.18 1.06 0.96
Udagawa and Yamada (8) 73
(0.05) (0.05) (0.09) (0.06)
1.35 1.35 1.07 1.00
Nast et al. (9) 3
(0.02) (0.02) (0.01) (0.01)
1.21 1.21 1.05 0.95
Aalberg and Larsen (10) 8
(0.04) (0.04) (0.18) (0.10)
Swanson and Leon (11) 1 1.18 1.18 0.99 0.88
1.14 1.13 1.10 0.98
Huns et al. (5) 10
(0.17) (0.19) (0.14) (0.13)
1.13 1.15 1.16 1.02
Mullin (12) 5
(0.04) (0.05) (0.03) (0.05)
Angles and Tees
1.00 1.01 0.89 0.87
Barthel et al. (13) 13
(0.04) (0.05) (0.04) (0.04)
Epstein (14) [only mean reported] 3 0.98 0.98 1.00 1.02
0.97 0.96 0.93 0.92
Gross et al. (15) 13
(0.05) (0.05) (0.07) (0.06)
1.09 1.12 1.09 1.08
Orbison et al. (16) [includes 9 tees] 12
(0.05) (0.07) (0.05) (0.05)
Coped Beams
Birkemoe and Gilmor (17) 1 1.17 0.95 1.18 0.90
1.22 1.03 1.25 0.96
Yura et al. (18) 3
(0.16) (0.13) (0.16) (0.13)
1.02 0.70 1.13 1.00
Ricles and Yura (19) 7
(0.10) (0.08) (0.06) (0.12)
1.33 1.13 1.21 1.01
Aalberg and Larsen (10) 8
(0.12) (0.08) (0.23) (0.11)
1.20 1.07 1.21 1.02
Franchuk et al. (20, 21) 17
(0.10) (0.13) (0.09) (0.12)
Note : The number in parentheses is the standard deviation.

326 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


COPED BEAM TESTS

In contrast to the number of gusset plate test results available, there are relatively few tests
of coped beams. Of the 36 tests shown in Table 1, ten had connections with two lines of
bolts and 26 had a single line. A recent test program at the University of Alberta (20, 21)
provided nearly one-half of the available test results and expanded the number of test
parameters by including several variables related to connection geometry and loading
conditions. Seventeen full scale tests were conducted on coped beams and included the
effect of beam end rotation, gross shear area, bolt end and edge distances, number of bolt
lines and bolt rows, bolt diameter, section depth, connection depth, and double copes vs.
single cope. Analysis of the load vs. deformation results showed that none of these
parameters affected connection capacity significantly, apart from the associated changes in
tension and shear areas. It was concluded that only the tension and shear areas need be
considered in a design equation to reasonably represent the failure mode. The test results
indicated that rupture on the tension plane occurs before shear rupture, as shown in Figure 2
(left). It was also evident from the tests that shear rupture does not take place on the net
shear area. As shown in Figure 2 (right), shear failure takes place on a plane that intercepts
the holes very near to their edges.

Tension
fracture

Figure 2. Block shear failure in coped beams (20, 21).

The ratio of test ultimate load to the AISC predicted ultimate load is significantly non-
conservative for the tests by Ricles and Yura (19), which were all tests on coped beams with
two lines of bolts. The average AISC test-to-predicted ratio for the three test specimens of
Franchuk et al. (20, 21) with two lines of bolts is 0.90, compared to 1.10 for the test
specimens with one line. Both the Canadian and European standards seem to predict the
test results conservatively, although a large standard deviation is observed in some cases.

RELIABILITY ANALYSIS

Although the test-to-predicted ratios presented in Table 1 provide an indication of the ability
of each equation to predict block shear capacity, they do not by themselves quantify the level
of safety being provided. Therefore, the suitability of the various design equations was
assessed using reliability methods. Since the equations presented above do not contain a
performance factor, the following analysis derives the performance factors required to
achieve pre-selected levels of safety. Because the desired level of safety is a matter of

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 327


continuing debate, the calculations are performed for three different safety levels.

In the development of a limit states design equation, a performance factor, φ, is determined


that is a function of the mean values and variabilities—expressed as the coefficients of
variation—of the relevant material and geometric properties, as well as the ability of the
equation itself to predict capacity. The performance factor is calculated using the following
equations from Ravindra and Galambos (22):
φ = ρR exp ( −αR β VR ) (9)

where β is the safety index, which is directly related to the probability of failure during the life
of the structure. It is desirable to have a higher safety index (lower probability of failure) for
connections than for structural members, such as beams, that tend to be more ductile. As
such, members are usually assigned a safety index of about 3.0, while connections have
typically been assigned a value of approximately 4.5 (22). The bias coefficient for resistance,
ρR , is given as:
ρR = ρM ρG ρP (10)

where ρM is the ratio of the mean measured to nominal material strength, ρG is the ratio of
mean measured to nominal connection geometric properties, and ρP is the professional
factor, or mean ratio of measured to predicted resistance, which reflects the ability of the
equation to predict the capacity. The coefficient of variation for the resistance, VR , is given
as:

VR = VM2 + VG2 + VP2 (11)

where VM , VG , and VP are the coefficients of variation associated with ρM , ρG , and ρP ,


respectively. Ravindra and Galambos (22) recommend that the separation variable, αR , in
equation (9) be taken as 0.55.

There is an interdependence between φ and the load factors, shown by Fisher et al. (23).
This means that the use of a safety index other than 3.0 in equation (9) requires that a
modification factor be applied. As shown in (20, 24), equation (9) must be modified to:

φ = ⎡0.0062 β2 − 0.131 β + 1.338 ⎤ ρR exp ( −αR β VR ) (12)


⎣ ⎦

Material Factor

The material factor reflects the difference between the nominal and measured material
strengths (yield or ultimate). The bias coefficient and coefficient of variation for the material
properties were obtained using data from Schmidt and Bartlett (25) for plates and rolled wide
flange sections. The material factors selected for gusset plates and angles and tees are
presented in Table 2. The values selected for coped beams are presented in Table 3.

Geometry Factor

The geometry factor accounts for the difference between the nominal and measured plate
thickness and hole layout dimensions. Since insufficient data exist to evaluate the geometry
factors for gusset plates, angles and tees, the bias coefficient and coefficient of variation for
the geometry factor proposed by Hardash and Bjorhovde (6) was used. The values selected
for coped beams were obtained from measurements on the 17 specimens tested by
Franchuk et al. (20, 21, 24) and data reported by Kennedy and Gad Aly (26). The values of
ρG and VG adopted are presented in Tables 2 and 3.

328 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Professional Factor

The professional factor (the ratio of the measured capacity, obtained either by laboratory
testing or from a validated finite element analysis, to the capacity predicted by the equation
using measured material properties and geometry) represents the ability of a model to
predict the block shear capacity. The mean and coefficient of variation of the test-to-
predicted ratios for the individual test programs presented in Table 1 have been consolidated
for gusset plates, angles and tees, and coped beams and are summarized in Tables 2 and 3.
Although the North American standards require the inclusion of a 2 mm allowance when
punched holes are used, it was not included in the professional factors presented in this
work. This is because in many cases it is not known how the holes were made. The effect of
this approximation was assessed for coped beams by Franchuk et al. (24) and found to be
relatively small. The performance factor, φ , was calculated for three levels of safety index,
namely, 3.5, 4.0, and 4.5.

Table 2. Reliability parameters for gusset plates and angles and tees.

Table 3. Reliability parameters for coped beams.

Using the performance factor adopted by each standard (0.9 or 0.75 for the Canadian and
American standards, respectively) the current safety indices for each type of connection can
be determined. Use of the Canadian standard gives a safety index of 4.5 for gusset plates,
3.4 for angles and tees, and 4.1 and 3.0 for one- and two-line coped beam connections. Use

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 329


of the AISC specification gives a safety index of 5.8 for gusset plates, 4.6 for angles and
tees, and 4.4 and 1.9 for one- and two-line coped beam connections. Both standards provide
an inconsistent level of safety. The level of safety in the Canadian standard is too low for
angles and tees and for two-line coped beam connections. The level of safety provided by
the AISC specification is too low for two-line coped beam connections. Both North American
standards provide an unacceptably low level of safety for the two-line connections. The
Eurocode approach, however, provides a consistent level of safety for one- and two-line
connections, although there is inconsistency when including gusset plates, angles, and tees,
as can be deduced from Tables 2 and 3. For comparison, the equations currently proposed
for the 2005 AISC Specification, with the associated performance factor of 0.75, give a safety
index of 5.9 for gusset plates, 5.5 for angles and tees, and 5.3 and 4.1 for one- and two-line
coped beam connections. Had the punched hole allowance been included, the safety indices
could be as much as 7% higher than the values presented above (24).

PROPOSED UNIFIED EQUATION

In order to address the shortcomings and inconsistencies of existing design equations, a new
equation for the design of gusset plates, angles, tees, and coped beams is proposed. It uses
laboratory observations of the failure mode as its basis. A single equation is considered
appropriate because the failure mode observed over a large variety of tests consists of
rupture on the tension plane after yielding has occurred on the shear plane, but prior to shear
rupture. Although most of the existing design equations predict the test capacity for angles
reasonably well, the capacity of gusset plates is generally under-predicted by a considerable
margin, and none of the equations adequately predict the block shear capacity of all coped
beams. It is also desirable to have a single equation rather than multiple equations, as is
currently used in the design standards. Although the stress on the shear plane is lower than
the rupture stress when rupture on the tension plane occurs, it is likely significantly greater
than the yield stress. Simply taking the mean of the yield and ultimate shear stresses
provides excellent correlation with test results. Another issue that must be taken into account
is that shear failure has been observed to occur on a plane that intercepts the holes very
near to their edges (see Figures 1 and 2). Therefore, the gross area is deemed to be more
appropriate for determining the strength of the shear components of the block.

The proposed unified equation combines effective stresses on both the net tension area and
the gross shear area:
⎛ Fy + Fu ⎞
Pu = R t Ant Fu + R v A gv ⎜ ⎟ (13)
⎜ 2 3 ⎟
⎝ ⎠
where R t and R v are the tension area and shear area stress correction factors, respectively,
as given in Table 4. These factors account for the non-uniform stress distributions that occur
in some cases on the tension and shear planes of the block shear failure surface. These
factors are semi-empirical in that they have been selected to optimize the consistency of the
safety indices, although they have support from finite element studies (20).

The proposed model is simplified over those of the existing North American and European
standards in that only one equation need be checked for a particular connection type.
Nevertheless, it not only provides an accurate prediction of connection capacity in an
average sense, with mean professional factors varying from 0.96 to 1.07 for all types of
connections considered, it also results in coefficients of variation that are among the lowest
of all procedures investigated. These values are tabulated in Tables 2 and 3. The excellent
correlation is attributed to the consistency of the structure of the proposed unified equation
with experimental observations.

Tables 2 and 3 give the performance factors, φ, required to achieve safety indices of 3.5, 4.0,

330 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


and 4.5. The tables indicate that a consistent performance factor is achieved with the unified
equation for gusset plates, angles, tees, and coped beams with one or two lines of bolts. The
performance factor varies from about 0.8 to 0.7 to achieve safety indices from 3.5 to 4.5 for
all connection types considered. As a point of reference, the traditional target safety index for
connections of 4.5 is achieved for all cases by adopting a performance factor of 0.67,
although it can legitimately be argued that a somewhat lower safety index would be sufficient
for this type of failure.

Table 4. Stress correction factors.


Rt Rv
Gusset plates 1.0 1.0
Angles and tees 0.9 0.9
One-line connections 0.9 1.0
Coped Beams
Two-line connections 0.3 1.0

SUMMARY AND CONCLUSIONS

A comprehensive research program was carried out to evaluate the level of safety currently
being provided by design standards for determining the block shear capacity of gusset
plates, angles, tees, and coped beams. A reliability analysis indicates that the current North
American standards do not provide an adequate level of safety for all coped beam
connections, as compared to the traditional target safety index for connections of 4.5. Both
standards provide a sufficient level of safety for gusset plates and coped beams with one line
of bolts. For angles and tees and for coped beam connections with two lines of bolts, the
level of safety in the Canadian standard is unacceptable. The safety index provided by the
1999 AISC Specification for coped beams with two lines of bolts is far below the target value
of 4.5 for connections and even lies well below the target for ductile member failure of 3.0.
On the other hand, the equations currently proposed for the 2005 AISC Specification result in
safety indices very much higher than 4.5 for most common connections, while the value for
coped beams with two lines of bolts is considerably lower. The Eurocode equation provides
an inconsistent level of safety when considering various connection types.

The proposed unified design model (Equation 13) is recommended for the prediction of block
shear capacity. It is simpler than existing procedures and reflects the failure mode observed
consistently in a wide variety of tests. Moreover, with a single performance factor the model
provides a consistent and adequate level of safety for all types of connections investigated in
this study.

ACKNOWLEDGEMENT

The research conducted at the University of Alberta was funded by the the Steel Structures
Education Foundation, the Natural Sciences and Engineering Research Council of Canada,
and the C.W. Carry Chair in Steel Structures Research.

REFERENCES

1. Canadian Standards Association (CSA). (2001). “Limit States Design of Steel


Structures,” CSA–S16–01, Toronto, ON, Canada.
2. American Institute of Steel Construction (AISC). (1999). "Load and Resistance Factor

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 331


Design Specification for Structural Steel Buildings," Chicago, IL, USA.
3. Kulak, G.L., and Grondin, G.Y. (2001). "Block Shear Failure in Steel Members – A
Review of Design Practice," Engineering Journal, AISC, 38(4), 199-203.
4. European Committee for Standardization (ECS) (1992). “Eurocode 3: Design of Steel
Structures,” ENV 1993–1–1, Brussels, Belgium.
5. Huns, B.B.S., G.Y. Grondin, and R.G. Driver (2002). "Block Shear Behaviour of Bolted
Gusset Plates," Structural Engineering Report No. 248, Department of Civil
Engineering, University of Alberta, Edmonton, Alberta.
6. Hardash, S.G., and R. Bjorhovde, (1984). "Gusset Plate Design Utilizing Block-Shear
Concepts," The Department of Civil Engineering and Engineering Mechanics, The
University of Arizona, Tucson, Arizona.
7. Rabinovitch, J.S., and Cheng J.J.R. (1993). "Cyclic Behaviour of Steel Gusset Plate
Connections," Structural Engineering Report 191, Department of Civil Engineering,
University of Alberta, Edmonton, Alberta.
8. Udagawa, K. and Yamada, T. (1998). "Failure Modes and Ultimate Tensile Strength of
Steel Plates Jointed with High-Strength Bolts," J. of Structural and Construction
Engineering, Architectural Institute of Japan, No. 505, pp. 115-122.
9. Nast, T.E., Grondin, G.Y., and Cheng, J.J.R. (1999). "Cyclic Behaviour of Stiffened
Gusset Plate-Brace Member Assemblies," Structural Engineering Report 229,
Department of Civil and Environmental Engineering, University of Alberta, Edmonton,
Alberta.
10. Aalberg, A. and Larsen, P.K. (1999). “Strength and Ductility of Bolted Connections in
Normal and High Strength Steels,” Department of Structural Engineering, Norwegian
University of Science and Technology, N-7034, Trondheim, Norway.
11. Swanson, J.A. and R.T. Leon (2000). "Bolted Steel Connections: Tests on T-Stub
Components," J. of Structural Engineering, ASCE, Vol. 126, No. 1., pp. 50-56.
12. Mullin, D. (2002). Unpublished test data. Department of Civil & Environmental
Engineering, University of Alberta, Edmonton, Alberta.
13. Barthel, R.J., Peabody, M.T., and Cash, U.M. (1987). "Structural Steel Transmission
Tower Angle and Rectangular Coupon Tension Testing," Western Area Power
Administration Report, Department of Energy.
14. Epstein, H.I. (1992). "An Experimental Study of Block Shear Failure of Angles in
Tension," Engineering Journal, AISC, 29(2), pp. 75-84.
15. Gross, J.M., Orbison, J.G., and Ziemian, R.D. (1995). "Block Shear Tests in High-
Strength Steel Angles," Engineering Journal, AISC, 32(3).
16. Orbison, J.G., Wagner, M.E., and Fritz, W.P. (1999). “Tension Plane Behavior in
Single-Row Bolted Connections Subject to Block Shear,” J. of Constructional Steel
Research, Vol. 49, pp. 225-239.
17. Birkemoe, P.C. and Gilmor, M.I. (1978). "Behavior of Bearing Critical Double-Angle
Beam Connections," Engineering Journal, AISC, 15(4), pp. 109-115.
18. Yura, J.A., Birkemoe, P.C., and Ricles, J.M. (1982). "Beam Web Shear Connections:
An Experimental Study," J. of the Structural Division, ASCE, 108(ST2).
19. Ricles, J.M. and Yura, J.A. (1983). "Strength of Double-Row Bolted-Web Connections,"
J. of the Structural Division, ASCE, 109(ST1).
20. Franchuk, C.R., Driver, R.G., and Grondin, G.Y. (2002). "Block Shear Behaviour of
Coped Steel Beams," Structural Engineering Report No. 244, Department of Civil and
Environmental Engineering, University of Alberta, Edmonton, AB, Canada.
21. Franchuk, C.R., Driver, R.G., and Grondin, G.Y. (2003). "Experimental Investigation of
Block Shear Failure in Coped Steel Beams," Canadian Journal of Civil Engineering,
NRCC, 30(5), pp. 871-881.

332 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


22. Ravindra, M.K., and Galambos, T.V. (1978). "Load and Resistance Factor Design for
Steel." J. of the Structural Division, ASCE, 104(ST9), pp. 1337-1353.
23. Fisher, J.W., Galambos, T.V., Kulak, G.L., and Ravindra, M.K. (1978). "Load and
Resistance Factor Design Criteria for Connectors." J. of the Structural Division, ASCE,
104(ST9), pp. 1427-1441.
24. Franchuk, C.R., Driver, R.G., and Grondin, G.Y. (2004). "Reliability Analysis of Block
Shear Capacity of Coped Steel Beams," J. of Structural Engineering, ASCE, (in press).
25. Schmidt, B.J. and Bartlett, F.M. (2002). "Review of Resistance Factor for Steel: Data
Collection," Canadian Journal of Civil Engineering, NRCC, 29(1), pp. 98-108.
26. Kennedy, D.J.L., and Gad Aly, M. (1980). "Limit States Design of Steel Structures–
Performance Factors," Canadian Journal of Civil Engineering, NRCC, 7(1), pp.45-77.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 333


334 Connections in Steel Structures V - Amsterdam - June 3-4, 2004
DUCTILITY REQUIREMENTS IN
SHEAR BOLTED CONNECTIONS
C. Pietrapertosa, Liège University, M&S Department, Belgium
E. Piraprez, Steel Solutions, Belgium
J.P. Jaspart, Liège University, M&S Department, Belgium

ABSTRACT

In most of the design codes, very few recommendations exist to ensure the
ductile behaviour of connections with non preloaded bolts under shear loading.
The engineers are therefore used to consider steel as a “good” material with
sufficient ductility, so allowing generally to reach at ultimate state a uniform
distribution of the internal forces consecutive to a plastic redistribution amongst
the bolt rows. But in reality this requires a sufficient deformation capacity from
the constitutive plates in bearing and connectors in shear. In the present paper,
all these aspects are discussed and ductility requirements allowing to reach
uniform distributions of internal forces in shear connections are proposed.
These investigations have to be seen as a preliminary step on the way to the
proposal of more general ductility criteria covering other important aspects as
high strength steels, tolerances of fabrication, oversized and slotted holes, …

DESCRIPTION OF THE PROBLEM

When a bolted connection is submitted to shear loading (Figure 1), a distribution of the
external applied force takes place amongst the connectors. Locally, forces are transferred
from one plate to another (to the others) by plate-to-bolt contact, so leading to bearing and
tension forces in the plates as well as shear forces in the connectors.

F
F

F F

F/2
F
F/2

Figure 1. Bolted connection under shear loading.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 335


F F<Fe In Figure 2, the evolution of
Fpl the internal forces versus the
Fe externally applied force F is
illustrated for a connection
C1 C2 C3 C2 C1 with five bolts (or five bolt-
rows). ∆ is the connection
∆ elongation.
In the elastic range of
F F=Fe
behaviour (F < Fe), internal
Fpl Rc Rc forces distribute according to
Fe
the local elastic stiffness
around the connectors, until
the local design resistance Rc
C1 C2 C3 C2 C1 of the plate or of the
∆ connector is reached in the
most loaded zone (F = Fe). If
F Fe<F<Fpl the first zone to be yielded
Fpl
Rc Rc possesses some plastic
Fe capacity of deformation, the
applied force F may be further
increased and plastic
C1 C2 C3 C2 C1 redistribution progressively

takes place (Fe < F). If
sufficient deformation may be
F provided around each
F=Fpl
Fpl Rc connector, a full plastic
Fe
redistribution may be
observed (F = Fpl). In all other
cases, the failure of the
C1 C2 C3 C2 C1 connection is reached by lack
∆ of ductility and the maximum
external force to be
transferred is lower than the
Figure 2. Distribution of internal forces at successive
full plastic resistance (Fe < F
loading steps.
< Fpl).
The evaluation of the resistance of shear connections therefore requires a study of the
distribution of the internal forces as well as the systematic control of the available local
deformation capacity around the bolts. So the objectives here are the following:
• follow the progressive redistribution of forces inside the connection as shown in Figure 2;
• develop a model for the evaluation of the ductility required to reach a full redistribution;
• assess the available ductility around each bolt;
• formulate design requirements for ductility.

BEHAVIOUR OF A SHEAR CONNECTION

Component approach
For the modelling and calculation of the connection shown on Figure 4.a, the component
method (1), prescribed by Eurocode 3 - Part 1.8 (2), can be efficiently applied. The
components to be considered for shear connections are presented in Table 1 (bolts in shear,
plates in bearing and plates in tension). These ones are subjected to an internal force R and
elongates accordingly by δ; their characteristic stiffness (Sc) and resistance (Rc) properties
(Figure 3.a), given in Part 1.8 of Eurocode 3, are reported in Table 1. On the contrary, no
information is provided as far as the available deformation capacity (δc) is concerned.

336 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Table 1. Stiffness (Sc) and resistance (Rc) of the basic components.
Sc Rc

Plate in
Spl = EA/pb Rpl=min(A fy; 0,9 Anet fu)
tension

Bolt in
Sb = 8 d² fub/dM16 Rb = αvfub Ab
shear

Plate in
Sp = 12 kb kt d fu Rp = k1αb fu d t
bearing

Equivalent Seq =
Req = min(Rb, Rp1, Rp2)
component (Sb-1 + Sp1-1 + Sp2-1)-1

E Young Modulus d diameter of the bolt


A gross area of the plate d0 diameter of the bolt hole
Anet net area of the plate dM16 nominal diameter of a M16 bolt
pb pitch distance (// to load transfer) e2 edge distance (⊥ to load transfer)
eb end distance (// to load transfer) p2 pitch distance (⊥ to load transfer)
fy yield strength of the plate kb = min(kb1;kb2)
fu ultimate strength of the plate kb1 = 0,25 eb/d + 0,5 but kb1 ≤ 1,25
t thickness of the plate kb2 = 0,25 pb/d + 0,375 but kb2 ≤ 1,25
Ab shear area of the bolt (nominal or kt =1,5 t / dM16 but kt < 2,5
stress area) αv = 0,5 or 0,6
fub ultimate strength of the bolt αb = min(eb/3d0; pb/3d0 – 0,25 ; fub/fu; 1,0)
k1 = min(2,8 e2/d0 – 1,7; 1,4 p2/d0 – 1,7; 2,5)

R R
Rc (p, pl or b)
Req

Sc (p, pl or b)
δ Seq
δc (p, pl or b) δ
δav
a – Individual component b- “Equivalent bolt zone” component

Figure 3. Component response.

In Figure 4.b, a mechanical model is built where all the constitutive components are
represented by extensional springs. When the forces pass through the bolts from the lower
plate to the upper plate, the bolt holes in the lower plate deform by bearing, the bolts are
sheared and finally bearing ovalise the holes in the upper plate. This transfer of forces is
modelled by three springs in series in Figure 4.b. These three springs may be merged

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 337


around each bolt into a single one (see Figure 3b); the behaviour of the latter depends of the
individual responses of the three “component” springs (Table 1). As a result, the connection
is finally represented by a simplified model with only two kinds of different components
(Figure 4.c): plate in tension and equivalent bolt zone (including bolt shearing and plate
bearing). In further studies, effects resulting from a partial bolt preloading or from the
existence of tolerances of fabrication (on bolt diameter, bolt alignment, …) will be
investigated by characterising this equivalent component by an appropriate behavioural law.
Finally it should be mentioned that single lap joint connections are covered in the present
paper, but that all developments may easily extended to double overlap joints.

a – Shear connection with three bolts


Sp2 Sp2 Sp2
Sb Sb Sb
Sp1 Sp1 Sp1
b – “Actual” mechanical model

Seq Seq Seq F

c – Simplified mechanical model


Figure 4. Mechanical model of behaviour based on the component approach.

Failure modes

The following possible failure modes may be observed: (i) plate in tension (in the net area or
gross area) or (ii) failure of an equivalent bolt zone in shear. The second one is linked either
to the bearing resistance of one of the connected plates, or to the resistance of a bolt in
shear. In the present study, the properties of the connected plates are considered to be such
that the failure by “plate in tension” is not at all relevant.

Distribution of the internal forces

To illustrate how internal forces distribute amongst the components when the connection is
progressively loaded, a simple lap connection with 5 bolts is considered (Table 2). Again an
elasto-plastic behaviour of the components is assumed (Figure 3.b).
Step 1: For a low level of load, all the components are in their elastic range of behaviour and
the two exterior components are more loaded than the interior ones.
Step 2: For a higher applied force, the exterior components reach their strength limit Req.
They can not support any further load. The other components are still elastic.
Step 3: When the applied load increases, plastic redistributions of forces take place as the
exterior components cannot transfer further load; forces are reported on the other
components C1 and C2 until the C2 components reach also Req. This requires plastic
deformation capacity from the C1 components.
Step 4: Plastic redistribution may progress as long as the deformation capacity is available in
the C1 and C2 components. Finally, the maximum resistance of C3 is reached. A complete
plastic redistribution is then obtained and the connection resistance equals 5Req. To reach
this ultimate load, the exterior components have to exhibit a sufficient deformation δreq.

338 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Components
Assembling
C1 C2 C3
F F<Fe R R R
Fpl
Req Req Req
Fe

C1 C2 C3 C2 C1

∆ δ δ δ
δav δav δav

R R R
F F=Fe
Fpl Req Req Req Req Req
Fe

C1 C2 C3 C2 C1
δ δ δ
∆ δav δav δav

F Fe<F<Fpl R R R
Fpl
Req Req Req Req Req
Fe

C1 C2 C3 C2 C1
∆ δ δ δ
δav δav δav

R R R
F F=Fpl
Fpl Req Req Req Req
Fe

C1 C2 C3 C2 C1
δ δ δ
∆ δreq δav δav δav

Table 2. Response of a “5-bolts” shear connection and of its constitutive components under increasing applied loading.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 339


Numerical validation

The here-above presented model has been run numerically, through the use of the Liège
home-made non-linear FEM software FINELG, for many different situations. Some
experimental tests on shear connections performed by Valtinat in Hamburg (3) have also
been successfully simulated.

Required deformation capacity from the equivalent components

If the components exhibit an infinite deformation capacity, a total plastic redistribution of the
internal forces can take place and each equivalent component can reach its maximum strength
Req. In such a case, the shear resistance of the connection is defined as equal to Fpl = n Rc.
If, on the contrary, the “equivalent bolt” components do not exhibit good deformation
capacity, the resistance of the connection will have to be limited to a lower value, which is the
one transferred by the connection just when the maximum available deformation capacity δav
is reached in the most heavily deformed component.
In other words, a shear connection with n bolts (or bolt-rows) only reaches its full plastic
capacity nReq if the “equivalent bolt” components possess a deformation capacity δav higher
than the required one, δreq:

δreq > δav (1)

EVALUATION OF THE REQUIRED DEFORMATION CAPACITY δreq

Particular case of a connection with 3 bolts (or bolt-rows)

The elastic distribution of internal forces amongst the components depends on the relative
stiffness of the steel plate Spl and that of the equivalent bolt components Seq (see Figure 5).

δ0 δ1 δ0

Figure 5. Elastic distribution of internal forces in a connection with 3 bolts (bolt-rows).

340 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Because of symmetry, the forces transmitted by the external components are both equal to
F1; as a consequence, the internal component supports a force equal to F-2F1. Each portion
of the plate located between two components elongates according to the force it transfers.
The same principle applies to the equivalent bolt component in shear. As a result:

δ0 = F1/Seq (2.a)
δ1 = (F-2F1)/Seq (2.b)

By expressing the compatibility of displacements between the plates and the bolt zones, the
distribution of shear forces amongst the components in the elastic range may be derived
(figure 5):
⎛ 1 p ⎞ 1
⎜ + b⎟
⎜S EA ⎟ S
F1 = F ⎝ ⎠ (3.a and 3.b)
eq eq
F − 2 F1 = F
⎛ 1 p ⎞ ⎛ ⎞
⎜3 +2 b ⎟ ⎜ 3 1 + 2 pb ⎟
⎜ S EA ⎠⎟ ⎜ S EA ⎟⎠
⎝ eq ⎝ eq

If the plates in tension are stiff (EA → ∞), the distribution of forces tends to be uniform; for a
fully flexible plate (EA → 0) , only the external components transfer the shear force.

When the external components reach their maximum strength (F1 = Req), the maximum
elastic resistance of the connection is obtained. The corresponding applied load is equal to:
⎛ 1 p ⎞
⎜3 +2 b ⎟
⎜ S EA ⎟⎠
Fel = Req ⎝
eq
(4)
⎛ 1 p ⎞
⎜ + b⎟
⎜S ⎟
⎝ eq EA ⎠
For this level of shear load, the elongations in the external and internal equivalent bolt
components are respectively equal to:

δ0 = Req / Seq (5.a)


1
S eq
δ 1 = Req (5.b)
⎛ ⎞
⎜ 1 + pb ⎟
⎜ S eq EA ⎟⎠

The maximum strength of the internal bolt is not yet exhausted and the additional force that it
can still transfer is:
1
S eq
Req (1 − (6)
⎛ 1 p ⎞
⎜ + b⎟
⎜S ⎟
⎝ eq EA ⎠

In order to reach a full plastic distribution of internal forces, to which would be associated the
following connection resistance:
VRd = 3 Req (7)
the internal bolt component has therefore to undergo an additional elongation:

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 341


⎛ 1 ⎞
⎜ ⎟
Req ⎜ S eq ⎟
δ 1+ = ⎜ 1− ⎟ (8)
S eq 1 p
⎜⎜ + b ⎟
⎝ S eq EA ⎟⎠

And finally, through the equations of displacement compatibility, the total elongation of the
external bolt components required to enable a full plastic redistribution to take place may be
expressed as follows:
Req pb
δ 0,req = + Req (9)
S eq EA

Generalisation to connections with more than 3 bolts (or bolt-rows)

The procedure followed in the previous section may be generalised to connections with n
bolts or bolt-rows. Table 3 gives values of the required deformation capacity up to n=8.

Table 3. Maximum required deformation capacity from the equivalent bolt components.

Number of rows Required deformation capacity δreq

Req
2 S eq

⎛1 p ⎞
Req ⎜ + b⎟
3 ⎜S ⎟
⎝ eq EA ⎠

⎛ 1 p ⎞
4 Req ⎜ +2 b ⎟
⎜S EA ⎟⎠
⎝ eq
⎛ 1 p ⎞
5 Req ⎜ +4 b ⎟
⎜S EA ⎟⎠
⎝ eq
⎛ 1 p ⎞
6 Req ⎜ +6 b ⎟
⎜S EA ⎟⎠
⎝ eq
⎛ 1 p ⎞
7 Req ⎜ +9 b ⎟
⎜S EA ⎟⎠
⎝ eq
⎛ 1 p ⎞
8 Req ⎜ + 12 b ⎟
⎜S EA ⎟⎠
⎝ eq

In a more general way, the required deformation capacity may be evaluated as follows:
⎛ 1 p ⎞
δ req = Req ⎜⎜ +ρ b ⎟ (10.a)
⎝ S eq EA ⎟⎠
n/2
with: ρ = ∑ (n − 2i) for an even value of n (10.b)
i =1
( n −1) / 2
ρ= ∑ (n − 2i)
i =1
for an odd value of n (10.c)

342 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


EVALUATION OF THE AVAILABLE DEFORMATION CAPACITY δav

In Eurocode 3, no guideline is given on how to evaluate the available deformation capacity of


the bolts in shear and plates in bearing. As a consequence, the available deformation
capacity δav of the equivalent bolt components can not be easily defined. More theoretical,
analytical and experimental works in this field are therefore still required. In the present
research, as a first step, series of tests carried out in Moscow some years ago (4) have been
analysed and approximate values of the deformation capacity of bolts in shear have been
derived. For the deformation capacity for plates in bearing, reference has been made to (5)
where analytical expressions are proposed. The obtained values are given in Table 4.

Table 4. First approximate values of deformation capacity for individual components.

Components Available deformation capacity

Bolts in shear 8.8 bolts:


δb = 3,0 Rb / Sb M16 bolts
δb = 3,6 Rb / Sb M20 bolts
δb = 4,3 Rb / Sb M24 bolts
10.9 bolts:
δb = 1,2 Rb / Sb M16 bolts
δb = 2,0 Rb / Sb M20 bolts
δb = 2,7 Rb / Sb M24 bolts

Plate in bearing δp = 10Rp / Sp

From these values, the available deformation capacity of the “equivalent bolt zone
component” may be derived as follows:

ƒ If the shear resistance of the bolt is significantly higher than that of the plate(s) in bearing,
the whole plastic deformation capacity results from that of the plate(s) in bearing and:

δ av = Req / S eq + (δ p − Req / S p ) with: Req = Rp for: Rb > 1,25 Rp (11)

ƒ If the shear resistance of the bolt is significantly lower than that of the plate(s) in bearing,
the whole plastic deformation capacity results from that of the bolt in shear and :

δ av = Req / S eq + (δ b − Req / S b ) with: Req = Rb for: Rp > 1,25 Rb (12)

ƒ In intermediate cases, the plastic deformation capacity results from both the bolt and
plate(s) deformations; as a rough approximation, the following formulae are suggested:

δ av = Req / S eq + ((( Rb − R p ) / 0,25 R p )δ p − Req / S p )


+(((1,25 R p − Rb ) / 0,25 Rb )δ b − Req / S b )
with: Req = Rp for: 1≤ Rb / Rp ≤ 1,25 (13)
δ av = Req / S eq + ((( R p − Rb ) / 0,25Rb )δ b − Req / S b )
+(((1,25 Rb − R p ) / 0,25 R p )δ p − Req / S p )
with: Req = Rb for: 1≤ Rp / Rb ≤ 1,25 (14)

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 343


WORKED EXAMPLE

The following worked example (Figure 6) illustrates the content of the paper and allows to
draw some first conclusions about the capacity of shear connections to undergo plastic
redistributions of internal forces. It is also the occasion to identify further research needs.

The numbers of bolt rows and of bolt lines are varied. All the other geometrical and
mechanical parameters are fixed. The variation of the required and available deformation
capacities with the number of bolt rows and files is reported in Figure 7.

Steel of the plates: Bolts: Geometrical properties:


E = 210000 N/mm² d = 16 mm eb = 35 mm
fub = 1000 N/mm² d0 = 18 mm pb = 80 mm
fy = 355 N/mm² Ab = 157 mm² p2 = 40 mm
fu = 510 N/mm² b = 216 mm

p2 = 40mm
b = 216mm n2=1 → 5
eb = 35mm pb = 80mm

Figure 6. Mechanical and geometrical data.

n1 = 6
practical criterion

15 d0 EC3 criterion n2 = number of bolt lines

Figure 7. Required and available deformation capacities according


to the number of bolt rows and bolt lines.

The values reported in Figure 7 are calculated for “ideal” connections, i.e. with no
geometrical lack-of-fit (exact position of the holes, same hole diameters for all the bolts,
equal hole and bolt diameters,...). To reach the full plastic redistribution and consequently the
full resistance of the connection, the curve δav has to be above the curve δreq.

It may be seen that δreq increases with the number of bolt rows (see also Formula 10). On the
contrary δav is not affected by this parameter as it only depends on the local properties of the
equivalent component. The required deformation capacity increases also with the number of
bolt lines n2. Indeed, this last parameter influences the values of Seq and Req in Formula 10.

344 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


In the technical literature on bolted connections, few ductility requirements are proposed
which limit the application of the “full plastic redistribution” concept: Two of them are:

ƒ n1 < 6 to 8 limitation of the number of bolt rows to 6 or 8


ƒ Lj < 15d0 limitation of the connection length Lj to 15d0 (d0 is the hole diameter)
suggested by Eurocode 3
These two criteria are reported in Figure 7. From the graph, these two empirical criteria
appear as safe requirements as long as an ideal connection without any lack-of-fit is
considered. Indeed, the connection analysed in the work example is rather demanding in
terms of deformation requirement (very thick plate with small bolts) and even for a
connection with 5 bolt lines, the two criteria are fulfilled. However, it has to be clearly stated
that these considerations are probably no more applicable to actual connections because of
the unavoidable lack-of-fit aspects mentioned earlier. Indeed, as a result, for instance, of the
bolt misalignment in one row, some bolts will bear against the plates before the others and
more deformation capacity will therefore be required locally. This can prevent a full plastic
redistribution of the internal forces and so significantly decrease the actual resistance of the
shear connection. As a conclusion, further studies have to be carried out with the aim to
include the effects of lack-of-fit in the model and to develop more general design
requirements.

CONCLUSIONS

In the present paper, a methodology is introduced to check whether sufficient ductility for full
plastic redistribution of internal forces is available in shear connections with non preloaded
bolts. The development and preliminary validation of a behavioural model for shear bolted
connections is presented. Design requirements are expressed and formulae to evaluate the
available deformation capacity in the vicinity of the bolts are proposed for connections with
normal bolt holes. Finally, the justification of the apparently safe character of the EC3 ductility
criteria (and others) has been demonstrated. However it has to be mentioned that the
influence of fabrication tolerances as well as the effect of the gap between the bolt shanks
and the hole edges may significantly influence the resistance of the “ideal” studied
connections. In further studies, the effects of fabrication tolerances will be implemented and
the calculation model presented in the paper will be extended to slotted and oversize holes.
Finally, it is expected that these works will lead to new improved ductility requirements to be
included in a later stage in Eurocode 3.

REFERENCES

(1) Jaspart J.P., “Design of structural joints in building frames”, Progress in Structural
Engineering and Materials, Vol. 4 N° 1, January-February 2002, pp. 18-34.
(2) Eurocode 3 “Design of steel structures” Part 1.8 “Design of joints”, prEN 1993-1-8,
CEN, Brussels, April 2002.
(3) Valtinat G. and Wilhem M., “Schrauben-verbindungen im Stahlbau“, Last-
Verschiebungs-Diagramme, Hamburg University, Germany, 1988 – doc. ECCS-TC10-
89-301.
(4) Karmalin V.V. and Pavlov A.B., “Load capacity and deformability of bearing and friction-
bearing connections”, Proceedings of the International Colloquium on Bolted and
Special Structural Connections, USSR, Moscow, May 15-20, 1989, pp. 52-60.
(5) Jaspart J.P., “Etude de la semi-rigidité des noeuds pouter-colonne et son influence sur
la résistance et la stabilité des ossatures en acier”, Ph.D. Thesis, M&S Department,
Liège University, Belgium, 1991.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 345


346 Connections in Steel Structures V - Amsterdam - June 3-4, 2004
AN IMPROVED APPROACH TO COMPUTE BLOCK SHEAR
CAPACITY OF STEEL ANGLES
Dr. Mohan Gupta, Bhilai Institute of Technology, Durg, India
Dr. L. M. Gupta, Visvesvaraya National Institute of Technology, Nagpur, India

ABSTRACT

Current practice for computing block shear capacity of steel angles with bolted
connections is to assume the stress distribution along the “block” gross shear
plane to be uniform, which may not always be true. Based on previously
published experimental data and nonlinear finite element analysis, it is
suggested here that area along “block” gross shear plane should be suitably
modified taking into account the stagger of bolts (keeping the “block” net
tension area unaltered). This improved approach leads to adequate prediction
of block shear capacity of single as well as double angles, with staggered as
well as non-staggered bolts.

INTRODUCTION

Steel tension members are probably the most common and efficient members in structural
applications. It is relatively easy to fabricate and erect structures, or a part thereof, comprising
of angles because of the basic simplicity of its cross-section. As such, steel angles are
frequently used as tension members in a majority of applications, in spite of their complex
behaviour.

Often, it is not practicable to connect both the legs of the angle with the gusset plate. Angles
are generally used as single or as a pair, symmetrically placed about a gusset plate that
passes between them. The connection between the angle and gusset may be made by
welding or by bolting. Bolting requires drilling or punching of holes, which results in reduced
area at the critical section. The bolts may be placed along one or more gauge lines, and may
or may not be staggered along the length of the connection. When staggered, well known
s2/4g rule is applied to compute the least net area.

Angles with bolted connections normally observe net section failure for relatively longer
connections. However, for relatively shorter connections, the mode of failure may be block
shear, wherein a ‘block’ of the connected element may separate from the remainder of the
element (Figure 1). The failure path in this case lies entirely within the connected leg;
outstanding leg is stressed very little. On account of this, the load carrying capacity is
significantly reduced when the mode of failure is block shear. The angle section must be
checked for block shear failure also, even if the end, edge and pitch distances are within the
specified limits.

Block shear mode of failure has drawn the attention of researchers from most parts of the
world. As a result of the ongoing research, the provisions for block shear have been modified
in almost every revision of various specifications. However, even the current provisions in
various specifications are not the perfect reflection of experimental results, as envisaged in
the recent research in this area. Thus, there is a scope for improvement in the way the block
shear strength of angles under tension is predicted.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 347


Figure 1. Block shear failure of steel angle with bolted connection (1).

American Institute of Steel Construction (AISC) has issued a separate specification


exclusively for the design of steel angles (2). It is intended to be compatible with, and a
supplement to the AISC specification for design of structural steel buildings (3). No specific
provisions for predicting block shear capacity of angles are specified in (2), though the failure
mechanism is quite different than that in some other connections such as coped beams.

In the present study, block shear model based on the failure mechanism observed during the
experiments is evaluated on the basis of relevant test results available in the literature.
These test results include single as well as double angles, having bolt holes in one or more
rows, and with staggered and non-staggered holes. An improvement in this approach of
computing block shear capacity of steel angles is then proposed.

BLOCK SHEAR FAILURE EQUATION

The failure mechanism observed during the experiments consists of rupturing of the block
net tension plane, accompanied by inelastic deformation of the block gross shear plane
(Figure 2).

Block net tension plane


Block gross shear plane “Block”

Figure 2. Typical block shear failure mechanism.

348 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


This mode of failure is best described by the equation:

Rb = fu . Ant + fys . Agv (1)

The yield strength in shear is taken as 1 / √3 times the yield strength in tension, according to
Von Mises yield criteria, so that, fys = fy / √3. Kulak and Grondin (4), (5) are of the view that
this equation be used for angles, on the rationale that it is based on the behaviour observed
during the experiments on angles. In their evaluation, they have included only the angles in
which bolts are not staggered and have shown that rules in AISC specification are in better
agreement with the angle tests than Equation (1). In their discussion, Gupta and Gupta (6)
included few angles that had staggered holes.

In the following, it is shown that this simple equation adequately predicts the block shear
capacity of angles with non-staggered bolts and with staggered bolts as well, after
improvement in the way the block gross shear area is computed in certain cases.

EVALUATION OF BLOCK SHEAR EQUATION

One of the simplest and straight forward ways to evaluate the prediction of block shear
strength by Equation (1) would be to work out the ratio of experimental failure load to the
predicted ultimate load, termed as professional factor, for previously published test results.
Professional factor less than 1.0 indicates non-conservative load prediction, while factor
exceeding 1.0 is produced by conservative prediction.

In Table 1a, the specimen numbers used are same as reported in the original research. In
this table, only those specimens from previously published literature have been included that
showed block shear failure. All these specimens follow all provisions regarding minimum
pitch, edge and end distances. The specimens which violate one or more of these provisions
are not included here. Angles composed of high strength steel are also not included here.
Specimens of Epstein (7) had double angles with two rows of bolts. In a majority of these
specimens, the bolts are staggered. Specimens of Gross et. al. (8), Orbison et. al. (9) and
Gupta and Gupta (10) are single angles with one row of bolts.

The professional factors are worked out in Table 1a, using the actual material properties of
each of these specimens. Average and standard deviation of professional factors have also
been worked out. The average professional factor of 1.04 for all tests indicates adequate
representation of block shear strength by Equation (1). The average professional factor as
per AISC specification (3) is 1.02 while it is as high as 1.26 for Eurocode3 (11).

In Table 1a, there are certain specimens in which the bolts are not placed along the standard
gauges. It is observed that the professional factors for few such specimens, as per Equation
(1) and AISC specification (3) are on the non-conservative side. There is an increase in the
in-plane eccentricity as the bolt line shifts towards the toe of the angle section, and it is
reconfirmed here that the block shear capacity decreases with an increase in the in-plane
eccentricity. More non-conservative predictions may be expected, if there is further increase
in the in-plane eccentricity. As such, angles in which the bolts are not placed along standard
gauges must be given due consideration in design and the capacity should be judiciously
reduced, depending upon the eccentricity.

This evaluation points that the block shear equation (Equation 1) predicts the capacity of
angles more conservatively than the AISC specification (3), when the mode of failure is block
shear. AISC approach involves comparatively more computations and in many cases, the
governing mechanism does not agree with the experimentally observed mechanism.
Eurocode3 (11) provisions predict this capacity quite conservatively.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 349


Table1a. Professional factors for different approaches.
Professional factors for different approaches
Source Specimen Equation (1) AISC LRFD Eurocode3
(1999) (2001)

[1] ** 0.88 0.87 1.00


[9] ** 1.00 0.98 1.13
[11] 1.02 1.01 1.17
Epstein
[25] ** 1.03 1.01 1.20
(7) [37] ** 1.10 1.08 1.28

Gross et. al. [A36-1] 1.10 1.04 1.37


(8)
[A36-2] ⇒ 1.00 0.95 1.30
[A36-3] ⇒ 0.92 0.91 1.21

[A-1] ⇒ 1.00 0.99 1.22


[A-2] 1.13 1.16 1.34
Orbison et. al.
[A-3] 1.11 1.09 1.29
(9)

Gupta and Gupta [65_2_1a] 1.13 1.09 1.45


(10) [65_2_1b] 1.10 1.03 1.39

Average 1.04 1.02 1.26


Std. Dev. 0.08 0.08 0.12

** Two rows of fastners in staggered holes


⇒ Bolts not placed along standard gauges (Bolt line is towards the toe of angle).

There are other specimens tested by Esptein (7), in which the mode of failure was partially
block shear. A majority of such specimens had bolt holes staggered in two rows such that the
end distance along the row of bolts towards the heel was more than the end distance along
the row of bolts towards the toe of the angle (Figure 3a).

e p p p e p p p

Agv Agv

Agv = ( p + p + p + e ) * t Agv = ( p + p + e ) * t
(a) (b)

Figure 3. Block gross shear plane area in different cases.

It is observed from results reported in Epstein (7) that, in many cases, the experimental
failure load of specimens with bolt hole configurations as in Figure 3a and Figure 3b, is about
the same. Whereas, the predicted block shear capacity, as per Equation (1), is much more in
specimens of Figure 3a than that of specimens of Figure 3b, because of increased block
gross shear plane area. The bolt hole configuration of specimens considered in Table 1b is

350 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


shown in Figure 4. It is observed from Table 1b that the professional factors as per Equation
(1) for almost all of these specimens are on the unconservative side, indicating that Equation
(1), in its present form, is not applicable in such cases.

Specimen Nos. Specimen Nos. Specimen No.


[ 5 ], [13], [21], [28], [32], [36] [ 2 ], [10], [18], [26], [30], [34] [38]

Figure 4. Bolt hole configuration of specimens of Table 1b.

IMPROVED BLOCK SHEAR EQUATION

Hardash and Bjorhovde (12), in their study on gusset plates, observed that the effect of
varying the connection length is important, and proposed that it must be incorporated in a
rational and complete gusset plate model. They have suggested a linear interpolation
function that is applied on shear stress acting on gross connection length area. A deep
insight into the bolt hole configurations of specimens of Table 1b reveals that the stress
distribution along the block gross shear plane may not be uniform. The stress is likely to be
more towards the lead bolt hole and the stress near the end of block gross shear plane may
be quite less. This non-uniform stress distribution is illustrated in Figure 5.

More stressed area


Lesser stressed area

Gross block shear plane

Figure 5. Stress distribution along block gross shear plane.

A nonlinear finite element model was prepared to verify this concept and the resulting shear
stress distribution along block gross shear plane is shown in Figure 6. This finite element
analysis is largely based on the methodology adopted by Kulak and Wu (13) for modelling of
angles under tension.

Outstanding leg

Minimum
Gusset
Plate Zero

Maximum
Connected leg

Figure 6. Non uniform stress distribution along block gross shear plane.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 351


This suggests that while computing the block shear capacity by Equation (1), the lesser
stressed part of the block gross shear plane may be ignored and only the more stressed part
of the block gross shear plane should be considered. A simple approach in such cases would
be to consider the length of block gross shear plane from centre of lead bolt hole to centre of
first bolt hole only, as shown in Figure 7. The effective block gross shear area, A*gv, as per
this improved approach will be somewhat less than the block gross shear area, Agv of
Equation (1). Equation (1) in its improved form is written as:

Rb = fu . Ant + fys . A*gv (2)

e p p p e p p p
Sample Calculations:
Agv = ( p + p + p + e ) * t
Agv A*gv
A*gv= ( p + p + p ) * t

Figure 7. Sample calculations for block gross shear area.

Table 1b. Professional factors for different approaches.


Professional factors for different approaches
Source Specimen Equation (1) AISC LRFD Eurocode3 Equation (2)
(1999) (2001)

[2] 0.85 0.80 0.95 0.97


[5] 0.89 0.84 1.00 1.01
[10] 0.95 0.89 1.06 1.08
[13] 0.84 0.79 0.94 0.95
[18] 0.82 0.77 0.92 0.93
[21] 0.97 0.91 1.09 1.10
[26] 0.90 0.84 1.02 1.04
Epstein
[28] 0.86 0.80 0.98 1.00
(7) [30] 0.94 0.84 1.07 1.09
[32] 0.97 0.87 1.09 1.11
[34] 0.95 0.88 1.08 1.10
[36] 0.97 0.89 1.10 1.12
[38] 1.01 1.00 1.07 1.25

Average 0.92 0.86 1.03 1.06


Std. Dev. 0.06 0.06 0.06 0.09

The professional factors with the AISC approach, Eurocode3 approach and improved
approach suggested herein (Equation 2) are also presented in Table 1b. The average
professional factor of these specimens as per AISC approach is only 0.86 and indicates
unconservative prediction of block shear capacity. The Eurocode3 provisions for such
specimens are not as conservative as for specimens of Table 1a. It is observed that the
professional factors as per Equation (2) represent the block shear capacity adequately in
most of the cases with a mean value of 1.06. Thus, there is a considerable improvement in
values of professional factors when the concept of effective block gross shear area is used in
the computations.

It is to be noted that the data available on block shear failure of angles with staggered holes
is from a single source only. More such tests may be conducted to further validate the

352 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


findings of this study. Likewise, a rigorous nonlinear finite element analysis to cover various
bolt hole configurations and other parameters may further be performed.

CONCLUSIONS

This study examined the block shear capacity of steel single as well as double angles, for
bolt holes in one or more rows, and with staggered and non-staggered holes. Only those
specimens that follow all provisions regarding minimum pitch, edge and end distances were
included in this study. Angles composed of high strength steel are not included in this study.
An improved approach, to compute the the block gross shear area of specimens that had
bolt holes staggered such that the end distance along the row of bolts towards the heel is
relatively more, is suggested. This area, A*gv, as per this improved approach is termed here
as effective block gross shear area and is somewhat less than the block gross shear area,
Agv, as per current practice. There is a considerable improvement in values of professional
factors when the concept of effective block gross shear area is used in the computations.

The following simple equation was found to give adequate results for single as well as double
angles, for bolt holes in one or more rows, and with staggered and non-staggered holes.

Rb = fu . Ant + fys . A*gv

The data available on block shear failure of angles with staggered holes is from a single
source only. More such tests may be conducted to further validate the findings of this study.

The proposed approach is specific to angles, and may likely be incorporated in the AISC
specification (2) on angles.

NOTATIONS

A*gv Effective Block Gross Shear Area


Agv Block Gross Shear Area
Ant Block Net Tension Area
e End Distance
fu Ultimate Tensile Stress
fy Tensile Yield Stress
fys Shear Yield Stress
p Pitch Distance
Rb Predicted Block Shear Capacity
t Thickness of Angle Section

REFERENCES

(1) Gupta Mohan, “Study of Effective Areas of Steel Angles under Tension”, Ph. D.
Thesis, Visvesvaraya National Institute of Technology, Nagpur, India, May 2003.
(2) AISC LRFD:2000, “Load & Resistance Factor Design Specification for Single Angle
Members”, Second Edition, American Institute of Steel Construction, Chicago, Illinois.
(3) AISC LRFD:1999, “Load & Resistance Factor Design Specification for Structural
Steel Buildings”, Third Edition, American Institute of Steel Construction, Chicago,
Illinois.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 353


(4) Kulak, G. L., and Grondin, G. Y., “Block Shear Failure in Steel Members - A Review
of Design Practice”, Proceedings of the Fourth International Workshop on
Connections in Steel Structures, October 22 - 25, Roanoke, VA, 2000.
(5) Kulak, G. L., and Grondin, G. Y., “AISC LRFD Rules for Block Shear - A Review”,
Engineering Journal, American Institute of Steel Construction, Vol. 38, No. 4, Fourth
Quarter, 2001, pp. 199 - 203.
(6) Gupta, Mohan and Gupta, L.M., "Discussion: AISC LRFD Rules for Block Shear in
Bolted Connections - A Review", Engineering Journal, American Institute of Steel
Construction (AISC), Vol. 39, No. 4, 4th Quarter 2002, Chicago, Illinois, USA, pp. 240.
(7) Epstein, H. I., ”An Experimental Study of Block Shear Failure of Angles in Tension”,
Engineering Journal, American Institute of Steel Construction, Vol. 29, No. 2, Second
Quarter, 1992, pp. 75 - 84.
(8) Gross, J. M., Orbison, J. G., and Ziemian, R. D., “Block Shear Tests in High-Strength
Steel Angles”, Engineering Journal, American Institute of Steel Construction, Vol. 32,
No. 3, Third Quarter, 1995, pp. 117 – 122.
(9) Orbison, J. G., Wagner, M. E., and Fritz, W. P., “Tension Plane Behavior in Single
Row Bolted Connections Subject to Block Shear”, Journal of Constructional Steel
Research, Vol. 49, 1999, pp. 225 – 239.
(10) Gupta, Mohan and Gupta, L. M., “Block Shear Failure of Bolted Steel Single Angles in
Tension”, Advances in Structures - Steel, Concrete, Composite and Aluminum
(ASSCCA' 03), International Conference, 23 - 25 June 2003, Sydney, Australia, pp.
89 - 93.
(11) Eurocode 3: 2001, “Design of Steel Structures Part 1.8: Design of Joints”, Final Draft,
European Committee for Standardization, Brussels.
(12) Hardash, S. G., and Bjorhovde, R., “New Design Criteria for Gusset Plates in
Tension”, Engineering Journal, American Institute of Steel Construction, Vol. 22, No.
2, Second Quarter, 1985, pp. 77 – 94.
(13) Kulak G. L., and Wu, Y., “Shear Lag in Bolted Angle Tension Members”, Journal of
Structural Engineering, ASCE, Vol. 123, No. 9, 1997, pp. 1144 – 1152.

354 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


BEHAVIOUR OF FLANGE TIP CONNECTIONS: PRELIMINARY
TESTING AND ANALYSIS

H.H. Snijder, Eindhoven University of Technology, The Netherlands


J.C.D. Hoenderkamp, Eindhoven University of Technology, The Netherlands

ABSTRACT
Flange tip connections have to transfer primarily shear forces from a
secondary beam to a main beam through the main beam flanges since
stiffeners are omitted for economic reasons. In the flanges of the main beam,
longitudinal bending stresses occur due to beam behaviour. But
simultaneously also transverse bending stresses occur in these flanges at the
connection resulting in a complex stress state in the main beam flanges. The
work reported in the paper includes experimental, analytical and numerical
research on flange tip connections and is aiming at obtaining design rules for
these connections.

INTRODUCTION

Flange tip connections are a special type of beam-to-beam connections, for example in a
floor or roof structure, where a secondary beam is connected to a main beam. Usually these
connections have to transfer primarily shear forces from the secondary beam to the main
beam (figure 1). There are several alternatives in case the top of the upper flanges are to be
at identical levels. One of the economically attractive ones is by adding a connection plate to
the side of the flanges of the main beam and a (short) header plate to the end of the
secondary beam thus connecting both beams by bolts through the connection plate and
header plate. From point of view of force transfer it would be logical to add a stiffener
between flanges, web and connection plate of the main beam. However, this stiffener is
preferably left out for economic reasons and so the secondary beam is connected to the
flange tips of the main beam only, hence the name ‘flange tip connection’.

In the flanges of the main beam, longitudinal bending stresses occur due to beam behaviour.
But simultaneously also transverse bending stresses occur in these flanges at the connection
due to the presence of the secondary beam. Also shear stresses are present resulting in a
complex stress state in the connection and complex connection behaviour. In general there
is interaction between connection and beam behaviour. Several research projects have been
carried out at Eindhoven University of Technology (1,2,3) to analyse flange tip connections.
This is done by simplifying the problem to first focussing on the flange tip connection itself
and then on the interaction between connection and beam behaviour. The work reported in
the paper includes analytical, numerical and experimental research on flange tip connections
and is aiming at design rules for these connections. The work is still in progress so only
preliminary results are presented here.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 355


Main beam

Secondary beam

M line

Figure 1. Typical flange tip connection between main and secondary


beam and bending moment line in secondary beam.

APPROACH

A structural floor system or even a sub-system consisting of one main and two secondary
beams is relatively complex to investigate experimentally. Therefore, the secondary beams
are omitted and the load is applied to the connection plates connected to the flange tips of
the main beam (figure 2). Also, the problem is further simplified by testing as a first step the
connection only. This is done by supporting the main beam along its length under the web at
the connection. This situation is called a supported connection and only connection
behaviour is investigated. As a second step, the main beam is not supported under the web
at the connection. The main beam can be seen now as a simply supported beam in three-
point bending with the load applied at the connection plates. This situation is called an
unsupported connection and now the interaction between beam and connection behaviour is
investigated.

SUPPORTED CONNECTION - CONNECTION ONLY

Experimental research - Supported connection

Supported connections of three types with an IPE330 as main beam are investigated as
shown in figure 2 (2). In type 1 only the top flange is investigated. In type 2 a thick connection
plate is used forcing yielding into the flanges and distributing the load equally over top and

t = 12 mm
t = 40 t=6

a) type 1 b) type 2 c) type 3


Figure 2. Tests on supported connections.

356 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


bottom flange. In type 3 a thin connection plate is used where yielding may also occur in this
connection plate and where the load is expected to be unequally distributed over top and
bottom flanges. The test set-up is shown in figure 3 and some typical test results are shown
in figure 4. Experimental failure loads per flange tip (Fexp) are listed in table 1. The failure
load of test 3.1 is relatively small due to unequal load distribution between top and bottom
flange. Also large deformations occur in the connection plate as schematically indicated in
figure 5d. In the tests strains were measured at several locations: see e.g. inset in figure 13.

Figure 3. Test set-up for supported connections.

Figure 4. Typical test results for supported connections.

Theoretical research - Supported connection

Typical yield line failure mechanisms are shown in figure 5 (2,3). Using first order plastic
theory, the following failure loads per flange tip can be derived for the mechanisms 1, 2 and 3
of the figures 5a, 5b and 5c respectively:

Fthe,1= (5.29 β k/d) Mp,fl (1)

Fthe,2= (2 π + β k/d) Mp,fl (2)

Fthe,3= (5.64 + β k/d) Mp,fl (3)

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 357


The plastic moment per unit length of the flange is:

Mp,fl = ¼ t2 fy (4)

The distance between root radius and flange tip is:

d = (bf – tw – 2r)/2 (5)

The formulas (1-3) with β = 1 hold for type 1 connections. The three failure loads of these
mechanisms are very close, but formula (1) is the governing one. In case of type 2
connections another yield line forms along the connection plate and then formulas (1-3) with
β = 2 are valid. For type 3 an additional yield line develops in the connection plate resulting in
a modification of the formulas (1-3). A value of (k/d Mp,hp) has to be added to the failure loads
of formulas (1-3) with β = 1 ( fy = 308 N/mm2 for the connection plate material). The failure
loads calculated this way are listed in table 1. For type 3 the calculated theoretical failure
load overestimates the experimental failure load substantially. This is due to the fact that the
bottom flange does not contribute to the failure load and the connection plate fails instead of
the bottom flange. Therefore, the failure load was estimated on the basis of connection plate
failure (figure 5d) with mechanism 1 in the top flange resulting in the failure load indicated
between brackets in table 1.

a) mechanism 1 b) mechanism 2

c) mechanism 3 d)

Figure 5. Typical failure mechanisms.

Numerical research - Supported connection

Tests 1.1 and 2.1 have numerically been simulated (3) using the finite element method
(FEM) programme Ansys. The finite element model of test 1.1 is shown in figure 6. Figure 6a
shows the tested specimen, figure 6b shows the shell elements used in the section and
figure 6c gives an impression of the mesh used. Here also the theoretical failure mechanisms
1 and 2 are shown. The material model used is represented in figure 7. Hardening is
neglected. The calculations carried out include material and geometrical non-linearity. It
turned out that the effect of geometrical non-linearity is substantial.

358 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


a) b) c)
Figure 6. Finite element model for test 1.1.

Figure 7. Material model.

The load-displacement diagram calculated numerically is shown in figure 8. Unfortunately,


measured results are not available for comparison. The failure load calculated with FEM is
Fnum = 70 kN and is also presented in table 1. Equivalent stresses calculated at failure are
shown in figure 9. Large yield areas can be observed.
Load F (kN)

Vertical flange tip displacement (mm)

Figure 8. Numerical load – displacement diagram test 1.1.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 359


Figure 9. Equivalent stresses at failure (Fnum =70 kN) for test 1.1.

t=40

Figure 10. Finite element model for test 2.1.

The finite element model of test 2.1 is shown in figure 10 and is comparable with that of test
1.1 (figure 6). Again the material model of figure 7 is used. The calculations carried out
include material and geometrical non-linearity. However, in this case it turned out that the
effects of geometrical non-linearity can be neglected. The load-displacement diagram (3) is
not shown in this paper. The failure load calculated with FEM is Fnum = 93 kN and is also
presented in table 1. Equivalent stresses calculated at failure are shown in figure 11. Large
yield areas can be observed in top and bottom flange. Figure 12 shows qualitatively the
numerically determined displacements at failure. The numerically found failure mechanism
corresponds to theoretical mechanism 3 with lines of equal displacement not ending
perpendicular to the flange tip but under an angle of approximately 70o.

Figure 11. Equivalent stresses at failure (Fnum = 93 kN) for test 2.1.

360 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Figure 12. Numerical displacements at failure (Fnum = 93 kN) for test 2.1.

In figure 13 measured strains are compared with numerically determined strains and they
correspond reasonably well, with measured strains extending beyond calculated strains.

The numerically determined failure load underestimates the measured failure load
considerably and therefore the numerical model was refined according to figure 14. In figure
14a the original model is shown. The value d is used in the theoretical models. In the finite
element model, the weld thickness is neglected and the connection plate is shifted compared
to the test. The refined model is shown in figure 14b. Now the welds are represented by a
thicker shell element and the connection plate is in the right position. The numerically
determined failure load is now Fnum = 109 kN, listed between brackets in table 1.
Load F (kN)

FEM
Test

Strain in %
Figure 13. Comparison of equivalent strain at rosette R5 for test 2.1.

Theory

Experiment
Experiment

FEM refined
FEM original

a) b)

Figure 14. Refining the FE model for test 2.1.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 361


UNSUPPORTED CONNECTION - BEAM CONNECTION INTERACTION

Experimental research - Unsupported connection

Unsupported connections, to investigate beam connection interaction, were tested in three


point bending tests as shown in figure 15. These are so called type 4 tests. Several
preliminary tests have been done with different span lengths and yield stresses on IPE330
sections. The experimental failure loads per flange tip are listed in table 2.

Figure 15. Test set-up for unsupported connections – test type 4.

Theoretical research - Unsupported connection

In case the beam is very short, the behaviour will be similar to a supported connection and
connection failure will govern behaviour. For this case the theoretical failure load per flange
tip Fthe,1 can be calculated using formula (1) with β = 2.

In case the beam is very long, the behaviour will be governed completely by beam
behaviour. Using first order plastic theory assuming a plastic hinge at mid span, the failure
load of the beam can be calculated as follows:

M = Mp,b = ¼ Pp L (6)

Pp = 4 Mp,b / L (7)

And with 4 flange tips present:

Pp = 4 Fthe.b (8)

Combining formulas (7) and (8) results in the theoretical beam failure load per flange tip:

Fthe.b = Mp,b / L (9)

Both Fthe,1 and Fthe.b are upper bounds of the theoretical failure load that includes beam
connection interaction. These values are also listed in table 2.

362 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Numerical research - Unsupported connection

Numerical research on unsupported connections has not yet been carried out but is planned.

DISCUSSION

The failure loads for supported flange tip connections are summarised in table 1. In all cases
the lowest theoretical failure load is the one for mechanism 1 as could be expected
considering formulas (1-3). Therefore, the theoretical failure load according to mechanism 1
(formula (1)) will be used for comparisons.

Table 1. Failure loads per flange tip for supported flange tip connections.
Type Test fy [N/mm2] Fexp [kN] Fthe,1 [kN] Fthe,2 [kN] Fthe,3 [kN] Fnum [kN]
1 1.1(2) 340 65 72 82 76 70
2.1(2) 340 118 94 103 96 93 (109)
2.2(1) 340 93 97 108 101 n.a.
2
2.3 289 121 82 90 85 n.a.
2.4 289 101*) 82 90 85 n.a.
3 3.1(2) 340 61**) 78 (45) 88 82 n.a.
*) weld failure before flange failure **) connection plate failure

For test 1.1 the theoretical failure load is 11% unsafe with respect to the experimental failure
load. The numerical failure load is close to the theoretical failure load but also on the unsafe
side (8%) when compared to the experimental failure load.
For test 2.2 also the theoretical failure load is on the unsafe side but by just 4%.
For test 2.1 the theoretical failure load is on the safe side (20%). The numerical failure load
obtained with the original finite element model is however very close to the theoretical failure
load. Refining the finite element model taking the weld geometry into account (figure 14),
improves the numerical failure load, however, to be still on the safe side (8%).
For the tests 2.3 and 2.4 the results seem to be comparable with those of test 2.1, however,
the theoretical failure loads are even more conservative. Note that premature weld failure
governed the failure load of test 2.4.
For test 3.1 connection plate failure determined the failure load and this explains why the
theoretical failure load obtained with formula (1) is too high. Adopting a different failure
mechanism accounting for connection plate failure gives a far too low theoretical failure load.

It can be concluded that further research is necessary to improve the model (failure
mechanism) to obtain the theoretical connection failure load, making use of more tests and
finite element calculations and taking the weld geometry into account.

The failure loads for unsupported flange tip connections are summarised in table 2.

Test 4.1 is a relatively long specimen and it can be seen that beam behaviour governs the
failure mode. The theoretical beam failure load is very close to the experimental failure load, 5%
on the unsafe side. The theoretical connection failure load is relatively high so that the
connection is very strong and does not influence the behaviour.
Tests 4.2 and 4.3 are similar. Here, the theoretical connection failure load and the theoretical
beam failure load are relatively close so that both were expected to influence the behaviour in
this test, leading to an even lower experimental failure load than the theoretical upper bounds.
This is not confirmed by the tests since the experimental failure load exceeds the theoretical
upper bounds (22% and 8%). However, bearing in mind the fact that the theoretical connection

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 363


failure load is probably on the low side (see test 2.3 in table 1) and the real connection failure
load may be expected to be about 120 kN, beam failure has governed the test. And the
theoretical beam failure load is on the safe side (8%). Then, lower spans would be interesting to
investigate. These tests are planned. Also the beam failure load model has to be improved since
it is likely that the plastic mechanism does not form at mid-span but just besides the connection.
This becomes more important for shorter beams.

Table 2. Failure loads per flange tip for unsupported flange tip connections.
Type Test fy [N/mm2] L [m] Fexp [kN] Fthe,1 [kN] Fthe,b [kN]
4.1(1) 340 6.65 39 97 41
4 4.2 289 2.40 106 83 97
4.3 289 2.40 105 83 97

When two proper upper bounds are obtained after modifying the models for connection
failure and beam failure, the next step would be to develop a model predicting the combined
connection and beam interaction failure load, taking the complex stress state in the main
beam flanges at the connection into account.

SUMMARY AND CONCLUSIONS

Flange tip connections between main and secondary beams have been investigated
experimentally, theoretically and numerically to try to obtain design rules. A complex stress
state exists in the flanges of the main beam due to overall beam bending and local
connection behaviour. Simplifications have been made omitting the secondary beam and
investigating the connection only by supporting the main beam over its full length. Several
theoretical failure mechanisms have been considered and compared with experimental and
finite element results. Further research is necessary to improve the connection failure model,
making use of more tests and finite element calculations and taking the weld geometry into
account. Also, the interaction between beam and connection behaviour was investigated for
beams in three-point bending. So far, this interaction has not been found in the tests carried
out. It seems that this interaction only is relevant for relatively short beams. Further research
on short beams is therefore planned. A simple beam failure model was developed which
requires improvement. Further research is necessary to find a model predicting the combined
connection and beam interaction failure load.

NOTATION

a length, a=0.378d
b length, b=1.512d
bf width of the flange
d distance between root radius and flange tip
E Young's modulus
fy yield stress
F load on flange tip
Fexp experimental failure load per flange tip
Fnum numerical failure load per flange tip
Fthe,1 theoretical failure load per flange tip for mechanism 1
Fthe,2 theoretical failure load per flange tip for mechanism 2
Fthe,3 theoretical failure load per flange tip for mechanism 3
Fthe,b theoretical beam failure load per flange tip
k width of connection plate

364 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


L span length
M bending moment
Mp,b plastic moment of the beam
Mp,hp plastic moment per unit length of the connection plate
Mp,fl plastic moment per unit length of the flange
Pp failure load of the beam
r root radius
t thickness
tw web thickness
y axis
z axis
α angle
β factor
ε strain
σ stress

REFERENCES

(1) M.E. van den Munckhof, Onderzoek stalen flenstipverbinding – mechanicamodellen,


report CO 98.12 (in Dutch), Eindhoven Technical University, Faculty of Architecture
Building and Planning, Structural Design, 1998.
(2) J.J.W.J. Houben, Onderzoek stalen flenstipverbinding – proevenserie, report CO
98.15/1 (in Dutch), Eindhoven Technical University, Faculty of Architecture Building
and Planning, Structural Design, 1998.
(3) I. van der Sluis, Onderzoek stalen flenstipverbinding – numeriek onderzoek, report CO
02.01 (in Dutch), Eindhoven Technical University, Faculty of Architecture Building and
Planning, Structural Design, 2002.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 365


366 Connections in Steel Structures V - Amsterdam - June 3-4, 2004
APPLICABILITY OF PJP GROOVE WELDING TO BEAM-COLUMN
CONNECIONS UNDER SEISMIC LOADS

Yoshiaki Kurobane and Koji Azuma, Sojo University, Kumamoto, Japan


Yuji Makino, Kumamoto University, Kumamoto, Japan

ABSTRACT

The applicability of partial joint penetration (PJP) groove welding to


beam-column connections subjected to seismic loads is the main concern of
this study. PJP groove welding can be significantly more economical than
complete joint penetration (CJP) groove welding. However, the PJP groove
welded joints inevitably contain sharp notches at the roots of welds, which may
induce a non-ductile failure, especially when subjected to cyclic loading. Four
full-sized beam-to-column connections were tested under cyclic loads. When
the unfused regions created by PJP groove welds were reinforced by fillet
welds so that the welded joints have sufficient cross-sectional areas,
connections showed sufficient strength to withstand large inter-story drift angle
demands. Strains sustained at points around internal discontinuities were
found to be low because of greater cross-sectional areas of welded joints
compared with the cross-sectional area of the beam flanges. Both the test
results and the fracture mechanics-based assessment demonstrated that it is
unlikely to initiate brittle fracture at these discontinuities. One connection,
which failed prematurely due to a lack of penetration in the PJP welds, also
showed that a brittle fracture was unlikely because tips of unfused gaps were
at a low level of stress triaxiality.

INTRODUCTION

The 1994 Northridge and 1955 Kobe Earthquakes revealed that beam-to-column
connections in steel moment resisting frames were susceptible to brittle fracture. In
Northridge brittle fractures most frequently occurred at the beam bottom flange groove
welds with cracks initiating at roots of the welds. Unfused regions between steel
backup bars and column flanges created sharp notches at the weld roots. In Kobe
cracks frequently initiated at toes of the beam copes. The occurrences of these cracks
were found to be reduced by improving profiles of the beam copes. Cracks also
frequently emerged from notch roots formed by the steel weld tabs and beam flanges at
the terminations of CJP groove welded joints. The improvements of the weld tab regions
were found to be made by replacing the steel tabs by flux tabs or removing the steel tabs
after welding and then grinding smooth the ends of the welded joints. However, the
improvement of the weld tabs only is insufficient to avoid a brittle fracture initiating from
the terminations of the CJP welds at the beam ends. Further improvements in welded
joints are required by some means.

The shop-welded connections proposed here utilize PJP groove welds with reinforcing
fillet welds for the welded joints at the beam flange ends. The beam webs are
fillet-welded to the column flanges. Backing bars and beam copes are unnecessary for

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 367


these connections. After flux weld tabs are removed, superimposed fillet welds are
returned continuously around the edges of the beam flanges. Therefore, by using PJP
welding, improvements can be made on 3 representative locations where a brittle fracture
was susceptible to occur. The cost of these improvements is minimal. When an
appropriate size of reinforcing fillet welds is selected, the strength demand per unit area
on the welded joints can be lower than the strength demand on unreinforced CJP welded
joints. The past investigations into the applicability of PJP groove welded joints to
beam-to-column connections supported these possible advantages over conventional CJP
welded connections (1,2).

PJP groove welded joints, however, inevitably contain discontinuities that may act as
sharp notches. Thus, the design strength of PJP groove welds recommended in the
Eurocode 3 (3) and AISC LRFD Specification (4) are generally equivalent to the design
strength of fillet welds. Eurocode 3 allows that a tee-butt joints, consisting of a pair of
PJP groove welds reinforced by superimposed fillet welds, is calculated as a CJP groove
weld, if the total nominal throat thickness is not less than the thickness t of the part
forming the stem of the tee joint, although the stringent limitation that the unwelded gap is
not more than (t/5) or 3 mm is imposed. The AIJ LSD Recommendations (5), in contrast,
specify the design strength of PJP welds equivalent to that of CJP welds.

The design strength according to AIJ is based on extensive test results for plate-to-plate
welded joints. However, details of the joints and loading conditions are different between
the beam-to-column joints and plate-to-plate joints. Especially, the former joints are
subjected to cyclic loading. Further experimental investigations are required to establish
a reliable design methodology for PJP welded beam-to-column joints. This paper
presents test results for two connections, previously reported, and for two additional
connections. Based on these test results, the behavior and design of PJP welded
beam-to-column connections will be discussed hereafter.

SPECIMENS AND TESTING PROCEDURES

Full-size beam-to-column connections with PJP groove welded joints were tested to
failure under cyclic loads. Four specimens, two with wide-flange section columns,
designated as H1 and H2, and two with RHS columns, designated as R3 and R4, were
tested. All the specimens were of one-sided connections, reproducing beam-to-exterior
column connection assemblies (See Fig.1).

WF Column 9 RHS Column 8


(414x405x18x28) (400×400×12)
8
8 5

Beam flange
Beam flange
7 8

8
8

7 8
H1 R3
Stiffener Diaphragm
(PL-16) 12 (PL-25) 10
12 4 12

12 4 10

Beam flange
Beam flange
10
12

WF Beam WF Beam
(500x200x10x16) 12 (500×200×10×16) 10
H2 R4

Figure 1. Configuration of specimens.

368 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Beam top flange Column

Column flange
Beam top flange
flange

Beam web

(a) Specimen H1 (b) Specimen H2


Figure 2. Macro-etch cross sections of PJP welds.

The beams are wide-flange sections with the nominal dimensions of 500x200x10x16 mm
and with the steel grade JIS (Japanese Industrial Standard) SN490B (specified minimum
yield strength=325 N/mm2) for all the specimens. The beam top and bottom flanges
were PJP groove-welded to the column flanges using the single-bevel grooves with the
groove angle of 45 degrees. The depth of the grooves are 8 mm (one half of the nominal
flange thickness) for Specimens H1 and R3 and 4 mm (one quarter of the nominal flange
thickness) for Specimens H2 and R4. Reinforcing fillet welds were added all around the
perimeters of the beam sections contacting the column faces. The nominal dimensions
of the welds are shown in Fig.1.

Incomplete penetration of PJP groove welds was found in Specimens H1 and H2 after
testing. Sliced sections of the welded joints are shown in Fig. 2. Portions of the weld
metal near the root of the weld are not fused with the base metal in Specimen H1. The
lack of penetration from the root of the joint was of about 5 mm for Specimen H1 and less
than 1 mm for Specimen H2. These specimens were welded in a flat position.
Preliminary trials showed that better penetration of the weld metal was achieved by
welding in a horizontal position than in a flat position. For Specimens R3 and R4 the
beam ends were welded in a horizontal position after placing the columns horizontally.
The penetration of the weld metal was satisfactory for these specimens. The GMAW
(Gas Metal Arc Welding) electrodes used for fabricating Specimens H1 and H2 were of
the grade JIS YGW11 with the specified minimum yield strength of 390 N/mm2, while
those for fabricating Specimens R3 and R4 were of the grade JIS YGW18 with the
specified minimum yield strength of 430 N/mm2.

Table 1. Material properties.


Coupon Thicness Yield strength Ultimaate tensile Elongation
Specimen
extracted from (mm) (N/mm2) strength (N/mm2) (%)
H1, H2 Colum flange 28.6 374 518 27
H1, H2 Colum web 19.4 352 509 22
H1, H2 Beam flange 16.3 323 510 22
H1, H2 Beam web 10.4 361 514 19
R3, R4 Column 12.2 404 474 22
R3, R4 Beam flange 15.9 368 526 25
R3, R4 Beam web 9.7 402 541 20
R3, R4 Weld metal - 486 617 26

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 369


Table 2. Charpy impact test results.
Material VE0 (J) VTre (degree C) VEshelf (J)
Weld (H1, H2) 119 -8.6 194
Base(R1, R2) 221 -46 239
HAZ (R1, R2) 254 -102 246
Weld(R1, R2) 127 -47 147
E
V 0 = absorbed energy at 0 degree C
T
V re = energy transition temperature
VEshelf = absorbed energy at upper shelf

u1

L=2400

td u2
Hd

v1 v2

Figure 3. Test set-up.

Specimens H1 and H2 have continuity plates welded to the column flanges and webs at
the positions of beam flanges. Specimens R3 and R4 have internal diaphragms CJP
welded to the columns at the positions of beam flanges.

Material properties determined by tensile coupon tests and measured plate thicknesses
are summarized in Table 1. All-weld-metal tension testing was conducted for the
electrode YGW18, the results of which are also included in Table 1. The notch
toughnesses of materials used were measured by Charpy impact testing and summarized
in Table 2. Charpy specimens were taken from plates welded under the conditions close
to those used for fabricating connection specimens. Notch roots were located at the
base metal, weld metal and HAZ.

The both ends of the column were fixed to a strong floor, while the cyclic shear load was
applied at the end of the beam by a double action hydraulic ram. The loading
arrangements are shown in Fig. 3. Lateral bracing systems were provided at two
positions of the beam.

Displacement measurements taken were not only the horizontal displacement u1 at the
loading point but also the vertical displacements v1 and v2 at the column face where the
continuity plates or diaphragms exist and the horizontal displacement u2 at the column
end. The rotation of the beam θf was calculated by the following equation.

U1 − U2 V1 − V2
θf = − (1)
L Hd − t d

where L and (Hd-td) denote the distance from the loading point to the column face and the

370 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


distance between the two continuity plates or diaphragms, respectively.

The load was applied in the following sequences: firstly a few cycles of loads were applied
in the elastic region and then the amplitude of the beam rotation was increased as 2θp,
4θp, 6θp,,…with 2 cycles of load application at each amplitude, up to failure. θp is herein
defined as the elastic component of the beam rotation when the beam moment at the
column face reaches the fully plastic moment Mp. The elastic rotations measured were
slightly different from those calculated using a simple beam theory. The elastic beam
rotations were obtained from experimental M versus θ relationships. The values of Mp
and θp, calculated using the measured material properties, dimensions and rotations are
shown below.

Specimens H1 and H2 Mp = 708 kNm θp = 0.00659 radians


Specimens R3 Mp = 793 kNm θp = 0.00847 radians
Specimens R4 Mp = 793 kNm θp = 0.00882 radians

HYSTERETIC BEHAVIOR AND FAILURE MODES

The moment versus rotation hysteresis loops measured during the test are shown in Fig.
4. The moment is the moment at the column face Mf and is non-dimensionalized by the
full-plastic moment Mp of the beam. The moment takes a positive value when the bottom
flange is in tension. The rotation θf, which was determined by Eq. 1, is also
non-dimensionalized by θp of the beam. Several failure events observed during the test
are described below. The numbers shown in the hysteresis loops indicate the occasions
when these events occurred.

Specimen H1
The beam top flange ruptured along the weld toes (1). The load was reversed. A
ductile crack was found at the toes of the welds at the edge of the beam bottom flange (2).
Local buckles were found in the beam top flange (3). The cracks in the beam bottom
flange grew rapidly (4). Complete rupture of the beam bottom flange (5).
Specimen H2
The beam top flange buckled locally (1). Buckles of the beam top flange and web
became large (2). The beam bottom flange buckled extensively (3). Small cracks were
found along the weld toes in the bottom flange.
Specimen R3
The beam top flange buckled locally (1). A crack was initiated along the weld toes in the
top flange (2). The bottom flange buckled locally. The crack in the top flange extended
over 1 mm. (3). Local buckles of the top flange grew (4). Local buckles of the bottom
flange grew and cracks in the top flange extended significantly (5). The top flange
buckled accompanying lateral buckling (6). The bottom flange buckled accompanying
lateral buckling (7).
Specimen R4
The beam bottom flange buckled locally (1). Two cracks were found at the weld toes in
the beam bottom flange (2). A crack initiated at the weld toe in the bottom flange (3).
Local buckles of the top flange and web progressed. The cracks in the top flange
extended (4). The top flange buckled locally and laterally. The cracks in the bottom
flange grew extensively (5). Extensive lateral buckling of the beam (6).

Specimen H1 sustained a premature rupture of the welded joints owing to incomplete


fusion of the weld metal (See Fig. 5). For the remaining 3 specimens the ultimate limit
state was governed by combined local and lateral buckling of the beams. In these 3

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 371


specimens ductile cracks invariably initiated at the weld toes at the edges of the beam
flanges and grew stably along the toes of the welds on the beam flange side (See Fig. 6).
2 2
4 1.5 1 2
1.5 3
2 4
1
1
5
0.5
0.5

Mf/Mp
0
Mf/Mp

0
-0.5
-0.5
-1
1
-1 3
-1.5
-1.5 -2
-4 -2 0 2 4 6 8 10 -12 -8 -4 0 4 8 12
θ f /θ p θ f /θ p

(a) Specimen H1 (b) Specimen H2


1.5 1.5
1 4 3
6 5
1 1

0.5 0.5
Mf/Mp
Mf/Mp

0 0

-0.5 -0.5

6
-1 -1
7 2 1
3 4 2
5 -1.5
-1.5
-10 -8 -6 -4 -2 0 2 4 6 8 10 -10 -8 -6 -4 -2 0 2 4 6 8 10
θ f /θ p θ f /θ p

(c) Specimen R3 (d) Specimen R4


Figure 4. Moment versus rotation hysteretic curves.

Figure 5. Specimen H1 after failure.

372 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Figure 6. Ductile tear observed in R4.

ROTATION CAPACITY

Although several criteria have been proposed for the assessment of plastic deformation
capacities of beam-column assemblies, cumulative plastic deformation factors for the
connections discussed herein are evaluated and shown in Table 3. The values of
cumulative plastic deformation factors evaluated from the hysteresis loops are
comparable to those for connections with improved details (6) and meet ductility
requirements for the seismic design. Specimen H1 failed prematurely and is the
exception to the above statement. The definition of the cumulative plastic deformation
factor can be found in other literature (for example, reference (7)).

Table 3. Summary of test results and predictions.


Test results Prediction
Cumulative
Specimen Maximum moment Maximum moment Test/Prediction
plastic
at column face at column face
deformation
(kNm) (kNm)
factor
Top flange -784 -784 1.00
H1 20.5
Bottom flange 1016 781 1.30
Top flange -975 -988 0.99
H2 64.1
Bottom flange 1009 988 1.02
Top flange -1016 -959 1.06
R3 69.0
Bottom flange 1000 959 1.04
Top flange -1014 -959 1.06
R4 61.9
Bottom flange 1046 959 1.09

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 373


EVALUATION OF ULTIMATE STRENGTH OF CONNECTIONS

Two failure modes were identified in this series of tests: 1. combined local and lateral
buckling of beams; and 2. tensile failure of welded joints including ductile crack growth.
The ultimate strengths are evaluated focusing only on the latter failure mode.

The ultimate strengths of the welded joints at the beam flange ends are calculated mainly
using the recommendations by AIJ. The resistance factor is excluded. The AIJ
recommendations show no formula for calculating the strength of a welded joint with
combined PJP and superimposed fillet welds. Although Eurocode 3 shows a design rule
that this type of joint can be calculated as a fillet weld with deep penetration, this rule
gives rather conservative predictions as compared with test results.

a2

a1

Column Beam flange


flange

a3
Beam web

Figure 7. Effective throats of PJP and fillet welds.

The ultimate strength of the welded joint is calculated as the sum of the ultimate strengths
of one PJP weld and two fillet welds, each being calculated as an independent weld (See
Fig. 7). Namely, the ultimate tensile strength of the welded joint Pu is calculated by

1.4 1.4
Pu = a1l1 Fw,u + a 2l2 Fw,u + a3l3Fw,u (2)
3 3

where the symbols a and l denote the effective throat thickness and length of the weld,
respectively. The subscripts 1, 2 and 3 distinguish among the PJP weld, superimposed
fillet weld and fillet weld on the opposite side of the groove. The factor 1.4 in Eq. 2 is to
increase the design strength of the fillet weld when the axis of the weld is perpendicular to
the direction of loading. Fw,u represents the ultimate tensile strength of the weld metal.
The ultimate strength of the welded joint is governed also by the tensile strength of the
beam flange given by

Pu = BfltflFfl,u (3)

where Bfl, tfl and Ffl,u denote the width, thickness and tensile strength of the beam flange,
respectively. Pu takes the smaller value of those given by Eqs. 2 and 3.

374 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


The ultimate flexural strength of the connection at the column face Mf is calculated by

M f = Pu ( Hb − t f ) + mWweb,net Fweb ,y (4)

in which Wweb,net signifies the plastic section modulus of the net area of the beam web
considering reduction of the cross section due to the cope holes (no cope hole exists in
the present examples) and Fweb,y represents yield strength of the beam web, while the
symbol m represents the dimensionless moment capacity of the welded web joints. The
second term of the right-hand side of Eq. 4 represents the ultimate flexural load carried by
the welded web joint considering flexibility of the column flange. Further details of Eq. 4
are referred to elsewhere (7,8).

The ultimate flexural capacities calculated by Eq. 4 are summarized in Table 3. When
evaluating ultimate capacities of connections, the measured dimensions, including the
sizes of the welds, and material properties were used. All-weld-metal tension testing of
the electrode YGW11 was not conducted. However, extensive investigations on material
properties of weld metal have already been conducted (9). Especially, many data exist
for welded joints with the base metal of SN490B and the filler metal of YGW11. From
these data, the data for welded joints having the same heat-input as that for the
specimens H1 and H2 were selected. The material properties for the electrode YGW11
that were inferred from the existing data and used for the calculation are: Fw,y=450 N/mm2
and Fw,u=550 N/mm2. The ultimate capacity was governed by the strength of the weld
metal only in Specimen H1. The ultimate capacities of the Specimens H2, R3 and R4
were determined by the strength of the base metal (Eq. 3).

In Table 3, the rows “top flange” and “bottom flange” distinguish the cases when the top is
in tension and when the bottom flange is in tension. As for Specimen H1 the top flange
ruptured at the same load as the load predicted. When the load was reversed, the
connection supported a much higher load than the predicted load. The reason for this
increased resistance is not known. The neutral axis may have moved after failure of the
top flange. Specimen H2 showed about the same ultimate loads as the predicted
capacities. This specimen sustained extensive buckling but cracks found were small.
On the other hand, Specimens R3 and R4 sustained not only extensive buckling but also
significant crack growth. The flanges of these two specimens seemed to have nearly
reached their tensile capacities. The experimental ultimate loads are a little higher than
the predicted ultimate loads. This underprediction is understandable because Eq. 4 is
ignoring the effects of biaxial stress state at the beam flange ends and cyclic hardening of
materials on the flexural strength of connections.

ASSESSMENT OF EFFECTS OF DISCONTINUITIES ON BRITTLE FRACTURE

The possibility of a brittle fracture initiating from various discontinuities contained in


Specimen H1 and H2 was assessed using the CTOD design curve approach
recommended by JWES (10). The results of this assessment have already been
reported by Azuma and co-authors. (2). The assessment showed that the required
fracture toughness values for preventing a brittle fracture from the roots of the PJP welds
of these specimens are much lower than the fracture toughness of the material used.
Rather, the possibility of the brittle fracture is higher at the weld toes at the edges of the
beam flanges than at the weld roots, although the required notch toughness is still
significantly lower than the toughness of the material even at the beam flange edges. In
conclusion, the previous investigation indicated that a brittle fracture is unlikely to occur in
these specimens. Although the similar assessment on Specimens R3 and R4 has not

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 375


been completed yet, the same conclusions as those on Specimens H1 and H2 could be
drawn judging from the present test results and the material properties shown in Table 2.

However, the previous investigation used the nominal dimensions of the welds for
modeling the specimens. A significant lack of penetration was observed in the PJP
welds of Specimen H1 as was mentioned earlier. A reanalysis was conducted using
measured dimensions of the welds of Specimen H1. The software used is the ABAQUS
(2003) general purpose FE package.

Figure 8. Contour plot of von Mises’s equivalent stress in cross section on


the plane of symmetry.

Figure 9. Contour plot of equivalent plastic strain in cross section on plane of symmetry.

376 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


As seen in the von Mises’s equivalent stress distributions on the plane of symmetry shown
in Fig. 8, stresses concentrate at the two notch roots locating at the tip of the unfused
region. A large area of the base metal near the interface with the weld metal is fully
yielded at an early stage of loading. As the load is increased further, a slip band in which
the von Mises’s equivalent strains reach the order of 20 % is observed in the base metal
(See Fig.9). Note that the weld metal is overmatching by about 30 % in yield strength.
This band extends diagonally from the unfused region to the surface of the beam flange.
Figure 10 shows the J integral versus stress triaxiality Ts curve calculated at one of the
notch roots. The stress triaxiality Ts is defined as
σh
Ts = (5)
σ eq

in which σh denotes the hydrostatic stress and σeq denotes the Mises’s equivalent stress.
The triaxiality represents the level of plastic constraint at crack tips. As seen in this
figure, the stress triaxialiy first reaches a maximum value of about 2.0 under small scale
yielding conditions and then decreases as yielding progresses, showing that the
constraint is at a very low level under fully yielded conditions. The J integral reached
about 250 N/mm at the stage when the connection ruptured during the test.

2.5

1.5
Ts

0.5

0
0 50 100 150 200 250
J (N/mm)

Figure 10. Stress triaxiality Ts versus J integral curve.

The numerical analysis demonstrated that the roots of the PJP welds are nearly under a
plane stress state. The loss of constraint leads to enhanced resistance to the cleavage
fracture. Although the notch toughness of the base metal of Specimen H1 is not
measured, the J integral of 250 N/mm is not large enough to induce cleavage fracture
(See reference (11)). Therefore, it is unlikely to have a brittle fracture starting from the
roots of the PJP welds. If a ductile tensile failure occurs after extensive yielding of the
joint, the ultimate strength of the joint can be predicted based on a simple plastic analysis,
like the analysis presented in the previous section.

CONCLUSIONS

Testing of 4 full-scale beam-to-column connections was conducted under cyclic loads.


The connections have PJP welded joints with reinforcing fillet welds between the beam
flanges and column flanges and two sided fillet welds between the beam webs and

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 377


column flanges.

Three connections of these four showed sufficient strength and plastic deformation
capacity. Plastic hinge formed at the beam ends accompanying local and lateral
buckling of the beams. Ductile cracks initiated at the weld toes at the edges of the beam
flanges and grew stably. The previous investigation indicates that it is unlikely that brittle
fractures start from the roots of the PJP welds or cracks at the edges of the beam flanges.

One connection sustained a ductile tensile failure prematurely. This premature failure
was caused by the lack of penetration in the PJP welded joints. It was found important to
select an appropriate joint detail, welding position and other welding conditions for
achieving sufficient penetration in PJP welded joints.

The evaluation of the ultimate strength of the connections based on a simple plastic
analysis was found to be accurate and applicable to the connection design.

A non-linear FE analysis showed that tips of unfused gaps existing in PJP welded tee-butt
joints are not subjected to high plastic constraint. The stresses were not high enough to
induce a brittle fracture starting from the tips.

ACKNOWLEDGEMENTS

The proposals made here are based on joint experimental programs between Sojo and
Kumamoto Universities. The authors wish to thank H. Shinde, M.Sc. student and the
other 4th year students for their hard work in laboratories. This work was partly supported
by the Japanese Society for the Promotion of Science Grant-in-Aid for Scientific Research
under the number 13650645.

REFERENCES

(1) Azuma, K., Kurobane, Y. and Makino, Y., (2000). Cyclic testing of beam-to-column
connections with weld defects and assessment of safety of numerically modeled
connections from brittle fracture. Engineering Structures, Vol. 22, No. 12,
pp.1596-1608.
(2) Azuma, K., Kurobane, Y., Dale, K. and Makino, Y., (2003). Full-scale testing of
beam-to-column connections with partial joint penetration groove welded joints.
Tubular Structures X, M.A. Jurrieta, A. Alonso and J.A. Chica eds., Balkema, Lisse,
The Netherlands, pp. 419-427.
(3) CEN, (1992). Eurocode 3: Design of steel structures, ENV 1993-1-1: Part 1.1
General rules and rules for buildings. Comité Européen de Normalisation, Brussels,
Belgium.
(4) AISC, (2000). Load and resistance factor design specification for structural steel
buildings. American Institute of Steel Construction, Chicago, Ill., USA.
(5) AIJ, (1998). Recommendation for limit state design of steel structures. Architectural
Institute of Japan, Tokyo, Japan. (in Japanese)
(6) Kurobane, Y., Azuma, K. and Makino, Y., (2003). Fully restrained beam-to-RHS
column connections with improved details. Tubular Structures X, M.A. Jurrieta, A.
Alonso and J.A. Chica eds., Balkema, Lisse, The Netherlands, pp. 439-446.
(7) Kurobane, Y., Packer, J.A., Wardenier, J. and Yeomans, N.F., (2004). Design guide
for structural hollow section column connections. CIDECT/Verlag TÜV Rheinland,

378 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Köln, Germany. in press
(8) AIJ, (2001). Recommendations for the design of structural steel connections.
Architectural Institute of Japan, Tokyo, Japan. (in Japanese)
(9) Mukai, A., Nakano, T., Okamoto, H. and Morita, K.,( 2000). Investigation on MAG
welding wires for building. Journal of Constructional Steel, Japanese Society of Steel
Construction, Vol. 7, No. 26, pp. 13-25. (in Japanese)
(10) JWES, (2000). Method for assessment of brittle fracture in steel weldments subjected
to cyclic and dynamic large straining, WES-TR2808, The Japan Welding Engineering
Society, Tokyo, Japan. (in Japanese)
(11) Iwashita, T., Kurobane, Y., Azuma, K. and Makino, Y., (2003). Prediction of brittle
fracture initiating at ends of CJP welded joints with defects: study into applicability of
failure assessment diagram approach. Engineering Structures, Vol. 25, Issue 14, pp.
1815-1826.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 379


380 Connections in Steel Structures V - Amsterdam - June 3-4, 2004
SHEAR LAG EFFECTS IN FILLET-WELDED TENSION
CONNECTIONS OF CHANNELS AND SIMILAR SHAPES
Matthew J. R. Humphries, Halsall Associates Ltd., Toronto, Canada
Peter C. Birkemoe, University of Toronto, Canada

ABSTRACT

The strength behavior of tension members is largely governed by connection


parameters. For many steel tension members the phenomenon known as
shear lag is often the principal reduction of the ultimate member strength. The
sections were designed so that a comparison of the results would illustrate the
effect of various design parameters on the shear lag behavior.
A review of previous experimental research suggested that some common
parameters had not been examined. These, including weld strength, weld size,
member length, member size and connection symmetry were studied in an
experimental series of tensile specimens that were designed using the current
CAN/CSA S.16-01 standard. These rules attempt to treat shear-lag reduction
in both welded and bolted construction similarly. The test results permit
comparison of the behaviors of single plate connections to channels, angles
and hollow structural sections in various configurations.
A portion of this study sought to develop a simple means of measuring high
strain near fracture using photometric techniques in an attempt to characterize
the state of strain at ultimate load to permit better analytical modeling of
rupture behavior.
The results of this research suggest that current design standards are
generally conservative with a few cases of unconservatism; recommendations
for more explicit interpretation of the design requirements are made to improve
the estimation of ultimate strength as a limit state.

INTRODUCTION

The design philosophy for tension members is largely an attempt at replication of the basic
tensile behavior found in a tensile coupon and early tension members were made as pinned
ended where the reduced member length yielded and eventually ruptured under increasing
axial load. While this behavior may not always be achieved in modern designs of bolted or
welded members, the treatment of the limit state of gross member yielding as a goal and a
secondary limit of ultimate rupture or fracture of the member itself as an additional limit state
is common practice. A connected member may have its behavior controlled by rupture as
influenced by connection parameters and never achieve the gross yield limit state albeit at a
reduced rated capacity or resistance for design.

In addition to the necessary or purposeful reduction of the net cross sectional area of a
member, the additional concerns that are common to bolted and welded connections are
those associated with “shear lag,” “block shear” and “eccentricity.” When the long
connections and the eccentricities of force at the connection are minor, the problems
associated with reduced section capacity are minimized or absent, but shorter connection
lengths tend to be controlled by these reduced capacity failure modes.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 381


When a member develops a cross sectional fracture other than that associated with the
whole tensile cross section, the member is said to fail by “block shear” or “patterned tear
out.” These failures generally precipitate as a combination of shearing and tensile ruptures
that are usually initiated in tension; this failure can have a strong interaction with eccentric
effects. This mode is important because as the length of a connection increases there is a
transition to the “shear lag” mode. When a member fails through the net tensile area, at less
than the full section tensile strength, it is generally the result of “shear lag;” the member
geometry could not develop the full tensile ultimate strength and has suffered a shear lag
reduction. “Shear lag” will also interact with eccentricity if it is large.

The current study is examines the behaviour of simple single plate, fillet welded, connections
to single and paired channels used as tension members. The results of full scale tests on
channels are studied here and compared to tests on members of similar geometry but
consisting of slotted hollow structural sections. An immediate goal is to evaluate these results
to characterize the behavior for design, but a longer term aspect is the measurement of
strain fields during the fracture at ultimate and the anticipated development of reasonable
analytical limits for advanced analysis applied to more complex connection configurations.

In the examination of past work on this and closely related problems the examination of
parameters that were not addressed or varied intentionally, led to a series of tests that
demonstrated the effect of those parameters and helped explain some of the variations in
historical results. This paper will emphasize the parametric effects that were found to
potentially have the largest impact on design and will introduce the ongoing development of
large strain measurement in steel near rupture.

EXPERIMENTAL APPROACH

In all cases, the design of the specimens followed the design or proportioning of practical
tension members. This entails the design of tension members that are of similar length,
strength and cross section as those commonly used in current steel construction. However,
the constraints of the testing equipment imposed a maximum specimen length of about 2
metres. This length was thought to be representative of tension members commonly used in
small trusses. While many tension members would be longer, it was considered reasonable
to assume that the additional length would have an insignificant effect on the results of the
investigation. The length of the tension member was one of the parameters that had been
overlooked in many earlier works while the member and connection proportions were
realistic; economy and equipment limitations were likely influences.

A channel section, (C150 x 12) was selected for two reasons: its strength would be sufficient
that it could be reasonably used in various real applications, and its geometry had a
theoretically significant shear lag effect.

The connection length was chosen as a balance between two competing factors: realism and
efficiency. Excessively short connections did not seem realistic as they did not seem to be on
the order of what might be reasonably designed in practice. To ensure, however, that shear
lag governed the design, the connection length had to be short. A compromise was reached
with a connection length equal approximately to the depth (d) of the member. This produced
reasonable design and a calculated efficiency that would cause the net section strength limit
state strength to govern

The specific designs of specimens were done so that differences in the connection efficiency
could be assumed to result from the specific changes to specimen orientation, symmetry or
geometry. Figure 1 illustrates the variation in specimen design and the philosophy behind

382 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


the experimental program. As well as changes to orientation, symmetry or geometry,
specific changes were also made to weld metal properties and sizes in an attempt to
determine their influence on connection efficiency.

Figure 1. Cross section variation for channel-like tension members to single plates.

All specimens were similar on either connection end and thus possessed a symmetry about
the mid-length. All cross sections were symmetric about at least one axis; the eccentricity of
a member centroid was varied by orientation with the connecting plate as seen in the figure.

TEST SPECIMENS, TEST RIG AND MEASUREMENTS

The test matrix was designed so that for any of the variables being examined, specimens
were tested that varied the parameter of interest at a rational extreme (for example, there
were specimens with high weld metal strength and low weld metal strength). All other details
of the specimens were identical. In this way, a variation in strength between otherwise
comparable specimens was reasonably attributed to the chosen parameter variation.

Two test series were planned; the second was intended to be guided by the results of the
first. A simple naming convention for each series is outlined as follows:

Series 1
The naming convention used the following template: prefix-a-suffix-A or B where
prefix: MH1 for all series 1 specimens
a: 6 or 8, the design weld size
suffix: 480 or 550, the design weld metal strength
A or B: used to identify replicate specimens

Series 2
A modified naming convention used the following template: prefix-abc-suffix-A/B where
prefix: MH2 for all series 2 specimens
a: C or H, section type, C for channels, H for HSS
b: T, B or X, orientation, T for toe to toe channel sections, B for back to back channel
sections and X for HSS sections

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 383


c: S, U, H or L, symmetry/eccentricity, S for symmetric channel specimens, U for
unsymmetric channel sections, H for HSS oriented to maximize eccentricity areas either
side of the connection plate and L for HSS oriented to minimize it.
suffix: 12, 19 or 64 identifier in group, 12 or 19 for channels (section weight in kg/m) and
64 for HSS (thickness in mm *10)
A/B: used to identify replicates of the same design

The baseline connection was a pair of longitudinal welds of length equal to the nominal
channel section depth (150mm). For two specimens (MH1-6-480A and MH2-CBS-19B), an
additional transverse weld was added across the web/gusset at the end of the sections.
HSS sections were slotted and connected similarly to pairs of channels with a pair of
longitudinal welds on each side of the gusset plate. A close up of the baseline connection is
shown as built in Figure 2.

Figure 2. Typical connection “as built”.

Test rig and measurements

All specimens were prepared for a photometric strain measurement technique which was
being evaluated in this study. A visible grid was applied to the surfaces of the specimen in
the region of the connection using contrasting inked dots on a painted surface. Before
loading high resolution pictures of the grid were then taken as a zero reference. Resistance
type strain gauges were also applied to the specimens at their centres at this time. These
strain gauges were used to verify gross section yielding during the testing.

Connection specimens were mounted in an MTS 3000 kN universal test frame using a
special gusset plate assembly to maximize total specimen length. Linearly variable
differential transformers (LVDT) were mounted over the connection length. Load was
applied under displacement control until the initiation of failure that is typically defined by the

384 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


peaking of the applied load. At this point, when possible, the test was held, and a second set
of high-resolution digital pictures was taken of the displaced grid.

A view of one of the connections (boxed channels) showing the initiation of the net section
rupture (top left), the “strain grid,” and indications of yielding.

Figure 3. Close up of boxed channel at failure showing fracture initiation (upper left).

TEST RESULTS

Table 1 summarizes the results of the experimental programme.

Discussion of results

On the basis of these results, several comparisons were undertaken to quantify and observe
the effects of specific variations of connection parameters.

Current design standards do not consider weld size (or connected area) in their estimation of
connection efficiency. This reflects their basis in experimental results from bolted and riveted
specimens. In this research programme specimens were tested whose design efficiencies were
identical, but whose weld sizes were significantly varied. Specimens MH1-6-480B and MH1-6-
480C were connected with 6 mm welds where specimens MH1-8-480A and MH1-8- 480B were
identical except that they were connected using 8 mm welds. Their global behavior is very
consistent as seen in the Load-Displacement plots of Fig. 4.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 385


Table 1. Results.
Design Design CAN/CSA Exper. Observed Exper. Exper./
Specimen Strength Failure S16.1 Ultimate Failure Efficiency Design
Mode Design Strength Mode Efficiency
Efficiency
(kN) (%) (kN) (%)
MH1-6-480B 436 WD 63 749 NS 87 1.38
MH1-6-480C 436 WD 63 802 NS 93 1.47
MH1-6-550A 500 WD 72 804 G-NS 94 1.30
MH1-6-550B 500 WD 72 747 NS 87 1.20
MH1-8-480A 582 NS 82 823 G-NS 96 1.17
MH1-8-480B 582 NS 82 818 NS (top) + 94 1.15
MH2-CBU-12A 582 NS 82 724 BM/NS 93.3 1.14
MH2-CBU-12B 582 NS 82 734 BM/NS 93.1 1.13
MH2-CTS-12A 479 * NS 69.2 * 644 BM/NS 81.3 1.17
351 NS 50.6 1.61
MH2-CTS-12B 479 * NS 69.2 * 642 BM/NS 80.8 1.17
351 NS 50.6 1.60
MH2-CTU-12A 479 * NS 69.2 * 456 NS 57.3 0.83
351 NS 50.6 1.13
MH2-CTU-12B 479 * NS 69.2 * 434 NS 55.3 0.80
351 NS 50.6 1.09
MH2-HXL-64A 962 * NS 72.2 * 610 WD-PRY 79.1 1.10
691 NS 51.9 1.52
MH2-HXL-64B 962 * NS 72.2 * 624 WD-PRY 86.6 1.20
691 NS 51.9 1.67
MH2-HXH-64A 962 * NS 72.2 * 554 WD-PRY 70.5 0.98
691 NS 51.9 1.36
MH2-HXH-64B 962 * NS 72.2 * 576 WD-PRY 72.0 1.00
691 NS 51.9 1.39
MH2-CBS-19A 997 WD 90.4 1094 NS (top) + 84.1 1.10
MH2-CBS-19B 962 NS 73.9 1107 NS 85.0 1.15
MH2-CBU-19B 997 WD 90.4 1222 BM/NS 94.9 1.05
Notes : * (top figure) efficiency calculated based on distance between welds equal to
perpendicular distance between welds; the second values (lower figures) are
efficiencies based on the perimeter between the welds as a suggested interpretation
in some literature.
+ the failure occurred at the top connection which had a considerably greater design
strength
key to failure modes: WD – weld, NS – net section rupture, G-NS – gusset plate net
section, BM/NS – base metal failure beginning at the net section, WD-PRY – prying
failure of the weld (failure began at distal end of welds)

There is clearly a small increase in the efficiency of the connection with increasing weld size.
In fact, the increase shown above (from an average efficiency of 90% to 95%) is the
minimum increase as neither of the MH1-8-480 series specimens failed at the weaker end as
indicated by design but practically reached the ultimate strength of the member section.

Current design standards that account for shear lag reduction in welded tension members do
not account for member symmetry with respect to connection loading. For example, the
design strength of specimens comprised of back-to-back channels using CAN/CSA S16-01
(1) is the same as that for specimens comprised of a single similarly connected channel.
These design codes assume that the effect of the eccentricity at the connection will be small

386 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


or accounted for by the inherent restraint. Several sets of specimens were designed to
determine the effects of this member symmetry on connection efficiency.

Single member specimens without axial symmetry experience realignment (through bending)
of their sections and connections so that the eccentricity of the section and connection is
minimized. Since they have no net eccentricity, the symmetric members do not experience
this realignment to the same degree. Realignment of the section is proposed to have two
competing contributions to the shear lag effect: a decrease in the degree of stress
redistribution required to shed the load into the connection and an increase in extreme fiber
stresses and strains caused by the bending. For small eccentricities, unsymmetric
specimens can have higher net section efficiencies as a result of this realignment. If the
eccentricity of the section is high however, the extreme fiber stress/strain caused by bending
can have a substantial reduction on net section efficiency.
1800

1600

1400

1200

1000
Load (kN)

800 MH1-8-480B
MH1-8-480A
600
MH1-6-480B
MH1-6-480C
400

200

0
0 20 40 60 80 100 120 140 160
Global Displacement (mm)

Figure 4. Load vs. global displacement for specimens with varying weld sizes.

As illustrated by the following results, Fig. 5, overall member symmetry with respect to the
connecting plate can have significant effects on the efficiency in welded tension connections.

The eccentricity of the force remains a significant value even after some realignment and relief.
The amount of reduction in strength is a function of the stiffness of the member and the restraint
that is provided by the connecting ends. The connections here have little rotational restraint
since they are only loaded and attached at the edge of the connecting plate.

The use of slotted HSS sections is popular in modern steel design. It was decided that a
study of the comparative efficiencies of slotted HSS members and similar boxed channels
would provide more information on very similar configurations. Specimens of the series
MH2-HXL-64 and MH2-CTS-12 have this global configuration similarity. It was postulated
that the less restrained boxed channel sections may deform (move towards each other) and
achieve a better efficiency. The results would demonstrate any effect of the apparent
restraint given by the continuity of the HSS that is absent between the two channels as they
enter the connection.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 387


700

600

500
Load (kN)

400 MH2-CTU-12A
MH2-CTU-12B
300 MH2-CTS-12A
MH2-CTS-12B

200

100

0
0 10 20 30 40 50 60 70 80 90
Global Displacement (mm)

Figure 5. Load per channel vs. global displacement for (double channel) symmetric and
(single channel) unsymmetric connections using identical members.

These slotted HSS members did not deform plastically as did the double boxed channel
sections.. This is clearly evident in their efficiency (load) vs. global displacement graphs
shown in Fig. 6. Their stiffness is nearly linear until failure. However, the boxed channel
sections clearly show a high degree of plastic deformation before failure. Their freedom to
deform inward just out side of the connection permits redistribution and this inward
movement was visibly apparent during testing. The efficiency here is the percentage of the
full member yield that was achieved. Further comparisons of efficiencies related to design
are found in Table 1.
1.4

1.2

1
Efficiency (Load/[(Ag)(Fy)])

0.8
MH2-HXL-64B
0.6 MH2-HXL-64A
MH2-CTS-12B

0.4
MH2-CTS-12A

0.2

0
0 10 20 30 40 50 60 70 80 90 100

Global Displacement (mm)

Figure 6. Comparison of the efficiency for slotted HSS and boxed channel sections.

388 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


A weld across the web end of the section provides complete attachment of the web at the
connection according to CSA S16-01 (1). When the section is broken down into elemental
plate sub-sections for shear lag calculation, the web considered to be 100% effective.
However, the weld across the web of the section tends to reduce the ductility of the section in
that it restricts the ability of the web to strain and redistribute stress/strain concentrations in
the section. Adding a weld across the web of the section therefore decreases the relative
portion of the section that is affected by shear lag, but increases the effect of shear lag on
the portions of the section that are affected. On the basis of the provisions of CAN/CSA-S16-
01 (1), the addition of the weld across the web of the section should increase connection
efficiency.

Figure 7. Comparison of load vs. global displacement for geometrically identical


sections with varying weld arrangements.

It is clear from the results of this series of tests that the reduction in ductility caused by the
weld across the web end of the channel had a similar effect as the increase in connected
area caused by the additional weld. In all cases, the sections with the additional weld
exhibited very similar efficiencies to those without it. Clearly the ability for the section to
distribute stress/strain across the section through plastic deformation is as important as the
weld length. The important aspect of this comparison is that the addition of a weld across the
web of the section did not have a significant effect on the efficiency of the connection.

CONCLUSIONS

This experimental study has demonstrated the influence of a few of the parameters in welded
connection design that have an influence on the strength and ductility of tension members.
This data and the detailed displacement/strain distribution data that was not reported here
provide insight and support for new analytical modeling capabilities of the nonlinear response
of complex geometries up to the development of fracture and ultimate strength.

An increase in weld size causes an increase in efficiency in shear lag effected welded
tension connections. This finding seems intuitive since connections with larger welds have

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 389


more connected area, and therefore more area through which to transmit the load into the
connection. Current design standards do not account for the effect of a change of the weld
size except as it affects the weld strength.

Member symmetry (with respect to the load or connection plate) can have a profound effect
on efficiency in shear lag effected welded tension connections. Member sections oriented so
that their individual eccentricity was maximized (boxed channels) with respect to the gusset,
and exhibited large reductions in efficiency when tested singly as a non-symmetric member.
But when tested as part of a symmetric member (double boxed channels) the efficiency was
increased. Many current design standards do not prescribe variations in shear lag reductions
to account for this effect of global symmetry on efficiency. The current standards penalize the
symmetric members in that their design efficiencies are representative of the lower
experimental values for single unsymmetrically loaded member.

Slotted HSS sections although geometrically similar to boxed channel sections in shear lag
affected welded tension connections show distinctively different behavior. The continuity of
the sides of the HSS section beyond the slot restrains shape of the section; the more ductile
behaviour of the channel sections occurred in part as a result of the freedom of the individual
channel halves to realign a small amount. The limited results, shown here, were for HSS that
were not welded at the end of the slot.

In a simple comparison of the results of this series with the simplified empirical shear lag
reduction factor (1-x/L), the results of this study (6) of welded connections are plotted in Fig.
8 with a selection of past research (2, 3, 4, 5) results for various details of a similar nature.
This compilation is not complete but does indicate a trend as well as a significant scatter.
Improvement by reduction of scatter has been accomplished by reinterpreting and changing
the rules of application of the reduction factor. Further analytical and experimental work is
underway to find a non-empirical solution to this problem that will lead to the improved
advanced analysis of strength and deformation for this and for more complex connection
details.

ACKNOWLEDGEMENT

The substantial portion of the support for this study is from a grant from the SSEF (Steel
Structures Education Foundation) and from the NSERC (Natural Sciences and Engineering
Research Council of Canada). All of the tests were performed in the large Structural Test
Facilities at the University of Toronto, Department of Civil Engineering. Support in the form
of material, fabricated specimens and practical technical interaction is greatly appreciated
and crucial to this effort; CANRON Eastern Division Ltd., M&G Steel Ltd, Leroux Steel Inc.,
CoSteel LASCO Ltd., and Hobart Inc. are especially cited for their generous contributions.

REFERENCES

1. Limit States Design of Steel Structures CAN/CSA-S16-01, Canadian Standards


Association, Mississauga, Ontario L4W5N6.
2. Davis, R.P. and Boomslitter G.P. "Tensile Tests of Welded and Riveted Structural
Members," Journal of the American Welding Society, Vol 13(4), 1934, pp 21-27.
3. Easterling W.S. and Gonzalez Giroux, L. "Shear Lag Effects in Steel Tension Members,"
Engineering Journal, AISC, 3rd Quarter, 1993, pp. 77-89.
4. Sarvinis, P. "Shear Lag and Tear-out Considerations in Fillet Welded Connections of
Double Angles," Civil Engineering Bachelors Thesis, University of Toronto, 1989.

390 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


5. Newell, M. P. “Net Section Design for Bolted and Welded Connections: Considerations
of Shear-lag and Tear-out Modes,” Civil Engineering Bachelors Thesis, University of
Toronto, 1989.
6. Humphries, M. “An Investigation of the Shear Lag Effect in Welded Channel Tension
Connections,” Civil Engineering M.A.Sc. Thesis, University of Toronto, 2003.

110
Gonzalez Giroux and Easterling - 1989

Sarvinis - 1989
100
Davis and Boomslitter - 1934

Zanon (Newell 1989)


90
Humphries - 2003
Efficiency (%)

80

70

60

50
0,5 0,6 0,7 0,8 0,9 1
1-xbar/L

Figure 8. Experimental efficiency correlation with empirical shear lag reduction factor.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 391


392 Connections in Steel Structures V - Amsterdam - June 3-4, 2004
EXPERIMENTAL BEHAVIOUR OF STEEL JOINTS
UNDER NATURAL FIRE

František Wald, Czech Technical University in Prague, Czech Republic


Luís Simões da Silva, University of Coimbra, Portugal
David Moore, Building Research Establishment, United Kingdom
Aldina Santiago, University of Coimbra, Portugal

ABSTRACT
Current design models for fire resistance of structures are based on isolated
member tests subjected to standard fire conditions. Such tests provide raw
data for numerical models and to develop design methods, however, they do
not reflect the behaviour of a complete building. Many aspects of behaviour
occur due to the interaction between members and cannot be predicted or ob-
served in isolated tests.
In order to study this global structural behaviour a research project was con-
ducted on the 8-storey steel-concrete frame building at the Cardington labora-
tory. This paper summarises the experimental programme and presents the
thermal and mechanical results of the studied joints.

INTRODUCTION

Significant developments have been made in the analysis of the behaviour of steel framed
structures under fire conditions over the last ten years. Due to the high cost of full-scale fire
tests and size limitations of existing furnaces, these studies are based on the observation of
real fires and on tests performed on isolated elements subjected to standard fire regimes,
which serve as reference heating, but do not model the natural fire. However, the failure of
the World Trade Centre on 11th September 2001 and, in particular, of building WTC 7,
alerted the engineering profession to the possibility of connection failure under fire condi-
tions. Many aspects of behaviour occur due to the interaction between members and cannot
be predicted or observed from isolated tests, such as global or local failure of the structure,
stresses and deformations due to the restraint to thermal expansion by the adjacent struc-
ture, redistribution of internal forces, etc. Otherwise, it is known that even nominally ‘simple’
connections can resist significant moments at large rotation. At the severe deformation of a
structure in fire, moments are transferred to the connection and to the adjacent members,
and hence they may have a beneficial effect on the survival time of structure.

Unlike the standard fire curve, a natural fire is characterized by three phases: a growth
phase, a fully developed phase and a decay phase. It is necessary to evaluate not only the
effect on the structural resistance during the heating phase, but also the high cooling strains
in the joints induced by distortional deformation of the heated elements during the decay
phase.

In order to study this global structural behaviour a collaborative research project (Tensile
membrane action and robustness of structural steel joints under natural fire, European
Community FP5 project HPRI - CV 5535) was conducted on the 8-storey steel-concrete

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 393


composite frame building at the Cardington laboratory during a BRE large-scale fire test. This
collaborative research project involved the following institutions: Czech Technical University
(Czech Republic), University of Coimbra (Portugal), Slovak Technical University (Slovak Re-
public) and Building Research Establishment (United Kingdom). It is the objective of this pa-
per to summarise the experimental programme and to present and to explain the thermal and
mechanical results of the studied joints.

EXPERIMENTAL RESEARCH ON THE FIRE PERFORMANCE OF STEEL JOINTS

Steel joints fire tests

The experimental results on the response of steel connections under fire conditions are relatively
recent and limited, partly because of the high cost of the fire tests and the limitations on the size
of furnace used. Only a few connections tests have been performed and they have concentrated
on obtaining the moment-rotation relationships of isolated connections. It is doubtful whether
these results will be useful when dealing with the behaviour of frame connection. The develop-
ment of the Cardington Laboratory of the Building Research Establishment (BRE) has provided
the opportunity to carry out several research projects that included full-scale fire tests. A brief
state-of-art description of the steel joints fire tests is presented in Table 1.

Table 1. Summary of steel joints fire tests.


Authors Year Objectives
Isolated tests
Six connections types from “flexible” to “rigid” were tested to establish
Kruppa (1) 1976 the performance of high strength bolts. The results indicated that bolt
failure does not occur before gross deformation of the other elements.
A “rigid” moment-resisting form of cleated connections was tested.
BS476 (2) 1982 The results suggested that bolts and their connected elements can
undergo considerable deformation in fire
A programme of fire tests on eight beam-to-column connections, at
elevated temperatures, was developed. The principal aim of those
Lawson et al. (3) 1990
tests was to demonstrate the beneficial effect of moments transferred
to connection in fire.
Leston-Jones A programme was performed to develop moment-rotation relation-
1997
et al. (4) ships for steel beam-to-column connections at elevated temperature.
This work extended the scope of the previous work (5) to include fur-
Al-Jabri et al. (5,6) 1997 ther parameters such as the influence of member size, connection
type and different failure mechanisms.
An experimental programme was conducted to assess the component
Spyrou and
2001 method (9), in order to include the influence of axial stresses due to
Davidson (7) the thermal expansion and contraction.
Fire test of four header plate bolted connections directed to establish
Beneš (8) 2003
the behaviour of the equivalent T stub in tension.
Full-Scale Tests
Seven large-scale fire tests at various positions within the experimen-
Cardington Labo- 1993 - tal building were conducted. The main objective of the compartment
ratory 2003 fire tests was to assess the behaviour of structural elements with real
restraint under a natural fire (9, 10).

The Cardington laboratory

The BRE’s Cardington Laboratory is a unique worldwide facility for the advancement of the
understanding of whole-building performance. This facility is located at Cardington, Bedford-
shire, UK and consists of a former airship hangar with dimensions 48 m x 65 m x 250 m. The
Cardington Laboratory comprises three experimental buildings: a six storey timber structure,
a seven storey concrete structure and an eight storey steel structure.

394 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


The steel test structure was built in 1993. It is a steel framed construction using composite
concrete slabs supported by steel decking in composite action with the steel beams. It has
eight storeys (33 m) and is five bays wide (5 x 9 m = 45 m) by three bays deep (6 + 9 + 6 =
21 m) in plan, see figure 1. The structure was built as non-sway with a central lift shaft and
two end staircases providing the necessary resistance against lateral wind loads. The main
steel frame was designed for gravity loads, the connections consisting of flexible end plates
for beam-to-column connections and fin plates for beam-to-beam connections, designed to
transmit vertical shear loads. The building simulates a real commercial office in the Bedford
area and all the elements were verified according to British Standards and checked for com-
pliance with the provisions of the Structural Eurocodes (11).

A B C D E F
9000 9000 9000 9000 9000
4
6000
3

Fire compartment
9000

2
6 950
6000
10 900
1

Figure 1. The Cardington fire test (9) with identification of the fire compartment.

STRUCTURAL INTEGRITY TEST PROGRAM

Fire compartment

The fire test was carried out in a centrally located compartment of the building, enclosing a
plan area of 11 m by 7 m on the 4th floor.

The mechanical load was simulated using sandbags, each weighing 1100 kg. In addition to
the self-weight of the structure, sand bags represented the following mechanical loadings:
remaining permanent actions, 100% of variable permanent actions and 56% of live actions.

The fire load was provided by 40 kg/m2 of wooden cribs (moisture contents < 14 %) covering
the compartment floor area. The fire compartment was bounded with three layers of plaster-
board (15 mm + 12.5 mm + 15 mm) with a thermal conductivity around 0.19-0.24 W/mK. In
the external wall (gridline 1) the plasterboard was fixed to a 1.5 m high brick wall. An opening
1.27 m high and 9 m long simulated an open window to ventilate the compartment and to al-
low the observation of the behaviour of the various elements. The columns, external joints
and connected beam (1.0 m from the joints only) were fire protected to prevent global struc-
tural instability. The material protection used was 15 mm of Cafco300 vermiculite-cement
spray, with a thermal conductivity of 0.078 W/mK.

Structural arrangement

The steel structure exposed to fire consists of two beam sections (356x171x51 UB for the
edge beams and the 6 m primary beams and 305x165x40 UB for the internal secondary
beams) and two columns (305x305x198 UC and 305x305x137 UC) as shown in figure 2a

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 395


(12). The joints were of a cruciform arrangement of a single column with either three or four
beams connected, respectively, to the column flange or web. The composite behaviour was
achieved by a concrete slab over the beams cast on shear studs. The measured sections
geometry is presented in figure 2b and table 2. Table 3 reproduces the material properties at
ambient temperature of steel and steel connectors.

D E wp c
End plate connection

eob
eo
P8-260x140 Secondary beam

p1 e1
305x165x40UB

eop
2
End plate connection
P8-260x150 Primary beam
aw aw

hp
356x171x51UB
Primary beam
356x171x51UB Secondary beam
305x165x40UB
Fin plate connection g
P10-260x100 Fin plate connection
P10-260x100
e2 e2
Secondary beam N
356x171x51UB
1 End-plate Fin-plate

Figure 2. (a) Arrangement of members in the fire compartment; (b) connections geometry.

Table 2. Connections measured geometry (in mm) (13).


Connection ID hp wp tp e0 e1 p1 e2 g c aw
D1.5 259 100 10 22 27 40 60 50 10 20 6
Fin plate
E1.5 260 100 10 25 25 40 60 50 9 20 7
D2 maj. 260 150 8 20 40 60 30 - - 6
D2 min. 262 140 8 25 40 60 30 - - 6
End plate
E2 maj. 259 150 8 22 40 60 30 - - 6
E2 min. 259 140 8 25 40 60 30 - - 6
Bolts M20

Table 3. Material properties of steel and steel connectors at ambient temperature (12).
Material Ultimate Stress (MPa) Yield Stress (MPa)
nominal measured nominal measured
S275 430 469 275 303
Steel
S355 510 544 355 396
Plate Grade 43 430 ----- 275 -----
Bolts 8.8 800 869 640 -----

Laboratory equipment and instrumentation

The main requirements of the instrumentation were to measure the temperature, the distribu-
tion of internal forces, the deflected shape of the floor and main structural elements. The in-
strumentation included thermocouples, strain gauges and displacement transducers (figure
3). Additionally, ten video cameras and two thermo imaging cameras recorded the fire and
smoke development, the structural deformations and the temperature distribution with time.

396 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


East view North view

460 13 463 454


469 457
466 15 455
461 11
467 464 17
470 100 458
468 50 50 456
471 459
462 9 19
465

Thermocouples (TC)
D2 major axis TC + HSG in Bolt D2 minor axis
HT Strain Gauges (HSG)

Figure 3. Instrumentation of end-plate connections in location D2 (13).

EXPERIMENTAL OBSERVATIONS

Temperature variation

Figure 4 compares the temperatures recorded in the compartment with the parametric curve
presented in Eurocode 1, Annex A (11). The quantity of thermal load and the dimensions of
the opening on the facade wall were designed to achieve a representative fire in the office
building. The maximum recorded compartment temperature near the wall (250 mm from D2)
was 1107,8 ºC after 54 minutes, while the maximum temperature predicted by the parametric
curve was 1078 °C after 53 min., see figure 4.

Temperature, °C
Prediction prEN 1991-1-2, Annex A
1200 1108
Back in fire compartment c d
300
1000 1078 c
500 500
800
Fire compartment (DE, 1-2)
600
Average temperature
400
In front of fire compartment
53 54 d
200
Time, min
0
0 15 30 45 60 75 90 105 120 135 150

Figure 4. Compartment temperature.

Measurements of the temperature in the connections were taken on the beam adjacent to
the connection, in the plate and in the bolts. A summary of the temperatures recorded in the
connections is presented in figures 5, 6 and 7.

In the heating phase, the joint temperature is significantly lower than the remote bottom flan-
ge, which is usually the critical element that defines the limiting temperature of the beam; in
contrast, the cooling down in the joints was slower. At the maximum temperature, the joint
temperature was around 200 ºC lower than the limiting temperature of the beam. The first
bolt row from the top was significantly cooler than the lower bolts, because of shielding by
the adjacent slab and column. The end-plate was hotter than the bolts at the same level due
to the ratio of the bolt diameter to the end-plate thickness (20 mm).

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 397


Temperature, °C D2 E2
1000
c
800 e b
f
ghi N
600
b Midspan bott. flange D1 E1
i Bott. flange
400 f Plate 4th row
c Plate 1st row
200 g Upp. flange
Time, min.
0
0 15 30 45 60 75 90 105 120 135 150

Figure 5a. Temperature variation within the beam-to-beam fin plate connection D1/2-E1/2.

+ 12 mm (elongation in first bolt hole after test)

g
c 30 min.
plate 63
bolts 106 min. 45 min. 90 min. 75 min. 60 63 min. 60 min.
h beam
e
end
f
i

120 400 °C 600 °C 800 °C Temperature, °C


700 °C
- 10 mm (elongation in fourth bolt hole after test)

Figure 5b. Temperature variation in connection D1/2-E1/2;


in 63 min max. temperature of bolt 907.2 °C.

Temperature, °C Midspan beam bottom flange D2 E2


1000 i

defgh i
800
c
600
N
bott. flange g
400 plate 4th row f
D1 E1
4th bolt f
200 upp. flange h
1st bolt c
plate 1st row c Time, min
0
0 15 30 45 60 75 90 105 120 135 150

Figure 6a. Temperature variation within the beam-to-column


minor axes end plate connection D2-E2.

398 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


h

c
60 min. 90 min. 75 min.
d
e plate
bolts 45 min. 106 min.
f
beam 60 min.
106 min.
end
g 45 min. 75 min.
Temperature, °C
400 °C 500 °C 600 °C 700 °C 800 °C

Figure 6b. Temperature variation in connection D2-E2;


in 75 min max. temperature of bolt 811.7 °C.

Temperature, °C g Beam bott. flange D2 E2


1000 f 4th bolt row
c 1st bolt row cdefgh
Crack on one side
800 of end-plate D2 - D1

600 N
D1 E1
400

200

Time, min
0
0 15 30 45 60 75 90 105 120 135 150

Figure 7a. Temperature variation within the beam-to-column


major axes end plate connection D2-D1.

h
c
plate
d bolts
e beam
end
45 min. 106 60 min. 90 min. 75 min.
f

500 °C 600 °C 700 °C 800 °C Temperature, °C


Fracture in HAZ of end plate, see Fig. 10

Figure 7b. Temperature variation in connection D2-D1;


in 75 min max. temperature of bolt 779.1 °C.

Behaviour of the joints

Local buckling of the beam lower flange was one of the main mechanisms. It is observed in
the lower beam flange and web adjacent to the joints, see figure 8, the concrete slab having
restrained the upper flange. This local buckling occurs during the heating phase after about
23 min. of fire (observed by thermo imaging camera), due to the restraint to thermal elonga-
tion provided by the adjacent cooler structure and the structural continuity of the test frame.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 399


The heated lower flange of the beam is unable to transmit the high normal forces generated
in the lower flange of the beam to the adjacent beams/columns after closure of the gap in the
lower part of the connections. The beam could be assumed to behave as ‘simply supported’,
allowing larger mid-span deflections to develop. As temperature and the associated deforma-
tions increase, the shear resistance of the beam web was also reached, see figure 9b.

D2 E2

D1 E1

Figure 8. Local buckling of beam lower flange.

D2 E2 D2 E2

N N

(a) D1 E1 (b) D1 E1

Figure 9. (a) Beam web in shear; (b) buckling of column flange in compression.

The local buckling of the column flange in compression was observed in the major axis
beam-to-column joints, see figure 9b. This behaviour results from the small column flange
thickness (t = 21.4 mm) and the small distance between the bolts, the bolted end-plate be-
having as a welded joint. This behaviour was observed in both columns flanges of the two
beam-to-column joints (D2; E2).

Fracture of the end-plate along the welds was observed, caused by the horizontal tensile
forces during cooling of the connected beam under large rotations associated with flexible
end-plate joints, see figure 10. The fracture occurred along one side of the connection only,
while the other side remained intact. After one side has fractured, the increased flexibility al-
lowed larger deformations without further fracture. This behaviour was observed in both the
major axis beam-to-column joints (D2-D1; E2-E1) and the minor axis beam-to-column joints
(D2-C2).

The elongation of the holes (in top bolt row +12 mm and in bottom bolt row -10 mm in con-
nection D1/2-E1/2) in the beam web in the tension/compression part of the fin plate connec-
tion is due to the associated large rotations, see figure 11. The elongation of the holes oc-
curred on the web of connected beam, while the fin-plate remained intact: the beam web
thickness (6 mm) is smaller than the fin-plate (10 mm); again, the elongation of the holes of
the beam web leads to increased joint flexibility, allowing larger deformations without further
fracture.

400 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


D2 E2
Fracture in
connection D2 - C2 Fracture in
Fracture in connection E2 - E1
connection D2 - D1

(a) connection E2-E1 (b) cracked connections


Figure 10. Fracture of the end-plate along the welds.

D2 E2

D1 E1

Figure 11. Elongation of holes in the beam web in fin plate connection.

CONCLUSIONS

Globally, we did not succeed to reach the collapse of structure or its parts was not reached
2
for the fire load of 40 kg/m , which represents the fire load in a typical office building, to-
gether with a mechanical load greater than standard approved cases. The structure showed
good structural integrity. The test results supported the concept of unprotected beams and
protected columns as a viable system for composite floors.

More specifically, with respect to the performance of the joints, local buckling of the lower
flanges of beams was observed after 23 minutes of fire. Fracture of the end plates occurred
under cooling in the heat affected zones of welds without losing the shear capacity of the
connections. The fin plate connections behaved in a ductile fashion due to elongation of
holes in bearing.

The detailed behaviour of the beams and connections in the real boundary conditions is cur-
rently being investigated to refine the analytical and numerical prediction models, preliminary
published results being available in (14, 15). Further experimental tests on subassemblies
using the boundary conditions measured on the Cardington frame test are currently being
prepared.

ACKNOWLEDGEMENTS

The project has been supported by the grant of European Community FP5 HPRI - CV 5535

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 401


and COST C12. Paper was prepared as a part of project 103/04/2100 of the Czech Grant
Agency

REFERENCES

(1) Kruppa, J. “Resistance au feu des assemblages par boulons haute resistance”, St.
Remy-les-Chevreuse, Centre Technique Industriel de la Construction Metalique,
France, 1979.
(2) “The performance of beam/column/beam connections in the BS476: art 8 fire test”, Brit-
ish Steel (Swinden Labs), Reports T/RS/1380/33/82D and T/RS/1380/34/82D.
(3) Lawson, R.M., “Behaviour of steel beam-to-column connections in fire”, The Structural
Engineer, 68(14), London, pp. 263-271, 1990.
(4) Leston-Jones, L.C., Burgess, I.W., Lennon, T. and Plank, R.J.: “Elevated temperature
moment-rotation tests on steelwork connections”. Proceedings of Institution of Civil
Engineers. Structures & Buildings; 122(4): pp. 410-419, 1997.
(5) Al-Jabri, K.S., Burgess, I.W. and Plank, R.J.: “Behaviour of steel and composite beam-
to-column connections in fire - volume 1”. Research Report DCSE/97/F/7, Department
of Civil and Structural Engineering, University of Sheffield, 1997.
(6) Al-Jabri, K.S., Lennon, T., Burgess, I.W. and Plank, R.J., “Behaviour of steel and com-
posite beam-column connections in fire, Journal of Constructional Steel Research, 46,
pp. 1-3, 1998.
(7) Spyrou, S. and Davidson, J.B., “Displacement measurement in studies of steel T-stub
connections”, Journal of Constructional Steel Research, 57, pp. 647-659, 2001.
(8) Beneš, M.: Equivalent T stub in tension under fire, Ph.D, thesis, CTU in Prague, 2004,
p. 124.
(9) Wald, F., Simões da Silva, L., Moore, D., Lennon, T., Chladná, M., Santiago, A., Beneš
M. and Borges, L.: Experimental behaviour of steel structure under natural fire, The
Structural Engineer (submitted for publication, 2004).
(10) Bailey, C.G., Lennon, T., Moore, D.B.: The behaviour of full-scale steel-framed building
subject to compartment fires, The Structural Engineer, Vol.77/No.8, 1999, p. 15-21.
(11) Draft prEN - 1991-1-2: 200x, Part 1.2: General actions – Actions on structures exposed
to fire, Eurocode 1: Actions on structures, Final Draft, 2002, CEN, European Commit-
tee for Standardization, Brussels, 2002.
(12) Bravery, P.N.R.: Cardington Large Building Test Facility, Construction details for the
first building, Building Research Establishment, Internal paper, Watford 1993, p. 158.
(13) Wald F., Santiago, A., Chladná, M., Lennon, T., Burgess, I.W. and Beneš, M.: Tensile
membrane action and robustness of structural steel joints under natural fire, Internal
report, Part 1 - Project of Measurements; Part 2 - Prediction; Part 3 – Measured data;
Part 4 – Behaviour, BRE, Watford, 2002-2003.
(14) Santiago, A., Simões da Silva, L., Vila Real, P., Franssen, J.M.: Effect of cooling on the
behaviour of a steel beam under fire loading including the end joint response, in Pro-
ceedings of the 9th International Conference on Civil and Structural Engineering Com-
puting, ed. Topping, B.H.V., Civil-Comp Press, Stirling, United Kingdom, paper 65,
2003.
(15) Sokol, Z., Wald, F., Pultar, M. and Beneš, M.: Numerical simulation of Cardington fire
test on structural integrity, in Mathematical and computer modelling in science and en-
gineering, ed. Kočandrlová M., Kelar V., CTU, 27-30.1.03, p. 339-343.

402 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


STRENGTH AND STIFFNESS OF RHS BEAM TO RHS
CONCRETE FILLED COLUMN JOINTS
J. K. Szlendak, Bialystok Technical University, Poland

ABSTRACT
Composite connection made with RHS chord or column filled by concrete and
branches with RHS steel profile are studied herein. The aim of this paper is
deriving a simple theoretical formula for calculating the strength and stiffness
of such joints. Test results of twelve connections in natural scale are
described. Geometry and material properties of the tested joints are given.
Theoretical solution of the joint strength and stiffness are proposed and the
comparisons between theoretical and experimental results are presented.

INTRODUCTION

European Code EC 4 (1) makes possibility to design much more effective structures which
combined advantages of steel structural sections and concrete structures. However many of
structural problems are not included in this regulation. If the steel Vierendeel girder should be
loaded by the significant load the interesting solution is such design where RHS or box
chords section are concrete filling. From the structural point of view the box chords section
ought to have the possibly large dimensions and their wall thickness ought to as small as
possible. However in such situation two problems arise. Local instability of section walls
leads to degradation of chord resistance and very thin walls decrease the strength and
stiffness of joints. It leads to decreasing the overall carrying capacity of such structure. The
strengthening of joints is possible by the steel plate welded to the face of chord. This
however, does not strength the slender webs of box section. The other possibility is concrete
filling of hollow section. Such filling leads also to increasing the thermal capacity of structure
and its fire resistance. The comparison of these two ways of strengthening is given in (2).
Strength and stiffness of T concrete filled joints made with RHS are the aim of this paper.

TEST RIG, TEST SPECIMENS AND MEASUREMENTS

Test rig is shown in figure 1. Twelve joints in natural scale were tested here up to failure.Ten
of specimens, made with RHS, have the concrete filling chords. Two additional specimens
are not concrete filling and are used for comparison how the concrete filling is effective
compare with the pure steel RHS joints. The compression load equal to 420 kN, simulating
the load in real structure, was applied to chord before the branch was loaded. Therefore in
several steps the branch was loaded up to the reach the failure load. After each loading step,
the joint was unloaded to measure the permanent deformations of the tested specimen.
Typical type of joint failure was the inelastic deformation of the flange in the tension zone and
finally cracking of welds, see figure 2.

In Table 1 the geometry of the specimens, mechanical properties, and failure moment are
given. The mechanical properties are the medium value from three tension coupons tests.
The concrete mechanical properties were obtained from tests of five concrete standard
cubes 100x100x100mm. Results obtained shown that the filling concrete has characteristic
stress 42 MPa. Thickness of welds was equal to a = 1,2 tn.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 403


Four LVDT gauges were used to measure the displacements and rotations, see figure 2.
Registrations of the results were made permanently during full loading and unloading
process, up to failure. After each loading step the joint was unloaded to measure the
permanent deformations. For the control of obtained data from LVDT, the additional dial
gauges were used, see figure 1.

Table 1. Geometrical dimensions and mechanical properties.


Geometrical dimensions Yield stress Parameters
RHS RHS chord branch length ultimate
chord branch wall wall branch chord of failure
No of
b o x ho b n x hn thick thick fyn fyo β η λo branch moment
joint
to tn
mm mm mm mm MPa MPa m kNm
BS1 140x140 80x80 5,2 4,3 400 479 0,57 0,57 26,9 0,415 8,51
S2
140x140 100x100 7,1 5,1 380 457 0,71 0,71 19,7 0,415 20,23
steel
BS3 140x140 120x120 7,05 5,15 369 404 0,86 0,86 19,9 0,41 32,80
BS4 140x140 100x100 5,1 5,1 373 457 0,71 0,71 27,5 0,415 14,94
BS5 140x140 100x100 7,05 5,15 380 457 0,71 0,71 19,9 0,405 23,49
BS6 140x140 80x80 7,1 4,4 392 479 0,57 0,57 19,7 0,408 13,06
BS7 140x140 100x100 5,35 4,3 373 457 0,71 0,71 26,2 0,407 14,25
BS8 140x140 100x100 7 4,3 369 457 0,71 0,71 20 0,4 20,00
BS9 140x140 120x120 7 5,2 375 404 0,86 0,86 20 0,41 31,78
BS10 140x140 120x120 5,2 5,2 373 404 0,86 0,86 26,9 0,41 26,24
BS11 140x140 80x80 7 4,15 400 479 0,57 0,57 20 0,411 14,39
S12
140x140 100x100 7,15 4,15 380 460 0,71 0,71 19,6 0,408 16,73
steel

Figure 1. Specimen during test. Figure 2. Joint failure (crack of welds).

404 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


THEORETICAL ESTIMATIONS

Strength prediction

For prediction the theoretical strength of filled joints, from the observations which were done
during experimental tests, the following assumptions are adopted:

xb o

M ip,1,Ed f 2

f f to
bo bn 3 1

d Legend: Φ
φ 1, φ 2 , φ 3
- virtual
0,65h n f rotations in
2
plastic hinges
δ- virtual
ho displacement
hn concrete 0,65hn - range
of the tension
zone

Figure 3. Failure model of joint – yield line mechanism.

1. Yield line mechanism, which is created in the tension zone of joint, is deceived. Erasing
inelastic deformations leads finally to situation that steel loaded flange looses the contact
with filled concrete.
2. In compression zone the connection is almost absolutely stiff. So, for the simplicity could
be assumed that this part of joint is compact.
3. In tension zone range of yield line mechanism is larger then in compression one. From
the tests the assumption is adopted that in tension zone range of yield line mechanism is
equal to 0,65hn

For the prediction of theoretical strength the yield line mechanism is proposed, similar to that
as for unstrengthen steel joints (3). Proposed theoretical model is shown in figure 3.
From the equation that the virtual work dissipated in the hinges by inner forces on the virtual
rotations and deformations is equal to outer forces work on the virtual displacements the
formula to predict strength of joints is given

M ip ,1, Rd 2 8
= + ( x + 0 ,65 ) + 3,08 (1)
η bo m pl x 1− β

From the condition dMip,1,Rd/dx = 0 occurs

1− β
x= (2)
2

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 405


After substitution (2) to (1) the design formula is obtained

M ip ,1, Rd 8 0 ,65
= (1 + ) + 3,08 ( 3)
η bo m pl 1− β 1− β

Initial stiffness

Initial stiffness Sj,ini is a coefficient in the linear function between the bending moment applied
to the joint and its local rotation (M = Sj,ini Φ).

For pure steel joints the power function is assumed to predict the initial stiffness of the joints
when β > 0,4, see (4). Analysis of the influence of particular parameters leads to the
following formula:

Sj,ini = ks E to3β y4η y5λoy6 (4)

After the numerical simulation the followings exponents were obtained: y4 = 2, y5 = 3, y6 = 1.


For eliminating the false results the Chauvenet rule was used (4).
For assumption that the level of confidence will be 0,95 and when coefficient γM5 = 1,1 the
coefficient ks = 6 was obtained. Then, the design value of the joints initial stiffness could be
calculated as below:

Sj,ini = 6 E to3 β 2η 3 λo (5)

However, for the concrete filled joints, after the numerical simulation, the increasing
coefficient 1,3 is suggested and the design value of the joints initial stiffness could be
calculated as below:

Sj,ini = 7.8 E to3 β 2η 3 λo (6)

Secant stiffness

According the recommendations which are given in EC-3, see part 5.1.2 (5), as a
simplification, the rotational, secant stiffness may be taken as Sj,ini /η in the analysis for all
values of the design moment. Therefore, the secant stiffness of concrete filled joints is
suggested to be calculated using coefficient η = 2, see Table 5.2 (5), as below:

Sj,sec = 3.9 E to3 β 2η 3 λo (7)

COMPARISON OF EXPERIMENTAL RESULTS AND THEORETICAL ESTIMATIONS

In figure 4 to 15 the moment-rotation curves (M - Θ) for each tested joints are presented.
They are shown not only loading but also unloading curves registered by LVDT and dial
gauges. Unloading curves gives possibility to obtain the end of its elastic behaviour and
show the arising of the joint permanent deformations. In Table 2 the comparison between the
theoretical prediction and the test results is presented.

406 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


BS1
10 BS3
35

Moment kNm
30
8

Moment kNm
25

6 LVDT Loading
20
LVDT Unloading
LVDT Loading
Dial gauge Loading
LVDT Unloading 15
4
Dial gauge Loading Dial gauge Unloading

Dial gauge Unloading 10 Welds failure

2 Initial stiffness (6) Initial stiffness (6)

Secant stiffness (7) 5 Secant stiffness (7)

Design load (3)


0 0
0 Rotation
1,5 x 10-2 rad
3 4,5 6 0 Rotation
1,5 x 10-23rad 4,5 6

Figure 4. Joint BS1 β = 0.57, λo=26.9. Figure 5. Joint BS3 β = 0.86, λo=19.9.
BS5
BS4
16 27
Moment kNm
Moment kNm

24

21
12
18 LVDT Loading

15 LVDT Unloading

8 Dial gauge Loading


LVDT Loading
12
LVDTUnloading Dial gauge Unloading

Dial gauge Loading 9 Welds failure

4 Dial gauge Unloading Initial stiffness (6)


6
Initial stiffness (6)
Secant stiffness (7)
Secant stiffness (7) 3
Design load (3)
Design load (3)
0 0
0,0 Rotation x1,5
10-2 rad 3,0 4,5 0 1,5 x 10-2
Rotation 3 rad 4,5 6 7,5

Figure 6. Joint BS4 β = 0.71, λo=27.5. Figure 7. Joint BS5 β = 0.71, λo=19.9.
BS6 BS7
15 15
Moment kNm
Moment kNm

12 12

LVDT Loading
LVDT Loading
9 9 LVDT Unloading
LVDT Unloading
Dial gauge Loading
Dial gauge Loading
Dial gauge Unloading
6 Dial gauge Unloading 6
Welds failure
Welds failure
Initial stiffness (6)
Initial stiffness (6)
3 3 Secant stiffness (7)
Secant stiffness (7)
Design load (3)
Design load (3)
0 0
0 Rotation
1,5x 10-2 rad3 4,5 6 0 Rotation
1,5 x 10-23rad 4,5 6 7,5

Figure 8. Joint BS6 β = 0.57, λo=19.7. Figure 9. Joint BS7 β = 0.71, λo=26,2.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 407


BS8 BS9
21 35

Moment kNm

Moment kNm
18 30

15 25 LVDT Loading
LVDT Loading
LVDT Unloading
LVDT Unloading
12 20
Dial gauge Loading
Dial gauge Loading
9 Dial gauge
Dial gauge Unloading 15
Unloading
Welds failure
Welds failure
6 10 Initial stiffness (6)
Initial stiffness (6)
Secant stiffness (7)
3 Secant stiffness (7) 5
Design load (3)
Design load (3)
0 0
0 Rotation
1,5 x 10-23 rad 4,5 6 7,5 0 Rotation
0,5 x 10-21rad 1,5 2 2,5

Figure 10. Joint BS8 β = 0.71, λo=20. Figure 11. Joint BS9 β = 0.86, λo=20.
BS10 BS11
27 15
Moment kNm

Moment kNm
24

12
21
LVDT Loading LVDT Loading
18
LVDT Unloading LVDT Unloading
9
15
Dial gauge Loading Dial gauge
Loading
12 Dial gauge Dial gauge
Unloading 6
Unloading
9 Welds failure Welds failure
Initial stiffness (6) Serie6
6 3
Secant stiffness (7) Serie7
3
Design load (3) Serie8
0 0
0 1,6x 10-2 rad
Rotation 3,2 4,8 6,4 0 1,5
Rotation x3 10-24,5
rad 6 7,5 9 10,5

Figure 12. Joint BS10 β = 0.86, λo=26,9. Figure 13. Joint BS11 β = 0.57, λo=20.
S2 S12
22 21
Moment kNm

Moment kNm

18

16,5
15 LVDT Loading
LVDT Loading
LVDT Unloading
LVDT Unloading
12
Dial gauge
11 Dial gauge Loading
Loading Dial gauge
Dial gauge 9 Unloading
Unloading Welds failure
Welds failure
6 Initial stiffness (5)
5,5 Initial stiffness (5)
Secant stiffness
Secant stiffness 3
Design load
Design load
0 0
0 1,5x 10-2 rad
Rotation 3 4,5 6 0 1,5 x 10-2
Rotation 3 rad 4,5 6 7,5

Figure 14. Joint S2 β = 0.71, λo=19,7. Figure 15. Joint S12 β = 0.71, λo=19,6.

408 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Table 2. Comparison between theoretical and experimental strength of joints.
Theoretical Theoretical
No Experimental
strength strength of Mexp / Mode of
of strength
of joint (3) welds Mip,1,Rd failure
joint Mexp kNm
Mip,1,Rd kNm Mw,Rd kNm
BS1 5,91 13,83 5,75 0,97 yield of chord face
BS3 33,89 30,87 - - welds cracking
BS4 8,70 24,34 9,90 1,14 yield of chord face
BS5 16,94 24,60 17,1 1,01 yield of chord face
BS6 10,81 14,19 10,45 0,97 yield of chord face
BS7 9,57 20,10 10,00 1,05 yield of chord face
BS8 16,21 20,38 16,50 1,02 yield of chord face
BS9 33,95 31,19 - - welds cracking
BS10 18,64 31,19 19,5 1,05 yield of chord face
BS11 10,72 14,19 11,2 1,05 yield of chord face
S2 12,8 24,34 15,2 1,18 yield of chord face
S12 12,98 19,12 14,5 1,11 yield of chord face

Theoretical strength of pure steel joints S2 and S12 is estimated from the formula presented
in (3) but initial stiffness is calculated from formula (5). Secant stiffness of these joints are
equal to Sj,ini /η , where η = 2.

CONCLUSIONS

a. Table 2 shown that formula (3) good predicts the strength of T RHS joints which chords
are filled by concrete.
b. Filled joints could be classified as the joints with full strength (see joints BS 3 and BS 9 in
Table 2), where the parameter β < 0,85 and if the wall slenderness of the chord section
λo is not slender then 20. With regard to the unfilled, pure steel joints, the joints could be
classified as the full strength if they are more compact i.e. parameter β < 1 and λo < 16.
c. As it could be expected, it was noticed that the rotation capacity of filled joints is much
smaller than the adequate steel joint. However, all the tested joints have the rotation
capacity over 1,5 x 10-2 rad, what guarantee to reach the serviceability limit of beam. For
very flexible joints, for example BS 1, the rotation capacity is only equal to 6 x 10-2 rad,
see figure 4, when for the adequate unfilled steel joint it is much larger and equal at least
to 20 x 10-2 rad.
d. After the opening the RHS it was observed very good condition of the concrete. It was not
worse then the similar concrete curing in more wet environment.
e. As it is shown, see formula (6), that the stiffness of filled joints is about 30 % larger then
adequate steel joint. The test results show that such estimation could be accepted.
However, more tests are needed for confirmation of this data.

ACKNOWLEDGEMENT

This research project No W/IIB/10/01 was financially supported by Bialystok Technical


University, Poland

NOTATION

fy yield stress of (fyn – branch member, fyo – chord member)


ks coefficient

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 409


mpl plastic moment of resistance per unit length in the chord face (mpl = fyoto2/4)
y1, y2, y3 unknown exponents
β branch to chord width ratio (β = bn/bo)
η branch depth to chord width ratio (η = hn / bo)
λo slenderness of chord face ( λo = bo/to)

REFERENCES

(1) prEN 1994-1-1 "Design of composite steel and concrete structures". Part 1.1 "General
rules and rules for buildings", CEN, 18 March 2002.
(2) Szlendak J. Improve the joints strength in steel frames with RHS Columns by concrete
filling. Proc. Int. Conf. on “Steel Structures of the 2000’s”, Istanbul, 2000, pp.345-352.
(3) Szlendak J., Bródka J. "Investigation into the static strength of welded T moment
unreinforced joints in rectangular hollow sections“, International Institute of Welding,
IIW-Doc. XV 538-83, March 1982.
(4) Szlendak J.K.: Design models of welded joints in steel structures with rectangular
hollow sections. DSc thesis, Bialystok Technical University Press, 2004 (in Polish)
(5) prEN 1993-1-8 Eurocode 3: Design of steel structures: Part 1.8: Design of joints, CEN,
31 January 2003 (Stage 34 draft).

410 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


STRENGTH AND STIFFNESS OF 3D PLATES TO
RHS COLUMN PIN JOINTS
J. K. Szlendak, Bialystok Technical University, Poland

ABSTRACT
Nominally pinned joints where longitudinal plates are welded to the walls of
RHS column are studied herein. The earlier results of such experiments made
by Jarrett, where tension loading is applied to the joints, are reminded. A
design formula for prediction the strength of joints, more optimistic, than one
given by Jarrett is proposed. Results of experimental tests, where connections
are loaded by the shear forces and secondary bending moments, are further
discussed. Three types of joint failure were observed in tests. Models for these
joint failures are given. For inelastic failure of chord face, proposed formulas
for prediction the strength and stiffness of the joint have not reference to EC-3.

INTRODUCTION

Braced frameworks with nominally pinned joints could be the proper solution for the saving of
labour, manufacturing and erection costs. If in such structures the columns are made with
box section (RHS), and the beam is the I - type profile, one of the easiest joint between them
is the longitudinal plate welded to column, figure 1. This plate usually is connected with web
of beam by using a bolted connection. Such joints are given in Table 7.13 (1). It ought to be
remember that they are very flexible (2, 3) and sensitive on the cracks in the tension area.
So, they should be rather used for the structures with predominantly static loading. Such
joints are usually loaded by shear forces and secondary bending moments. Furthermore,
they should be also able to carry the incidental tension load (2). If a carrying capacity for
such loading is too small the failure of joint, as in figure 2, is available.

When the semi-continuous framing is considered, than the design formulas to predict the
strength, stiffness and rotation capacity of joints ought to be developed. From the reason of
complex behaviour of described here joints such formulas are often semi-empirical.

TENSION LOADING OF JOINTS

Test results

Tests of joints, where the tension load Nt is applied from the I beam web by the steel
longitudinal gusset plate to the wall of RHS, are rather rare. One of it as in figure 2, for 13
joints, has been undertaken by Jarrett (2). In figure 3 results of these experiments for two
joints sign 6T and 12T are shown. For joint 6T n = 0 but for 12T n = 0,52. It is easy to noticed
that the prestressing of chord n = No/Aofyo decreased the ultimate load of joint. It could be
also seen that such joints have a significant “overstrength”, which increases for large
deformations of the loaded flanges of joint, see figures 2 and 3.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 411


Strength prediction

Prediction of the joint strength is given from the yield line bending-squash mechanism, which
occurs on the face walls of the RHS. In this mechanism the initial energy is dissipated not
only by the bending moments but also by the membrane forces in the yield lines, see
Groeneveld (4), Szlendak and Brodka (5) and Szlendak (6).

Figure 1. Geometry of joint with Figure 2. Typical failure of joint (2).


longitudinal plate (2).

250

Joint 6T ( n=0)
200
Joint 12T ( n = 0,52)

150
jt,ult

t
N jt from (10)
e
N

for joint 6T
100

50
t
N jt from (10) for
joint 12T
0
0 5 10 15 20 25 30 35 40 45
Deformation of loaded wall "w" [mm]

Figure 3. Experimental ultimate load Nejt, ult as a function of (w) 6T and 12T, (2) and
theoretical strength Ntjt estimated from (10).

Strength of joint Ntjt, as in figure 1, is estimated from formula:

(
N t jt = 4m pl η C n + 2 Dn 2(1 − n 2 ) ) (1)
where coefficients Cn and Dn are calculated as below:

412 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


for value of parameter
2λo − 1.5
β<
2λo
2 + (G 2 − 1) 2 (1 − β ) 2 λo
2
Cn = ( 2)
G 2 (1 − β )
1
2 + (G 2 − 1) 2 (1 − β ) 2 λo
2

Dn = 3 (3)
G 2 (1 − β )
2.5
G2 = 1 + ( 4)
(1 − β ) 2 λo
2

but when
2λo − 1.5 17λo − 10
≤β<
2λo 17λo
1 + 2 B(1 − β )λo
Cn = (5)
(1 + B )(1 − β )
1
1 + B(1 − β )λo +
3B(1 − β )λo
Dn = ( 6)
(1 + B )(1 − β )
1 + 1 + (1 − β )λo [17(1 − β )λo − 10]
B= ( 7)
(1 − β )λo [17(1 − β )λo − 10]
Furthermore, when
17λo − 10
< β ≤1
17λo
Cn = 2λo , Dn = λo (8)
When the pure bending model occurs, what is the typical assumption (1), the particular
solution is obtained for which Cn is calculated as below:
2
Cn = ( 9)
1− β
If a deformation of loaded wall “w” arises then the strength of the joint increases due to the
membrane effect. As a good approximation could be assumed that for deformations such
that w < to/2, the coefficient increasing the strength of the joint due to the membrane effect is
equal to (1+4(w/to)2), (4). If one assumes that the maximum deformations of the loaded walls
is equal to 1% bo, (3), this coefficient could be written as (1 + 4 (0,01bo/to)2) = (1 + 0,0004
λo2). Then, more optimistic estimation of the joint strength is suggested. Proposed below
formula is the extension of formula (1) and includes the increasing coefficient from the
membrane effect:

[ ](
N t jt = 4m pl η C n + 2 Dn 2(1 − n 2 ) 1 + 0,0004 λ o
2
) (10)
In figure 4 the comparison between this theoretical estimation and the experimental strength
of joints tested by Jarrett (2), is shown. The experimental strength Nejt,pl, for the moment of
creating the yield line mechanism and ultimate strength Nejt,ult, for large deformations and full
membrane action is given for the joints 6T and 12T. As the illustration the results of ultimate
strength Nejt,ult for the other from 13 joints are also included.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 413


140

120

tj,pl/mpl
100
e
Experimental results N

80
Jarrett (2) - ultimate load (13 joints)

60 Joint 6T - design load (10)

Joint 12T - design load (10)


40
Joint 6T - ultimate load

20 Joint 12T - ultimate load

0
0 20 40 60 80 100 120 140
t
Theoretical estimation N tj/mpl

Figure 4. Comparison: theoretical estimation (10) and the experimental strength from (2).

SHEAR LOADING OF JOINTS

Test results

Test rig is shown in figure 5. Six joints were tested here in natural scale.

Figure 5. Test rig.

414 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Three specimens have two gusset plates (on opposite sides of RHS, sign 2D) and three
other four gusset plates (on each side of RHS, sign 3D). Thickness of gusset plates were tn
= 6 mm and the fillet welds 3,5 - 4 mm. Plates were connected, with using 3 M16 bolts grade
8.8., to more stiff plates, which simulate the webs of beams, see figure 8. Step by step
compression force is applied by hydraulic jack to the RHS column, which was inside the rig
tube diameter 406/8.8 mm. Because stiff plates were supported on strengthen walls of rig
tube, the gusset plates of joint were loaded by shear load from the reactions, as in the real
structure.

Eight LVDT gauges were used to measure the displacements and the rotations of each
gusset plate. Registrations of the results were made permanently (one registration per one
second) during the full loading and unloading process up to failure. After each loading step
the joint was unloaded to measure the permanent deformations of the tested specimen.
Three types of joint failure, as described below, were observed in tests.

Table 1. Geometrical dimensions and mechanical properties.


Geometrical dimensions Mechanical properties Parameters
No RHS chord Plate chord chord plate plate
speci- b o x h o x to bn x hn x tn fyo fuo fyn fun β η λo
2 2 2 2
men mm mm N/mm /mm N/mm /mm
3D/1 150x150x10 130x60x6 394 516 308 435 0,04 0,87 15
2D/1 150x150x10 130x60x6 395 513 308 435 0,04 0,87 15
3D/2 180x180x8 130x60x6 374 509 308 435 0,033 0,72 22,5
2D/2 180x180x8 130x60x6 377 512 308 435 0,033 0,72 22,5
3D/3 150x150x5 130x60x6 410 549 308 435 0,04 0,87 30
2D/3 150x150x5 130x60x6 408 542 308 435 0,04 0,87 30

40 Ø17

35

25

Figure 6. Bearing of gusset plate – view. Figure 7. Welds failure – view.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 415


Bearing of plate material failure occurred for
specimens’ 3D/1 and 2D/1. These failure
occurs when column section is compact, here
λo = 15, and welds has the adequate strength.
In such situation connection failed by the
shear of bolts or bearing of gusset plate, see
figure 6 and 9, where the example of moment
- rotation curve is shown. It is obtained for
one of the gusset plates of the joint 3D/1.

Welds failure occurred for specimens 3D/2


and 2D/2. These failure occurs when column
section is compact enough, here λo = 22,5,
and strength of bolts is larger then the
strength of welds. In such situation
connection failed by cracking welds in the
corners, in tension zone of gusset plate, see
figure 7 and 9, where the moment-rotation
curve for joint 3D/2 is shown. Figure 8. Inelastic deformations of RHS
flanges – view.

Flange yield failure occurred for specimens’ 3D/3 and 2D/3. This failure occurs when walls of
section are slender, here λo = 30. Inelastic mechanism arises on the loaded flanges and
sufficient membrane action was noticed, see figure 8. Large permanent deformations were
observed during the unloading process. In figure 10 the examples of moment - rotation
curves are shown. They are obtained for three of the gusset plates of joint 3D/3.

Joint 3D/1 and 3D/2

13
12
11
10
9
Moment kNm

8 Test results for joint 3D/1


7 Test results for joint 3D/2
6 Initial stiffness for joint 3D/1

5 Secant stiffness for joint 3D/1

4 Initial stiffness for joint 3D/2

3 Secant stiffness for joint 3D/2

2 Plate theoretical strength from (12)

1 Weld theoretical strength from (13)

0
0 25 50 75 100 125 150 175 200 225
Local rotation x mrad

Figure 9. Moment-rotation curves (bearing of gusset plate material and welds cracking).

416 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Joint 3D/3

6
Moment kNm

5
Results for plate nr 1
4 Results for plate nr 3
Results for plate nr 4
3
Initial stiffness (19)
2 Secant stiffness (20)
Design load (16)
1

0
0 50 100 150 200 250 300
Local rotation x mrad

Figure 10. Moment-rotation curves (yielding of the loaded flanges of RHS).

Strength prediction

Model 1: Bearing of plate material

Formulas for design resistance for individual fasteners subjected to shear are shown in Table
1, (1). For Category A (bearing type) connections the below relations occur:
Fv,Ed ≤ Fv,Rd
Fv,Ed ≤ Fb,Rd
where,
α v f ub A
Fv,Rd = (11)
γ M2
For strength grades 4.6, 5.6 and 8.8: αv = 0,6
and,
k1 a b f u d t
Fb,Rd = (12)
γ M2
f ub
where αb is the smaller of αd; or 1,0;
fu
Model 2: Welds failure

According the formula (4.1) part 4.5 (1), the resistance of the fillet weld will be sufficient if the
following are both satisfied:
[σ┴2 + 3 (τ┴2 + τ║2)] 0,5 ≤ fu / (βw γM2 ) and σ┴ ≤ fu / γM2 (13)

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 417


Model 3: Yield failure of loaded flanges

Estimation of strength the RHS joints with longitudinal gusset plates is included, loaded by
normal forces, is given in EC-3 (see Table 7.13 (1). However, up to now, the design formula
for such joints, loaded by shear forces and bending moments, is not proposed there.
Theoretical investigations into the static strength of RHS to RHS beam-column welded joints
were undertaken in earlier works see e.g. (3, 6). The new design formula, presented in (6),
could be used in case of joints studied here, when β < 0,2. Proposed formula has easy form
for the direct calculation by the designers as below:

f yo
M t jy , pl = k M f M ( β ,η , λo ) M n pl , y (14)
f yn

From different possible functions fM the power function is chosen. Then, the strength of joint
Mtjy,pl could be calculated from formula:

f yo
M t jy , pl = k M β y1η y 2 λo
y3
M n pl , y (15)
f yn

After the numerical simulations the following exponents were obtained: y1=1/6, y2=1/2, y3= -
4/3. For eliminating the false results the Chauvenet rule was used, see (6).

For assumption that the level of confidence will be 0,95 the coefficient kM = 27,5 was
obtained. Furthermore, when the coefficient γM5 = 1,1 ( in EC-3 (1) it is assumed γM5 = 1)
then the design value of the joints strength could be calculated as below:

βη 3 f yo n
M t
jy , pl = 25 6 M pl , y (16)
λo 8 f yn

Comparison of this estimation with 186 test results of the static strength of RHS to RHS
beam-column welded connections, collected in the data-bank (7), is given in (6). Results of
27 tests are ignored as false if level of confidence will be 0,95. This simplified estimation
good predicts the experimental results (6).

Initial stiffness

Initial stiffness Sj,ini is a coefficient in the linear function between the bending moment applied
to the joint and its local rotation (M = Sj,ini Φ).
The power function is assumed to predict the initial stiffness of the joints when β > 0,4, see
(6). Analysis of the influence of particular parameters leads to the following formula:

Sj,ini = ks E to3β y4η y5λoy6 (17)

After the numerical simulations the following exponents were obtained: y4 = 2, y5 = 3, y6 = 1.


For eliminating the false results the Chauvenet rule was used (6).

For assumption that the level of confidence will be 0,95 and when coefficient γM5 = 1,1 the
coefficient ks = 6 was obtained. Then, the design value of the joints initial stiffness could be
calculated as below:

Sj,ini = 6 E to3 β 2η 3 λo (18)

418 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Comparison of this estimation with 202 test results of the static strength of RHS to RHS
beam-column welded connections, collected in the data-bank (7), shows that this simplified
estimation good predicts the experimental results (6).

However, for the longitudinal gusset plate joints parameter β should follow the condition β <
0,2; see Table 7.13 (1). Even if the thickness of gusset plates changed the value of β is still
very small and influence of that parameter on the initial stiffness of joint is negligible. So, in
the formula (18) it is assumed that β2 = 0,04. Therefore for 0,03 < β < 0,2 the initial stiffness
of longitudinal gusset plate joints could be calculated as below:

Sj,ini = 0.24 E to3 η 3 λo (19)

Secant stiffness

According the recommendations which are given in EC-3, see part 5.1.2 (1), as a
simplification, the rotational, secant stiffness may be taken as Sj,ini /η in the analysis for all
values of the design moment. From Table 5.2 (1) the stiffness modification coefficient is
equal to η = 2. Therefore for 0,03 < β < 0,2 the secant stiffness of longitudinal gusset plate
joints could be calculated as below:

Sj,sec = 0.12 E to3 η 3 λo (20)

Joint 3D/3 (focus)

3,0

2,5

2,0
Moment kNm

1,5
Results for plate nr 1
Results foir plate nr 3
1,0
Results for plate nr 4
Initial stiffness (19)
0,5 Secant stiffness (20)
Design load (16)
0,0
0 5 10 15 20 25 30 35 40 45 50
Local rotation x mrad

Figure 11. Moment-rotation curve (yielding of the loaded flange of RHS)


– details of assumed stiffness and design load of joint.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 419


CONCLUSIONS

a. New design formula (10) for joints loaded by tension load Nt, is suggested. Strength of
the joint increases due to the membrane effect. Proposed estimation is more optimistic,
than one given by Jarrett (2) and in Table 7.13 EC 3 (1). For compact RHS, when λo = 10
this strengthening is negligible and equal to 4%. However for slender wall RHS, when λo
= 30 it is significant and equal to 36%.
b. Joints loaded by shear load and by additional bending moment failed from: bearing of
plate material, shear of bolts and welds cracking. However, the yield failure of loaded
flanges was specially research in this paper. From figure 9 and 10 could be easy noticed
that the rotation of joints is large enough to reach the serviceability limit state. It is
reminded that for typical load the rotation of beam supports about 15 mrad leads to
exceeding this limit. Deflection of beam could be calculated with including the initial
stiffness of beam supports, which is given by (19).
c. For ultimate limit state the minimum strength of joints calculated from formula (11-13)
and (16) ought to be checked. Moreover, when the semi-continuous framing is
considered the secant stiffness of joints given by the formula (20) could be included.
Design model of such joint is given in figure 11.

NOTATION

a throat thickness of welds


fM function
fy yield stress of (fyn – gusset plate, fyo – chord member)
kM , ks coefficient
mpl plastic moment of resistance per unit length in the chord face (mpl = fyoto2/4)
n dimensionless prestressing in chord (n = No/Aofyo),
y1, y2, y3 unknown exponents
Mnpl,y plastic moment of resistance of branch member (gusset plate)
β branch to chord width ratio (β = bn/bo)
η branch depth to chord width ratio (η = hn / bo)
λo slenderness of chord face ( λo = bo/to)

ACKNOWLEDGEMENT

This research project No W/IIB/10/01 was financially supported by Bialystok Technical


University, Poland

REFERENCES

(1) European Committee for Standardisation (CEN): Eurocode 3: Design of steel


structures: Part 1.8: Design of joints, European Standard, prEN 1993-1-8:2003, 31
January 2003 (Stage 34 draft).
(2) Jarrett N.D., Malik A.S.: Fin plate connections between RHS columns and I beams.
Proceedings of the 5th International Symposium on Tubular Structures, Nottingham,
Ed. M.G. Coutie and G. Davies, 1993.
(3) Wardenier J.: Hollow section joints, Delft University Press, 1982.
(4) Groeneveld H.: Rigid plastic 2nd order calculation of horizontally restrained beam and
plates. Delft University of Technology, 1981.

420 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


(5) Szlendak J., Bródka J.: Yield and buckling strength of T, Y and X joints in rectangular
hollow section trusses. Proceedings of Institution of Civil Engineers, Part 2, 79, Mar.,
167-180, 1985.
(6) Szlendak J.K.: Design models of welded joints in steel structures with rectangular
hollow sections. DSc thesis, Bialystok Technical University Press, 2004 (in Polish)
(7) Szlendak J. Broniewicz M.: Data bank of connections. Beam to column welded
connections. Part 1: RHS column to RHS beam. Bialystok Technical University, June
1995.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 421


422 Connections in Steel Structures V - Amsterdam - June 3-4, 2004
AN EFFECTIVE EXTERNAL REINFORCEMENT SCHEME
FOR CIRCULAR HOLLOW SECTION JOINTS
Y. S. Choo, National University of Singapore, Singapore
J. X. Liang, National University of Singapore, Singapore
G. J. van der Vegte, Delft University of Technology, The Netherlands

ABSTRACT

This paper presents an effective external reinforcement scheme for


circular hollow section joints. The collar plate reinforcement is a scheme
which may be applied to newly fabricated or existing joints which are
found to be under-strength. The paper first introduces the structural
scheme and then presents results of extensive numerical studies on the
static strength of circular hollow section (CHS) joints reinforced with a
collar plate. The results show that significant strength enhancement of
the reinforced joints can be achieved through proper proportioning of the
reinforcement plate.

INTRODUCTION

This paper presents an external reinforcement scheme, termed a collar, for strengthening
circular hollow section (CHS) joints. Choo et al. (1) first investigated plate reinforcement
schemes which may be used for field installation of auxiliary structures for offshore
structures. The collar plate may be suitable to provide reinforcement to a pre-fabricated joint
that is found to be under-designed. This concept may also find potential applications for
reinforcing joints in older offshore platforms and large span structures.

Fig. 1 illustrates the schematic arrangement for the collar reinforcement for a X-joint which
may be found to be under-strength. In the figure, the collar plate reinforcement, assumed to
be square in this case, is shown to be placed outside the foot-print of the brace-chord
intersection, with thickness tc and length lc. The usual notations for the outside diameter and
wall thickness of the brace (d1 and t1) and chord (d0 and t0), and associated geometric ratios
are also indicated. The details 1 and 2, with additional weld shown hatched in Fig. 1, are
indicative welding arrangement to connect the collar plate to the brace and chord. The
edges of collar plate can be profiled to accommodate the existing full penetration weld at the
brace-chord intersection.

For the externally placed collar reinforcement plate which may be bent to be compatible with
the chord curvature, Fig. 2 shows three possible schemes: 4 parts, 2 parts (parallel) and 2
parts (perpendicular). For the 4 part scheme, for example, the solid lines indicated in Fig. 2
denote the lines of weld connecting the collar plate to the brace and chord. Indicative weld
details shown in Fig. 1 (details 1 and 2) can be sized appropriately for the design
requirements. For a joint loaded predominantly by in-plane bending, the 2 parts (parallel)
arrangement may be an option if the welding requirement needs to be minimized. For a joint
loaded predominantly by brace axial load or out-of-plane bending, the 2 parts
(perpendicular) arrangement may be considered. For collar plates with large lc/d1 ratio,
additional slot welds may be placed within the boundaries to provide supplementary ties
between the collar plate and the chord.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 423


d1 α = 2l0/d0 τ = t1/t0 d1

t1 β = d1/d0 τc = tc/t0 t1
2γ = d0/t0 lc/d1
lc 1

tc
t0
2
lc
d0

d0 collar chord
plate brace

Detail 1 Detail 2
Figure 1. Collar plate reinforced CHS X-joint.

A B

4 parts 2 parts (parallel) 2 parts (perpendicular)


Figure 2. Arrangement of collar plate parts.

This paper presents results of numerical studies on the behaviour of CHS T- and X-joints
with collar plate reinforcement. The accuracy of the numerical results is verified against the
T-joint tests reported by Choo et al. (1, 2). The results show that significant strength
enhancement for collar reinforced joints can be achieved through proper proportioning of the
reinforcement plate. Selected plots are presented to demonstrate the strength enhancement
of X-joints under brace axial compression, in-plane and out-of-plane moments.

COMPARISON WITH REFERENCE TEST RESULTS

Reference information on T-joint Tests

Choo et al. (1) presented results from an experimental programme investigating the strength
enhancement to a simple T-joint by provision of reinforcement around the intersection
region, in the form of a doubler plate or a collar plate. The experimental programme
consisted of eight tests with brace axial load with four pairs of tests, each pair with brace
compression and tension. The chord length was chosen such that joint failure occurred prior
to chord member failure, with particular reference to recommendations by Zettlemoyer (3).

424 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Detailed investigations into the behaviour of the test specimens and strength enhancement
offered by the doubler and collar reinforcement for T-joints, are presented by Choo et al. (2)
and van der Vegte et al. (4).

In this paper, the experimental result for the collar reinforced T-joint specimen EX-03 and the
calibration of the nonlinear finite element model are provided for illustration. Details can be
referenced in our papers (2, 4).

Mesh densities and element type

For a particular joint subjected to given loading, an analyst can consider the appropriate
symmetry in geometry, loading and boundary conditions to determine the finite element (FE)
model for analysis. For a X-joint subjected to brace axial load, only one-eighth of the joint
modelled (as shown in Fig. 3) with appropriate symmetry conditions and load specification is
required. For each FE model, more refined mesh is generated where stress gradient is more
critical. The automatic mesh generator for reinforced joints in this study is an extension of
that presented by Qian et al. (5).
Y

Figure 3. FE model for one eighth of a collar plate reinforced CHS X-joint
2γ = 50.8, β = 0.64, lc/d1 = 1.50 and τc = 1.0.

For the present FE models, two layers of 20-noded solid elements, type C3D20R with
reduced integration in ABAQUS (6), are specified through the thickness of all members to
provide good description of possible non-linearity in the thickness direction. Depending on
the actual joint geometry, 500 to 1000 elements are created to represent one-eighth of a
whole joint. Such mesh density has been proven to be able to produce results with good
accuracy (7).

Weld geometries

As three-dimensional solid elements are used in the FE models, it is possible to simulate the
weld geometries with high accuracy. The actual geometric definition of the welds is included
in all FE models. The geometry of the penetration weld between the brace and the chord is
modelled following the American Welding Society (8) recommendations. The depth of the
fillet welds between the reinforcing plate and the chord surface is taken the same as the
thickness of the plate, with two layers of finite elements specified. The welds connecting the
collar plate and chord (along the chord circumferential or longitudinal directions) are not
explicitly modeled. These are reflected in the FE model by specifying the appropriate
spatially common nodes to be tied.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 425


Geometric and material specifications

The geometrical non-linearity is included to predict possible buckling in the chord wall
through the NLGEOM parameter in the *STEP option in ABAQUS input file. The material
nonlinearity is specified using the “true stress” and associated logarithmic strain to define the
plasticity with isotropic hardening (6).

Contact interaction

When a collar plate reinforced joint is loaded, contact may occur between the bottom of the
collar parts and the chord outer surface. The contact interaction plays an important role in
the load transferring mechanism of plate reinforced joints and thus non-linear contact
analysis is required. Since both of the reinforcing plate and the chord wall are deformable
bodies, a deformable-deformable contact interaction was defined using a “master-slave”
algorithm in the numerical analysis (6).

Comparison between test and FE results

Fig. 4a shows the cut-section of the collar-reinforced Specimen EX-03 after completion of
the test. It can be observed that the collar reinforcement has relocated the chord plastic
hinges away from the brace-chord intersection, and that the brace has deformed extensively
adjacent to the intersection. Fig. 4b shows the deformed shape predicted by the nonlinear
FE analysis, and very good agreement with the experimental result is observed.

a b

Figure 4. Comparison between test and FE results, (a) Cut-section of EX-03 after test, and
(b) FE prediction.

The load-ovalisation curves (in which ovalisation at particular load level is based on the
change in diameter of the chord section) for Specimen EX-03 are shown in Fig. 5. The
numerical prediction is found to correspond very closely with the experimental curve, and
this serves to verify the accuracy of the numerical method.

Programme set-up

Parametric studies to investigate the static strength of collar plate reinforced X-joints have
been conducted by the authors. The chord diameter of all joints was taken as do=508 mm,
with β varying from 0.25 to 0.80 (β= 0.25, 0.43, 0.64 and 0.80), α = 12, and 2γ= 31.8 and
50.8. The brace-to-chord thickness ratio τ =1.0, and the brace length was kept at 4d1. The
thickness of the reinforcing plate was assumed equal or larger than the chord wall thickness
t0. For each combination of 2γ and β ratios, three values of plate thickness parameter
(τc=1.00, 1.25 and 1.60) and five values of plate length parameter (lc /d1=1.25, 1.5, 2.0, 2.5
and 3.0) have been considered. The corresponding un-reinforced joints were also included
to provide the appropriate reference strength. A total of 8 un-reinforced joints and 120 collar

426 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


reinforced joints were analyzed, with each joint subjected to brace axial compression, in-
plane moment or out-of-plane moment separately.

The un-bent collar plate was assumed to be square in shape, except for large β cases,
where the plate width exceeded half the perimeter of the chord section, and for this case, the
plate width was limited to half the chord perimeter with welds along its edges.

In the following sections, selected results shown for the various loading conditions are
focussed on joints with 2γ = 50.8 and β = 0.25 and 0.64.

β=0.54, 2γ=50.6, Collar, Compression


Experimental
Numerical
500

400

300
Load [kN]

200

100

0
0 20 40 60 80 100 120
Ovalisation [mm]
Figure 5. Experimental and numerical load-ovalisation curves for EX-03.

STRENGTH OF REINFORCED X-JOINT UNDER AXIAL COMPRESSION

Failure mechanisms and load-indentation curves

Fig. 6a to 6b show the deformed shapes of two collar reinforced X-joints subjected to axial
brace compression. Due to the weld at the brace-chord intersection, and the collar-chord
segments along the longitudinal (crown) and circumferential (saddle) segments, the collar
plate is effective in stiffening the chord and enhancing the load transfer from the brace.

β = 0.25 β = 0.64
τ = 1.00 τ = 1.00
lc = 2.00d1 lc = 2.00d1
τc = 1.00 τc = 1.00

Figure 6. Deformed shapes of collar reinforced X-joints with lc=2.0d1 and 2γ=50.8 with
(a) β = 0.25, (b) β = 0.64.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 427


In Fig. 7a and 7b, the non-dimensionalised loads F/fy0t02 for the joint with β= 0.25 and 0.64
and plate sizes lc=1.25d1 to 2.5d1 are plotted against the displacement δ/d0, where δ is the
indentation of the chord wall at the crown position. It can be seen that significant strength
enhancement is achievable for plate reinforced joints. For a joint with β= 0.64 and lc=2.5d1,
a “jump” in joint strength can be observed when the collar plate width reaches half of the
chord section perimeter due to a more direct and effective load transfer mechanism through
the welds.
Unreinforced joint Unreinforced joint
lc/d1=1.25 α =12.0 lc/d1=1.25 α =12.0
lc/d1=1.50 2γ=50.8 lc/d1=1.50 2γ=50.8
lc/d1=2.00 τ c=1.00 lc/d1=2.00 τ c=1.00
lc/d1=2.50 β =0.25 lc/d1=2.50 β =0.64
15 40

30
10
F/fy0*t02

F/fy0*t02
20

5
10

0 0
0.00 0.02 0.04 0.06 0.08 0.00 0.02 0.04 0.06 0.08
δ/d0 δ/d0

Figure 7. Normalised load-indentation curves for collar reinforced X-joints with different plate
width to brace diameter ratios (a) β = 0.25 (b) β = 0.64.

The deformation limit proposed by Yura et al. (9), which is defined as 60fyd1/E, is adopted to
determine the ultimate strength of a joint without a pronounced peak value in the load-
displacement curve. It is noted that the collar plate reinforcement can provide substantial
strength enhancement to the joint.

Effects of τc and lc /d1

Fig. 8a and 8b present the strength enhancement due to provision of collar plate for joints
with 2γ=50.8 and β=0.25 and 0.64, with the corresponding un-reinforced joint strength as
reference strength. Each of the strength ratios is plotted against the plate parameters τc and
lc/d1 in a three-dimensional diagram for each β. As noted in Fig. 8b, the reinforced joint
strength, obtained by the provision of an appropriately dimensioned collar plate, can be up to
3 times of the strength of an un-reinforced joint. The strength of a collar plate reinforced joint
may be improved either by increasing the collar plate length or by using a thicker plate. For
joints with small values of lc/d1, the effect of the plate thickness is insignificant. The effect of
plate thickness becomes more important as the collar plate length increases.

BEHAVIOUR OF REINFORCED X-JOINTS UNDER IN-PLANE BENDING

In this section, the failure mechanisms for un-reinforced and collar reinforced X-joints under
in-plane bending are presented to highlight the differences. The geometric parameters of the
joints considered are 2γ = 50.8 and β = 0.25 and 0.64. More details are reported by Choo et
al. (10).

428 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


2γ=50.8 β =0.25 2γ=50.8 β =0.64

2.2 3.5

2.0 3.0
1.8
2.5
Fu,c /Fu,u

Fu,c /Fu,u
1.6
2.0
1.4 3.0 3.0
1.2 2.5 1.5 2.5

1.0 2.0 1.0 2.0

1
1

ld /d
lc /d
1.0 1.0
1.2 1.5 1.2 1.5
1.4 1.25 1.4 1.25
τc 1.6 τc 1.6

Figure 8. The effects of τd and ld/d1 on the strength of axially loaded collar plate reinforced
X-joints with 2γ = 50.8 (a) β = 0.25 (b) β = 0.64.

Failure mechanisms

Fig. 9a and 9d show the deformed shapes of collar plate reinforced joints with different
combination of β and lc/d1. The collar plate reinforced joint is observed to fail with relatively
large plastic zones formed near the brace-chord intersection. Because of the welds between
the collar plate parts and the chord surface parallel to the chord axis, the collar plate acts
closely with the chord wall on both compressive and tensile sides.

For joints with short collar plates (Fig. 9a and 9c), plastic hinges are observed near the
welds between the collar plate and the chord. The strength enhancement due to the short
collar plate may be regarded as an equivalent increase in β. No obvious plastic hinge is
found for a joint with long collar plates (Fig. 9b and 9d).

a β = 0.25 c β = 0.64
lc = 1.25d1 lc = 1.25d1

b β = 0.25 d β = 0.64
lc = 2.00d1 lc = 2.00d1

Figure 9. Deformed shapes of collar plate reinforced X-joints under in-plane bending.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 429


Effects of τc and lc /d1

Fig. 10a and 10b present the strength enhancement due to provision of collar plate for joints
with 2γ=50.8 and β=0.25 and 0.64. As noted in Fig. 10b, the reinforced joint strength,
obtained by the provision of an appropriately dimensioned collar plate can be up to 2.8 times
of the strength of an un-reinforced joint. For joints with small values of lc/d1, the effect of the
plate thickness is insignificant. The effect of plate thickness becomes more important as the
collar plate length increases and more deformation of the collar plate takes place.

2γ=50.8 β =0.25 2γ=50.8 β =0.64

3.0 3.0

2.5 2.5
Mi,u,c /Mi,u,u

Mi,u,c /Mi,u,u
2.0 2.0
3.0 3.0
1.5 2.5 1.5 2.5

1.0 2.0 1.0 2.0

1
1

ld /d
lc/d

1.0 1.0
1.2 1.5 1.2 1.5
1.4 1.25 1.4 1.25
τc 1.6 τc 1.6

Figure 10. The effects of τc and lc /d1 on the strength of collar plate reinforced X-joints under
IPB with 2γ = 50.8 (a) β = 0.25 (b) β = 0.64.

STRENGTH OF REINFORCED X-JOINTS UNDER OUT-OF-PLANE BENDING

Failure mechanisms

Fig. 11a and 11b show the deformed shapes of collar plate reinforced joints loaded by out-
of- plane bending. It can be observed that the weld connecting the collar plate to the chord,
from the saddle positions along the chord circumferential direction is effective in transferring
the brace moment.

Figure 11. Deformed shapes of collar plate reinforced X-joints under out-of-plane bending
with 2γ = 50.8 (a) β = 0.25 (b) β = 0.64.

430 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Effects of τc and lc /d1

Fig. 12a and 12b show the potential strength enhancement for collar plate reinforced X-
joints. It can be seen that the strength ratio of the reinforced joint to the corresponding un-
reinforced joint varies from 1.6 to 3.6. The plate thickness parameter τc and length parameter
lc /d1 have significant effects on the strength of the reinforced joints for cases with large lc/d1
ratios. Equivalent strength enhancement can be obtained by either increasing the plate
length or by using a thicker collar plate.

2γ=50.8 β =0.25 2γ=50.8 β =0.64

4.0 4.0

3.0 3.0
Mo,u,c /Mo,u,u

Mo,u,c /Mo,u,u
2.0 3.0 2.0 3.0
2.5 2.5

1.0 2.0 1.0 2.0

1
1

ld /d
lc/d

1.0 1.0
1.2 1.5 1.2 1.5
1.4 1.25 1.4 1.25
τc 1.6 τc 1.6

Figure 12. The effect of τd and ld/d1 on the strength of collar plate reinforced X-joints under
OPB with 2γ = 50.8 (a) β = 0.25 (b) β = 0.64.

SUMMARY AND CONCLUSIONS

Extensive numerical studies have been conducted to evaluate the behaviour of circular
hollow section (CHS) X-joint reinforced with a collar plate, subjected to axial brace
compression, in-plane bending or out-of-plane bending respectively. From the presented
results of un-reinforced and collar plate reinforced CHS T- and X-joints, the following may be
concluded:

1. The collar plate is an effective reinforcement scheme, and can improve the static strength
of CHS T- and X-joints considerably.
2. Each of the parameters: the brace-to-chord diameter ratio β, the plate-to-chord wall
thickness ratio τc, and the plate length-to-brace diameter ratio lc/d1 have significant
influence on the strength of collar plate reinforced joints.
3. For a reinforced joint with fixed brace and chord dimensions, equivalent strength
enhancement can be obtained by either appropriately increasing the plate length or using
a thicker reinforcement plate.

RECOMMENDATIONS FOR FUTURE RESEARCH

Based on the present studies, the following are possible recommendations for future
research on the collar plate reinforced joints:

1. Since only part of the geometric parameters and loading conditions have been covered in
the current study, more extensive parametric studies will provide a comprehensive
understanding of collar plate reinforced joints.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 431


2. Fatigue analyses of collar plate reinforced joints are proposed. The current study
concentrated on the static strength of plate reinforced joints. It is important to investigate
the behavior of plate reinforced joints under fatigue loading.
3. Experimental investigations on plate reinforced joints subjected to different loading cases
will provide reliable reference results for parametric numerical investigation. Due to lack of
experimental data on plate reinforced joints, this study used available test results on
reinforced T-joints and published numerical results to verify the numerical methods.

ACKNOWLEDGEMENTS

The authors wish to record their appreciation to Dr Nick Zettlemoyer of ExxonMobil


Upstream Research (USA) for initiating the research on reinforced joints in the National
University of Singapore. They like to thank Professor Jaap Wardenier in Delft University of
Technology and Professor Richard Liew in National University of Singapore for their
contributions towards the studies.

REFERENCES

1. Choo, Y.S., B.H. Li, G.J. van der Vegte, N. Zettlemoyer & J.Y.R. Liew (1998). Static
strength of T-joints reinforced with doubler plate or collar plate. Tubular Structures VIII:
Proceedings Eighth International Symposium on Tubular Structures, Singapore, pp. 139-
145.
2. Choo, Y.S., G.J. van der Vegte, B.H. Li, N. Zettlemoyer & J.Y.R. Liew (2005). Static
strength of T-joints reinforced with doubler or collar plates - Part I: Experimental
investigations. Journal of Structural Engineering, ASCE, Vol. 131, No. 1, pp. 119-128.
3. Zettlemoyer, N. (1988). Developments in ultimate strength technology for simple tubular
joints. Proc. Offshore Tubular Joints Conference (OTJ’88), Surrey, UK.
4. van der Vegte, G.J., Y.S. Choo, J.X. Liang, N. Zettlemoyer and J.Y.R.Liew (2004). Static
strength of T-joints reinforced with doubler or collar plates - Part II: Numerical
simulations. Journal of Structural Engineering, ASCE (accepted for publication).
5. Qian X.D., Romeijn A., Wardenier J. and Choo Y.S. (2002). An automatic FE mesh
generator for CHS tubular joints. Proc. 12th International Offshore and Polar Engineering
Conference. Kita-Kyushu, Japan.
6. Abaqus/Standard User’s Manual Version 6.2 (2001). Hibbitt, Karlsson and Sorensen
Inc., Rhode Island, USA.
7. van der Vegte, G.J. (1995). The static strength of uniplanar and multiplanar tubular T-
and X-joints. PhD thesis. Delft University Press.
8. A.W.S. (1996). Structural Welding Code, AWS D1.1-96. American Welding Society Inc.,
Miami, USA.
9. Yura, J.A., N. Zettlemoyer & I.F. Edwards (1980). Ultimate capacity equations for tubular
joints. Proc. Offshore Technology Conference, Paper OTC 3690, Houston, U.S.A.
10. Choo Y.S., Liang J.X., van der Vegte G.J., Liew J.Y.R. (2004). Static strength of collar
plate reinforced CHS X-joints loaded by in-plane bending. Journal of Constructional
Steel Research, Vol. 60, No. 12, pp. 1745-1760.

432 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


THE INFLUENCE OF BOUNDARY CONDITIONS ON
THE CHORD LOAD EFFECT FOR CHS GAP K-JOINTS

G.J. van der Vegte


Kumamoto University, Japan / Delft University of Technology, The Netherlands
Y. Makino, Kumamoto University, Japan
J. Wardenier, Delft University of Technology, The Netherlands

ABSTRACT
In the framework of a larger programme to establish new chord load functions
for circular hollow section joints, this study evaluates the effects of various sets
of boundary conditions and chord pre-load on the static strength of axially
loaded gap K-joints. The influence of boundary conditions on the chord stress
contours is made clear for four different combinations of the geometric
parameters β and 2γ. It is concluded that a better understanding of the effects
of chord pre-stress on the strength of K-joints is obtained by considering the
maximum chord stress as the governing variable, instead of the chord stress
due to externally applied pre-loads.

INTRODUCTION

In current design rules, insufficient emphasis is put on the consistency of various design
equations. For circular hollow section (CHS) joints, the external chord “pre-load” (i.e. the
additional load in the chord which is not necessary to resist the horizontal components of the
brace forces) is used to account for the effects of chord loading. However, for rectangular
hollow section (RHS) joints, the chord stress formulation is based on the maximum chord
stress i.e. the stresses as a result of axial forces and (where applicable) bending moments.

To a designer, it is confusing that different approaches should be used for different


categories of joints, which may lead to misinterpretations and errors. Hence, it has been
proposed in the framework of a CIDECT (Comité International pour le Développement et
l’Étude de la Construction Tubulaire) programme, to re-analyse the effects of chord stress on
the ultimate strength of tubular joints in order to establish a chord stress formulation as a
function of the maximum chord stress, consistent for CHS and RHS joints.

As reported by van der Vegte and Makino (1), in the past, research into the effects of chord
pre-load on the strength of tubular joints was limited. Three experimental studies are
available in the literature regarding the effects of pre-load on the ultimate strength of tubular
X-joints. Although for CHS K-joints, the number of experiments outnumbers the data
available for uniplanar X-joints, in most of the K-joint tests, chord stress was simply a result
of horizontal equilibrium loads and was not meant as a prime variable. Only a few
researchers e.g. Kurobane and Makino (2) and de Koning and Wardenier (3) explicitly
applied a chord pre-load to the joints. More recent investigations into the effects of chord pre-
load on CHS joints were conducted by Dier and Lalani (4) and Pecknold et al (5, 6).

Since numerical tools offer the flexibility to vary various parameters and at the same time

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 433


exclude the scatter usually observed between different series of tests, additional data on
axially loaded uniplanar K-joints subjected to chord pre-load were generated by van der
Vegte et al (7) using numerical methods. Twelve K-joint configurations were analysed, using
the software package ABAQUS/Standard (8). The brace-to-chord diameter ratio β ranged
from 0.25 to 0.67, while the chord thinness ratio 2γ was taken as 25.4 or 63.5. The brace
angle θ was set to 45˚ or 60˚. For each of the K-joints considered, the uniformly applied
chord stress varied from –0.9 fy to +0.9 fy (tensile pre-stresses are referred to as positive).
The boundary conditions employed for the set of twelve K-joint configurations are shown in
figure 1. The chord stress contours obtained from the numerical analyses were presented not
only as a function of the chord pre-stress due to external loads but also as a function of the
maximum chord stress. Both contours appeared to be considerably different for the joints
where large chord loads were introduced to maintain horizontal equilibrium.

Figure 1. Boundary conditions for K-joints used by van der Vegte et al (7).

Various researchers assessed the effects of boundary conditions on the ultimate strength of
uniplanar K-joints and came to the conclusion that the influence of restraints could be
significant. A brief overview of some of the investigations is presented in the section
hereafter. Additional FE analyses were conducted to evaluate the effects of various sets of
boundary conditions on the static strength of uniplanar K-joints subjected to chord pre-load.
The current study addresses the research programme, the FE strategy and the failure
criteria. Finally, the numerical results are presented as a function of either the chord pre-
stress due to external loads or the maximum chord stress.

PREVIOUS RESEARCH INTO THE EFFECT OF BOUNDARY CONDITIONS ON CHS AND


RHS K-JOINTS

In 1989, Connelly and Zettlemoyer (9) performed numerical research into the static
behaviour of various overlap and gap K-joint configurations. Each joint was analysed twice :
at first the K-joints were restrained and loaded in a manner consistent with laboratory tests
on such joints. In the second analysis, the K-joint models were mounted into a braced frame
with the load being applied directly to the frame instead of the K-joint. Connelly and
Zettlemoyer found that for this specific configuration with β = 1.0, the frame-mounted K-joints
showed axial capacities which were between 11 to 26 % higher compared to the isolated
joints. The authors suggested that, if possible, future tests on isolated joints should consider
a more accurate replication of the boundary conditions found in actual frames.

In 1992, Bolt et al (10) conducted numerical analyses on a single gap K-joint geometry using
different boundary conditions for chord and braces. Variations in capacities of up to 10 %
were observed while the post-peak load-deformation responses also varied significantly
among the cases considered. However, Bolt reported that confidential research on K-joint
configurations with different sets of geometric parameters suggested an even much greater
dependency of boundary conditions.

As part of a larger study into the behaviour of overlapped K-joints, Healy (11) performed
various numerical simulations to assess the effect of chord and brace end restraints on the
axial capacity. Healy evaluated two sets of boundary conditions often used in experiments.

434 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


(a) “single” restraints (b) “double” restraints

Figure 2. “Single” and “double” boundary conditions for K-joints.

Figure 2a illustrates “single” boundary conditions : one chord end is left free, while one brace
end reacts the load applied to the other. Figure 2b depicts the “double” set of boundary
conditions where both chord ends are restrained. Healy concluded that, when lateral
movements of the braces are restricted, the differences between both sets of boundary
conditions are negligible for the joints considered. On the other hand, Healy mentioned that
experiments carried out by Bjornoy (12) revealed a strong dependency of boundary
conditions on the ultimate strength of K-joints, especially for eccentric overlapped joints.

Bjornoy’s conclusion was supported by preliminary research on overlapped K-joints carried


out by Dexter et al (13). “Single” type restraints were found to give significantly higher
capacities compared to exactly balanced loading conditions, especially for the more heavily
overlapped joints.

In 1998, Liu et al (14, 15) investigated the effect of boundary conditions and chord load on
the capacity of selected uniplanar and multiplanar RHS gap K-joints. Similar to the findings of
the previous researchers, the authors concluded that boundary conditions have a significant
influence on the ultimate strength of RHS K-joints.

N2 N1 α = 2l0/d0
t2 t1 β = d1/d0
d2 d1 2γ = d0/t0
g ξ = g/d0
t0 l2 θ2 θ1 l1
+Nop
d0 e

l0

Figure 3. Dimensions and non-dimensional geometric parameters for uniplanar K-joints.

SCOPE OF NUMERICAL RESEARCH

The configuration of uniplanar K-joints and the definition of the geometric parameters are
shown in figure 3. The geometric parameters β and 2γ analysed in van der Vegte’s numerical
study (7) are summarized in Table 1. Although the programme of twelve K-joints considered
two values of the brace angle θ (45˚ and 60˚), only the six configurations referring to θ = 45˚
are presented in Table 1.

The current study focuses on the following four K-joint geometries shown in Table 1 : β =
0.48 and 0.67, 2γ = 25.4 and 63.5 (joints K3 to K6). The K-joints with β = 0.25 are not further
studied, as the influence of the horizontal reaction forces on the ultimate load is not as
pronounced as for the other configurations due to the relatively small failure loads of K-joints

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 435


with small β values.

Since the current programme is limited to K-joints with positive gap-values, for some
configurations, eccentricities have been introduced to avoid overlapped joints. For each of
the K-joints with eccentricities, the gap size g was taken as 32 mm, corresponding to t1 + t2
for the thick walled joints. The values of the gap ratio ξ (= g/d0), g/t0 and e/d0 are also
presented in Table 1. The non-dimensional chord length parameter α (= 2l0/d0) is held at 16,
whereas d0 = 406.4 mm. The length of the braces is set to 5d1. The steel grade used for the
tubular members is S355 with fy = 355 N/mm2 and fu = 510 N/mm2.

The boundary conditions investigated are summarized in figure 4. These boundary


conditions are commonly used in experiments and numerical analyses. In model 1, loads
have been applied to the compression brace end only. This force is primarily reacted by the
pinned end of the tension brace whereas the resulting horizontal forces are reacted at the
pinned end of the chord. However, because of geometric non-linear effects, this arrangement
may lead to unequal forces in the compression and tension braces. In models 2 to 4, equal,
opposite loads have been applied to both the compression and tension braces, using the
Riks algorithm.

In models 1 and 2, the horizontal component of the brace forces causes compressive
stresses in the left side of the chord, while in model 4, tensile stresses occur in the right side
of the chord. In model 3, the horizontal components of the brace loads are distributed
between both chord ends.

Model 1 Model 2 Model 3 Model 4


Figure 4. K-joint models with different boundary conditions.

In line with the analyses of the parametric study on K-joints, for each of the configurations
considered, nine values of the external chord pre-load N0p have been analysed, giving the
following chord pre-load ratios N0p/A0fy0 : +0.9, +0.8, +0.6, +0.3 , 0.0, -0.3, -0.6, -0.8, -0.9
(positive values refer to tensile pre-load).

Table 1. Geometric parameters of the four K-joints analysed (K3 to K6).


joint t0 (mm) 2γ ξ g/t0 e/d0
β = 0.25 K1 16 25.4 0.65 16.4 0.0
(d1 = 101.6 mm) K2 6.4 63.5 0.65 41.0 0.0
β = 0.48 K3 16 25.4 0.33 8.3 0.0
θ = 45˚
(d1 = 193.7 mm) K4 6.4 63.5 0.33 20.7 0.0
β = 0.67 K5 16 25.4 0.079 2.0 0.015
(d1 = 273.1 mm) K6 6.4 63.5 0.079 5.0 0.015

Remarks : - for all joints, d0 = 406.4 mm


- the numerical analyses of K-joints K1 and K2, modelled with boundary
conditions 2 are reported by van der Vegte et al (7)

436 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


FINITE ELEMENT ANALYSES

FE modelling aspects

The numerical analyses were carried out with the finite element package ABAQUS (8). The
joints are modelled using twenty noded solid elements employing reduced integration
(ABAQUS element C3D20R). Two layers of elements are modelled through the thickness of
each member. Due to symmetry in geometry and loading, only one half of each joint has
been analysed. The appropriate boundary conditions are applied to the nodes located in the
plane of symmetry.

For all joints, the geometry of the welds at the brace-chord intersection has been modelled.
The dimensions of the welds in the numerical model are in accordance with the
specifications recommended by the AWS (16).

Both ends of the braces and the chord ends have rigid diaphragms. The length of the
members is considered to be sufficient to exclude any influence of the end caps on the static
response of the joints.

Since the incorporation of material- and geometric non-linearity in ABAQUS requires the use
of true stress-true strain relationships, the engineering stress-strain curve is modelled as a
multi-linear relationship and subsequently converted into a true stress-true strain relationship.
The hardening rule proposed by Ramberg-Osgood has been used to describe the true
stress-true strain behaviour after the peak stress in the engineering stress strain curve is
reached.

In order to validate the numerical model, comparisons were made with experimental
evidence. In 1981, de Koning and Wardenier (3) conducted a series of static tests on
uniplanar CHS K-joints. Out of this programme, two thin walled gap K-joints (β = 0.33 and
0.65, 2γ ≈ 56, θ = 45˚, e = 0) were chosen to serve as a basis for validation of the current FE
model. As described by van der Vegte et al (7), for both geometries good agreement was
observed between the numerical and experimental load-deformation responses, not only for
the initial stiffness but also for the peak load and the post-peak behaviour.

Loading and boundary conditions of the K-joints

In the first step of the numerical analyses, the chord end is pre-loaded with uniformly
distributed axial forces using the load-control method. During this first step, the chord ends
are roller supported i.e. free to move laterally.

In the second step of the loading procedure, the appropriate boundary conditions are applied
to the chord ends whereas axial brace loads are applied to the nodes of the brace tip.
Meanwhile, the axial forces at the chord ends are maintained at the same level as at the end
of the first step.

In the second step of the joints modelled with the boundary conditions of model 1, the
displacement of the tip of the compression-loaded brace is prescribed, i.e. the joint is loaded
employing the displacement-control method. In the second step of the loading history of
models 2 to 4, equal but opposite loads are applied to both the compression and tension
braces, using the Riks algorithm, enabling to monitor the load-indentation behaviour for
declining brace loads.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 437


NUMERICAL RESULTS AND DISCUSSION

Failure criteria

Ultimate load is defined as the force on the compression brace first exceeding one of the
following four failure criteria :

(1) peak load in the load-displacement diagram.


(2) for the joints which load-displacement curves do not show a clear peak load, the value
at Lu’s deformation limit (= indentation of 0.03d0) is taken as the “ultimate” load (17).
(3) for joints with small gaps, large tensile strains are observed in the chord wall at the weld
toe of the tension brace. Although the load-indentation curves can be extended well
beyond this point, it is assumed that cracks initiate if the strain at the integration point of
the chord element closest to the weld toe of the tension brace, exceeds 20 %. While the
value of 20 % is arbitrary, the use of such a “crack initiation” criterion enables a
comparison among the different K-joints and to identify those joints which are
vulnerable for fracture.
(4) early termination of the numerical analysis, indicating member failure (i.e. reaching
squash load) or the onset of chord buckling.

Effect of boundary conditions

Figure 5 shows the chord stress contours obtained for K-joints K5 and K6 (β = 0.67)
modelled with various boundary conditions. Although not presented, the contours for joints
K3 and K4 with β = 0.48 look very similar. For each joint, the ultimate load is normalized by
the corresponding capacity for N0p = 0, where N0p refers to the externally applied chord pre-
load. The diagrams shown in figure 5a display the normalized capacities versus the chord
pre-stress ratio n’ = N0p/A0fy0, while the contours in figure 5b are based on the actual chord
stress ratio n, including the horizontal brace load components. Depending on the boundary
conditions shown in figure 4, n is defined as n’-2N1,u cos θ /A0fy0 for models 1 and 2 (left side
of the chord), while n = n’+2N1,u cos θ /A0fy0 for model 4 (right side of the chord).

Since the K-joints modelled with boundary conditions 3 are statically indeterminate, the chord
stresses can not be captured by a formula and are obtained from the FE analyses. For these
joints, figure 5b presents the chord stress contours for both sides of the chord.

As presented in the section hereafter, the differences between the ultimate capacities of
models 1 and 2 are small. To avoid overlap of the chord stress contours and to enhance
clarity of the diagrams, the curves for model 1 are not shown in the diagrams of figure 5b.

Comparison between models 1 and 2.

Comparing the chord stress contours of the joints modelled with boundary conditions 1 and
2, displayed in the graphs of figure 5a, it becomes clear that the differences are small.

A better understanding of the effects of chord stress for the joints modelled with boundary
conditions 2 is gained after examining the diagrams of figure 5b, where the horizontal axis
depicts the actual chord stress ratio. Because of the applied boundary conditions, each
contour is shifted in horizontal direction towards the compression side, whereas the amount
of transition is determined by the magnitude of the equilibrium loads. From these diagrams it
becomes clear that the joints of model 2 subjected to large compressive pre-loads fail by
chord member failure. Although not shown in these chord stress contours, the joints of model
1 exhibit similar failure behaviour.

438 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Model 4 CHS K-joints Model 4 CHS K-joints
Model 3 β = 0.67 Model 3 - right side β = 0.67
Model 2 2γ = 25.4 Model 3 - left side 2γ = 25.4
Model 1 θ = 45 Model 2 θ = 45
1.2 1.2

1 1

f(n)
f(n')

0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2

0 0
-1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1 -1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1
n' n
Model 4 CHS K-joints Model 4 CHS K-joints
Model 3 β = 0.67 Model 3 - right side β = 0.67
Model 2 2γ = 63.5 Model 3 - left side 2γ = 63.5
Model 1 θ = 45 Model 2 θ = 45
1.2 1.2

1 1
f(n')

f(n)

0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2

0 0
-1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1 -1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1
n' n
(a) based on chord pre-stress ratio n’ (b) based on actual chord stress ratio n
Figure 5. Chord stress contours for K-joints K5 and K6 modelled
with various boundary conditions.

Comparison between models 2 and 4.

Comparing the chord stress contours of the joins modelled with boundary conditions 2 and 4,
it is found that for the joints under large tensile chord pre-loads, the ultimate capacities of the
joints of model 4 are significantly lower, caused by member failure of the chord. Chord
member failures are easily detected from the chord stress contours with the actual chord
stress being displayed on the horizontal axis (diagrams displayed in figure 5b).

For the K-joints modelled with boundary conditions 4 and pre-loaded by compression or
under zero pre-load, the ultimate capacities of the joints are higher than for the
corresponding joints of model 2. The unfavourable combination of compression pre-load and
compressive reaction forces leading to member failure (i.e. chord buckling or reaching the
squash load) observed for the joints of model 2, does not occur in the joints of model 4 under
compression pre-load. For the joints under zero pre-load and modelled with boundary
conditions 4, the gap area is still subjected to tensile stresses, giving a higher ultimate
strength in comparison with the joints where the gap area is loaded in compression, as is the
case for the joints of model 2.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 439


Discussion of results for model 3.

The chord stress contours for the joints of model 3, based on the pre-load ratio n’, show that
the effect of chord pre-load on the ultimate capacity is much smaller than for the other
models. Especially for the joints pre-loaded by compression, the reduction in strength is
much less pronounced. This may be explained by considering the chord stress distribution as
a result of both chord pre-load and reaction forces. For this purpose, K-joint K5 (θ = 45˚, β =
0.67 and 2γ = 25.4) is examined in detail.

K-joint K5 under compressive pre-load K-joint K5 under tensile pre-load


(n’ = -0.9) (n’ = 0.9)

Step 1 : Chord pre-load

Step 2a : Small brace loads

Step 2b : Reversal of left or right chord reaction force

Step 2c : Ultimate load

Remarks :
- The length of each arrow is proportional to the magnitude of the force
- Open block arrow : chord pre-load
- Black arrow : brace loads and chord reaction forces
Figure 6. External pre-load, brace loads and chord reaction forces of K-joint K5
modelled with boundary conditions 3.

In figure 6, the external chord pre-load and the reaction forces of K-joint K5 modelled with
boundary conditions model 3 are schematically illustrated for large compressive (n’ = -0.9)
and tensile (n’ = 0.9) chord pre-stress. In each of the diagrams, the length of the arrows is
proportional to the magnitude of the forces.

The loading procedure considers two steps. In the first step, the chord pre-load is applied
while the chord ends are free to move horizontally. The first row of figure 6 illustrates the
external forces applied to both chord ends. At the start of the second load step, in which the
brace loads are applied, the chord is restrained in horizontal direction, causing reaction

440 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


forces to develop as soon as the braces are loaded. For the two load cases considered (n’ =
-0.9 and 0.9), the second step is also visualized in figure 7, displaying the non-dimensional
chord load at either side of the K-joint as a function of the brace load.

For both chord pre-load cases, the following three stages can be distinguished when the
braces are loaded up to failure :

STEP 2A : For small brace loads, the chord reaction forces at either side of the K-joint are
almost equal in magnitude, pointing in the direction opposite to the brace loads (see second
row of figure 6). For the K-joint under pre-compression, this leads to a further increase of
compressive chord stress at the left side of the chord, while the chord stress in the right side
reduces. For the K-joint under pre-tension, the tensile stress in the right side of the chord
becomes larger. This trend continues until the left chord side of the K-joint under pre-
compression and the right side of the K-joint under pre-tension come close to yield, resulting
in the initiation of load-redistribution.

STEP 2B : For the K-joint subjected to n’ = -0.9, further loading of the braces causes the
chord reaction force at the left end to decrease and reverse. For the K-joint subjected to n’ =
0.9, a similar remark can be made for the chord reaction force at the right chord end. The
reversal of the chord reaction forces is illustrated by step 2b in figure 6.

STEP 2C : Continuous loading of the braces up to failure will then increase all chord reaction
forces. Failure of the K-joints is visualized by the “ultimate load” data points in figure 7, which
are also part of the chord stress contours displayed in figure 5b.

Left chord side (under tension brace)


Right chord side (under compression brace)

4000
Ultimate load
N1 [kN]

3000

2000 Reversal of chord


reaction force

1000

Small brace load

0
-1.0 -0.8 -0.6 -0.4 -0.2 0.0 0.2 0.4 0.6 0.8 1.0
chord load / A0 fy0

Figure 7. Chord load at either side of the chord for K-joint K5 under compressive (n’ = -0.9)
or tensile (n’ = 0.9) pre-load as a function of brace load.

For the thick walled K-joint (2γ = 25.4) under pre-compression, the chord is subjected to such
large reaction forces at failure that the resulting chord stress in the right side of the chord has
turned tensile. For the pre-tensioned joint, a similar redistribution is observed. These aspects
also become clear from the chord stress contours based on the actual chord stress, shown in
figure 5b. While the initial chord stresses due to the external loads vary between -0.9 fy0 and
+0.9 fy0, at failure, the actual chord stresses at both sides of the chord are significantly less,
thus explaining the less pronounced effect of external chord pre-load.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 441


For the thin walled joints (2γ = 63.5), a similar redistribution is found although less significant.
For these joints, the non-dimensional chord reaction forces (i.e. made non-dimensional after
dividing by A0fy0) are much smaller than for thick walled joints.

A comparison between the contours based on the chord pre-load ratio n’, shown in figure 5a,
reveals no relation between the curves of model 3 and the contours of models 2 and 4.
However, after looking at the diagrams in figure 5b based on the actual chord stress, it
becomes clear that the stress contours of the “compression” (left) side of the chord of model
3 are slightly below the contours obtained for model 2, while the chord stress contours of the
“tension” (right) side of the chord of model 3 are in good agreement with the stress contours
of model 4. This means that, when the actual chord stress is considered, the contours for
boundary conditions 3 are closely related to the contours of the two other models. This
clearly confirms the need to describe the chord stress effects as a function of the actual
chord stress rather than the stress due to externally applied chord loads.

CONCLUSIONS

Numerical analyses have been carried out into the strength of four axially loaded uniplanar
CHS gap K-joints subjected to axial chord pre-loading, with the main variables being the
geometric parameters β and 2γ and various sets of boundary conditions. Based on the
results of this selected set of K-joints, the following conclusions can be drawn :

a. In line with the data obtained for uniplanar X-joints, it is found that compressive chord
stresses have a detrimental effect on the ultimate capacity of axially loaded uniplanar K-
joints. For K-joints under tensile chord pre-load, the capacity of the joints either
increases or decreases compared to the ultimate strength of the corresponding joints
under zero pre-load, dependent on the value of β and 2γ, the amount of pre-load and
the boundary conditions.
b. The influence of boundary conditions on the ultimate capacity of K-joints can be
significant. The differences between the chord stress contours obtained for the joints
with the tensile brace end being pinned (model 1) and the joints where the tensile brace
is roller-supported to enable equal brace loads (model 2) are negligible. For K-joints
with both chord ends being pinned (model 3), the reduction in strength due to chord
pre-load diminishes, contrary to the behaviour exhibited by the joints modelled with the
other sets of boundary conditions, for which the strength reducing effects due to chord
pre-load are more pronounced.
c. The chord stress contours generated for K-joints clearly show that a better
understanding of the effects of chord pre-load is attained by considering the maximum
chord stress as the governing variable, rather than the chord stress due to external pre-
loads.
d. In future publications, the FE data generated in this study will be combined with
available data on other types of CHS and RHS joints and proposals for new chord
stress functions will be made.

ACKNOWLEDGEMENT

The first author would like to thank the Centennial Anniversary Foundation of the Faculty of
Engineering, Kumamoto University, Japan for the opportunity to carry out the research
reported herein.

442 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


REFERENCES

(1) Vegte, G.J. van der and Makino, Y., (2001). The Effect of Chord Stresses on the Static
Strength of CHS X-Joints. Memoirs of the Faculty of Engineering, Kumamoto University,
Vol. 46, No. 1.
(2) Kurobane, Y. and Makino, Y., (1965). Local Stress in Tubular Truss Joints. Research
Report, Kyushu Branch of Architectural Institute of Japan, No. 4, pp. 75-80 (in
Japanese).
(3) Koning, C.H.M. de and Wardenier, J., (1981). The Static Strength of Welded CHS K-
Joints. TNO-IBBC Report BI-81-35/63.5.5470, Stevin Report 6-81-13, Delft, The
Netherlands.
(4) Dier, A.F. and Lalani, M., (1998). New Code Formulations for Tubular Joint Static
Strength. Proc. 8th International Symposium on Tubular Structures, Singapore, pp. 107-
116.
(5) Pecknold, D.A., Ha, C.C. and Mohr, W.C., (2000). Ultimate Strength of DT Tubular
Joints with Chord Preloads. Proc. 19th International Conference on Offshore Mechanics
and Arctic Engineering, New Orleans, U.S.A.
(6) Pecknold, D.A., Park, J.B. and Koppenhoefer, K.C., (2001). Ultimate Strength of Gap K
Tubular Joints with Chord Preloads. Proc. 20th International Conference on Offshore
Mechanics and Arctic Engineering, Rio de Janeiro, Brazil.
(7) Vegte, G.J. van der, Makino, Y. and Wardenier, J., (2002). The Effect of Chord Pre-load
on the Static Strength of Uniplanar Tubular K-joints. Proc. 12th International Offshore
and Polar Engineering Conference, Kitakyushu, Japan, Vol. IV, pp. 1-10.
(8) ABAQUS/Standard, (2000). Version 6.1, Hibbitt, Karlsson & Sorensen, U.S.A.
(9) Connelly, L.M. and Zettlemoyer, N., (1989). Frame Behaviour Effects on Tubular Joint
Capacity. Proc. 3rd International Symposium on Tubular Structures, Lappeenranta,
Finland, pp. 81-89.
(10) Bolt, H.M., Seyed-Kebari, H. and Ward, J.K., (1992). The Influence of Chord Length and
Boundary Conditions on K-Joint Capacity. Proc. 2nd International Offshore and Polar
Engineering Conference, San Francisco, U.S.A., Vol. IV, pp. 347-354.
(11) Healy, B.E., (1994). A Numerical Investigation into the Capacity of Overlapped Circular
K-joints. Proc. 6th International Symposium on Tubular Structures, Melbourne, Australia,
pp. 563-571.
(12) Bjornoy, O.H., (1993). Static Strength of Tubular Joints, Phase II, Analyses and Tests of
Gap and Overlap K-Joints. Veritec Report No 91-3393, AS Veritec.
(13) Dexter, E.M., Lee, M.M.K. and Kirkwood, M.G., (1994). Effect of Overlap on Strength of
K-joints in CHS Tubular Members. Proc. 6th International Symposium on Tubular
Structures, Melbourne, Australia, pp. 581-588.
(14) Liu, D.K., Yu, Y. and Wardenier, J., (1998). Effect of Boundary Conditions and Chord
Preload on the Strength of RHS Uniplanar Gap K-Joints. Proc. 8th International
Symposium on Tubular Structures, Singapore, pp. 223-230.
(15) Liu, D.K., Yu, Y. and Wardenier, J., (1998). Effect of Boundary Conditions and Chord
Preload on the Strength of RHS Multiplanar Gap K-Joints. Proc. 8th International
Symposium on Tubular Structures, Singapore, pp. 231-238.
(16) AWS American Welding Society, (1992). Structural Welding Code. AWS D1.1-92.
(17) Lu, L.H., Winkel, G.D. de, Yu, Y. and Wardenier, J., (1994). Deformation Limit for the
Ultimate Strength of Hollow Section Joints. Proc. 6th International Symposium on
Tubular Structures, Melbourne, Australia, pp. 341-347.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 443


444 Connections in Steel Structures V - Amsterdam - June 3-4, 2004
SHEAR LAG IN SLOTTED GUSSET PLATE
CONNECTIONS TO TUBES
S. Willibald, University of Toronto, Canada
J.A. Packer, University of Toronto, Canada
G. Martinez Saucedo, University of Toronto, Canada
R.S. Puthli, Universität Karlsruhe, Germany

ABSTRACT

Hollow Structural Sections (HSS) are commonly used as bracing members in


steel-framed buildings or as web members in roof trusses. The tubes are
frequently slotted onto gusset plates to simplify fabrication and avoid profiling.
Under tension loading, these gusset plate connections can be susceptible to
shear lag failure of the HSS since only a part of the tube cross-section is
connected to the plate. This paper compares current design proposals found
in research, design guides and specifications against recent experimental
and numerical work by the Authors, which comprised of gusset plate
connections for round hollow sections with varying fabrication details.

INTRODUCTION

Gusset plates can be found in almost any type of steel building. As hollow sections have
become more popular due to their exceptional properties in compression and torsion, the
combination of both gusset plates and hollow sections can be found in numerous
applications. These gusset plate connections can be used to splice hollow section members
or to connect web members to the chords in roof trusses (see Figure 1). Three possible
fabrication details are shown in Figure 2. Slotting the hollow section (Figures 2 (b) and (c))
is the most common version of this connection type. Various failure modes are possible
under tension loading with shear lag being one of them. Shear lag of the hollow section can
occur as the circumference of the hollow section is connected to the gusset plate only at two
points on opposing sides. The unconnected
circumference of the hollow section is not fully engaged
and contributes only in part to the resistance of the
member. In addition, local stress peaks at the slot ends
can cause initiating cracks that may result in an early
failure.

The presented study focused on shear lag failure for


round hollow sections. An experimental study on six
connections under tensile loading has been carried out,
followed by a numerical analysis, which will be used to
substitute for further experimental work. Additionally,
recent research and international specifications on this
topic are evaluated and "best practice"
recommendations are made for application to round Figure 1. Metro Toronto
hollow section members. Convention Centre.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 445


(a) (b) (c)
Figure 2. Fabrication details of tested gusset plate connections.

LITERATURE STUDY
40000
Of the failure modes that 90
can occur in gusset plate θ
connections, shear lag is 30000 0 180
one of the most ill-defined.
Strain [ µ m]

Shear lag fracture of the


hollow section takes place 20000
500 kN
due to the uneven stress 750 kN
distribution around the 1000 kN
10000
circumference of the 1032 kN
hollow section. The
stresses peak at the points 0
where the hollow section is 0 22.5 45 67.5 90 112.5 135 157.5 180
connected to the plate or Angle θ [ o ]
weld and become less as
the distance to the weld Figure 3. Strain distribution around circumference of the
increases (see Figure 3). round hollow section (specimen 1, slotted HSS, no
Therefore, the uncon- weld return).
nected circumference only
contributes in part to the capacity of the member.

International specifications and design guides

For tension loaded connections, Eurocode 3 (1) provides shear lag provisions that are only
applicable for bolted connections. For connections with welds, no specific design method is
provided in the Eurocode. Otherwise, shear lag is mentioned only in connection with locally
introduced shear loads causing bending moments in longitudinally stiffened plated
structures.

The North American specifications address shear lag for welded connections under tension
loading. Unfortunately, the American and Canadian specifications (AISC 2, CSA 3) differ in
their design methods for this limit state. In addition, changing formulae in old and new
specifications (e.g. AISC 4 versus AISC 5) indicate a lack of certainty with this connection
type. The Japanese specification (6) excludes shear lag by providing minimum connection
lengths. The following paragraphs briefly introduce the different design methods of the
various specifications.

446 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


The current American Specification for Hollow Structural Sections (AISC 2) uses the concept
of "effective net area". The design method provided is based on research carried out by
Chesson and Munse (7). An eccentricity factor U is calculated from the connection
eccentricity (see Figure 4) and the connection length Lw:
x
U = 1- L ≤ 0.9 (1)
w

D
with: x = for round hollow sections (2)
π
The effective net area is then calculated as:
Ae = An · U (3)

centre of gravity
of top half
w
tsl
x
D
tp
t
Lw wp

Figure 4. Parameters of the experimental and numerical study.

The current LRFD Specification (AISC 4, Equation B3-2) uses the gross area Ag of the
member to calculate the effective net area Ae (Ae = Ag · U) which can result in considerably
different design strengths for gusset plate connections where the hollow section is slotted
(An = Ag - 2tp·t). In practice, the slot width tsl is usually greater than tp to allow ease of
fabrication, and in such cases An = Ag - 2tsl·t.

Recently, a general examination of the AISC LRFD shear lag design provisions has been
made by Kirkham and Miller (8). Based on recent studies, it was concluded that the existing
design approaches are overly conservative and further research was necessary. The draft
version (5) of the upcoming AISC Specification in 2005 now uses the net area of the
member, An, in its formula but no longer has an upper limit of 0.9 for the eccentricity factor U
(see Equation 1). For round HSS with Lw ≥ 1.3D, the factor U becomes equal to unity.
Connection lengths Lw less than D are not covered.

The current Canadian Standard (CSA 3) addresses shear lag in elements connected by a
pair of welds parallel to the load by calculating the "effective net area" (Clause 12.3.3) based
on an efficiency factor that depends on the ratio of the distance between the welds around
the hollow section perimeter, w, and the connection length, Lw (see Figure 4). The efficiency
factor given is:
1.0 for Lw/w ≥ 2.0;
0.5 + 0.25 Lw/w for 2.0 > Lw/w ≥ 1.0;
0.75 Lw/w for Lw/w < 1.0.
A similar approach based on the former Canadian standard CAN/CSA-S16.1-94 (CSA 9), as
well as research done by Korol et al. (10), is given in the design guide for hollow structural
sections by Packer and Henderson (11). The recommended efficiency factor there is:
1.0 for Lw/w ≥ 2.0;
0.87 for 2.0 > Lw/w ≥ 1.5;
0.75 for 1.5 > Lw/w ≥ 1.0;
0.62 for 1.0 > Lw/w ≥ 0.6.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 447


For values smaller than Lw/w = 0.6 shear rupture of the base metal along the weld line is
supposed to govern. As the experimental study by Korol et al. (10) was done with gusset
plate connections for square hollow sections, it is possible that the results might not be fully
applicable for shear lag in round hollow sections.

The Japanese recommendations for the design and fabrication of tubular truss structures in
steel (6) exclude shear lag by providing a minimum connection length of Lw ≥ 1.2D for gusset
plate connections. To account for uncertainties in fabrication of these connections, the
connection capacity is restricted to 90% of the unslotted (gross) member strength. Also, AIJ
avoid use of the net area of the slotted tube by means of a specific fabrication detail. Table
1 gives an overview of all the above shear lag provisions.

Table 1. Shear lag design provisions for round hollow sections.


Specification or Effective net Range of
Shear lag coefficients
design guide area validity
AISC (4) Ae = Ag · U x D no
U = 1- ≤ 0.9 with x =
AISC (2) Lw π restrictions
x
Ae = An · U U = 1-
AISC (5) Lw Lw ≥ D
U=1 for Lw ≥ 1.3D
AIJ (6) Ae = Ag · U U = 0.9* Lw ≥ 1.2D
U = 1.0 for Lw/w ≥ 2.0
no
CSA (3) U = 0.5 + 0.25 Lw/w for 2.0 > Lw/w ≥ 1.0
restrictions
U = 0.75 Lw/w for Lw/w < 1.0
Ae = An · U U = 1.0 for Lw/w ≥ 2.0 shear lag
Packer and U = 0.87 for 2.0 > Lw/w ≥ 1.5 not critical
Henderson (11) U = 0.75 for 1.5 > Lw/w ≥ 1.0 for
U = 0.62 for 1.0 > Lw/w ≥ 0.6 Lw < 0.6w
α = 1.0 for Lw/w ≥ 1.2 shear lag
α = 0.4 + 0.5 Lw/w for 1.2 > Lw/w ≥ 0.6 not critical
Korol (17) Ae = An· α ·U x for
U = 1 – 0.4
Lw Lw < 0.6w

*) The factor 0.9 accounts for uncertainties in fabrication.

Recent research on shear lag in HSS

A number of studies on gusset plate connections have been carried out in the last few years.
The latest study, on a special type of gusset plate connection, the so-called hidden joint
connection, by Willibald (12) showed that shear lag was not critical for square HSS in this
specific connection type but can become critical for rectangular HSS. The results of the
parametric study supported the use of the American specifications (2, 4, 5) but indicated a
generally overly conservative approach in all current design methods. In an experimental
study by British Steel (Swinden Laboratories 13) on slotted end plate connections for
circular, square and rectangular hollow sections, 13 of the 24 specimens failed by shear lag.
The results of an experimental as well as numerical investigation on shear lag failure for
slotted circular hollow sections were given by Cheng et al. (14). Nine tests on gusset plate
connections to CHS tension members were performed, but none of the specimens failed by
shear lag. However, the experimental and numerical investigations showed that
considerable stress concentrations occur at the slot ends. Comparing the results of the study
with the then current Canadian (Cheng et al. 15) as well as American specifications (Cheng
and Kulak 16), it was shown that neither code accurately represented the behaviour of
slotted circular hollow section connections. In contrast to the specifications, Cheng and
Kulak (16) concluded that shear lag failure was not critical for round HSS if the connection

448 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


length was longer than 1.3 times the diameter of the circular hollow section (U = 1.0),
however all their tests but one were performed with the slot end welded. The results of this
research will be incorporated in the upcoming AISC (5) specification (see Table 1).

A study on shear lag in slotted square and rectangular hollow sections has been performed
by Korol et al. (10). A total of 18 specimens was tested under tensile loading with seven
specimens failing by shear lag. The authors concluded that for six of the seven specimens
that failed by shear lag, all with Lw/w ≈ 1.0, the connection capacity was nearly equal to the
tensile capacity of the hollow section, Nu = An · Fu. One specimen, where Lw/w = 0.61, failed
very prematurely due to shear lag. For specimens with Lw/w-ratios smaller than 0.6, base
metal shear failure of the hollow section governed. The influence of the eccentricity x on
the connection capacity was found to be only minor. Based on the results of the earlier
study, Korol (17) proposed a slightly modified approach for the calculation of the effective
shear lag net section area. Instead of using the efficiency factors as given in the Canadian
or American specifications, less conservative formulae were provided:
α = 1.0 for Lw/w ≥ 1.2 (net/gross section failure governs) (4a)
α = 0.4 + 0.5 Lw/w for 1.2 > Lw/w ≥ 0.6 (shear lag failure governs) (4b)
non-applicable for Lw/w < 0.6 (block shear tear-out governs).
The eccentricity factor U was then calculated by:
x
U = 1.0 – 0.4 L (5)
w

The effective shear lag net section was then given by:
Ae = An · α · U (6)

EXPERIMENTAL STUDY

Scope of testing

The experimental study comprised of six gusset plate connections for round hollow sections.
The specimens had varying fabrication details (see Figure 2): slotted versus unslotted HSS,
slot end welded versus no weld return. A further parameter in the test series was the weld
or connection length Lw with the Lw/w-ratio varying between 0.66 and 0.88 (with
w = 0.5 · π · D - tsl or w = 0.5 · π · D – tp). Standard cold-formed 168 x 4.8 mm Class C
hollow sections with a specified yield stress of 350 MPa (CSA 18) were used. The gusset
plate was made out of 1" (25.4 mm) Grade 300W steel. Table 2 shows the dimensional, and
Table 3 the material, properties of the tested specimens. The welds connecting the hollow
section and the gusset plate were standard 10 mm fillet welds using E480XX electrodes
(CSA 19). Each specimen was equipped with 10 linear strain gauges measuring the
longitudinal strain distribution on the hollow section (see Figure 5). The displacement of the
connection was measured by four LVDTs (Linear Variable Differential Transformers). The
results of these measurements were later on used to verify the numerical models of the
tested specimens.
22.5o
50 50 50 22.5o
8
7 22.5o
6 4 2
10 5 3 1 22.5o
9

Figure 5. Strain gauge locations on tested specimens.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 449


1400

Connection Load [kN]


1200
1000
800 Sp. 1
600 Sp. 2
Shear lag
Sp. 3
Tear-out 400 Sp. 4
200 Sp. 10
Sp. 11
0
0 5 10 15 20 25
Displacement [mm]
Figure 7. Load versus Displacement curves of the
tested specimens.
Figure 6. Specimen 1 at failure.

Table 2. Measured dimensional properties of test specimens.


tsl w Lw tp
Specimen HSS wp [mm] Connection type
[mm] [mm] [mm] [mm]
Slotted tube, slot end
1 156 197
not filled
2 168.5 x 4.9 169 197 Slotted tube, slot end
27.0 238
3 Ag = 2513 mm2 208 198 filled
25.7
with Slotted tube, slot end
4 192 198
x = 53.6 mm not filled
10 162 2 x 74.3
- 239 Slotted plate
11 195 2 x 75.5

Table 3. Measured material properties of test specimens.


E [MPa] Fy [MPa] Fu [MPa] εu [%]
HSS 196000 498* 540 25.9
Plate 201000 358 482 28.0
)
* Using the 0.2% offset method, as material was cold-formed.

Test results

Failure of all six specimens was caused by either shear lag (specimens 3, 4, 10, 11)), tear-
out of the HSS base material along the weld (specimen 2) or both failures taking place at the
same time (specimen 1, see Figure 6). Shear lag failure causes the HSS to fail circum-
ferentially while block shear tear-out happens along the weld. Table 4 shows the ultimate
connection strength, the failure mode as well as the predicted failure loads according to
current design methods. Generally, the specimens with the shorter connection lengths
(specimens 1, 2 and 10) had a reduced connection capacity. Before failure, all specimens
showed ovalization in the hollow section, especially pronounced in the specimens with the
unslotted tube and the slotted plate (see Figure 2 (a)). In all specimens, the strain gauge
readings showed very high stresses at the strain gauges closest to the weld at the beginning
of the connection (strain gauge 5, see Figures 3 and 5). The strain gauge furthest away
from the weld (strain gauge 8) reported either negligible or even negative strains (specimens
10 and 11) at ultimate load. The strain distribution along the weld and beyond could be

450 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


observed by comparing strain gauges 1, 3, 5 and 10. Again, strain gauge 5 had the highest
strain levels in each specimen.

Considerable displacement took place in all tested specimens (see Figure 7). The
specimens having an unslotted tube and a slotted plate showed the largest deformation
before reaching their ultimate loads. Considering these sizable displacements, it might be
necessary to define a deformation limit which, for some connections, will then govern their
capacity. Having comparable connection lengths, specimens 1,2 and 10 as well as
specimens 3, 4 and 11 have only slightly different capacities. This indicates that the
fabrication detail only has a little influence on the connection strength. Currently most codes
do not specify the use of a certain detail but provide one design method to cover all three
cases. However, with increasing gusset plate thickness, the difference in connection
strength between the various fabrication details might be more pronounced. Generally, the
design methods found in the American specifications (AISC 4,5) show the best agreement
with the tests but further research seems necessary.

Table 4. Actual and predicted connection strength of test specimens.


Packer and
Failure AISC AISC CSA Korol
Specimen Test Nux/AnFu Nux/AgFu Henderson
Mode* (4) (5) (3) (17)
(11)
Nux or Nu [kN] 1032 SL, 890 796 597 754 762
1 0.85 0.76
Nux/Nu - TO 1.16 1.30 1.73 1.37 1.35
Nux or Nu [kN] 1087 881 829 647 754 801
2 TO 0.89** 0.80
Nux/Nu - 1.17 1.31 1.68 1.44 1.36
Nux or Nu [kN] 1211 973 902 798 754 913
3 SL 1.00** 0.89
Nux/Nu - 1.20 1.34 1.52 1.61 1.33
Nux or Nu [kN] 1154 979 876 737 754 868
4 SL 0.95 0.85
Nux/Nu - 1.18 1.32 1.57 1.53 1.33
Nux or Nu [kN] 1107 907 907 688 842 869
10 SL 0.81
Nux/Nu - 1.22 1.22 1.61 1.31 1.27
Nux or Nu [kN] 1196 984 984 829 842 975
11 SL 0.88
Nux/Nu - 1.22 1.22 1.44 1.42 1.23
*) SL stands for shear lag failure and TO stands for block shear tear-out failure along the weld;
**) As slot end was welded, it might be also appropriate to assume An = Ag;
Nu = Ae · Fu.

NUMERICAL STUDY

For further study of gusset plate connections, a numerical study has been started. The final
goal of this numerical study will be a parametric study concentrating on the influences of
several variables: weld length, hollow section diameter, hollow section wall thickness, the
eccentricity of the top or bottom part of the HSS, and fabrication details.

The Finite Element program ANSYS 5.7 (Swanson Analysis System Inc. 20) has been used
for the numerical study. A geometric and material non-linear analysis was performed for all
specimens. 8-noded, large strain solid elements (solid45) with reduced integration and
hourglass control were used throughout. The material properties were input as a multi-linear
curve with the engineering stress and strain converted to the true stress and strain values.
To simulate cracking in the models, a maximum equivalent plastic strain limit was used. The
so-called "birth and death" elements allow the user to significantly reduce the stiffness of the
elements, or "kill them", if an equivalent plastic strain is reached.

Due to the symmetry of the connection it was only necessary to model an eighth of the
connection (see Figure 8). The welds were fully modelled. A gap between the gusset plate

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 451


Figure 8. Numerical model of specimen 2.

Table 5. Comparison between test and numerical results.


Specimen Nux [kN] NFE [kN] Nux/ NFE
2 1087 1073 1.01
3 1211 1190 1.02

1400
1400
1200
1200
Connection Load [kN]

Connection Load [kN]

1000 1000

800 Spec. 2 800 Strain Gauge 5


600
FE Spec. 2 Spec. 2
600
Spec. 3 FE Spec. 2
400 FE Spec. 3 400
Spec. 3
200 200 FE Spec. 3

0 0
0 2 4 6 8 10 12 14 16 0 5000 10000 15000 20000 25000 30000
Displacement [mm] Strain [µm]
Figures 9a and 9b. Comparison between test and numerical results.

and the hollow section was modelled to prohibit any direct stress transfer between the plate
and the hollow section thus forcing load transfer to occur only via the welds. Symmetry
boundary conditions were employed along the planes of symmetry (translations normal to
the plane of symmetry were fixed) and the nodes at the HSS end were fixed. The finite
element models were then loaded by displacing the nodes at the end of the gusset plate.

Specimens 2 and 3, which have a slotted HSS with the slot end welded (fabrication detail
(c), see Figure 2) have been numerically modelled and show very good agreement with the
tests. The predicted ultimate loads for these two specimens are both within 2% of the actual
ultimate loads for each test (see Table 5). Figure 9a compares the load-displacement
curves for specimens 2 and 3 with the respective results of the numerical models. For both
specimens the agreement is very good up to peak load. Unfortunately, the numerical
models had problems converging beyond a certain point. At this load step, a high number of
elements are “killed” which causes a sudden change in stiffness thus causing severe
convergence problems. Yet, due to the high number of lost elements it is safe to assume
that the ultimate load of the FE-model has been reached and subsequent calculation steps
would result in a lower connection load. The comparison of the most critical strain gauge
(strain gauge 5, see Figure 5) also shows good agreement between test and FE model (see

452 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Figure 9b). For the future, it is planned to have comparable models for the other two
fabrication details and to use these finite element models in a parametric study.

CONCLUSIONS

An experimental, numerical as well as literature study on shear lag in round hollow section
gusset plate connections under tension loading has been carried out. The experimental
study showed that shear lag can indeed become critical in gusset plate connections. The
connection length had the largest effect on the connection capacity, whereas the fabrication
detail of the connection (see Figure 2) only had a minor influence on the capacity. For some
specimens, large displacements could be observed before failure, which could become
critical if deformations are restricted by a deformation limit.

The numerical study showed that it is possible to generate very good finite element models
of these connections. The numerical study will soon be extended to do further parametric
studies to finally provide suitable design methods against shear lag failure. The design
methods that can be found in current international specifications have been introduced in the
literature study and have been evaluated against the experimental research carried out. At
present, the American specification (AISC 5) seems to be best suited to design against
shear lag failure under quasi-static tension loading, but all design methods are overly
conservative and additional research is still necessary.

For the future, further research is currently planned on shear lag in gusset plate connections
under cyclic loading, as can be found in earthquake situations. With these connections,
special attention will be paid to the fabrication and refined connection details will be
considered.

ACKNOWLEDGEMENTS

Financial support for this project has been provided by CIDECT (Comité International pour le
Développement et l’Etude de la Construction Tubulaire) Programme 8G and NSERC
(Natural Sciences and Engineering Research Council of Canada). IPSCO Inc. and Walters
Inc. (Hamilton, Canada) generously donated steel material and fabrication services,
respectively.

NOTATION

Ag = gross cross-sectional area of hollow section


Ae = effective net cross-sectional area of hollow section
An = net cross-sectional area of hollow section
D = outside diameter of round hollow section
E = modulus of elasticity
Fy = yield tensile stress
Fu = ultimate tensile stress
Lw = weld length
NFE = connection strength as predicted by numerical model
Nu = calculated connection strength
Nux = measured connection strength
U = coefficient for shear lag net section fracture calculation
t = wall thickness of hollow section
tp = thickness of gusset plate
tsl = width of slot in hollow section

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 453


wp = width of gusset plate
w = distance between the welds, measured around the perimeter of the HSS
x = eccentricity ratio
α = coefficient for shear lag net section failure calculation
εu = ultimate strain at rupture
θ = angle between gusset plate centre line and radial line of hollow section

REFERENCES

(1) Eurocode 3, (1993). Design of steel structures - General rules - Part 1-8: Design of
joints. Draft version. British Standards Institute, London, England.
(2) AISC, (2000). Load and Resistance Factor Design Specification for Steel Hollow
Structural Sections. American Institute of Steel Construction, Chicago, USA.
(3) CSA, (2001). Limit States Design of Steel Structures. CAN/CSA-S16-01. Canadian
Standards Association, Toronto, Canada.
(4) AISC, (1999). Load and Resistance Factor Design Specification for Structural Steel
Buildings. American Institute of Steel Construction, Chicago, USA.
(5) AISC, (2003). Specification for Structural Steel Buildings. Draft (December 1, 2003)
version of the forthcoming (2005) Specification. American Institute of Steel
Construction, Chicago, USA.
(6) AIJ, (2002). Recommendations for the Design and Fabrication of Tubular Truss
Structures in Steel. (in Japanese) Architectural Institute of Japan, Japan.
(7) Chesson E., Jr., and Munse, W.H. (1963). Riveted and bolted joints: Truss type
tensile connections. Journal of the Structural Division, ASCE, 89(1), 67-106.
(8) Kirkham, W.J., and Miller, T.H. (2000). Examination of AISC LRFD shear lag design
provisions. Engineering. Journal, AISC, 3rd Quarter, 83-98.
(9) CSA, (1994). Limit States Design of Steel Structures. CAN/CSA-S16.1-94. Canadian
Standards Association, Toronto, Canada.
(10) Korol, R.M., Mirza, F.A., and Mirza, M.Y. (1994). Investigation of shear lag in slotted
HSS tension members. Proceedings, 6th International Symposium on Tubular
Structures, Melbourne, Australia, 473-482.
(11) Packer, J. A., and Henderson, J. E. (1997). Hollow structural section connections and
trusses - A design guide. 2nd Ed., Canadian Institute of Steel Construction, Toronto,
Canada. ISBN: 0-88811-086-3.
(12) Willibald, S. (2003). Bolted Connections for rectangular hollow sections under tensile
loading. PhD thesis, University of Karlsruhe, Germany.
(13) Swinden Laboratories, (1992). Slotted end plate connections. Report No.
SL/HED/TN/22/-/92/D. British Steel Technical, Rotherham, England.
(14) Cheng, J.J.R., Kulak, G.L., and Khoo, H. (1996). Shear lag effect in slotted tubular
tension members. Proceedings, 1st CSCE Structural Specialty Conference,
Edmonton, Alberta, Canada, 1103-1114.
(15) Cheng, J.J.R., Kulak, G.L., and Khoo, H. (1998). Strength of slotted tubular tension
members. Canadian Journal of Civil Engineering, Vol. 25, 982-991.
(16) Cheng, J.J.R., and Kulak, G.L. (2000). Gusset plate connection to round HSS tension
members. Engineering Journal, AISC, 4th Quarter, 133-139.
(17) Korol, R.M. (1996). Shear lag in slotted HSS tension members. Canadian Journal of
Civil Engineering, Vol. 23, 1350-1354.

454 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


(18) CSA, (1998). General Requirements for Rolled or Welded Structural Quality Steel/
Structural Quality Steel. CAN/CSA-G40.20/G40.21-98. Canadian Standards
Association, Toronto, Canada.
(19) CSA, (2003). Welded Steel Construction (Metal Arc Welding). CAN/CSA-W59-03.
Canadian Standards Association, Toronto, Canada.
(20) ANSYS Release 5.7. (2000). Swanson Analysis System Inc., Houston, USA.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 455


456 Connections in Steel Structures V - Amsterdam - June 3-4, 2004
REVIEW OF TUBULAR JOINT CRITERIA

P.W. Marshall, Moonshine Hill Proprietary, Texas

ABSTRACT
This note reviews and compares four sets of tubular connection design criteria
for axially loaded circular tubes. The four criteria are AWS D1.1, API RP2A,
ISO/WD 15-1.2, and ANSI/AISC 360-05.

INTRODUCTION

The existing American design codes for welded tubular connections are AWS D1.1 Structural
Welding Code (1990 thru 2002) and the substantially identical AISC Specification for the
Design of Steel HSS (1997), the basis of which is documented in the author’s book (1).
Three new sets of design criteria are in the works. They are:

(1) Proposed update to API RP2A, Design... of Fixed Offshore Platforms, based on research
conducted by Prof. Pecknold at the University of Illinois, and sponsored the API Offshore
Tubular Joint Research Consortium. A lengthy Commentary is included to self-document
these criteria. Extensive nonlinear finite element analyses were used to extend the
experimental data base, particularly in the areas of overlapped K-joints, a moment-free
baseline for T-joints, and chord stress interaction for a wide variety of joint types and
loadings. In view of reduced scatter compared to existing criteria, a reduced WSD safety
factor of 1.6 is proposed. This update has been approved for publication in the 22nd edition
of RP2A, and is in the final stages of editing.

(2) Static Strength Procedure for Welded Hollow Section Joints, IIW doc XV-E-03-279,
based on CIDECT research. This is also on the fast track to becoming an international
standard as ISO/WD 15-1.2, with IIW commission XV as secretariat. The immediate purpose
of this note was to provide comments for a Sept. 2003 meeting of IIW s/c XV-E.

(3) ANSI/AISC 360-05, Standard Specification for Structural Steel Buildings (draft of August
20, 2003), Chapter K, HSS Connections, being prepared by an ad hoc task group under the
direction of Larry Kloiber. This is essentially the same as the CIDECT-based IIW document,
except that it gives the characteristic ultimate strength without hiding a partial safety factor
therein. Separate safety factors are then given for LRFD and ASD.

COMPARISON OF THE DESIGN EQUATIONS

Principal results of this review are shown in the Tables and Figures. Table 1 gives a side-by-
side tabulation of the design criteria for different types of circular joints. The square bracket
term is Qu in API, and simply written out in IIW and AISC. AWS criteria have been converted
to this format for comparison.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 457


Table 1. Design equations.
CIRCULAR AWS D1.1-92 API RP 2A-WSD CIDECT AISC 360-05
TUBE JOINTS and AISC ‘97 HSS Offshore Platforms IIW-XV-E-03-279 Chapter K
-- AXIAL LOAD Connection Manual 22nd edition ISO/WD 15-1.2 Kloiber draft 8/03
Pu sin θ = to2 Fyo [*] Qf
General format FS Pa sin θ = T2 Fyo [*] Qf Pn sin θ = to2 Fyo [*] Kp Pu sin θ = t2 Fyo [*] Qf
[*] defined below
[*] = 2.8 + (20 + 0.8γ) β1.6
T & Y joint [*] = (18.8β + 3.4) √Qβ [*] = (2.8 + 14.2β2) γ0.2 [*] = (3.1 + 15.6β2) γ0.2
but ≤ 2.8 + 36β1.6
[*] = 32β/α + 3.4 with [*] = (16 +1.2γ) β1.2 Qg ≤ 40 β1.2 Qg [*] = (1.8 + 10.2β) Kg
K & N gap [*] = (2.0 + 11.33β) Qg
α = 1.0 + 0.7 g/di ≤ 1.7 Qg = 1+ 0.2(1-2.8g/D)3 **
[*] = (32 β + 3.4) f 1 (q) [*] as for gap with ⎡ 0.024γ 1.2 ⎤ with Qg same as
K g = γ 0.2 ⎢1 + ⎥
K & N overlap
+ 3.2τβγ
FEXX
f 2 (q )
Qg = 0.13 + 0.65 φ √γ ∗∗ ⎢⎣ 1 + exp(0.5g/t 0 − 1.33) ⎥⎦ CIDECT Kg and
Fy and φ = (tb/Fyb) / (toFyo) g negative for overlap

[*] = (13.3β + 3.4) Qβ [*] = [2.8 + (12 + 0.1γ) β] Qβ [*] = 5.2


[*] = 5.7
X joint
Also can length effect Stronger alternate for tension 1 − 0.81β 1 − 0.81β
As for K with eff. β = ∑ βi / 3
See general Use gap between
K-T joint Special combo rules Not covered
multiplanar loaded diagonals
& missing cases
K-K joint See general multiplanar Check chord gap for shear + axial
Not covered Not covered
(delta truss) (no increase over K joint) (no increase over K joint)
General [*] = (32β/α +3.4) Qβ0.7(α -1) Commentary reference X-X joints covered
Not covered
multiplanar with α per Annex L To AWS method General case not covered
Chord load Tension or compression Comp.(+), tens.(-) at footprint Compression only
effect Qf = 1 - 0.03 λ γ Ū2 P M 2 Kp = 1 – 0.3 Ū (1+ Ū) Same as CIDECT
Q f = 1 − C1 − C2 − C3 U
Qf or Kp 2 2
Ū = [(P/Py) +(M/My) ] 2 Py My Ū = P/Py + M/My
0.3
Qβ = forβ > 0.6 ** Qg not defined for Pu is characteristic
Other terms β (1 − .833β ) Pn includes resistance factor
0.05 > g/D > -0.05 ultimate

458 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Details of the AWS conversion to international format for overlap K&N joints are given below,
where q = percent overlap as a fraction. In the existing Code, vp is the allowable punching
shear and the allowable capacity normal to the chord is given by...

Allowable Pn sin θ = vp to L1 + 2 vw tw L2

For conversion, the following substitutions are made...


L1 = π β D f1(q) (partial footprint arc) vw = 0.3 FEXX (allowable)
L2 = β D f2(q) (lap weld ┴ to chord) vw = 0.8 FEXX (ultimate)

with the resulting international format given by...


FEXX
Ultimate Pu sin θ = to2 Fyo [*] Qf with [*] = (32 β + 3.4 ) f1 ( q ) + 3.2τβγ f 2 (q )
Fy

In Figure 1, f1 and f2 shown for 45º N-joint with β < 0.5. These values were obtained
graphically from a scale layout of joints with varying degrees of overlap.

1.0
f1(q)

0.5

0.0
0.0 0.5 1.0 L2
q
L
1

1.0

q(βD)
f2(q)

0.5

0.0
0.0 0.5 1.0
q

Figure 1. Layout and functions f1 and f2 for overlapping 45º N-joints with β=0.5.

It may be noted from Table 1 that the AWS criteria cover a wider variety of design situations
than the others, with particular reference to the general multi-planar case. The proposed
AISC criteria cover the least, in a deliberate effort to minimize complexity.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 459


RESULTS

Figures 2-5 show parametric comparison of the criteria for T&Y joints and X joints, which was
performed on an Excel spreadsheet. The base case is beta of 0.5, tau of 0.5, and gamma of
20. The upper plots show the effect of varying beta, with the other parameters kept at the
base case. The author’s 1969 and 1975 criteria were subsequently shown to be un-
conservative for X-joints, but they are very close to the latest OTJRC results for T&Y joints.
The lower plots show the effect of varying gamma; there is no effect if the T2Fy format tells
the whole story. There was no effect of varying tau in any of these cases.

variations for X joints

AWS & AISC Hdbk API proposal AISC proposal

40

30
Qu

20

10

0
0 0.2 0.4 0.6 0.8 1 1.2

Figure 2. Effect of β for X-joints with γ=20.

variations for X joints

AWS & AISC Hdbk API proposal AISC proposal

20

15
Qu

10

0
0 5 10 15 20 25 30 35

Figure 3. Effect of γ for X-joints with β=0.5.

460 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


variations for T & Y joints

AWS & AISC hDBK API proposal AISC proposal

40

30
Qu

20

10

0
0 0.2 0.4 0.6 0.8 1 1.2

Figure 4. Effect of β for T & Y joints with γ=20.

variations for T & Y joints

AWS & AISC Hdbk API proposal AISC proposal

20

15
Qu

10

0
0 5 10 15 20 25 30 35
γ

Figure 5. Effect of γ for T & Y joints with β=0.5.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 461


Figures 6-8 plot plots results for K&N joints. The large-scale plot in Figure 6 shows the effect
of gap or overlap for base case beta, tau, and gamma. The single expression in the
CIDECT-based AISC proposal, covering both gap and overlap, matches AWS in the gap
range, and matches API finite element results in the overlap range, at least for the base case
parameters.

K & N joints

AWS & AISC Hdbk API proposal AISC proposal

40

30
Qu

20

10

0
-1 -0.5 0 0.5 1
(overlap) g/D (gap)

Figure 6. Effect of gap/overlap for 45º N-joint with β=0.5, τ=0.5, and γ=20.

Figures 7 and 8 show the effect of varying the parameters β, γ, and τ at 60% overlap and
0.1D gap, respectively. Here we see that the proposed AISC (and IIW/CIDECT) criteria
completely miss the strong effect of tau in the overlap region, as predicted by the AWS
strength-of-materials approach and confirmed by the API finite element studies. They also
appear to under-predict the beneficial effect of large beta.

462 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


overlap joints

AWS & AISC Hdbk API proposal AISC proposal

80
60
Qu

40
20
0
0 0.2 0.4 0.6 0.8 1 1.2

overlap joints

AWS & AISC Hdbk API proposal AISC proposal

50
40
30
Qu

20
10
0
0 5 10 15 20 25 30 35

overlap joints

AWS & AISC Hdbk API proposal AISC proposal

40
30
Qu

20
10
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Figure 7. Effects of β, γ, and τ for overlap N-joints with g/D = -0.3.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 463


gap joints

overlap joints API proposal AISC proposal

80
60
Qu

40
20
0
0 0.2 0.4 0.6 0.8 1 1.2

gap joints

AWS & AISC Hdbk API proposal AISC proposal

30

20
Qu

10

0
0 5 10 15 20 25 30 35

gap joints

AWS & AISC Hdbk API proposal AISC proposal

30

20
Qu

10

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Figure 8. Effect of β, γ, and τ for gapped N-joints with g/D = 0.1.

464 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Finally, Figure 9 shows the chord load effect, with the reduction factor called Qf in US specs
and kp internationally. All criteria pass through unity at zero chord load. Based on Yura’s
tests, AWS criteria show the effect of chord buckling tendencies at large gamma, a feature
which is missing from the other criteria. The API-OTJRC finite element parameter studies
show quite different results for different brace types and loadings, and are represented in the
criteria by a quadratic expression with tabulated coefficients; linear terms tilt the parabolas.
AISC and IIW proposals simplify this to a single line, which is reasonable representation of
the effect on joint strength at chord service loads, for most joint types. The notable exception
is equal diameter X joints with compression in the branches, where biaxial membrane
stresses at the saddle position make tensile chord loads detrimental and compressive chord
loads slightly beneficial.

new API K-axial axial new API K-axial IPB


new API T/Y axial axial new API T/Y comp IPB
new API X β<.9 axial new API X β=1 com ax ial
new API bending axial new API bending IPB
AWS γ = 20 axial combined AWS γ = 20 IPB combined
AWS γ = 20 OPB combined new AISC all cases all cases

1.2

0.8

0.4

0
-1 -0.5 0 0.5 1

Figure 9. Reduction factor (kp or Qf) versus chord utilization.

DISCUSSION

Q. What are conclusions of the comparison?

A. When comparing existing AWS-AISC criteria for circular tubular connections to CIDECT,
both in 1992 and today, neither criteria appear to have significantly different errors on the
unsafe side. Thus, one may ask the following questions:
“Why churn the Code by adopting essentially similar criteria but in a different format?”
“Why not look at new API results having a more significant impact on reliability?”
The issue is not simply whether or not to maintain the American status quo. It is important to
keep the Codes evergreen in the sense that they reflect the latest data, with researchers still

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 465


active into the future who remember where it all comes from.

Q. There are obviously parameters that are treated very differently (tau overlapped and
overlap) and others that vary much less between the standards. Why?

A. Tau: All the criteria capture the effect of tau (t/T) in the same way, but without explicit
expression in point load criteria for T, Y, and X connections. This is not to say that tau is
unimportant; indeed, its primary importance was more obvious to the user in the old AWS
punching shear format. When one gets called in after the fact on structural failures and
tubular projects in trouble, one of the first things to look for is excessive tau ratios.

Tau effect in overlapped connections: In the old AWS-API-AISC criteria, this is captured by
mechanistic consideration of both punching at the partial footprint (L1) and shear in the
overlap weld (L2). In Pecknold’s new API criteria, this is based on an extensive inelastic
finite element parameter study, with totally separate Qg expressions for gap and overlap
joints. Pecknold’s indicated higher strength for large beta and large tau is also consistent
with the unexpectedly good performance of same-size overlapped K-bracing in Hurricane
Andrew. In CIDECT-IIW-ISO criteria, overlap is treated as an extension of the behaviour of
gap connections in which tau has no effect; a single Kg expression for both is curve-fit to the
smaller empirical test data base. This makes computerized design easier, but gives the
designer no insight into the physical mechanism of load transfer.

Effect of overlap amount: Both Pecknold (new API) and CIDECT agree that there is a
significant, but nearly constant, beneficial effect beyond g/D of -0.1. The old AWS-API
approach (dating back to 1975) was apparently intuitively appealing but wrong, while
remaining on the safe side for moderate amounts of overlap. Pecknold gives no equations
for |g/D| smaller than 0.05, and suggests interpolation in this region. In older codes, this was
a prohibited zone, due to concern over creating a weak hard spot and awkward welding
conditions. The smooth transition shown in the CIDECT strength criteria (and in Efthymiou’s
SCF criteria) may simply be an artefact of curve-fitting. This issue needs to be re-examined,
seeking data in the prohibited zone. Welded connections with tensile strains over 2% in
inelastic finite element solutions may be considered vulnerable to fracture.

Effect of chord loading: The 1972 AWS Code included a modest Qf penalty for compressive
chord loads, based on Japanese data. Existing AWS-AISC criteria reflect further effects of
gamma and chord load type, based on Yura’s X-joint data. CIDECT criteria are simpler, with
Kp close to being on the safe side of Yura. Pecknold’s criteria, based on extended finite
element results, reflect effects of both chord and brace load type (but not gamma). This is
reasonably consistent with CIDECT for K and N connections, but its better prediction for
other types of connections has a significant effect on reliability, prompting a modest reduction
in working stress design safety factor in API. One common design case in which both
CIDECT and AWS-AISC are significantly on the unsafe side is equal size X-braces with
equal but opposite loads and no joint can. A caveat for this case is urgently needed.

Q. Do we need more data to choose the right direction for the parameters that vary widely?

A. Pecknold’s API data base includes both the CIDECT physical tests and his extended
finite element results. It could be readily compared to CIDECT-IIW-ISO criteria, to quantify
the reliability consequences of adopting these into AISC. Gathering additional research data
must be left to the future. In order to meet its ambitious publication schedule, AISC should
select one set of criteria and stick with it – no mix-and-match. For circular connections at this
point in time, Pecknold’s criteria have not been as widely vetted as CIDECT, while CIDECT
does not have the extensive set of worked examples and familiarity to American designers
as the existing AWS-AISC criteria.

466 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Q. Are the difference due to the wide difference in the type of structures the standards are
based on and do these variances really show variances due to exceeding testing limits i.e.
do the API people test 48 dia 1 inch connections while the CIDECT people test 4 in x 1.4
tube?

A. The differences are not always that great. D/t of 48 is quite typical for offshore structures,
so punching at the material shear strength rarely governs. Most CIDECT tests are in the D/t
range of standard weight structural tubing, which ranges from 6 to 34. European bridge
designers seem to favor the 3 to 13 D/t range of double extra strong. American designers
use D/t up to 120 for fabricated large diameter chord members, with much thicker joint cans
at the nodes. None of the design criteria show an adverse size or thickness effect on static
strength, although fracture mechanics and some of the tension test data suggest one. AWS
criteria give no bonus for tension.

Q. What is the effect on joints designed and built routinely?

A. People generally want to know the impact on cost and safety before they change a
design code. AISC should get a feel for this while re-working all the example problems and
tabulated results in the HSS manual. Bridges and offshore structures are also influenced by
fatigue, so the impact of a static strength change is muted.

CONCLUSION

Although this paper may be regarded as a “stream of consciousness” examination of ongoing


design code developments, it is already drawing worthwhile discussion.

ACKNOWLEDGEMENTS

The foregoing Q&A discussion was prompted by thoughtful review of an earlier draft of this
note from Tom Schlafly at AISC. Drafting of the figures has been performed by Dakang Liu
at TU Delft. The author is grateful to the conference organizers for accommodating this
paper.

REFERENCE

(1) P. W. Marshall, Design of Welded Tubular Connections, Developments in Civil


Engineering #37, Elsevier Science Publishers, Amsterdam, 1992 (limited availability at
civilbooks.com)

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 467


468 Connections in Steel Structures V - Amsterdam - June 3-4, 2004
CONCLUSIONS

J.W.B. Stark, Delft University of Technology, The Netherlands

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 469


CONNECTIONS IN STEEL STRUCTURES V

CONCLUSIONS

24 individual countries

56 attendees

EU 32
America (N&S) 16
Asia 5
Others 3

June 2004 ECCS/AISC workshop : Connections in Steel Structures V 1

STRUCTURAL CONNECTIONS
conception
conception

design calculation
design calculation

code check
code check
process
fabrication
fabrication

erection
erection

maintenance
maintenance

June 2004 ECCS/AISC workshop : Connections in Steel Structures V 2

470 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


TRENDS IN STRUCTURAL CONNECTIONS
conception
conception

1960
tubular connections

cast steel nodes


simple connections
sell steel plug & play connections

steel-to-concrete connections

composite connections
2004

June 2004 ECCS/AISC workshop : Connections in Steel Structures V 3

TRENDS IN STRUCTURAL CONNECTIONS


design calculation
design calculation

1960
plastic design

non-linear analysis
numerical
calculations 3-D analysis

component method

integration design - fabrication


2004

June 2004 ECCS/AISC workshop : Connections in Steel Structures V 4

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 471


TRENDS IN STRUCTURAL CONNECTIONS
code check
code check

1960
ISO ; IIW ; CEN

globalisation limit state design

statistical calibration

detailed rules

2004

June 2004 ECCS/AISC workshop : Connections in Steel Structures V 5

TRENDS IN STRUCTURAL CONNECTIONS


fabrication
fabrication

1960
gas shielded metal-arc welding

drilling – sawing equipment


N-C tools
automatic burning

punching

2004 integration design - fabrication

June 2004 ECCS/AISC workshop : Connections in Steel Structures V 6

472 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


TRENDS IN STRUCTURAL CONNECTIONS
erection
erection
1960
mobile cranes

off-site production
health & safety
requirements
friction grip bolting

injection bolts

2004 plug & play connections

June 2004 ECCS/AISC workshop : Connections in Steel Structures V 7

TRENDS IN STRUCTURAL CONNECTIONS


maintenance
maintenance
1960
new painting systems

weathering steel
environmental
requirements
stainless steel

aluminium

2004

June 2004 ECCS/AISC workshop : Connections in Steel Structures V 8

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 473


STRUCTURAL CONNECTIONS
input conception
conception

knowledge design calculation


design calculation

codes code check


code check

fabrication
fabrication
tools
products
erection
erection

maintenance
maintenance

June 2004 ECCS/AISC workshop : Connections in Steel Structures V 9

STRUCTURAL CONNECTIONS
conception
conception

knowledge design calculation


design calculation

codes code check


code check education

fabrication
fabrication innovation
tools
products
erection
erection

maintenance
maintenance

June 2004 ECCS/AISC workshop : Connections in Steel Structures V 10

474 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


CONNECTIONS IN STEEL STRUCTURES V
Where is your interest ?

June 2004 ECCS/AISC workshop : Connections in Steel Structures V 11

CONCLUSIONS FROM SESSIONS


Session 1: Design codes, design and education

- Duncan : New AISC design standard to be issued in 2005


LSD based on nominal strength criteria and alternatively ASD
Connection design criteria detailed in 2 chapters
Presentation of new criteria for bolted and welded connections

- Geschwinder : History of AISC steel design standards


Discussion of shear lag and block shear provisions

- Evers : Importance of communication between architect, engineer, fabricator


Detailing to be done early in the design process
Connections account for a large percentage of the total cost
Better use heavier steel shapes and cheap connections

- Wald : Large project on continuing education for steel connection design


Various educational materials (textbook with many examples),
including CD and internet facilities

June 2004 ECCS/AISC workshop : Connections in Steel Structures V 12

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 475


CONCLUSIONS FROM SESSIONS
Session 2: Modelling

- 3 Papers on the behaviour of the equivalent T-stub in tension (steel and aluminium)
All proposing methodologies for the assessment of this component
Special emphasis on ductility
under monotonic (2 papers) and
cyclic loading conditions (2 papers)

- 2 Papers on sophisticated non-linear numerical modelling of complete joints


One addressing finite element issues
The other more concerned with practical application
Both papers comparing numerical results with experimental results

- 1 Paper with a simple approach (4 parameter model) aimed at the characterization


of the moment-rotation response of end-plate joints, collected from a database of
tests, for use in system analysis

June 2004 ECCS/AISC workshop : Connections in Steel Structures V 13

CONCLUSIONS FROM SESSIONS


Session 3: Modelling, deformation capacity, seismic

- Charney : Simplified models capable of handling panel zone deformations.


Through the use of displacement participation factors and
sensitivity indices it was shown that the scissors model (with only
4 DOF) was capable of providing accurate results.

- Beg & Monte Carlo simulation aimed at determining the influence of


material properties on the M-θ curve for bolted connections
Simões da Silva : Both presentations indicated that further work was needed
before final recommendations could be made.

- Dubina : Characterization of the deformation capacity under cyclic loads


based on the results of both monotonic and cyclic tests.
The results indicate that the cyclic deformation for T- and clip
angle steel and composite connections are roughly ½ of that
given by monotonic tests.

June 2004 ECCS/AISC workshop : Connections in Steel Structures V 14

476 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


CONCLUSIONS FROM SESSIONS
Session 4: Seismic

- 5 Papers on seismic design in this session

- A good seismic resistant structure must provide a good balance between


strength, stiffness and very importantly ductility
For connections, this balance is even more important !
Although focussing particular problems, the papers in this session proved,
once more the importance of :
9 Proper design criteria ( Ricles )
9 Accuracy of theoretical models used for design checks ( Leon )
9 Quality of conception and detailing ( Dexter & Dunai )
and that there is still place for innovative solutions ( Stratan )

June 2004 ECCS/AISC workshop : Connections in Steel Structures V 15

CONCLUSIONS FROM SESSIONS


Session 5: Connections in thin walled sections,
Eccentrically loaded bolt groups
- 2 Papers on connections in rack structures
Aguirre : Experimental results for beam-to-column connections under static
and cyclic loading. Rack structures should be considered as
Partially Restrained Construction. The connections should not be
used for seismic loads.
Kozlowski : Development of the component method for connections in racking.

- 2 Papers on connections in building systems with cold-formed elements


Dunai : Behaviour of connections with screws
Dubina : Behaviour of connections with bolts

- 2 Papers on the design of excentrically loaded bolt groups


One paper showing that the revised rules in the AISC code lead to uneconomical
results. The other paper showing that the model in the AISC code is not correct in
all cases because it does not satisfy equilibrium.

June 2004 ECCS/AISC workshop : Connections in Steel Structures V 16

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 477


CONCLUSIONS FROM SESSIONS
Session 6: Bolted connections in shear

Despite the fact that much research has been carried out on bolted connections
in shear, the papers addressed several new topics.
- The influence of connection behaviour is a major topic of research. It was
shown that connection deformation may induce serviceability problems.
- The negative influence of hot dip galvanising and punching of holes on fatigue
behaviour can be overcome by applying friction grip bolts.
- A new type of bolt with a rounded head was investigated for use in repair and
renovation of historical buildings.
- A proposal was presented for a single harmonised equation for block shear in
coped beams.
- Tests on single plate connections were reported to investigate the influence of
moments in combination with shear.
- The influence of stiffness and ductility aspects for the distribution of forces in
bolted shear connections was demonstrated once again

June 2004 ECCS/AISC workshop : Connections in Steel Structures V 17

CONCLUSIONS FROM SESSIONS


Session 7: Welded connections, fire
- 1 Paper on bolted flange tip connection (Snijder)
Experimental and theoretical research. Development of FEM model.
Analytical models to be improved in next step of the project.

- 1 Paper on partial joint penetration welding for connections under seismic loads
Experimental research project.
The possibility of brittle fracture was investigated and it was concluded that brittle
fracture is unlikely. Ultimate strength can be predicted by simple plastic analysis.

- 1 Paper on shear lag effects in welded tension connections


Test program to determine influence of parameters not examined in previous
research such as weld strength, weld size, member size, member length and
symmetry.

- 1 Paper on behaviour of joints in fire


Test program on 8 storey steel-concrete frame building under natural fire in the
Cardington test facility.

June 2004 ECCS/AISC workshop : Connections in Steel Structures V 18

478 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


CONCLUSIONS FROM SESSIONS
Session 8: Tubular connections

The session dealt with :


- semi-rigid and pin ended beam to column connections
- reinforced joints of circular hollow sections
- the effect of chord loading on the joint strength
- gusset plate connections
- review of tubular joint criteria

The papers provide not only information on topics not yet included in the
recommendations and codes but also cover topics which are still in
discussion for the revision of the IIW/Cidect recommendations.

June 2004 ECCS/AISC workshop : Connections in Steel Structures V 19

STRUCTURAL CONNECTIONS
3
25 conception
conception
5
knowledge design calculation
design calculation 1
2 4
codes code check
code check education
1
1 fabrication
fabrication innovation
tools
products
erection
erection
1
maintenance
maintenance

June 2004 ECCS/AISC workshop : Connections in Steel Structures V 20

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 479


DISTRIBUTION OF PAPERS REFLECTS
IMPORTANCE OF TOPICS ?
3
25 conception
conception
5
knowledge design calculation
design calculation 1
2 4
codes code check
code check education
1
1 fabrication
fabrication innovation
tools
products
erection
erection
1
maintenance
maintenance
June 2004 ECCS/AISC workshop : Connections in Steel Structures V 21

ECCS = FABRICATORS ORGANISATION !


conception

knowledge design calculation

codes code check education

1 1
fabrication
fabrication innovation
tools
products
erection
erection
1
maintenance
maintenance

June 2004 ECCS/AISC workshop : Connections in Steel Structures V 22

480 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


CONNECTIONS IN STEEL STRUCTURES V
Innovative Steel Connections

conception

knowledge design calculation

codes code check education

fabrication
fabrication innovation
tools
products
erection
erection 0

maintenance
maintenance
June 2004 ECCS/AISC workshop : Connections in Steel Structures V 23

CONNECTIONS IN STEEL STRUCTURES V

A good connection
is better than a joint !

Thanks for your contribution

June 2004 ECCS/AISC workshop : Connections in Steel Structures V 24

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 481


482 Connections in Steel Structures V - Amsterdam - June 3-4, 2004
PARTICIPANTS ECCS - AISC WORKSHOP

Aasen, B. Norconsult AS Norway


Aguirre, C.A. Universidad Técnica Federico Santa Maria Chile
Baniotopoulos, C.C. Institute of Steel Structures Greece
Beg, D. University of Ljubljana Slovenia
Belder, E. Iv-Bouw & Industrie The Netherlands
Beuzekom, S. van Bouwen met Staal The Netherlands
Bijlaard, F.S.K. Delft University of Technology The Netherlands
Birkemoe, P.C. University of Toronto Canada
Bjorhovde, R. The Bjorhovde Group USA
Brettle, M.E. BRE United Kingdom
Bucak, Ö. Munich University of Applied Sciences Germany
Charney, F.A. Virginia Tech USA
Choo, Y.S. National University of Singapore Singapore
Dexter, R.J. University of Minnesota USA
Driver, R.G. University of Alberta Canada
Dubina, D. The Politehnica University of Timisoara Romania
Dunai, L. Budapest University of Technology and Economics Hungary
Duncan, C.J. American Institute of Steel Construction USA
Evers, H.G.A. ECCS bv The Netherlands
Gendebien, G. ECCS Belgium
Geschwindner, L.F. American Institute of Steel Construction USA
Girão Coelho, A.M. Polytechnic Institute of Coimbra Portugal
Gresnigt, A.M. Delft University of Technology The Netherlands
Huhn, H. Technical University Hamburg - Harburg Germany
Kishi, N. Muroran Institute of Technology Japan
Komuro, M. Muroran Institute of Technology Japan
Kouhi, J.E. VTT Building Technology AB Finland
Kozłowski, A. Rzeszów University of Technology Poland
Kurobane, Y. Sojo University Japan
Lanser, A. Iv-Bouw & Industrie The Netherlands
Leon, R.T. Georgia Tech USA
Lutke Schipholt, R. Bouwen met Staal The Netherlands
Makino, Y. Kumamoto University Japan
Mandara, A. Second University of Naples Italy

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 483


Marshall, P.W. Moonshine Hill Proprietary USA
Moore, D.B. BRE United Kingdom
Naessens, J. Arcelor Building & Construction Support Luxemburg
Nussbaumer, A.C. ICOM/EPFL Switzerland
Orton, A.H. Corus Tubes United Kingdom
Packer, J.A. University of Toronto Canada
Pietrapertosa, C. University of Liège Belgium
Piraprez, E.L.G. Steel Solutions Belgium
Puthli, R.S. University of Karlsruhe Germany
Rassati, G.A. The University of Cincinnati USA
Rathbone, A.J. CSC (UK) Ltd. United Kingdom
Ricles, J.M. Lehigh University USA
Ryan, M.O. CTICM France
Simões da Silva, L. University of Coimbra Portugal
Snijder, H.H. Eindhoven University of Technology The Netherlands
Sosic, I. IGH d.d. Croatia
Stark, J.W.B. Delft University of Technology The Netherlands
Stratan, A. The Politehnica University of Timisoara Romania
Sumner, E.A. North Carolina State University USA
Szlendak, J.K. Technical University of Bialystok Poland
Thornton, W.A. Cives Steel Company USA
Valtinat, G. Technical University Hamburg - Harburg Germany
Vegte, G.J. van der Delft University of Technology The Netherlands
Veljkovic, M.V. Luleå University of Technology Sweden
Wald, F. Czech Technical University Czech Republic
Wardenier, J. Delft University of Technology The Netherlands
Willibald, S. University of Toronto Canada

484 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


AUTHOR INDEX

Aguirre C.A. 233 Jaspart J.P. 335


Aste C. 45
Azuma K. 367 Kishi N. 99, 111
Komuro M. 99, 111
Baniotopoulos C.C. 287 Kovács N. 191
Beg D. 131 Kozłowski A. 253
Bijlaard F.S.K. 53 Kulak G.L. 323
Birkemoe P.C. 381 Kurobane Y. 367
Bjorhovde R. 11
Borges L. 155 Lee D. 177
Lehman D.E. 167
Calado L. 191 Leon R.T. 201
Charney F.A. 121 Liang J.X. 423
Chen W.F. 99, 111 Lu L.W. 211
Choo Y.S. 423
Ciutina A. 141, 263 Maatje F. 27
Clemente I. 77 Makino Y. 89, 367, 433
Mandara A. 65
De Matteis G. 65 Marshall P.W. 457
Della Corte G. 65 Mazzolani F.M. 65
Dexter R.J. 177 Moore D. 37, 393
Downs W.M. 121 Muir L.S. 273, 281
Driver R.G. 323
Dubina D. 141, 223, 263 Nagy Z. 263
Dunai L. 191, 243 Noé S. 77
Duncan C.J. 1
Packer J.A. 445
Eliášová M. 37 Pietrapertosa C. 335
Evers H.G.A. 27 Piraprez E. 335
Puthli R.S. 445
Fisher J.W. 211
Fóti P. 243 Rassati G.A. 77
Fulop L. 263 Ricles J.M. 211
Roeder C.W. 167
Gervásio H. 155
Geschwindner L.F. 21 Santiago A. 393
Girão Coelho A.M. 53 Saucedo G.M. 445
Glatzl A. 45 Schrauben C. 201
Grecea D. 141 Simões da Silva L. 53, 155, 393
Grondin G.Y. 323 Ślęczka L. 253
Gupta L.M. 347 Snijder H.H. 355
Gupta M. 347 Stark J.W.B. 469
Stratan A. 141, 223, 263
Hajjar J.F. 177 Szlendak J.K. 403, 411
Hoenderkamp J.C.D. 355
Hu J.W. 201 Thornton W.A. 273, 281
Huber G. 45
Huhn H. 297 Valtinat G. 297, 311
Humphries M.J.R. 381 Vegte G.J. van der 89, 423, 433
Veljkovic M. 37

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 485


Wald F. 37, 393
Wardenier J. 433
Willibald S. 445

Yoo J.H. 167

Zhang X. 211
Zupančič E. 131
Zygomalas M.D. 287

486 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


KEYWORD INDEX

AISC 273, 281, 457 ductile crack 367


API 433, 457 ductility 53, 155
AWS 457 ductility requirements 335
aluminium connections 37, 65
analytical method 131 EC3 procedure 191
analytical modelling 53 Eurocode 3 37
Eurocode 9 65
beam-to-column connections 89, 155, 367 Eurocodes 155
bearing 1 earthquake resistance 223
block shear 21, 323, 347, 381 eccentric bolts 273, 281
bolt bearing 273 eccentrically braced frames 223
bolt tear-out 273 eccentricities 27
bolted connections 21, 37, 65, 263, 297, education in connection design 37
335, 347 end-plate connection 99
bolted links 223 experimental research 355
bolts 1, 89 experiments 65, 77, 211, 243, 253, 355
boundary conditions 433 external reinforcement 423
braced frames 167
brittle fracture 367 failure modes 167, 191
buckling 167 fatigue 297
finite element analysis 89, 111, 121, 355,
Charpy 177 423, 433
CIDECT 457 fire design 37
cantilevering steel-composite platforms 45 flange tip 355
channels 381 flange tip connection 355
chord pre-load 433 four-parameter power model 99
circular hollow section 423 fracture 21, 177
cold-formed steel sections 243, 263 frame 243
collar plate 423 frequently asked questions 37
column stiffener 177 friction 89
component method 131, 155, 253 friction coefficients 297
composite floor slab 211 full scale tests 201, 393, 403, 411
concrete tower 45 full strength bolted connections 287
connection capacity 1
connection database 99 groove weld 367
connection design 21, 27 gross section 323
connection optimization 273 gross shear plane 347
constructable 27 gusset plates 323, 445
cope 323
cost-effective 27 half round head HV-bolts 311
curve-fitting models 201 high temperature 393
cyclic behaviour 65 high-strength steel grade 53
cyclic loading 77, 141, 201 historic steel constructions 311
cyclic tests 191, 263 hollow section trusses 45
hollow structural sections 445
deformation capacity 53 hot dip galvanized 297
design cases 37
design stages 27 IIW 457
design standards 11 ISO 433, 457
docking connection devices 45 implicit and explicit analyses 89
double angles 347

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 487


instantaneous center of rotation method reinforcing fillet weld 367
273, 281 reliability 323
internet/cd lessons 37 repair of riveted connections 311
resistance 53
joint characteristics 141 rotation capacity 131, 141, 367
round tubes 445
K-joints 433 rupture 323
Krawinkler model 121
Scissors model 121
lattice towers 297 S-N curve 297
longitudinal gusset plate to RHS chord seismic 177
joints 403 seismic design 37, 167
low-cycle fatigue 177 seismic design code 141
seismic load 367
Monte Carlo method 155 self-drilling screw 243
MR beam-to-column joints 141 semi-rigid connections 99, 111, 233, 243,
material behaviour 11 253
material saving 27 shear lag 21, 381, 445
mixed building technology 45 shear loading 335
moment connections 131, 177 single angles 347
moment resisting frames 121 slip and creep tests 297
moment-rotation curve 99, 111 slip resistant 1
monotonic loading 77, 141, 201, 263 specifications 1, 21
multi-dimensional response 11 splice bolted connections 287
staggered holes 347
natural fire 393 static behaviour 403, 411
net section 323 static strength 433, 457
net tension plane 347 statistical analysis 131
non-linear behaviour 233 steel angles 347
non-linear material models 53 steel buildings 1
non-linear programming 273 steel connections 53
non-uniform stress distribution 347 steel design 233
notch toughness 177 steel joints 37, 155, 393
numerical research 355 steel structures 131, 223, 253
steel tension connections 381
old grand halls of railway stations 311 steel-concrete construction 45
stiffeners 27
panel zone 177 storage pallet racks 253
panel zone deformations 121 strength and stiffness prediction 403, 411
parametric study 191 strength prediction 423
partial joint penetration 367 stress triaxiality 367
partially restrained beam-to-column structural engineering 155, 393
connections 77 structural integrity 393
performance factor 323 structural steel 37
performance-based design 223 structural systems 233
portal frames 263
preliminary design model 403 T-stub 53, 65, 77
preloaded fasteners 297 tear-out 381
punched holes 297 tension members 347
theoretical research 355
RHS branch beam-column joints 411 threaded rods 1
RHS concrete filled column 411 tightening behaviour of bolts 311
rack structures 233 top- and seat-angle type connection 111,
reduced beam sections 211 201
reinforced joint 423 truss 27

488 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


tubular joints 433

ultimate strength 381


ultimate strength design 367
uniaxial tension test 11

webplates 27
welded connections 37
welded joints 65
welded tubular connections 457
welding 381
welds 1, 177
wide flange column 211

yield mechanisms 167

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 489


490 Connections in Steel Structures V - Amsterdam - June 3-4, 2004

You might also like