You are on page 1of 122

AS/NZS 4600 Supp1:1998

AS/NZS 4600
Supplement 1:1998

Cold-formed steel structures—


Commentary

(Supplement 1 to AS/NZS 4600:1996)


Accessed by BHP BILLITON on 27 Jun 2002
AS/NZS 4600 Supp1:1998

This Joint Australian/New Zealand Standard was prepared by Joint Technical


Committee BD/82, Cold-formed Steel Structures. It was approved on behalf of the
Council of Standards Australia on 11 September 1998 and on behalf of the
Council of Standards New Zealand on 11 September 1998. It was published on
5 November 1998.

The following interests are represented on Committee BD/82:


Association of Consulting Engineers Australia
Australian Building Codes Board
Australian Chamber of Commerce and Industry
Australian Institute of Steel Construction
Bureau of Steel Manufacturers of Australia
CSIRO, Building, Construction and Engineering
Institution of Professional Engineers New Zealand
Metal Building Products Manufacturers Association
Metal Trades Industry Association of Australia
New Zealand Heavy Engineering Research Association
New Zealand Metal Roofing and Cladding Manufacturers Association
New Zealand Structural Engineering Society
The University of Queensland
University of Sydney
Welding Technology Institute of Australia
Accessed by BHP BILLITON on 27 Jun 2002

Review of Standards. To keep abreast of progress in industry, Joint Australian/


New Zealand Standards are subject to periodic review and are kept up to date by the
issue of amendments or new editions as necessary. It is important therefore that
Standards users ensure that they are in possession of the latest edition, and any
amendments thereto.
Full details of all Joint Standards and related publications will be found in the Standards
Australia and Standards New Zealand Catalogue of Publications; this information is
supplemented each month by the magazines ‘The Australian Standard’ and ‘Standards
New Zealand’, which subscribing members receive, and which give details of new
publications, new editions and amendments, and of withdrawn Standards.
Suggestions for improvements to Joint Standards, addressed to the head office of either
Standards Australia or Standards New Zealand, are welcomed. Notification of any
inaccuracy or ambiguity found in a Joint Australian/New Zealand Standard should be
made without delay in order that the matter may be investigated and appropriate action
taken.
AS/NZS 4600 Supp1:1998

AS/NZS 4600
Supplement 1:1998

Cold-formed steel structures—


Commentary

(Supplement 1 to AS/NZS 4600:1996)

First published as AS/NZS 4600 Supplement 1:1998.


Accessed by BHP BILLITON on 27 Jun 2002

Published jointly by:

Standards Australia
1 The Crescent,
Homebush NSW 2140 Australia

Standards New Zealand


Level 10, Radio New Zealand House,
155 The Terrace,
Wellington 6001 New Zealand
ISBN 0 7337 2237 7
AS/NZS 4600 Supp1:1998 2

PREFACE

This Commentary was prepared by the Joint Standards Australia / Standards New Zealand
Committee BD/82, Cold-formed Steel Structures.
The objective of this Commentary is to provide users with background information and
guidance to AS/NZS 4600:1996.
The Standard and Commentary are intended for use by design professionals with
demonstrated engineering competence in their fields.
In Australia, the Australian Standard for the design of cold-formed steel structural
members was first issued in permissible stress format as AS 1538—1974 (Standards
Australia, 1974), based mainly on the 1968 edition of the AISI Specification (AISI, 1968),
but with modifications to the beam and column design curves to keep them aligned with
the Australian Steel Structures Code at that time, ASCA1 (Standards Australia, 1968). It
was revised and published in 1988 as AS 1538—1988 (Standards Australia, 1988) and
was based mainly on the 1980 edition of the AISI Specification (AISI, 1983), but
included some material from the 1986 edition of the AISI Specification (AISI, 1986). In
1996, a limit states version of the cold-formed steel structures Standard was produced by
Standards Australia and Standards New Zealand. It was based mainly on the 1996 edition
of the AISI Specification (AISI, 1996) but with amendments where necessary to reflect
Australian/New Zealand practice.
In this Commentary, AS/NZS 4600:1996 is referred as ‘the Standard’.
To be consistent with the AISI Commentary, referenced documents provided in
Appendix A are listed in alphabetical order.
The clause numbers and titles used in this Commentary are the same as those in
AS/NZS 4600, except that they are prefixed by the letter ‘C’. To avoid possible confusion
between the Commentary and the Standard, a Commentary clause is referred to as
‘Clause C....’ in accordance with Standards Australia/Standards New Zealand policy.

© Copyright STANDARDS AUSTRALIA / STANDARDS NEW ZEALAND


Accessed by BHP BILLITON on 27 Jun 2002

Users of Standards are reminded that copyright subsists in all Standards Australia and Standards New Zealand publications and software.
Except where the Copyright Act allows and except where provided for below no publications or software produced by
Standards Australia or Standards New Zealand may be reproduced, stored in a retrieval system in any form or transmitted by any means
without prior permission in writing from Standards Australia or Standards New Zealand. Permission may be conditional on an
appropriate royalty payment. Australian requests for permission and information on commercial software royalties should be directed to
the head office of Standards Australia. New Zealand requests should be directed to Standards New Zealand.
Up to 10 percent of the technical content pages of a Standard may be copied for use exclusively in-house by purchasers of the
Standard without payment of a royalty or advice to Standards Australia or Standards New Zealand.
Inclusion of copyright material in computer software programs is also permitted without royalty payment provided such programs
are used exclusively in-house by the creators of the programs.
Care should be taken to ensure that material used is from the current edition of the Standard and that it is updated whenever the Standard
is amended or revised. The number and date of the Standard should therefore be clearly identified.
The use of material in print form or in computer software programs to be used commercially, with or without payment, or in commercial
contracts is subject to the payment of a royalty. This policy may be varied by Standards Australia or Standards New Zealand at any time.
3 AS/NZS 4600 Supp1:1998

CONTENTS
Page

INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

SECTION C1 SCOPE AND GENERAL


C1.1 SCOPE . . . . . . . . . . . . . . . . . . . . . .............. . . . . . . . . . . . . . . 7
C1.2 REFERENCED DOCUMENTS . . . . .............. . . . . . . . . . . . . . . 7
C1.3 DEFINITIONS . . . . . . . . . . . . . . . . .............. . . . . . . . . . . . . . . 7
C1.4 NOTATION . . . . . . . . . . . . . . . . . .............. . . . . . . . . . . . . . . 12
C1.5 MATERIALS . . . . . . . . . . . . . . . . . .............. . . . . . . . . . . . . . . 12
C1.6 DESIGN REQUIREMENTS . . . . . . .............. . . . . . . . . . . . . . . 21
C1.7 NON-CONFORMING SHAPES AND CONSTRUCTION . . . . . . . . . . . . . . 29

SECTION C2 ELEMENTS
C2.1 SECTION PROPERTIES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
C2.2 EFFECTIVE WIDTHS OF STIFFENED ELEMENTS . . . . . . . . . . . . . . . . 33
C2.3 EFFECTIVE WIDTHS OF UNSTIFFENED ELEMENTS . . . . . . . . . . . . . . 40
C2.4 EFFECTIVE WIDTHS OF UNIFORMLY COMPRESSED ELEMENTS
WITH AN EDGE STIFFENER OR ONE INTERMEDIATE STIFFENER . . . 43
C2.5 EFFECTIVE WIDTHS OF EDGE-STIFFENED ELEMENTS WITH ONE
OR MORE INTERMEDIATE STIFFENERS OR STIFFENED ELEMENTS
WITH MORE THAN ONE INTERMEDIATE STIFFENER . . . . . . . . . . . . 44
C2.6 ARCHED COMPRESSION ELEMENTS . . . . . . . . . . . . . . . . . . . . . . . . . 46
C2.7 STIFFENERS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

SECTION C3 MEMBERS
C3.1 GENERAL . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
C3.2 MEMBERS SUBJECT TO TENSION . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
C3.3 MEMBERS SUBJECT TO BENDING . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
C3.4 CONCENTRICALLY LOADED COMPRESSION MEMBERS . . . . . . . . . . 65
C3.5 COMBINED AXIAL LOAD AND BENDING . . . . . . . . . . . . . . . . . . . . . . 77
C3.6 CYLINDRICAL TUBULAR MEMBERS . . . . . . . . . . . . . . . . . . . . . . . . . 80

SECTION C4 STRUCTURAL ASSEMBLIES


C4.1 BUILT-UP SECTIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
C4.2 MIXED SYSTEMS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
C4.3 LATERAL RESTRAINTS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
Accessed by BHP BILLITON on 27 Jun 2002

C4.4 WALL STUDS AND WALL STUD ASSEMBLIES . . . . . . . . . . . . . . . . . . 91

SECTION C5 CONNECTIONS
C5.1 GENERAL . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
C5.2 WELDED CONNECTIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
C5.3 BOLTED CONNECTIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
C5.4 SCREWED CONNECTIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
C5.5 BLIND RIVETED CONNECTIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
C5.6 RUPTURE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
AS/NZS 4600 Supp1:1998 4

Page

SECTION C6 TESTING
C6.1 TESTING FOR DETERMINING MATERIAL PROPERTIES . . . . . . . . . . 106
C6.2 TESTING FOR ASSESSMENT OR VERIFICATION . . . . . . . . . . . . . . . 106

APPENDIX A REFERENCED DOCUMENTS . . . . . . . . . . . . . . . . . . . . . . . . . . . 109


Accessed by BHP BILLITON on 27 Jun 2002
5 AS/NZS 4600 Supp1:1998

INTRODUCTION

Cold-formed steel members have been used economically for building construction and
other applications (Winter 1959a, 1959b; Yu 1991; Hancock 1998). These types of
sections are cold-formed from steel sheet, strip, plate or flat bar in roll-forming machines
or by press brake or bending operations. The thicknesses of steel sheets or strip generally
used for cold-formed steel structural members range from 0.4 mm to about 6.4 mm. Steel
plates and bars as thick as 25 mm can be cold-formed successfully into structural shapes
and their design is covered by the Standard.
In general, cold-formed steel structural members can offer the following advantages for
building construction (Winter, 1970; Yu, 1991):
(a) Light members can be manufactured for relatively light loads or short spans, or
both.
(b) Unusual sectional configurations can be produced economically by cold-forming
operations and consequently favourable strength-to-weight ratios can be obtained.
(c) Load-carrying panels and decks can provide useful surfaces for floor, roof and wall
construction, and in some cases they can also provide enclosed cells for electrical
and other conduits.
(d) Panels and decks not only withstand loads normal to their surfaces, but they can
also act as shear diaphragms to resist forces in their own planes if they are
adequately interconnected to each other and to supporting members.
The use of cold-formed steel members in building construction began in about the 1850s.
However, in the United States such steel members were not widely used in buildings until
the publication of the first edition of the American Iron and Steel Institute Specification
in 1946 (AISI, 1946). This first design Standard was primarily based on the research work
sponsored by AISI at Cornell University since 1939. It was revised subsequently by the
AISI Committee in 1956, 1960, 1962, 1968, 1980 and 1986 to reflect the technical
developments and the results of continuing research. In 1991, AISI published the first
edition of the Load and Resistance Factor Design Specification for Cold-formed Steel
Structural Members (AISI, 1991). Both allowable stress design (ASD) and load and
resistance factor design (LRFD) specifications were combined into a single document in
1996.
In Australia, the Australian Standard for the design of cold-formed steel structural
members was first issued in permissible stress format as AS 1538—1974 (Standards
Australia, 1974) based mainly on the 1968 edition of the AISI Specification (AISI, 1968)
but with modifications to the beam and column design curves to keep them aligned with
the Australian Steel Structures Code at that time, ASCA1 (Standards Australia, 1968). It
was revised and published in 1988 as AS 1538—1988 (Standards Australia, 1988) and
was based mainly on the 1980 edition of the AISI Specification (AISI, 1983) but included
Accessed by BHP BILLITON on 27 Jun 2002

some material from the 1986 edition of the AISI Specification (AISI, 1986). In 1996, a
joint Australian/New Zealand limit states version of the cold-formed steel structures
Standard was produced by Standards Australia and Standards New Zealand. It was based
mainly on the 1996 edition of the AISI Specification (AISI, 1996) but with amendments,
where necessary, to reflect Australian/New Zealand practice.
During the period from 1958 through to 1983, AISI published Commentaries on several
editions of the AISI design specification, which were prepared by Professor George
Winter of Cornell University in 1958, 1961, 1962, and 1970. In the 1983, 1986 and 1996
editions, the format used for the AISI Commentary has been changed in that the same
section numbers are used in the AISI Commentary as in the AISI Specification. As for
AS/NZS 4600 Supp1:1998 6

previous editions of the AISI Commentary, this document contains a brief presentation of
the characteristics and the performance of cold-formed steel members. In addition, it
provides a record of the reasoning behind and the justification for various provisions of
the Standard. A cross-reference is provided between various design provisions and the
published research data. This Commentary to AS/NZS 4600:1996 is based mainly on the
1996 edition of the Commentary to the AISI Specification, and has been amended, where
necessary, to reflect the differences between the AISI Specification and AS/NZS 4600.
Accessed by BHP BILLITON on 27 Jun 2002
7 AS/NZS 4600 Supp1:1998

STANDARDS AUSTRALIA / STANDARDS NEW ZEALAND

Australian / New Zealand Standard


Cold-formed steel structures—Commentary
(Supplement 1 to AS/NZS 4600:1996)

S E C T I O N C 1 S C O P E A N D G E N E R A L

C1.1 SCOPE The cross-sectional configurations, manufacturing processes and


fabrication practices of cold-formed steel structural members differ in several respects
from those of hot-rolled steel shapes. For cold-formed steel sections, the forming process
is performed at, or near, room temperature by the use of bending brakes, press brakes or
roll-forming machines. Some of the significant differences between cold-formed sections
and hot-rolled shapes are—
(a) absence of the residual stresses caused by uneven cooling due to hot-rolling;
(b) lack of corner fillets;
(c) presence of increased yield strength with decreased proportional limit and ductility
resulting from cold forming;
(d) presence of cold-reducing stresses when cold-rolled steel stock has not been fully
annealed;
(e) prevalence of elements having large width-to-thickness ratios and, hence, subject to
local buckling in compression;
(f) rounded corners; and
(g) stress-strain curves can be either sharp-yielding type or gradual-yielding type.
AS/NZS 4600 (Standards Australia/Standards New Zealand, 1996) is limited to the design
of steel structural members cold-formed from carbon or low-alloy sheet, strip, plate or
bar. The design is to be in accordance with the limit states design method.
The Standard is applicable only to cold-formed sections not more than 25 mm in
thickness. Research conducted at the University of Missouri-Rolla (Yu, Liu, and
McKinney, 1973b and 1974) has verified the applicability of the Standard’s provisions for
such cases.
In view of the fact that most of the design provisions have been developed on the
experimental work subject to static loading, the Standard is intended for the design of
cold-formed steel structural members to be used for load-carrying purposes in buildings.
For structures other than buildings, appropriate allowances should be made for dynamic
effects. The Standard does not apply to the design of structures subject to fire or fatigue
Accessed by BHP BILLITON on 27 Jun 2002

since insufficient data was available on these phenomena for cold-formed members during
its preparation.

C1.2 REFERENCED DOCUMENTS The Standards listed in Appendix A are subject


to revision from time to time and the current issue should always be used. The currency
of any Standard may be checked with Standards Australia or Standards New Zealand.

C1.3 DEFINITIONS Many of the definitions in Clause 1.3 are self-explanatory. In


New Zealand, terms used in limit state design may differ from those used in Australia.
These are included in brackets. Those definitions that are not self-explanatory or that are
not defined in Clause 1.3 are briefly discussed in this Clause.

COPYRIGHT
AS/NZS 4600 Supp1:1998 8

C1.3.13 Distortional buckling — the mode of distortional buckling is included for the
first time in AS/NZS 4600. Figure C1.3(4)(a) shows distortional buckling for a
compression member, and Figure C1.3(b) for a flexural member. Appendix D gives
equations to compute elastic distortional buckling stresses.
C1.3.14 Effective design width — the effective design width is a concept that takes
account of local buckling and post-buckling strength for compression elements. The effect
of shear lag on short, wide flanges is also handled by using an effective design width.
These matters are treated in Section 2, and the corresponding effective widths are
discussed in the Commentary on that Section.
C1.3.17 Flexural-torsional buckling — the 1968 edition of the AISI Specification and
AS 1538 — 1974 pioneered methods for calculating column loads of cold-formed steel
sections prone to buckle by simultaneous twisting and bending. This complex behaviour
may result in lower column loads than would result from primary buckling by flexure
alone (Trahair, 1993).
C1.3.19 Limit state — limit states design (LSD) is a method of designing structural
components such that the applicable limit state is not exceeded when the structure is
subjected to all appropriate load combinations as specified in Clause 1.6.1.
C1.3.23 Multiple-stiffened elements — multiple-stiffened elements of two sections are
shown in Figure C1.3(1). Each of the two outer sub-elements shown in Figure C1.3(1)(a)
are stiffened by a web and an intermediate stiffener while the middle sub-element is
stiffened by two intermediate stiffeners. The two sub-elements shown in Figure C1.3(1)(b)
are stiffened by a web and the attached intermediate middle stiffener.
Accessed by BHP BILLITON on 27 Jun 2002

FIGURE C1.3(1) MULTIPLE-STIFFENED COMPRESSION ELEMENTS

COPYRIGHT
9 AS/NZS 4600 Supp1:1998

C1.3.38 Stiffened or partially stiffened compression elements — stiffened compression


elements of various sections are shown in Figure C1.3(2), in which sections shown in
Figures C1.3(2)(a) to C1.3(2)(e) are for flexural members, and sections shown in
Figures C1.3(2)(f) to C1.3(2)(i) are for compression members. Sections shown in
Figures C1.3(2)(a) and C1.3(2)(b) each have a web and a lip to stiffen the compression
element, i.e. the compression flange, the ineffective portion of which is shown shaded.
For the explanation of these ineffective portions, see Clause C1.3.1 and Section C2.
Figures C1.3(2)(c) to C1.3(2)(e) show compression elements stiffened by two webs.
Figures C1.3(2)(f) and C1.3(2)(h) show edge-stiffened flange elements that have a vertical
element (web) and an edge stiffener (lip) to stiffen the elements while the web itself is
stiffened by the flanges. Figure C1.3(2)(g) shows four compression elements stiffening
each other, and Figure C1.3(2)(i) has each stiffened element stiffened by a lip and the
other stiffened element.
Accessed by BHP BILLITON on 27 Jun 2002

COPYRIGHT
AS/NZS 4600 Supp1:1998 10
Accessed by BHP BILLITON on 27 Jun 2002

FIGURE C1.3(2) STIFFENED COMPRESSION ELEMENTS

COPYRIGHT
11 AS/NZS 4600 Supp1:1998

C1.3.46 Thickness — in calculating section properties, the reduction in thickness that


occurs at corner bends is ignored unless the manufacturing process warrants consideration
of a more accurate method (see Clause 2.1.2.1), and the base steel thickness of the flat
steel stock, exclusive of coatings, is used in all calculations for load-carrying purposes.
C1.3.49 Unstiffened compression element — unstiffened elements of various sections
are shown in Figure C1.3(3), in which sections shown in Figures C1.3(3)(a) to C1.3(3)(d)
are for flexural members and sections shown in Figures C1.3(3)(e) to C1.3(3)(h) are for
compression members. Sections shown in Figures C1.3(3)(a) to C1.3(3)(c) have only a
web to stiffen the compression flange element. The legs of the section shown in
Figure C1.3(3)(d) provide mutual stiffening action to each other along their common
edges. Sections shown in Figures C1.3(3)(e) to C1.3(3)(g), acting as columns, have
vertical stiffened elements (webs) which provide support for one edge of the unstiffened
flange elements. The legs of the section shown in Figure C1.3(3)(h) provide mutual
stiffening action.
Accessed by BHP BILLITON on 27 Jun 2002

FIGURE C1.3(3) UNSTIFFENED COMPRESSION ELEMENTS

COPYRIGHT
AS/NZS 4600 Supp1:1998 12

FIGURE C1.3(4) DISTORTIONAL BUCKLING MODES

C1.4 NOTATION The basis of the notation is in accordance with ISO 3898, as much
as possible.

C1.5 MATERIALS
C1.5.1 Structural steel
C1.5.1.1 Applicable steels The Australian and New Zealand Standards are the basic
source of steel designations for use with the Standard. Clause 1.5.1.1 provides a list of
Australian, New Zealand and Australian/New Zealand Standards for steels that are
acceptable by the Standard.
The important material properties for the design of cold-formed steel members are yield
stress, tensile strength and ductility. Ductility is the ability of a steel to undergo sizeable
plastic or permanent strains before fracturing and is important both for structural safety
and for cold forming. It is usually measured by the elongation in a 50 mm gauge length.
The ratio of the tensile strength to the yield stress is also an important material property.
This is an indication of strain-hardening and the ability of the material to redistribute
stress.
For the listed Australian and New Zealand Standards, the yield stresses of steels range
from 200 to 550 MPa and the tensile strengths vary from 300 to 550 MPa. The
elongations are no less than 8%. Exceptions are AS 1397 — G550 steel with a specified
minimum yield stress of 550 MPa, a specified minimum tensile strength of 550 MPa, and
with a minimum elongation of 2% in a 50 mm gauge length. This low ductility steel has
limits on its applicability for structural framing members and so its application without
restriction is limited to steels not less than 0.9 mm thick. Nevertheless, they are being
Accessed by BHP BILLITON on 27 Jun 2002

used successfully for specific applications, such as decks, panels and in steel-framed
housing, as structural members. The conditions for the use of AS 1397—G550 steel less
than 0.9 mm thick are outlined in Clause 1.5.1.5.
C1.5.1.2 Other steels Although the use of Australian and Australian/New Zealand
Standards listed in Clause 1.5.1.1 is encouraged, other steels that satisfy the ductility
requirements of Clause 1.5.1.5 may also be used in cold-formed steel structures.
C1.5.1.3 Strength increase resulting from cold forming The mechanical properties of
the flat steel sheet, strip, plate or bar, such as yield stress, tensile strength, and elongation
may be substantially different from the properties exhibited by the cold-formed steel
sections. Figure C1.5.1.3(1) shows the increase of yield strength and tensile strength from

COPYRIGHT
13 AS/NZS 4600 Supp1:1998

those of the virgin material at the section locations in a cold-formed steel channel section
and a joist chord (Karren and Winter 1967). This difference can be attributed to cold
working of the material during the cold-forming process.
The influence of cold-work on mechanical properties was investigated by Chajes, Britvec,
Winter, Karren, and Uribe at Cornell University in the 1960s (Chajes, Britvec, and
Winter, 1963; Karren, 1967; Karren and Winter, 1967; Winter and Uribe, 1968). It was
found that the changes of mechanical properties due to cold-stretching are caused mainly
by strain-hardening and strain-ageing, as shown in Figure C1.5.1.3(2) (Chajes, Britvec,
and Winter 1963). In this Figure, Curve A represents the stress-strain curve of the virgin
material. Curve B is due to unloading in the strain-hardening range, Curve C represents
immediate reloading, and Curve D is the stress-strain curve of reloading after
strain-ageing. It is interesting to note that the yield points of both Curves C and D are
higher than the yield point of the virgin material and that the ductilities decrease after
strain-hardening and strain-ageing.
Accessed by BHP BILLITON on 27 Jun 2002

COPYRIGHT
AS/NZS 4600 Supp1:1998 14
Accessed by BHP BILLITON on 27 Jun 2002

FIGURE C1.5.1.3(1) (in part) EFFECT OF COLD-WORK ON MECHANICAL


PROPERTIES IN COLD-FORMED STEEL SECTIONS

COPYRIGHT
15 AS/NZS 4600 Supp1:1998
Accessed by BHP BILLITON on 27 Jun 2002

FIGURE C1.5.1.3(1) (in part) EFFECT OF COLD-WORK ON MECHANICAL


PROPERTIES IN COLD-FORMED STEEL SECTIONS

COPYRIGHT
AS/NZS 4600 Supp1:1998 16

FIGURE C1.5.1.3(2) EFFECT OF ELONGATION HARDENING AND ELONGATION


AGEING ON STRESS-ELONGATION CHARACTERISTICS

Cornell research also revealed that the effects of cold-work on the mechanical properties
of corners usually depend on the following:
(a) The type of steel.
(b) The type of stress (compression or tension).
(c) The direction of stress with respect to the direction of cold-work (transverse or
longitudinal).
(d) The fu/fy ratio.
(e) The inside-radius-to-thickness ratio r i/t.
(f) The amount of cold-work.
Among Items (a) to (f), the fu/fy and ri/t ratios are the most important factors to affect the
change in mechanical properties of formed sections. Virgin material with a large f u/fy ratio
possesses a large potential for strain-hardening. Consequently, as the ratio increases, the
effect of cold-work on the increase in the yield point of steel increases. Small inside-
radius-to-thickness ratios (ri/t) correspond to a large degree of cold-work in a corner and,
therefore, for a given material, the smaller the ri/t ratio, the larger the increase in yield
point.
Investigating the influence of cold-work, Karren derived the following equations for the
ratio of corner yield strength to virgin yield strength (Karren, 1967):
Accessed by BHP BILLITON on 27 Jun 2002

(i) . . . C1.5.1.3(1)

(ii) . . . C1.5.1.3(2)

(iii) . . . C1.5.1.3(3)

COPYRIGHT
17 AS/NZS 4600 Supp1:1998

where
fyc = tensile yield stress of bends
fyv = tensile yield stress of unformed section
Bc = constant
m = constant
fuv = tensile strength of unformed section
ri = inside bend radius
t = sheet thickness
With regard to the full section properties, the tensile yield stress of the full section may
be approximated by using a weighted average as follows:
. . . C1.5.1.3(4)
where
fya = average design yield stress of the steel in the full section of compression
members
C = ratio of bend area to total cross-sectional area. For flexural members having
unequal flanges, the one giving a smaller C value is considered to be the
controlling flange
fyc = average tensile yield stress of bends
= . . . C1.5.1.3(5)

fyf = average tensile yield stress of flats


Good agreement between the calculated and the tested stress-strain characteristics for a
channel section and a joist chord section were demonstrated by Karren and Winter
(Karren and Winter, 1967).
In the last two decades, additional studies were made by numerous investigators. These
investigations dealt with the cold-formed sections having large r i/t ratios and with thick
materials. They also considered residual stress distribution, simplification of design
methods, and other related subjects. For details, see Yu (1991).
In 1962, the AISI Specification permitted the utilization of cold-work of forming on the
basis of full section tests. Since 1968, the AISI Specification has allowed the use of the
increased average yield point of the section (fya) to be determined by—
(A) full section tensile tests;
(B) stub column tests; or
(C) calculated in accordance with Equation C1.5.1.3(4).
Accessed by BHP BILLITON on 27 Jun 2002

However, such a strength increase is limited only to relatively compact sections designed
in accordance with Clause 3.3 (bending strength excluding the use of inelastic reserve
capacity), Clauses 3.4, 3.5, 3.6 and 4.4.
In some cases, when evaluating the effective area of the web, the effective width factor
(ρ), in accordance with Clause 2.2, may be less than unity but the sum of be1 and b e2 of
Figure 2.2.3 may be such that the web is fully effective, and the cold-work of forming
may be used.
C1.5.1.4 Effect of welding Welding may affect the mechanical properties of a member,
particularly in the heat-affected zone (HAZ). The designer should make allowances for
this on the basis of testing.

COPYRIGHT
AS/NZS 4600 Supp1:1998 18

C1.5.1.5 Ductility The nature and importance of ductility and the ways in which this
property is measured are briefly discussed in Clause C1.5.1.1.
Low-carbon sheet and strip steels with specified minimum yield points from 250 MPa to
500 MPa need to meet Australian and New Zealand Standards specified minimum
elongations in a 50 mm gauge length of at least 8%. However, for AS 1397—G550, for
which the specified minimum yield stress is 550 MPa, the elongation requirement is 2%
on a 50 mm gauge length for steels greater than or equal to 0.60 mm thick, and no
elongation is specified for thinner steels. G550 steel less than 0.9 mm thick, differs from
the array of steels listed Clause 1.5.1.1.
In 1968, because new steels of higher strengths were being developed, sometimes with
lower elongations, the question of how much elongation is really needed in a structure
was the focus of a study initiated at Cornell University. Steels that had yield strengths
ranging from 310 to 690 MPa, elongations in 50 mm ranging from 50 to 1.3%, and
tensile-to-yield strength ratios ranging from 1.51 to 1.00 were studied (Dhalla, Errera and
Winter, 1971; Dhalla and Winter, 1974a; Dhalla and Winter, 1974b). The investigators
developed elongation requirements for ductile steels. These measurements are more
accurate but cumbersome to make. Therefore, the investigators recommended the
following determination for adequately ductile steels:
(a) The tensile-to-yield strength ratio should not be less than 1.08.
(b) The total elongation in a 50 mm gauge length should not be less than 10%, or not
less than 7% in a 200 mm gauge length.
Also, the Standard limits the use of Sections 2 to 5 to adequately ductile steels. In lieu of
the tensile-to-yield strength limit of 1.08, the Standard permits the use of elongation
requirements using the measurement technique as given by Dhalla and Winter (1974a)
(Yu, 1991). Because of limited experimental verification of the structural performance of
members using materials having a tensile-to-yield strength ratio less than 1.08 (Macadam
et al., 1988), the Standard limits the use of this material to purlins and girts meeting the
elastic design requirements of Clauses 3.3.2.3, 3.3.3.2, 3.3.3.3 and 3.3.3.4. Thus, the use
of such steels in other applications (compression members, tension members, other
flexural members including those whose strength is based on inelastic reserve capacity,
and the like) is prohibited. However, in purlins and girts, concurrent design axial forces of
relatively small magnitude are acceptable providing the requirements of Clause 1.5.1.5 of
the Standard are met and N*/φRu does not exceed 0.15.
AS 1397—G550 steel less than 0.9 mm thick does not have adequate ductility as specified
in Clause 1.5.1.5(a). Its use has been limited as specified in Clause 1.5.1.5(b) to particular
configurations. The limit of the design yield stress to 75% of the specified minimum yield
stress, and the design tensile strength to 75% of the specified minimum tensile strength,
or 450 MPa, whichever is lower, introduces a higher safety factor, but still allows low
ductility steels, such as AS 1397—G550 less than 0.9 mm thick, to be used by the results
of load tests that are permitted as an alternative to making this reduction. It is also
permitted to use higher design stresses than specified in Clause 1.5.1.5(b)(i), provided it is
established that material ductility does not affect the strength, stability and serviceability
Accessed by BHP BILLITON on 27 Jun 2002

of the member or structural configuration to which the member belongs. This provision is
in AS/NZS 4600 but not in the AISI 1996 specification.
C1.5.1.6 Acceptance of steels Sheet and strip steels, both coated and uncoated, may be
ordered to nominal or minimum thickness. If the steel is ordered to minimum thickness,
all thickness tolerances are over (+) and nothing under (−). If the steel is ordered to
nominal thickness, the thickness tolerances are divided equally between over and under.
Therefore, in order to provide the similar material thickness between the two methods of
ordering sheet and strip steel, it was decided to require that the delivered thickness of a
cold-formed product be at least 95% of the design thickness. Thus, it is apparent that a
portion of the factor of safety may be considered to cover minor negative thickness
tolerances.

COPYRIGHT
19 AS/NZS 4600 Supp1:1998

Generally, thickness measurements should be made in the centre of flanges. For decking
and sheeting, measurements should be made as close as practicable to the centre of the
first full flat of the section.
The responsibility of meeting this requirement for a cold-formed product is that of the
manufacturer of the product, not the steel producer.
C1.5.1.7 Unidentified steel Clause 1.5.1.7 is carried over from AS 1538 and is not
included in the AISI specification.
C1.5.2 Design stresses The strength of cold-formed steel structural members depends
on the yield stress, except in those cases where elastic local buckling or overall buckling
is critical. Because the stress-strain curve of steel sheet or strip can be either sharp-
yielding type (see Figure C1.5.2(1)(a)) or gradual-yielding type (see Figure C1.5.2(1)(b)),
the method for determining the yield stress for sharp-yielding steel and the yield stress for
gradual-yielding steel are based on AS 1391 (Standards Australia, 1991). As shown in
Figure C1.5.2(2)(a), the yield stress for sharp-yielding steel is defined by the stress level
of the plateau. For gradual-yielding steel, the stress-strain curve is rounded out at the
‘knee’ and the proof stress is determined by either the non-proportional elongation method
(see Figure C1.5.2(2)(b)) or the total elongation method (see Figure C1.5.2(2)(c). The
term yield stress used in the Standard applies to either yield stress or yield strength.
The strength of members that are governed by buckling depends not only on the yield
stress but also on the modulus of elasticity (E) and the tangent modulus (Et). The modulus
of elasticity is defined by the slope of the initial straight portion of the stress-strain curve
(see Figure C1.5.2(1)). The measured values of E on the basis of the standard methods
usually range from 200 to 207 GPa. A value of 200 GPa is used in the Standard for
design purposes. The tangent modulus is defined by the slope of the stress-strain curve at
any stress level, as shown in Figure C1.5.2(1)(b).
For sharp-yielding steels, Et is equal to E up to the yield stress, but with gradually
yielding steels, Et equals to E only up to the proportional limit (Fpr). Once the stress
exceeds the proportional limit, the tangent modulus (Et) becomes progressively smaller
than the initial modulus of elasticity.
Various buckling provisions of the Standard have been written for gradually yielding
steels whose proportional limit is not lower than about 70% of the specified minimum
yield point.
Determination of proportional limits for information purposes can be done simply by
using the offset method shown in Figure C1.5.2(2)(b) with the distance (om) equal to
0.0001 length/length (0.01% offset) and calling the stress (R), where mn intersects the
stress-strain curve at the proportional limit.
Clause 1.5.2 stipulates the need for the yield stresses and tensile strengths not to exceed
those given in Table 1.5.
Accessed by BHP BILLITON on 27 Jun 2002

COPYRIGHT
AS/NZS 4600 Supp1:1998 20

FIGURE C1.5.2(1) STRESS-ELONGATION CURVES OF CARBON STEEL SHEET


OR STRIP
Accessed by BHP BILLITON on 27 Jun 2002

FIGURE C1.5.2(2) STRESS-ELONGATION DIAGRAMS SHOWING


METHODS OF YIELD POINT AND YIELD STRENGTH DETERMINATION

COPYRIGHT
21 AS/NZS 4600 Supp1:1998

C1.5.3 Fasteners and electrodes In Clause 1.5.3 of the Standard, the relevant
Australian and Australia/New Zealand Standards for steel bolts, nuts, washers, welding
consumables and screws are given. ANSI/AWS D1.3 may also be used to specify welding
consumables.

C1.6 DESIGN REQUIREMENTS


C1.6.1 Loads and load combinations In Australia, the load combinations are to be
calculated in accordance with AS 1170.1 (1989). Dead loads and live loads are also given
in AS 1170.1, wind loads are given in AS 1170.2, snow loads are given in AS 1170.3 and
earthquake loads are given in AS 1170.4. In New Zealand, load combinations, dead and
live loads, wind loads, snow loads and earthquake loads are to be calculated in accordance
with NZS 4203.
Recognized engineering procedures should be employed to reflect the effect of impact
loads on a structure. For building design, reference may be made to AISC publications
(AISC, 1989; AISC 1993).
When gravity and horizontal loads produce forces of opposite sign in members,
consideration should be given to the minimum gravity loads acting in combination with
wind or earthquake loads.
C1.6.2 Structural analyses and design
C1.6.2.1 General A limit state is the condition at which the structural usefulness of a
load-carrying element or member is impaired to such an extent that it becomes unsafe for
the occupants of the structure, or the element no longer performs its intended function.
Typical limit states for cold-formed steel members are excessive deflection, yielding,
buckling and attainment of maximum strength after local buckling, (i.e. post-buckling
strength). These limit states have been established through experience in practice or in the
laboratory, and they have been thoroughly investigated through analytical and
experimental research. The background for the establishment of the limit states is
extensively documented (Winter, 1970; Peköz, 1986b; and Yu, 1991), and a continuing
research effort provides further improvement in understanding them.
The following types of limit states are specified in Clause 1.6.2:
(a) The limit state of the strength required to resist the extreme loads during the
intended life of the structure.
(b) The limit state of the structure as a whole to prevent overturning, uplift or sliding.
(c) The limit state of the ability of the structure to perform its intended function during
its life.
These three limit states are usually referred to as the strength (ultimate) limit state, the
stability limit state and the serviceability limit state. AS/NZS 4600 focuses on the limit
state of strength in Clause 1.6.2.2, the limit state of stability (which in accordance with
NZS 4203 is part of the ultimate limit state) in Clause 1.6.2.3 and the limit state of
Accessed by BHP BILLITON on 27 Jun 2002

serviceability in Clause 1.6.2.4.


C1.6.2.2 Strength [ultimate] limit state For the limit state of strength [ultimate], the
general format of the limit states method is expressed as follows:
S* ≤ Rd . . . C1.6.2.2(1)
where
S* = design action effects (design actions)
Rd = design capacity

COPYRIGHT
AS/NZS 4600 Supp1:1998 22

The design action effects [design actions] (S*) may be determined by an elastic structural
analysis (normally linear elastic) or a plastic structural analysis if the plastic hinges have
adequate strength and ductility. Testing in accordance with Section 6 may be used in lieu
of analysis to establish member strength.
The design capacity (Rd) can be determined as the product of the nominal capacity
(strength reduction) (Ru) with the capacity [strength reduction] factor, i.e. R d = φRu, or by
testing in accordance with Section 6.
The principal purpose of the capacity [strength reduction] factor (φ) is to compensate for
uncertainties inherent in the design, fabrication, or erection of building components, as
well as uncertainties in the estimation of applied loads.
The nominal capacity (Ru) is the strength of the element or member for a given limit state,
calculated for nominal section properties and for minimum specified material properties in
accordance with the appropriate analytical model which defines the strength. The capacity
[strength reduction] factor (φ) accounts for the uncertainties and variabilities inherent in
Ru, and is usually less than unity.
The design action effects [design actions] (S*) are the forces on the cross-section
(i.e. bending moment, axial force, or shear force) determined from the specified nominal
loads by structural analysis.
The advantages of limit states design are as follows:
(a) The uncertainties and the variabilities of different types of loads and capacities are
different (e.g. dead load is less variable than wind load), and so these differences
can be accounted for by use of multiple load factors.
(b) By using the probability theory, designs can ideally achieve a more consistent
reliability. Thus, limit states design provides the basis for a more rational and
refined design method than is possible with the previous permissible stress design
method. The probabilistic basis for the limit state design of cold-formed steel
structures is as follows:
(i) Probabilistic concepts Safety factors or load factors are provided against
the uncertainties and variabilities that are inherent in the design process.
Structural design consists of comparing nominal design action effects [design
action] (S) to nominal capacities (R), but both S and R are random
parameters (see Figure C1.6.2.2(1)). A limit state is violated if R is less than
S. While the possibility of this event ever occurring is never zero, a
successful design should, nevertheless, have only an acceptably small
probability of exceeding the limit state. If the exact probability distributions
of S and R were known, then the probability of (R – S) being less than zero
could be exactly determined for any design. In general, the distributions of S
and R are not known, and only the means, Sm and Rm, and the standard
deviations, σS and σ R are available. Nevertheless it is possible to determine
relative reliabilities of several designs by the scheme shown in
Figure C1.6.2.2(2). The distribution curve shown is for ln (R/S) and a limit
Accessed by BHP BILLITON on 27 Jun 2002

state is exceeded when ln (R/S) is less than or equal to zero. The area under
ln (R/S) being less than or equal to zero is the probability of violating the
limit state. The size of this area is dependent on the distance between the
origin and the mean of ln (R/S). For given statistical data Rm, S m, σR and σ S,
the area under ln (R/S) is less than or equal to zero can be varied by
changing the value of β (see Figure C1.6.2.2(2)), since βσ ln(R/S) is equal to ln
(R/S)m, and β is calculated as follows:

. . . C1.6.2.2(2)

COPYRIGHT
23 AS/NZS 4600 Supp1:1998

where
VR = coefficient of variation of R

VS = coefficient of variation of S

The index β is called the ‘reliability index’ and it is a relative measure of the
safety of the design. When two designs are compared, the one with the larger
β is more reliable.
The concept of the reliability index can be used for determining the relative
reliability inherent in current design and can be used in testing out the
reliability of new design formats, as given by the following example of
simply supported, braced beams subjected to dead and live loading.
The limit state design requirement of the Standard for such a beam is as
follows:

(using AS 1170.1) . . . C1.6.2.2(3a)

(using NZS 4203) . . . C1.6.2.2(3b)

where
φ = capacity [strength reduction] factor for bending
= 0.9
Ze = elastic section modulus based on the effective section
fy = specified yield stress
ls = span length
s = beam spacing
G = dead load
Q = live load
The mean resistance (Rm) is calculated as follows (Ravindra and
Galambos, 1978):
. . . C1.6.2.2(4)
Accessed by BHP BILLITON on 27 Jun 2002

where
Rn = nominal resistance
Rn = Ze fy . . . C1.6.2.2(5)
Ze = elastic section modulus based on the effective section
fy = yield stress of beam
Rn is the nominal moment predicted on the basis of the post-buckling
strength of the compression flange and the web. The mean values Pm, Mm and
Fm and the corresponding coefficients of variation VP, VM and VF are the
statistical parameters which define the variability of the resistance.

COPYRIGHT
AS/NZS 4600 Supp1:1998 24

Pm = mean ratio of the experimentally determined moment to the


predicted moment for the actual material, and cross-sectional
properties of the test specimens
Mm = mean ratio of the actual yield point to the minimum specified
value
Fm = mean ratio of the actual section modulus to the specified
(nominal) value
The coefficient of variation of R is calculated as follows:

VR = . . . C1.6.2.2(6)

The values of these data were obtained from examining the available tests on
beams having different compression flanges with partially and fully effective
flanges and webs, and from analysing data on yield stress values from tests
and cross-sectional dimensions from many measurements. This information
was developed from research (Hsiao, Yu and Galambos, 1988 and 1990;
Hsiao, 1989) and is given as follows:
(A) Pm = 1.11
(B) VP = 0.09
(C) Mm = 1.10
(D) VM = 0.10
(E) Fm = 1.0
(F) VF = 0.05
(G) Rm = 1.22Rn
(H) VR = 0.14
Accessed by BHP BILLITON on 27 Jun 2002

FIGURE C1.6.2.2(1) RANDOMNESS S* AND R

COPYRIGHT
25 AS/NZS 4600 Supp1:1998

FIGURE C1.6.2.2(2) RELIABILITY INDEX β

The mean load effect (Sm) is calculated as follows:

Sm = . . . C1.6.2.2(7)

VS = . . . C1.6.2.2(8)

where
Gm = mean dead load intensity
Qm = mean live load intensity
VG = coefficient of variation of G
VQ = coefficient of variation of Q
Load statistics were analysed in a study of the National Bureau of Standards
(NBS) (Ellingwood et al., 1980), where it was shown that—
Gm = 1.05G
VG = 0.1
Qm = Q
VQ = 0.25
The mean live load intensity equals the code live load intensity if the
tributary area is small enough so that no live load reduction is included.
Accessed by BHP BILLITON on 27 Jun 2002

Substitution of the load statistics into Equations C1.6.2.2(7) and C1.6.2.2(8)


gives the following:

Sm = . . . C1.6.2.2(9)

VS = . . . C1.6.2.2(10)

COPYRIGHT
AS/NZS 4600 Supp1:1998 26

Sm and VS thus depend on the dead-to-live load ratio. Cold-formed steel


beams typically have small G/Q, and for the purposes of checking the
reliability of these strength limit state criteria, it will be assumed that—

and so

From Equations C1.6.2.2(3) and C1.6.2.2(5) the nominal resistance (R n) can


be obtained for G/Q equal to 1/5 and φ equal to 0.9 as follows:

. . . C1.6.2.2(11)

In order to determine the reliability index (β) from Equation C1.6.2.2(2) the
Rm/Sm ratio is required by considering Rm equal to 1.22Rn as follows:

. . . C1.6.2.2(12)

Therefore, from Equation C1.6.2.2(2)—

. . . C1.6.2.2(13)

If β is equal to 2.66 for beams having different compression flanges with


partially and fully effective flanges, and webs designed by the Standard are
compared to β for other types of cold-formed steel members, and to β for
designs of various types from hot-rolled steel shapes or even for other
materials, then it is possible to say that this particular cold-formed steel
beam has about an average reliability (Galambos et al., 1982).
(ii) Basis for limit state design of cold-formed steel structures A great deal of
work has been performed for determining the values of the reliability index
(β) inherent in traditional design as exemplified by the current structural
design specifications such as the AISC Specification for hot-rolled steel, the
AISI Specification for cold-formed steel, the ACI Code for reinforced
concrete members, AS 4100 and NZS 3404 for steel structures in Australia
Accessed by BHP BILLITON on 27 Jun 2002

and New Zealand, and the like. The studies for hot-rolled steel are
summarized by Ravindra and Galambos (1978), where also many papers,
which contain additional data, are referenced. The determination of β for
cold-formed steel elements or members is presented in several research
reports of the University of Missouri-Rolla (Hsiao, Yu, and Galambos, 1988;
Rang, Galambos and Yu, 1979a, 1979b, 1979c and 1979d; Suporn-
silaphachai, Galambos and Yu, 1979), where both the basic research data as
well as the βs inherent in the AISI Specification are presented in great detail.
The βs calculated in the above-referenced publications were developed with
slightly different load statistics than those of this Commentary, but the
essential conclusions remain the same.

COPYRIGHT
27 AS/NZS 4600 Supp1:1998

In Australia, the underlying philosophy behind the limit state codes and their
calibration is set out in Leicester, Pham and Kleeman (1985) and Pham
(1985). No detailed calibration of AS/NZS 4600 has been performed.
However, calibration of the R-factor design procedures for purlins under
wind uplift has been performed by Rousch and Hancock (1996a, 1996b).
The entire set of data for hot-rolled steel and cold-formed steel designs, as
well as data for reinforced concrete, aluminium, laminated timber and
masonry walls was re-analysed by Ellingwood, Galambos, MacGregor and
Cornell (Ellingwood et al., 1980; Galambos et al., 1982; Ellingwood et al.,
1982) using updated load statistics and a more advanced level of probability
analysis, which was able to incorporate probability distributions and to
describe the true distributions more realistically. The details of this extensive
reanalysis are presented by the investigators. Only the final conclusions from
the analysis are summarized below.
The values of the reliability index (β) vary considerably for the different
kinds of loading, the different types of construction, and the different types
of members within a given material design specification. In order to achieve
more consistent reliability, it was suggested by Ellingwood et al. (1982) that
the following values of β would provide this improved consistency while at
the same time would give, on the average, essentially the same design by the
limit state design method as is obtained by current design for all materials of
construction. These target reliabilities (βo) for use in limit state design are
given as follows:
(A) For basic case (gravity load) βo . . . . . . . . . . . . . . . . . . . . . . . . 3.0.
(B) For connections βo . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.5.
(C) For wind load βo . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.5.
The target reliability indices are the ones inherent in the load factors
recommended in the ASCE 7-95 Load Standard (ASCE, 1995).
For simply supported, braced, cold-formed steel beams with stiffened
flanges, which were designed in accordance with the 1996 AISI allowable
stress design method or to any previous version of that specification, it was
shown that for the representative dead-to-live load ratio of 1/5, the reliability
index β is equal to 2.79. Considering the fact that for other such load ratios,
or for other types of members, the reliability index inherent in current cold-
formed steel construction could be more or less than this value of 2.79, a
somewhat lower target reliability index of βo equal to 2.5 is recommended as
a lower limit for the AISI LRFD Specification. The capacity factors (φ) were
selected such that βo equal to 2.5 is essentially the lower bound of the actual
β values for members. In order to assure that failure of a structure is not
initiated in the connections, a higher target reliability index of βo equal to
3.5 is recommended for joints and fasteners. These two targets of 2.5 and 3.5
Accessed by BHP BILLITON on 27 Jun 2002

for members and connections, respectively, are somewhat lower than those
recommended by the ASCE 7-95, but they are essentially the same targets as
are the basis for the AISC LRFD Specification (AISC, 1993). For wind loads
the same ASCE target value of βo equal to 2.5 is used in the AISI LRFD
Specification.
In the development of the AISI LRFD Specification, the following statistical
data on material and cross-sectional properties were developed by Rang,
Galambos and Yu (1979a and 1979b) for use in the derivation of capacity
[strength reduction] factors (φ):

COPYRIGHT
AS/NZS 4600 Supp1:1998 28

= 1.10fy; Mm = 1.10; Vfy = Vm = 0.11

= 1.10fya; Mm = 1.10; Vfyn = Vm = 0.11

= 1.10fu; Mm = 1.10; Vfu = Vm = 0.08

= 1.00; VF = 0.05

where
m = mean value
V = coefficient of variation
M = ratio of the mean-to-the nominal material property
F = ratio of mean to nominal cross-sectional property
fy = specified minimum yield point
fya = average yield point including the effect of cold-forming
fu = specified minimum tensile strength.
These statistical data are based on the analysis of many samples (Rang et al.,
1978) and they are representative properties of materials and cross-sections
used in the industrial application of cold-formed steel structures.
C1.6.2.3 Stability limit state This limit state is included in the Standard although it is
not included in the AISI Specification. In the New Zealand Loadings Standard, NZS 4203
(1992), it is treated as part of the ultimate limit state.
C1.6.2.4 Serviceability limit states Serviceability limit states are conditions under
which a structure can no longer perform its intended functions. Safety and strength
considerations are generally not affected by serviceability limit states. However,
serviceability criteria are essential to ensure functional performance and economy of
design.
Common conditions which may require serviceability limits are as follows:
(a) Excessive deflections or rotations which may affect the appearance or functional use
of the structure. Deflections which may cause damage to non-structural elements
should be considered.
(b) Excessive vibrations which may cause occupant discomfort or equipment
malfunctions.
(c) Deterioration over time which may include corrosion or appearance considerations.
When checking serviceability, the designer should consider the appropriate service loads,
the response of the structure, and the reaction of building occupants.
Accessed by BHP BILLITON on 27 Jun 2002

Service loads that may require consideration include static loads, snow or rain loads,
temperature fluctuations, and dynamic loads from human activities, wind-induced effects,
or the operation of equipment. The service loads are actual loads that act on the structure
at an arbitrary point in time. Appropriate service loads for checking serviceability limit
states may only be a fraction of the nominal loads.
The response of the structure to service loads can normally be analysed assuming linear
elastic behaviour. However, members that accumulate residual deformations under service
loads may require consideration of this long-term behaviour.

COPYRIGHT
29 AS/NZS 4600 Supp1:1998

Serviceability limits depend on the function of the structure and on the perceptions of the
observer. In contrast to the strength limit states, it is not possible to specify general
serviceability limits that are applicable to all structures. The Standard does not contain
explicit requirements; however, guidance is generally provided by the applicable building
code. In the absence of specific criteria, guidelines may be found in Fisher and West
(1990), Ellingwood (1989), Murray (1991), Allen and Murray (1993). Serviceability limits
for domestic metal framing in Australia are given in AS 3623 (Standards Australia,
1993a) including both static and dynamic limits. For New Zealand, contact the National
Association of Steel Framed Housing (NASH), New Zealand.

C1.7 NON-CONFORMING SHAPES AND CONSTRUCTION Alternative shapes


and construction should comply with Section 6, particularly Clause 6.2.
Accessed by BHP BILLITON on 27 Jun 2002

COPYRIGHT
AS/NZS 4600 Supp1:1998 30

S E C T I O N C 2 E L E M E N T S

C2.1 SECTION PROPERTIES


C2.1.1 General In cold-formed steel construction, individual elements of steel
structural members are thin and the width-to-thickness ratios are large, when compared
with hot-rolled steel shapes. These thin elements may buckle locally at a stress level
lower than the yield point of steel when they are subjected to compression in flexural
bending, axial compression, shear, or bearing. Figure C2.1 illustrates some local buckling
patterns of certain beams and columns (Yu, 1991).
Because local buckling of individual elements of cold-formed steel sections is a major
design criterion, the design of such members should provide sufficient safety against
failure by local instability with due consideration given to the post-buckling strength of
structural components. Section 2 of the Standard contains the design requirements for
width-to-thickness ratios and the design equations for determining the effective widths of
stiffened compression elements, unstiffened compression elements, and elements with
edge stiffeners or intermediate stiffeners. Additional provisions are made for the use of
stiffeners.
Accessed by BHP BILLITON on 27 Jun 2002

FIGURE C2.1 LOCAL BUCKLING OF COMPRESSION ELEMENTS

C2.1.2 Design procedures The appropriate use of full and effective section properties
is explained in this Section. Full section properties are used for the determination of
buckling moments and stresses, as specified in Clause 3.1. Clause 2.1.2.1 specifies clearly
how the full section properties are to be calculated. For the more difficult property
calculations such as distortional buckling, shear centre position, warping constant and
monosymmetry parameters, the bends may be eliminated. The Standard has been
calibrated on this assumption.

COPYRIGHT
31 AS/NZS 4600 Supp1:1998

Effective section properties are used for the determination of section and member
capacities to allow for local instabilities of elements in compression, and for shear lag.
Clause 2.1.2.3 specifies where the reduced widths are to be located for both stiffened and
unstiffened elements, and elements under stress gradient and with edge stiffeners. These
requirements are the same as the ones shown in the appropriate figures later in the
Section.
C2.1.3 Dimensional limits
C2.1.3.1 Maximum flat-width-to-thickness ratios Clause 2.1.3.1 of the Standard
contains limitations on permissible flat-width-to-thickness ratios of compression flanges.
To some extent, these limitations are arbitrary. They do, however, reflect a long-time
experience and are intended to define practical ranges of application (Winter, 1970).
The limitation to a maximum b/t of 60 for compression flanges having one longitudinal
edge connected to a web and the other edge stiffened by a simple lip is based on the fact
that if the b/t ratio of such a flange exceeds 60, a simple lip with a relatively large depth
would be required to stiffen the flange (Winter, 1970). The local instability of the lip
would necessitate a reduction of the bending capacity to prevent premature buckling of
the stiffening lip.
The limitation to b/t equal to 90 for compression flanges with any other kind of stiffeners
indicates that thinner flanges with large b/t ratios are quite flexible and liable to be
damaged in transport, handling and erection. The same is true for the limitation to b/t
equal to 500 for stiffened compression elements with both longitudinal edges connected to
other stiffened elements and for the limitation to b/t equal to 60 for unstiffened
compression elements. The Note specifically states that wider flanges are not unsafe, but
that when the b/t ratio of unstiffened flanges exceeds 30 and the b/t ratio of stiffened
flanges exceeds 250, they are likely to develop noticeable deformation at the full design
strength, without affecting the ability of the member to develop the required strength. In
both cases the maximum b/t is set at twice that ratio at which first noticeable
deformations are likely to appear, based on observations of such members under tests.
These upper limits will generally keep such deformations to reasonable limits. In cases
where the limits are exceeded, tests in accordance with Section 6 are required.
C2.1.3.2 Flange curling In beams that have unusually wide and thin but stable flanges,
i.e. primarily tension flanges with large b/t ratios, there is a tendency for these flanges to
curl under bending. That is, the portions of these flanges most remote from the web
(edges of I-beams, centre portions of flanges of box or hat beams) tend to deflect toward
the neutral axis. An approximate, analytical treatment of this problem is given by Winter
(1948b). Equation 2.1.3.2 of the Standard permits one to compute the maximum
permissible flange width (b1) for a given amount of flange curling (cf).
It should be noted that Clause 2.1.3.2 does not stipulate the amount of curling, which can
be regarded as tolerable, but an amount of curling in the order of 5% of the depth of the
section is not excessive under usual conditions. In general, flange curling is not a critical
factor to govern the flange width. However, when the appearance of the section is
Accessed by BHP BILLITON on 27 Jun 2002

important, the out-of-plane distortion should be closely controlled in practice. Recently,


research on flange curling of profiled steel decks was completed by Bernard, Bridge and
Hancock (1996). A number of alternative curling models were developed for profiled steel
decks.
C2.1.3.3 Shear lag effects (short spans supporting concentrated loads) For beams of
usual shapes, the normal stresses are induced in the flanges through shear stresses
transferred from the web to the flange. These shear stresses produce shear strains in the
flange which, for ordinary dimensions, have negligible effects. However, if flanges are
unusually wide (relative to their length) these shear strains have the effect that the normal
bending stresses in the flanges decrease with increasing distance from the web. This
phenomenon is known as shear lag. It results in a non-uniform stress distribution across

COPYRIGHT
AS/NZS 4600 Supp1:1998 32

the width of the flange, similar to that in stiffened compression elements


(see Clause C2.2), though for entirely different reasons. The simplest way of accounting
for this stress variation in design is to replace the non-uniformly stressed flange of actual
width (b1) by one of reduced, effective width subject to uniform stress (Winter, 1970).
Theoretical analyses by various investigators have arrived at results which differ
numerically (Roark, 1965). The provisions of Clause 2.1.3.3 are based on the analysis and
supporting experimental evidence obtained by detailed stress measurements on eleven
beams (Winter, 1940). In fact, the values of effective widths in Table 2.1.2 are taken
directly from Curve A of Figure 4 of Winter (1940).
It should be noted that, in accordance with Clause 2.1.3.3, the use of a reduced width for
stable, wide flanges is required only for concentrated load, as shown in Figure C2.1.3.3.
For uniform load, it can be seen from Curve B of Figure C2.1.3.3 that the width reduction
due to shear lag for any practicable large width-span ratio is so small as to be effectively
negligible.
The phenomenon of shear lag is of considerable consequence in naval architecture and
aircraft design. However, in cold-formed steel construction it is infrequent that beams are
so wide as to require significant reductions in accordance with Clause 2.1.3.3.

FIGURE C2.1.3.3 ANALYTICAL CURVES FOR DETERMINING EFFECTIVE WIDTH


OF FLANGE OF SHORT SPAN BEAMS

C2.1.3.4 Maximum web depth-to-thickness ratio Prior to 1980, the maximum web
depth-to-thickness ratio (d1/tw) was limited to —
(a) 150 for cold-formed steel members with unreinforced webs; and
(b) 200 for members that are provided with adequate means of transmitting
Accessed by BHP BILLITON on 27 Jun 2002

concentrated loads or reactions into the web, or both.


Based on the studies conducted at the University of Missouri-Rolla in the 1970s (LaBoube
and Yu, 1978a, 1978b, and 1982; Hetrakul and Yu, 1978 and 1980; Nguyen and Yu,
1978a and 1978b), the maximum d1/tw ratios were increased to —
(i) 200 for unreinforced webs;
(ii) 260 for webs using bearing stiffeners; and
(iii) 300 for webs using bearing and intermediate stiffeners in the 1980 edition of the
AISI Specification.

COPYRIGHT
33 AS/NZS 4600 Supp1:1998

These d1/tw limitations are the same as those used in the AISC Specification (AISC, 1989)
for plate girders and are retained in the 1996 edition of the AISI Specification and
AS/NZS 4600. Because the definition for d1 was changed in the 1986 edition of the AISI
Specification from the ‘clear distance between flanges’ to the ‘depth of flat portion’
measured along the plane of web, the prescribed maximum d1/tw ratio may appear to be
more liberal. An unpublished study by LaBoube concluded that the present definition for
d1 had negligible influence on the web strength.

C2.2 EFFECTIVE WIDTHS OF STIFFENED ELEMENTS It is well known that the


structural behaviour and the load-carrying capacity of a stiffened compression element,
such as the compression flange of a hat section, depend on the b/t ratio and the supporting
condition along both longitudinal edges. If the b/t ratio is small, the stress in the
compression flange can reach the yield stress of steel and the strength of the compression
element is governed by yielding. For the compression flange with large b/t ratios, local
buckling (see Figure C2.2(1)) will occur at the following elastic critical buckling stress
(fol):

. . . C2.2

where
k = plate buckling coefficient (see Table C2.2)
= 4 for stiffened compression elements supported by a web on each longitudinal
edge
E = modulus of elasticity of steel
ν = Poisson’s ratio
= 0.3 for steel in the elastic range
b = flat width of the compression element
t = thickness of the compression element
When the elastic critical buckling stress calculated using Equation C2.2(1) exceeds the
proportional limit of the steel, the compression element will buckle in the inelastic range
(Yu, 1991).
Accessed by BHP BILLITON on 27 Jun 2002

COPYRIGHT
AS/NZS 4600 Supp1:1998 34

TABLE C2.2
VALUES OF PLATE BUCKLING COEFFICIENTS

Value of k for
Case Boundary condition Type of stress
long plate

(a) Compression 4.0

(b) Compression 6.97

(c) Compression 0.425

(d) Compression 1.277

(e) Compression 5.42

(f) Shear 5.34

(g) Shear 8.98

(h) Bending 23.9

(i) Bending 41.8


Accessed by BHP BILLITON on 27 Jun 2002

COPYRIGHT
35 AS/NZS 4600 Supp1:1998

FIGURE C2.2(1) LOCAL BUCKLING OF STIFFENED COMPRESSION FLANGE


OF HAT-SHAPED BEAM

Unlike one-dimensional structural members such as columns, stiffened compression


elements will not collapse when the buckling stress is reached. An additional load can be
carried by the element after buckling by means of a redistribution of stress. This
phenomenon is known as post-buckling strength of the compression elements and is most
pronounced for stiffened compression elements with large b/t ratios. The mechanism of
the post-buckling action of compression elements is discussed by Winter in earlier
editions of the AISI Commentary (Winter, 1970).
Imagine, for the sake of simplicity, a square plate uniformly compressed in one direction,
with the unloaded edges simply supported. Since it is difficult to visualize the
performance of such two-dimensional elements, the plate will be replaced by a model,
which is shown in Figure C2.2(2). The model consists of a grid of longitudinal and
transverse bars in which the material of the actual plate is thought to be concentrated.
Since the plate is uniformly compressed, each of the longitudinal struts represents a
column loaded by N/5 where N is the total axial force on the plate. As the force is
gradually increased, the compression stress in each of these struts will reach the critical
column buckling value and all five struts will tend to buckle simultaneously. If these
struts were simple columns, unsupported except at the ends, they would simultaneously
collapse through unrestrained, increasing horizontal deflection. It is evident, however, that
this cannot occur in the grid model of the plate. Indeed, as soon as the longitudinal struts
start deflecting at their buckling stress, the transverse bars that are connected to them
should stretch like ties in order to accommodate the imposed deflection. Like any
structural material, they resist stretch and, thereby, have a restraining effect on the
deflections of the longitudinal struts.
The tension forces in the horizontal bars of the grid model correspond to the so-called
Accessed by BHP BILLITON on 27 Jun 2002

membrane stresses in a real plate. These stresses, just as in the grid model, come into play
as soon as the compression stresses begin to cause buckling waves. They consist mostly
of transverse tension, but also of some shear stresses, and they counteract increasing wave
deflections, i.e. they tend to stabilize the plate against further buckling under the applied
increasing longitudinal compression. Hence, the resulting behaviour of the model is as
follows:
(a) There is no collapse by unrestrained deflections as in unsupported columns.
(b) The various struts will deflect unequal amounts, those nearest the supported edges
being held almost straight by the ties, those nearest the centre being able to deflect
most.

COPYRIGHT
AS/NZS 4600 Supp1:1998 36

FIGURE C2.2(2) POST-BUCKLING STRENGTH MODEL

In consequence of Item (a), the model will not collapse and fail when its buckling stress
(see Equation C2.2(1)) is reached, in contrast to columns it will merely develop slight
deflections but will continue to carry increasing load. In consequence of Item (b), the
struts (strips of the plate) closest to the centre, which deflect most, ‘get away from the
load’, and hardly participate in carrying any further load increases. These centre strips
may, in fact, even transfer part of their pre-buckling load to their neighbours. The struts
(or strips) closest to the edges, held straight by the ties, continue to resist increasing load
with hardly any increasing deflection. For the plate, this means that the hitherto uniformly
distributed compression stress re-distributes itself in a manner shown in Figure C2.2(3),
the stresses being largest at the edges and smallest in the centre. With additional increase
in load this non-uniformity increases further, as also shown in Figure C2.2(3). The plate
fails, i.e. refuses to carry any further load increases, only when the most highly stressed
strips, near the supported edges, begin to yield, i.e. when the compression stress (f max.)
reaches the yield stress (fy).
This post-buckling strength of plates was discovered experimentally in 1928, and an
approximate theory of it was first given by Th. v. Karman in 1932 (Bleich, 1952). It has
been used in aircraft design ever since. A graphic illustration of the phenomenon of
post-buckling strength can be found in the series of photographs shown in Figure 7 of
Accessed by BHP BILLITON on 27 Jun 2002

Winter (1959b).
The model shown in Figure C2.2(2) demonstrates the behaviour of a compression element
supported along both longitudinal edges, as does the flange shown in Figure C2.2(1). In
fact, such elements buckle into approximately square waves.
In order to utilize the post-buckling strength of the stiffened compression element for
design purposes, the AISI Specification has used the effective design width approach to
determine the sectional properties, and since 1946 and AS/NZS 4600 has adopted this
effective width approach. In Clause 2.2 of the Standard, design equations for calculating
the effective widths are provided for the following three cases:
(i) Uniformly compressed stiffened elements.

COPYRIGHT
37 AS/NZS 4600 Supp1:1998

(ii) Uniformly compressed stiffened elements with circular holes.


(iii) Webs and stiffened elements with a stress gradient.
The background information on various design requirements is discussed in the
subsequent sections herein.
C2.2.1 Uniformly compressed stiffened elements
C2.2.1.2 Effective width for capacity calculations In the ‘effective design width’
approach, instead of considering the non-uniform distribution of stress over the entire
width of the plate (b) it is assumed that the total load is carried by a fictitious effective
width (be) subject to a uniformly distributed stress equal to the edge stress (f max.) as shown
in Figure C2.2(3). The width (be) is selected so that the area under the curve of the actual
non-uniform stress distribution is equal to the sum of the two parts of the equivalent
rectangular shaded area with a total width (b e) and an intensity of stress equal to the edge
stress (fmax.).

FIGURE C2.2(3) STRESS DISTRIBUTION IN


STIFFENED COMPRESSION ELEMENTS

Based on the concept of ‘effective width’ introduced by von Karman et al. (von Karman,
Sechler and Donnell, 1932) and the extensive investigation on light-gauge, cold-formed
steel sections at Cornell University, the following equation was developed by Winter in
1946 for determining the effective width (be) for stiffened compression elements simply
supported along both longitudinal edges:

. . . C2.2.1.2(1)
Accessed by BHP BILLITON on 27 Jun 2002

Equation C2.2.1.2(1) can be written in terms of the ratio fol/fmax. as follows:

. . . C2.2.1.2(2)

COPYRIGHT
AS/NZS 4600 Supp1:1998 38

During the period from 1946 to 1968, the AISI design provision for the determination of
the effective design width was based on Equation C2.2.1.2(1). A long-time accumulated
experience has indicated that the following more realistic equation may be used for the
determination of the effective width (be) (Winter, 1970):

. . . C2.2.1.2(3)

The correlation between the test data on stiffened compression elements and
Equation C2.2.1.2(3) is illustrated by Yu (1991).
It should be noted that Equation C2.2.1.2(3) may also be rewritten in terms of the fol/fmax.
ratio as follows:

. . . C2.2.1.2(4)

Therefore, the effective width (b) can be determined as follows:


be = ρb . . . C2.2.1.2(5)
where
ρ = effective width factor

. . . C2.2.1.2(6)

In Equation C2.2.1.3(6), λ is a slenderness ratio and is calculated as follows:

. . . C2.2.1.2(7)

Figure C2.2(4) shows the relationship between φ and λ. It can be seen that if λ is less
than or equal to 0.673, φ is equal to 1.0.
Accessed by BHP BILLITON on 27 Jun 2002

COPYRIGHT
39 AS/NZS 4600 Supp1:1998

FIGURE C2.2(4) REDUCTION FACTOR (φ) VERSUS SLENDERNESS FACTOR (λ)

Based on Equations C2.2.1.2(5) to C2.2.1.2(7) and the unified approach proposed by


Peköz (1986b and 1986c), the 1986 edition of the AISI Specification adopted the
non-dimensional format in Clause 2.2 for determining the effective design width (be) for
uniformly, compressed stiffened elements. The same design equations are used in the
1996 edition of the AISI Specification and in the Standard.
In the Standard, fmax. in Equation C2.2.1.2(7) is taken as f *, the design stress in the
compression element calculated on the basis of the effective design width.
C2.2.1.3 Effective width for deflection calculations The effective design width
equations for load capacity determination discussed in Clause C2.2.1.2 can also be used to
obtain a conservative effective width (bed) for deflection calculation. It is included in
Clause 2.2.1.3 of the Standard as Procedure I.
For stiffened compression elements supported by a web on each longitudinal edge, a study
conducted by Weng and Peköz (1986) indicated that Equations 2.2.1.2(3) to 2.2.1.2(6) of
the Standard can yield a more accurate estimate of the effective width (bde) for deflection
analysis. These equations are given in Procedure II of the Standard for additional design
information. The design engineer has the option of using one of the two procedures for
determining the effective width to be used for deflection calculations.
C2.2.2 Uniformly compressed stiffened elements with circular holes In cold-formed
steel structural members, holes are sometimes provided in webs or flanges, or both, of
beams and columns for duct work, piping and other construction purposes. The presence
of such holes may result in a reduction of the strength of individual component elements
Accessed by BHP BILLITON on 27 Jun 2002

and the overall strength and stiffness of the members depending on the size, shape
arrangement of holes, the geometric configuration of the cross-section, and the mechanical
properties of the material.
The exact analysis and the design of steel sections having perforations are complex,
particularly when the shapes and the arrangement of holes are unusual. The limited design
provisions included in Clause 2.2.2 of the Standard for uniformly compressed stiffened
elements with circular holes are based on a study conducted by Ortiz-Colberg and Peköz
at Cornell University (Ortiz-Colberg and Peköz, 1981). For additional information on the
structural behaviour of perforated elements, see Yu and Davis (1973a) and Yu (1991).

COPYRIGHT
AS/NZS 4600 Supp1:1998 40

C2.2.3 Stiffened elements with stress gradient When a beam is subjected to bending
moment, the compression portion of the web may buckle due to the compressive stress
caused by bending. The theoretical critical buckling stress for a flat rectangular plate
under pure bending can be calculated using Equation C2.2(1), except that the depth-to-
thickness ratio (d1/tw) is substituted for the width-to-thickness ratio (b/t) and the plate
buckling coefficient (k) is equal to 23.9 for simple supports as given in Table C2.2.
In AS 1538 — 1988, the design of cold-formed steel beam webs was based on the full web
depth with the permissible bending stress (Fbw) specified in the Standard. In order to unify
the design methods for web elements and compression flanges, the ‘effective design
depth’ approach was adopted in the 1986 edition of the AISI Specification on the basis of
the studies made by Peköz (1986b), Cohen and Peköz (1987). This is a different approach
as compared with the past practice of using a full area of the web element in conjunction
with a reduced stress to account for local buckling and post-buckling strength (LaBoube
and Yu, 1982; Yu, 1985). The effective design depth approach is used in the Standard.

C2.3 EFFECTIVE WIDTHS OF UNSTIFFENED ELEMENTS Similar to stiffened


compression elements, the stress in the unstiffened compression elements can reach the
yield stress of steel if the b/t ratio is small. Because the unstiffened element has one
longitudinal edge supported by the web and the other edge is free, the limiting width-to-
thickness ratio of unstiffened elements is much less than that for stiffened elements.
When the b/t ratio of the unstiffened element is large, local buckling (see Figure C2.3(1))
will occur at the elastic critical stress calculated using Equation C2.2(1) with a value of k
equal to 0.43. This buckling coefficient is given in Table C2.2 for Case (c). For the
intermediate range of b/t ratios, the unstiffened element will buckle in the inelastic range.
Figure C2.3(2) shows the relationship between the maximum stress for unstiffened
compression elements and the b/t ratio, in which Line A is the yield point of steel, Line B
represents the inelastic buckling stress, Curves C and D illustrate the elastic buckling
stress. The equations for Curves A, B, C and D have been developed from previous
experimental and analytical investigations and used for determining the allowable design
stresses in the AISI Specification up to 1986 and AS 1538 — 1974 (Winter, 1970; Yu,
1991; Hancock, 1998). Also shown in Figure C2.3(2) is Curve E, which represents the
maximum stress on the basis of the post-buckling strength of the unstiffened element. The
correlation between the test data on unstiffened elements and the predicted maximum
stresses is shown in Figure C2.3(3) (Yu, 1991).
Accessed by BHP BILLITON on 27 Jun 2002

FIGURE C2.3(1) LOCAL BUCKLING OF UNSTIFFENED COMPRESSION FLANGE

COPYRIGHT
41 AS/NZS 4600 Supp1:1998

FIGURE C2.3(2) MAXIMUM STRESS FOR UNSTIFFENED


COMPRESSION ELEMENTS
Accessed by BHP BILLITON on 27 Jun 2002

FIGURE C2.3(3) CORRELATION BETWEEN TEST DATA


AND PREDICTED MAXIMUM STRESS

COPYRIGHT
AS/NZS 4600 Supp1:1998 42

Prior to 1986 (AISI) and 1988 (AS 1538), it had been a general practice to design
cold-formed steel members with unstiffened flanges by using the permissible stress design
approach. The effective width equation was not used in earlier editions of the AISI
Specification and AS 1538 — 1974 due to lack of extensive experimental verification and
the concern for excessive out-of-plane distortions under service loads.
In the 1970s, the applicability of the effective width concept to unstiffened elements
under uniform compression was studied in detail by Kalyanaraman, Peköz, and Winter at
Cornell University (Kalyanaraman, Peköz, and Winter, 1977; Kalyanaraman and Peköz,
1978). The evaluation of the test data using k equal to 0.43 was presented and
summarized by Peköz in the AISI report (Peköz, 1986b), which indicates that
Equation C2.2.1.2(6), developed for stiffened compression elements, gives a conservative
lower bound to the test results of unstiffened compression elements. In addition to the
strength determination, the same study also investigated the out-of-plane deformations in
unstiffened elements. The results of theoretical calculations and the test results on the
sections having unstiffened elements with b/t equal to 60 were presented by Peköz in the
same report. It was found that the maximum amplitude of the out-of-plane deformation at
failure can be twice the thickness as the b/t ratio approaches 60. However, the
deformations are significantly less under the service loads. Based on the above reasons
and justifications, the effective design width approach was adopted for the first time in
Section B3 of the 1986 AISI Specification and AS 1538 — 1988 for the design of
cold-formed steel members having unstiffened compression elements.
C2.3.1 Uniformly compressed unstiffened elements The effective widths (be) of
uniformly compressed unstiffened elements can be determined in accordance with
Clause 2.2.1.2 of the Standard with the exception that the plate buckling coefficient (k) is
taken as 0.43. This is a theoretical value for long plates. See Case (c) given in
Table C2.2. For deflection determination, the effective widths of uniformly compressed
unstiffened elements can only be determined in accordance with Procedure I of
Clause 2.2.1.3 of the Standard, because Procedure II was developed for stiffened
compression elements only.
C2.3.2 Unstiffened elements and edge stiffeners with stress gradient In
concentrically loaded compression members and in flexural members where the
unstiffened compression element is parallel to the neutral axis, the stress distribution is
uniform prior to local buckling. However, when edge stiffeners of the beam section are
turned in or out, the compressive stress in the edge stiffener is not uniform but varies in
proportion to the distance from the neutral axis.
There is a very limited amount of information on the behaviour of unstiffened
compression elements with a stress gradient. Cornell research on the behaviour of edge
stiffeners for flexural members has demonstrated that by using Winter’s effective width
equation (Equation C2.2.1.3(4)) with a k equal to 0.43, good correlation is achieved
between the tested and calculated capacity (Peköz, 1986b). The same trend is also true for
deflection determination. Therefore, in Clause 2.3.2 of the Standard, unstiffened elements
and edge stiffeners with stress gradient are treated as uniformly compressed elements for
Accessed by BHP BILLITON on 27 Jun 2002

the calculation of effective widths with stress (f*) to be the maximum compression stress
in the element.
The Standard allows two additional alternative procedures to those specified in the AISI
Specification. Firstly, the plate buckling coefficient (k) for each flat element may be
determined from a rational elastic buckling analysis of the whole section as a plate
assemblage subjected to the longitudinal stress distribution in the section prior to
buckling. This could be achieved using a finite strip buckling analysis such as that
described by Papangelis and Hancock (1995).
The second alternative approach allowed in the Standard is given in Appendix F. This
Appendix is based on Eurocode 3, Part 1.3, (1996) where buckling coefficients and
effective widths of unstiffened elements under stress gradient are given in Table 4.2. The

COPYRIGHT
43 AS/NZS 4600 Supp1:1998

use of this Table is assumed not to be iterative so that the stresses f *1 and f*2 used to
calculate the buckling coefficient (k) and effective width (be) are based on the stresses on
the full (gross) section. This approach is a higher tier in the Standard and further research
is required to fully validate its applicability for a range of sections.

C2.4 EFFECTIVE WIDTHS OF UNIFORMLY COMPRESSED ELEMENTS WITH


AN EDGE STIFFENER OR ONE INTERMEDIATE STIFFENER For cold-formed
steel beams such as hat, box or inverted U-type sections shown in Figures C1.3.2(c), (d)
and (e) the compression flange is supported along both longitudinal edges by webs. In this
case, if the webs are properly designed, they provide adequate stiffening for the
compression elements by preventing their longitudinal edges from out-of-plane
displacements. On the other hand, in many cases, only one longitudinal edge is stiffened
by the web, while the other edge is supported by an edge stiffener. In most cases, the
edge stiffener takes the form of a simple lip, such as in the channel and I-section as
shown in Figures C1.3.2(a) and (b).
The structural efficiency of a stiffened element always exceeds that of an unstiffened
element with the same b/t ratio by a sizeable margin, except for low b/t ratios, for which
the compression element is fully effective. When stiffened elements with large b/t ratios
are used, the material is not employed economically in as much as an increasing
proportion of the width of the compression element becomes ineffective. On the other
hand, in many applications of cold-formed steel construction, such as panels and decks,
maximum coverage is desired and, therefore, large b/t ratios are called for. For large b/t
ratios, structural economy can be improved by providing intermediate stiffeners between
webs. Such intermediate stiffeners provide optimum stiffening if they do not participate in
the wave-like distortion of the compression element, in which case they break up the
wave pattern so that the two strips to each side of intermediate stiffener distort
independently of each other, each in a pattern similar to that shown for a simple, stiffened
element in Figure C2.2(1). Compression elements furnished with such intermediate
stiffeners are designated as ‘multiple-stiffened segments’.
As far as the design provisions are concerned, the 1980 and earlier editions of the AISI
Specification and AS 1538 — 1988 included the requirements for the minimum second
moment of area of stiffeners to provide sufficient rigidity. When the size of the actual
stiffener does not satisfy the required second moment of area, the load-carrying capacity
of the beam has to be determined either on the basis of a flat element disregarding the
stiffener or through tests.
In 1986, the AISI Specification included the revised provisions in Clause 2.4 for
determining the effective widths of elements with an edge stiffener or one intermediate
stiffener on the basis of Peköz’s research findings in regard to stiffeners (Peköz, 1986b).
These design provisions were based on both critical local buckling and ultimate strength
criteria recognizing the interaction of plate elements. Also, for the first time, the design
provisions could be used for analysing partially stiffened and adequately stiffened
compression elements using different sizes of stiffeners. These provisions are included in
Accessed by BHP BILLITON on 27 Jun 2002

Clause 2.4 of the Standard.


C2.4.2 Elements with an intermediate stiffener The buckling behaviour of
rectangular plates with central stiffeners is discussed by Bulson (1969). For the design of
cold-formed steel beams using intermediate stiffeners, the 1980 AISI Specification and
AS 1538 — 1988 contained provisions for the minimum required second moment of area,
which was based on the assumption that an intermediate stiffener needed to be twice as
rigid as an edge stiffener. Subsequent research conducted by Desmond, Peköz, and Winter
(1981b) has developed expressions for evaluating the required stiffener rigidity based
upon the geometry of the contiguous flat elements.

COPYRIGHT
AS/NZS 4600 Supp1:1998 44

In view of the fact that for some cases the design requirements for intermediate stiffeners
included in the 1980 Specification could be unduly conservative (Peköz, 1986b), the AISI
design provisions were revised in 1986 according to Peköz’s research findings (Peköz,
1986b and 1986c). In this method, the buckling coefficient for determining the effective
width of sub-elements and the reduced area of the stiffener are to be calculated by using
the ratio Is/Ia. In the foregoing expression, Is is the actual stiffener second moment of area
and Ia is the adequate second moment of area of the stiffener determined from the
applicable equations.
C2.4.3 Elements with an edge stiffener An edge stiffener is used to provide a
continuous support along a longitudinal edge of the compression flange to improve the
buckling stress. Even though in most cases, the edge stiffener takes the form of a simple
lip, other types of edge stiffeners can also be used for cold-formed steel members. In
order to provide necessary support for the compression element, the edge stiffener must
possess sufficient rigidity. Otherwise it may buckle perpendicular to the plane of the
element to be stiffened.
Both theoretical and experimental studies on the local stability of compression flanges
stiffened by edge stiffeners have been carried out in the past. The design requirements
included in Clause 2.4.3 of the Standard are based on the investigations on adequately
stiffened and partially stiffened elements conducted by Desmond, Peköz and Winter
(1981a), with the additional research work by Peköz and Cohen (Peköz, 1986b). These
design provisions were developed on the basis of the critical buckling criterion and the
ultimate strength criterion.
Clause 2.4.3 recognizes that the necessary stiffener rigidity depends upon the slenderness
(b/t) of the plate element being stiffened. Thus, Cases I, II and III each contains different
definitions for an adequate stiffener second moment of area.
The interaction of the plate elements, as well as the degree of edge support, full or partial
is compensated for in the expressions for k, ds, and As (Peköz, 1986b).
In the 1996 edition of the AISI Specification (AISI, 1996) and AS/NZS 4600, the design
equations for buckling coefficient were changed for further clarity. In Case II, the
equation for k a equal to 5.25 − 5 (d 1/b) ≤ 4.0 is applicable only for simple lip stiffeners
because the term d1/b is meaningless for other types of edge stiffeners. It should be noted
that the provisions in this section were based on research dealing only with simple lip
stiffeners and extension to other types of stiffeners was purely intuitive. The requirement
of 140°≥ θ ≥ 40° for the applicability of these provisions was also decided on an intuitive
basis.
Test data to verify the accuracy of the simple lip stiffener design was collected from a
number of sources, both university and industry. These tests showed good correlation with
the equations in Clause 2.4.3. However, proprietary testing conducted in 1989 revealed
that lip lengths with a d/t ratio of greater than 14 gave unconservative results.
A review of the original research data showed a lack of data for simple stiffening lips
with d/t ratios greater than 14. Therefore, pending further research, an upper limit of 14 is
Accessed by BHP BILLITON on 27 Jun 2002

recommended.

C2.5 EFFECTIVE WIDTHS OF EDGE-STIFFENED ELEMENTS WITH ONE OR


MORE INTERMEDIATE STIFFENERS OR STIFFENED ELEMENTS WITH
MORE THAN ONE INTERMEDIATE STIFFENER As discussed in Clause C2.4, the
current design provisions for the effective widths of elements with an edge stiffener or
one intermediate stiffener were based on the results of previous Cornell research. Because
there has been insufficient research to further our understanding of the behaviour of
multiple-stiffened elements, the 1996 edition of the AISI Specification and AS/NZS 4600
have retained Equation 2.5(1) from previous editions of the Specification (AISI, 1986;
1991) and AS 1538 (1988) for evaluating the minimum required rigidity (I s, min.) of an

COPYRIGHT
45 AS/NZS 4600 Supp1:1998

intermediate stiffener for multiple-stiffened elements. If the actual second moment of area
of the full intermediate stiffener (Is) does not satisfy the minimum requirement of
Equation 2.5(1), the intermediate stiffener is disregarded for the determination of the
effective width of stiffened elements. The problem involved in the determination of the
load-carrying capacities of members having such inadequately stiffened compression
elements is complex because the buckling wave tends to spread across the intermediate
stiffener rather than being limited to individual waves occurring on both sides of the
stiffener. Once such a spreading wave occurs, the stiffened compression element is hardly
better than an element without intermediate stiffeners. For this reason, the sectional
properties of members having inadequately stiffened compression flanges are determined
on the basis of flat elements disregarding the intermediate stiffeners. The same is true for
edge-stiffened elements with intermediate stiffeners.
In addition, Clause 2.5(a) stipulates that if the spacing of intermediate stiffeners between
two webs is such that for the sub-element between stiffeners b e < b, only two intermediate
stiffeners adjacent to web elements shall be counted as effective. Additional stiffeners
would have two or more sub-elements between themselves and the nearest
shear-transmitting element (i.e. web) and, hence, could be ineffective. Clause 2.5(b)
applies the same reasoning to intermediate stiffeners between a web and an edge stiffener.
If intermediate stiffeners are spaced so closely that the sub-elements are fully effective,
i.e. be equal to b, no plate buckling of the sub-elements will occur. Therefore, the entire
assembly of sub-elements and intermediate stiffeners between webs behaves like a single
compression element whose rigidity is given by the second moment of area (I sf) of the full
area of the multiple-stiffened element, including stiffeners. Although the effective width
calculations are based upon an equivalent element having width (b2) and thickness (t s), the
actual thickness should be used when calculating section modulus.
With regard to the effective design width, results of tests of cold-formed steel sections
having intermediate stiffeners showed that the effective design width of a sub-element of
the multiple-stiffened compression elements is less than that of a single-stiffened element
with the same b/t ratio. This is true, particularly if the b/t ratio of the sub-element exceeds
about 60.
This phenomenon is due to the fact that, in beam sections, the normal stresses in the
flanges result from shear stresses between web and flange. The web generates the normal
stresses by means of the shear stress, which transfers to the flange. The more remote
portions of the flange obtain their normal stress through shear from those close to the
web. For this reason there is a difference between webs and intermediate stiffeners. The
latter is not a shear-resisting element and does not generate normal stresses through shear.
Any normal stress in the intermediate stiffener should be transferred to it from the web or
webs through the flange portions. As long as the sub-element between web and stiffener is
flat or is only very slightly buckled, this stress transfer proceeds in an unaffected manner.
In this case, the stress in the stiffener equals that at the web, and the sub-element is as
effective as a regular single-stiffened element with the same b/t ratio. However, for
sub-elements having larger b/t ratios, the slight buckling waves of the sub-element
Accessed by BHP BILLITON on 27 Jun 2002

interfere with complete shear transfer and create a ‘shear lag’ problem which results in a
stress distribution as shown in Figure C2.5.
For multiple-stiffened compression elements or wide stiffened elements with edge
stiffeners, the effective widths of sub-elements and the effective areas of stiffeners are
calculated using Equations 2.5(3) to 2.5(6) of Clause 2.5.

COPYRIGHT
AS/NZS 4600 Supp1:1998 46

FIGURE C2.5 STRESS DISTRIBUTION IN COMPRESSION FLANGE


WITH INTERMEDIATE STIFFENERS

C2.6 ARCHED COMPRESSION ELEMENTS An arched compression element is


shown in Figure 1.3(d) and defined in Clause 1.3.3 of the Standard. This provision was
carried forward from AS 1538 — 1988.

C2.7 STIFFENERS
C2.7.1 Transverse stiffeners Design requirements for attached transverse stiffeners
and for shear stiffeners were added in the 1980 AISI Specification and were unchanged in
the 1986 Specification and were also given in AS 1538 (1988). The same design equations
are retained in the 1996 AISI Specification and the Standard. The nominal strength
equation given in Item (a) of Clause 2.7.1 of the Standard serves to prevent end crushing
of the transverse stiffeners, while the nominal strength equation given in Item (b) of
Clause 2.7.1 is to prevent column-type buckling of the web-stiffeners. The equations for
calculating the effective areas (As1 and As2) and the effective widths (b1 and b2) were
adopted from Nguyen and Yu (1978a) with minor modifications.
The available experimental data on cold-formed steel transverse stiffeners were evaluated
by Hsiao, Yu and Galambos (1988). A total of 61 tests was examined. The capacity factor
of 0.85 used was selected on the basis of the statistical data.
C2.7.2 Shear stiffeners The requirements for shear stiffeners included in Clause 2.7.2
of the Standard were primarily adopted from the AISC Specification (1978). The
equations for calculating the minimum required second moment of area
(Equation 2.7.2(1)) and the minimum required gross area (Equation 2.7.2(2)) of attached
intermediate stiffeners are based on the studies summarized by Nguyen and Yu (1978a).
In Equation 2.7.2(1), the minimum value of (d1/50)4 was selected from the AISC
Specification (AISC, 1978).
The available experimental data on the shear strength of beam webs with shear stiffeners
were calibrated by Hsiao, Yu and Galambos (1988). The statistical data used for
Accessed by BHP BILLITON on 27 Jun 2002

determining the capacity factor were summarized in the AISI Design Manual
(AISI, 1991). Based on these data, the safety index was found to be 4.10 for φ equal
to 0.90.
C2.7.3 Non-conforming stiffeners Tests on rolled-in transverse stiffeners covered in
Clause 2.7.3 of the Standard were not conducted in the experimental program reported by
Nguyen and Yu (1978a). Lacking reliable information, the design capacities of members
are to be determined by tests in accordance with Section 6 of the Standard.

COPYRIGHT
47 AS/NZS 4600 Supp1:1998

S E C T I O N C 3 M E M B E R S

INTRODUCTION Section 3 provides the design requirements for the following:


(a) Tension members.
(b) Flexural members.
(c) Concentrically loaded compression members.
(d) Combined axial load and bending.
(e) Cylindrical tubular members.
To simplify the use of the Standard, all design provisions for a given specific member
type have been assembled in a particular Section within the Standard. In general, a
common nominal strength equation is provided in the Standard for a given limit state with
a required capacity [strength reduction] factor (φ) for limit states design.

C3.1 GENERAL The geometric properties of a member (i.e. area, second moment of
area, section modulus, radius of gyration and the like) are evaluated using conventional
methods of structural design. These properties are based upon either full cross-section
dimensions, effective widths or net section, as applicable.
For the design of tension members, the net section is employed when calculating the
nominal tensile strength of the axially loaded tension members.
For flexural members and axially loaded compression members, both full and effective
dimensions are used to calculate sectional properties. The full dimensions are used when
calculating the buckling load or moment, while the effective dimensions are used to
calculate the section and member capacities. For deflection calculation, the effective
dimension should be determined for the compressive stress in the element corresponding
to the service load. Peköz (1986a and 1986b) discusses this concept in more detail.

C3.2 MEMBERS SUBJECT TO TENSION There is very limited amount of data


regarding the capacity of cold-formed steel tension members. The provisions in Clause 3.2
of the Standard are the same as those in Clauses 7.2 and 7.3 of AS 4100/NZS 3404. The
Commentary to AS 4100 (AS 4100 Supp 1 — 1990) or to NZS 3404
(NZS 3404:Part 2:1997) should, therefore, be referred to for the explanation of
Clause 3.2.

C3.3 MEMBERS SUBJECT TO BENDING For the design of cold-formed steel


flexural members, consideration should be given to the following:
(a) Bending strength and deflection.
(b) Shear strength of webs and combined bending and shear.
Accessed by BHP BILLITON on 27 Jun 2002

(c) Web crippling strength and combined bending and web crippling.
(d) Bracing requirements.
For some cases, special consideration should also be given to shear lag and flange curling
due to the use of thin material. The design provisions for Items (a), (b) and (c) are
provided in Clause 3.3 of the Standard, while the requirements for lateral bracing are
specified in Clause 4.3. The treatments for flange curling and shear lag were discussed in
Clauses C2.1.3.2 and C2.1.3.3, respectively.

COPYRIGHT
AS/NZS 4600 Supp1:1998 48

C3.3.1 Bending moment Bending strengths of flexural members are differentiated


according to whether or not the member is laterally braced or the compression flange is
torsionally braced. If such members are laterally supported, then they are proportioned in
accordance with the nominal section moment capacity (see Clause 3.3.2 of the Standard).
If they are laterally unbraced, then the limit state is lateral buckling (see Clause 3.3.3.2 of
the Standard) or lateral-distortional buckling (see Clause 3.3.3.3(b) of the Standard). If the
compression flange is torsionally unbraced but the member is laterally braced, then the
limit state may be flange distortional buckling (see Clause 3.3.3.3(a) of the Standard). For
channel or Z-sections with the tension flange attached to deck or sheeting and with the
compression flange laterally unbraced, the bending capacity is less than that of a fully
braced member but greater than that of an unbraced member (see Clause 3.3.3.4 of the
Standard). The governing nominal bending capacity is the smallest of the values
determined from the applicable conditions.
C3.3.2 Nominal section moment capacity Clause 3.3.2 of the Standard includes two
design procedures for calculating the nominal section moment capacity. The first
procedure in Clause 3.3.2.2 of the Standard is based on initiation of yielding and the
second procedure in Clause 3.3.2.3 of the Standard is based on inelastic reserve capacity.
C3.3.2.2 Based on initiation of yielding In Clause 3.3.2.2 of the Standard, the nominal
section moment capacity (Ms) is the effective yield moment determined on the basis of the
effective areas of flanges and the beam webs. The effective width of the compression
flange and the effective depth of the webs can be calculated from the design equations
given in Section 2 of the Standard.
Similar to the design of hot-rolled steel shapes, the yield moment of a cold-formed steel
beam is defined as the moment at which an outer fibre (tension, compression, or both)
first attains the yield point of the steel. This is the maximum bending capacity to be used
in elastic design. Figure C3.3.2.2 shows several types of stress distributions for yield
moment based on different locations of the neutral axis. For balanced sections (see
Figure C3.3.2.2(a)), the outer fibres in the compression and tension flanges reach the yield
point at the same time. However, if the neutral axis is eccentrically located, as shown in
Figures C3.3.2.2(b) and (c), the initial yielding takes place in the tension flange for
Case (b) and in the compression flange for Case (c).
Accordingly, the nominal section strength for initiation of yielding is calculated using
Equation C3.3.2.2 as follows:
Ms = Ze fy . . . C3.3.2.2
where
Ze = elastic section modulus of the effective section calculated with the extreme
compression or tension fibre at f y
fy = design yield stress
For cold-formed steel design, Ze is usually calculated by using one of the following:
(a) If the neutral axis is closer to the tension than to the compression flange, the
Accessed by BHP BILLITON on 27 Jun 2002

maximum stress occurs in the compression flange and, therefore, the plate
slenderness ratio (λ) and the effective width of the compression flange are
determined using the b/t ratio and f* equal to f y. This procedure is also applicable to
those beams for which the neutral axis is located at the mid-depth of the section.
(b) If the neutral axis is closer to the compression than to the tension flange, the
maximum stress of fy occurs in the tension flange. The stress in the compression
flange depends on the location of the neutral axis, which is determined by the
effective area of the section. The latter cannot be determined unless the compressive
stress is known. A closed-form solution of this type of design is possible but would
be a very tedious and complex procedure. It is therefore customary to determine the
sectional properties of the section by iteration.

COPYRIGHT
49 AS/NZS 4600 Supp1:1998

For determining the design section moment capacity (φbMs), slightly different capacity
factors are used for the sections with stiffened or partially stiffened compression flanges
and the sections with unstiffened compression flanges. These φ b values were derived from
the test results and a dead-to-live load ratio of 1/5. They provide β values ranging from
2.53 to 4.05 (AISI, 1991; Hsiao, Yu and Galambos, 1988).

FIGURE C3.3.2.2 STRESS DISTRIBUTION FOR YIELD MOMENT

C3.3.2.3 Based on inelastic reserve capacity Prior to 1980, the inelastic reserve
capacity of beams was not included in the AISI Specification because most cold-formed
Accessed by BHP BILLITON on 27 Jun 2002

steel shapes have large width-to-thickness ratios which are considerably in excess of the
limits required by plastic design. Inelastic reserve capacity was introduced in
AS 1538 — 1988.
In the 1970s and early 1980s, research work on the inelastic strength of cold-formed steel
beams was carried out by Reck, Peköz, Winter, and Yener at Cornell University (Reck,
Peköz and Winter, 1975; Yener and Peköz, 1985a, 1985b). These studies showed that the
inelastic reserve strength of cold-formed steel beams due to partial plastification of the
cross-section and the moment redistribution of statically indeterminate beams can be
significant for certain practical shapes. With proper care, this reserve strength can be
utilized to achieve more economical design of such members.

COPYRIGHT
AS/NZS 4600 Supp1:1998 50

In order to utilize the available inelastic reserve strength of certain cold-formed steel
beams, design provisions based on the partial plastification of the cross-section were
added in the 1980 edition of the AISI Specification and the 1988 edition of AS 1538. The
same provisions are retained in the 1996 edition of the Standard. In accordance with
Clause 3.3.2.3 of the Standard, the nominal section moment capacity (M s) of those beams
satisfying certain specific limitations can be determined on the basis of the inelastic
reserve capacity with a limit of 1.25(My)e, where (M y)e is the effective yield moment. The
ratio of M s/(My)e represents the inelastic reserve strength of a beam cross-section.
The nominal section moment capacity (Ms) is the maximum bending capacity of the beam
by considering the inelastic reserve strength through partial plastification of the
cross-section. The inelastic stress distribution in the cross-section depends on the
maximum strain in the compression flange (εeu). Based on the Cornell research work on
hat sections having stiffened compression flanges (Reck, Peköz and Winter, 1975), the
design provision limits the maximum compression strain to be Cyey, where Cy is a
compression strain factor determined using the equations provided in Clause 3.3.2.3 of the
Standard, as shown in Figure C3.3.2.3.

FIGURE C3.3.2.3 FACTOR Cy FOR STIFFENED COMPRESSION ELEMENTS


WITHOUT INTERMEDIATE STIFFENERS

On the basis of the maximum compression strain (C yey) permitted in the Standard, the
Accessed by BHP BILLITON on 27 Jun 2002

neutral axis can be located using Equation C3.3.2.3(1) and the nominal moment (Ms) can
be determined using Equation C3.3.2.3(2) as follows:
. . . C3.3.2.3(1)

. . . C3.3.2.3(2)
where
σ = the stress in the cross-section.

COPYRIGHT
51 AS/NZS 4600 Supp1:1998

C3.3.3 Nominal member moment capacity


C3.3.3.2 Members subject to lateral buckling The bending capacity of flexural
members is not only governed by the strength of the cross-section, but is also limited by
the lateral buckling strength of the member if lateral and torsional braces are not
adequately provided. The design provisions for determining the nominal member moment
capacity of laterally unbraced segments are specified in Clause 3.3.3.2 of the Standard.
Clause 3.3.3.2(a) is based on the AISI Specification (1996) and Clause 3.3.3.2(b) of the
Standard is based on AS 1538 — 1988 converted to limit states format.
If an equal-flanged I-beam is laterally unbraced, it may fail in lateral (flexural-torsional)
buckling. In the elastic range, the elastic buckling moment (M o) can be determined using
Equation C3.3.3.2(1) as follows:

. . . C3.3.3.2(1)
where
E = modulus of elasticity
G = shear modulus
Iy = second moment of area about the y-axis
Iw = warping constant of torsion
J = St. Venant torsion constant
l = unbraced length
Equation C3.3.3.2(1) is given as a function of f oy and foz by Equation 3.3.3.2(7) in the
Standard except that the C b factor discussed later is included in the numerator. The terms
(foy) and (foz) are the elastic buckling stresses calculated using Equations 3.3.3.2(11)
and 3.3.3.2(12) in the Standard.
In Equation C3.3.3.2(1), the first term under the square root represents the strength due to
the St. Venant torsional rigidity, and the second term represents the strength due to the
warping rigidity. For thin-walled cold-formed steel sections, the second term usually
exceeds the first term considerably and so Equation C3.3.3.2(1) can be simplified as
follows:

. . . C3.3.3.2(2)

Equation C3.3.3.2(2) is given as Equation 3.3.3.2(21) in the Standard except that the Cm
factor discussed later is included in the denominator.

Equation C3.3.3.2(2) can be simplified by considering as follows:

. . . C3.3.3.2(3)
Accessed by BHP BILLITON on 27 Jun 2002

where
Iyc = second moment of area of the compression portion of the section.
Equation C3.3.3.2(3) was derived on the basis of a uniform bending moment and is
conservative for other cases. For this reason M o may be modified by multiplying the right-
hand side by a bending coefficient (Cb) as follows:

. . .C3.3.3.2(4)

COPYRIGHT
AS/NZS 4600 Supp1:1998 52

where Cb is the bending coefficient, which can conservatively be taken as unity, or


calculated as follows:

. . . C3.3.3.2(5)
where
Mmax. = absolute value of maximum moment in the unbraced segment
M3 = absolute value of moment at quarter point of unbraced segment
M4 = absolute value of moment at centerline of unbraced segment
M5 = absolute value of moment at three-quarter point of unbraced segment
Equation C3.3.3.2(5) was derived by Kirby and Nethercot (1979) and can be used for
various shapes of moment diagrams within the unbraced length under consideration
(equivalent to segment in AS 4100 and in NZS 3404). It gives accurate solutions for
fixed-end beams and moment diagrams that are not straight lines. Equation C3.3.3.2(5) is
the same as that being used in the AISC LRFD Specification (AISC, 1993).
Based on the elastic critical buckling stress of an I-section calculated using
Equation C3.3.3.2(4), the simplified elastic buckling moment for lateral buckling of point
symmetric Z-sections can be determined using Equation C3.3.3.2(6)
(i.e. Equation 3.3.3.2(16) in the Standard), which is simply derived by factoring
Equation C3.3.3.2(4) by ½ to allow for the inclined principal axes of the Z-section, as
follows:

. . . C3.3.3.2(6)

Equation C3.3.3.2(2) applies only to elastic buckling of cold-formed steel beams when the
calculated theoretical buckling stress is less than or equal to the proportional limit. When
the calculated stress exceeds the proportional limit, the beam behaviour will be governed
by inelastic buckling. The inelastic buckling stress can be calculated using
Equation C3.3.3.2(7) (Yu, 1991) as follows:

. . . C3.3.3.2(7)

Consequently, Equation C3.3.3.2(8) (i.e. Equation 3.3.3.2(3) in the Standard) can be used
to determine the inelastic critical moment for lateral buckling as follows:

. . . C3.3.3.2(8)
where
My = the moment causing initial yield at the entrance compression fibre of the full
section and equals Z f fy
The elastic and inelastic buckling moments for lateral buckling strength are shown in
Accessed by BHP BILLITON on 27 Jun 2002

Figure C3.3.3.2(1) (Yu, 1991). As specified in Clause 3.3.3.2(a) of the Standard, buckling
is considered to be elastic up to a moment equal to 0.56 My. The inelastic region is
defined by a Johnson parabola from 0.56 M y to (10/9) My at an unsupported length of
zero. The (10/9) factor is based on the partial plastification of the section in bending
(Galambos, 1963). A flat plateau is created by limiting the maximum moment to M y which
enables the calculation of the maximum unsupported length for which there is no moment
reduction due to lateral instability. This maximum unsupported length can be calculated
by setting My equal to the Johnson parabola. This inelastic lateral buckling curve for
singly-, doubly- and point-symmetric sections has been confirmed by research in beam-
columns (Peköz and Sumer, 1992) and wall studs (Kian and Peköz, 1994).

COPYRIGHT
53 AS/NZS 4600 Supp1:1998

FIGURE C3.3.3.2(1) ELASTIC AND INELASTIC CRITICAL MOMENTS


FOR LATERAL BUCKLING STRENGTH

In the Standard, the design equations from AS 1538 — 1988 were converted to limit states
format and included in Clause 3.3.3.2(b) of the Standard. In particular, the strength
Equations 3.3.3.2(17), 3.3.3.2(18) and 3.3.3.2(19) in the Standard were derived directly
from Clause 3.3.2 of AS 1538 with stresses (Fb, FY and Fo) replaced by equivalent
moments (Mc, My and M o), and with the factor of safety of 1.67 removed.
The above discussion deals only with the lateral buckling strength of locally stable beams.
For locally unstable beams, the interaction of the local buckling of compression elements
and the overall lateral buckling of beams may result in a reduction of the lateral buckling
strength of the member. The effect of local buckling on critical moment is considered in
Clause 3.3.3.2 of the Standard, in which the nominal member moment capacity is
determined as follows:
. . . C3.3.3.2(9)
where
Mc = elastic or inelastic critical moment, whichever is applicable
Zc = elastic section modulus of the effective section calculated at a stress Mc/Zf in
the extreme compression fibre
Zf = elastic section modulus of the full unreduced section for the extreme
compression fibre
In Equation C3.3.3.2(9), the ratio of Z c/Zf represents the effect of local buckling on lateral
buckling strength of beams.
Accessed by BHP BILLITON on 27 Jun 2002

Using the above nominal member moment capacity with a capacity factor of φb equal to
0.90, values of β vary from 2.4 to 3.8.
Lateral buckling problems discussed above deal with the type of lateral buckling of
I-beams, channels and Z-shaped sections for which the entire cross-section rotates and
deflects in the lateral direction. This is not the case for U-shaped beams and the combined
sheet-stiffener sections shown in Figure C3.3.3.2(2). For this case, when the section is
loaded in such a manner that the flanges of stiffeners are in compression, the tension
flange of the beam remains straight and does not displace laterally; only the compression
flange tends to buckle separately in the lateral direction, accompanied by out-of-plane
bending of the web, as shown in Figure C3.3.3.2(3), unless adequate bracing is provided.

COPYRIGHT
AS/NZS 4600 Supp1:1998 54

The precise analysis of the lateral buckling of U-shaped beams is rather complex. The
compression flange and the compression portion of the web act not only like a column on
an elastic foundation, but the problem is also complicated by the weakening influence of
the torsional action of the flange. For this reason, the design procedure outlined in
Chapter C of Supplementary Information of the AISI Design Manual for determining the
allowable design strength for laterally unbraced compression flanges is based on the
considerable simplification of an analysis presented by Douty (1962).
In 1964, Haussler presented rigorous methods for determining the strength of elastically
stabilized beams (Haussler, 1964). In his methods, Haussler also treated the unbraced
compression flange as a column on an elastic foundation and maintained more rigour in
his development.
A comparison of Haussler’s method with Douty’s simplified method indicates that the
latter may provide a smaller critical stress.
An additional study of laterally unbraced compression flanges was recently made at
Cornell University (Serrette and Peköz, 1992, 1994 and 1995) and an analytical procedure
was developed for determining the distortional buckling strength of the standing seam
roof panel.

FIGURE C3.3.3.2(2) COMBINED SHEET-STIFFENER SECTIONS


Accessed by BHP BILLITON on 27 Jun 2002

FIGURE C3.3.3.2(3) LATERAL BUCKLING OF U-SHAPED BEAM

C3.3.3.3 Members subject to distortional buckling Recent research conducted by


Ellifritt, Sputo and Haynes (1992) has indicated that when the unbraced length is defined
as the spacing between intermediate braces, the equations used in Clause 3.3.3.2 of the
Standard may be conservative for cases where one mid-span brace is used, but may be
unconservative where more than one intermediate brace is used.

COPYRIGHT
55 AS/NZS 4600 Supp1:1998

The above-mentioned research (Ellifritt, Sputo, and Haynes, 1992) and the recent study of
Kavanagh and Ellifritt (1993 and 1994) have shown that a discretely braced beam, not
attached to deck and sheeting, may fail either by lateral-torsional buckling between
braces, or by distortional buckling at or near the braced point. The distortional buckling
strength of C-sections and Z-sections has recently been studied extensively at the
University of Sydney by Lau and Hancock (1987); Hancock, Kwon and Bernard (1994);
and Hancock (1997).
Two types of distortional buckling are specified in Clause 3.3.3.3 of the Standard. These
are flange distortional buckling, which involves rotation of a flange and lip about the
flange/web junction of a C-section or Z-section, as shown in Figure C3.3.3.3(a), and
lateral-distortional buckling, which involves transverse bending of a vertical web as
shown in Figure C3.3.3.3(b).
The flange-distortional mode is specified in Clause 3.3.3.3(a) of the Standard. The elastic
distortional buckling stress (fod) can be based on a rational elastic buckling analysis, such
as that given in Paragraph D3 of Appendix D of the Standard, which is based on Hancock
(1997). The equation for critical moment (M c) is developed in Hancock, Kwon and
Bernard (1994), and is further supported by testing (Hancock, Rogers and Schuster
(1996)). Flange distortional buckling is most likely to occur in the unsheeted compression
flanges of purlins under wind uplift when the purlins are braced to prevent twisting and
lateral buckling.
The lateral-distortional mode is specified in Clause 3.3.3.3(b) of the Standard. It is most
likely to occur in beams, such as the hollow flange beam shown in Figure C3.3.3.3(b),
where the high torsional rigidity of the tubular compression flange prevents it from
twisting during lateral displacement. Equations for the elastic distortional buckling stress
are given in Pi and Trahair (1997). The equation for critical moment (M c) is based on the
Johnston parabola (Galambos, 1988) and, at this stage, has not been supported by testing.
Accessed by BHP BILLITON on 27 Jun 2002

FIGURE C3.3.3.3 FLEXURAL DISTORTIONAL BUCKLING MODES

C3.3.3.4 Beams having one flange through-fastened to sheeting For beams having the
tension flange attached to deck or sheeting and the compression flange unbraced, e.g. a
roof purlin or wall girt subjected to outward wind pressure, the bending capacity is less
than a fully braced member, but greater than an unbraced member. This partial restraint is
a function of the rotational stiffness provided by the panel-to-purlin connection. The
Standard contains factors that represent the reduction in capacity from a fully braced
condition. The factors in the AISI Specification (1996) are based on experimental results

COPYRIGHT
AS/NZS 4600 Supp1:1998 56

obtained for both simple and continuous span purlins (Peköz and Soroushian, 1981 and
1982; LaBoube, 1986; Haussler and Pahers, 1973; LaBoube, et al., 1988; Haussler, 1988).
In the Standard the R factors were calibrated by Johnston and Hancock (1994), based on
testing in the vacuum test rig at the University of Sydney (Hancock, Celeban and Healy
(1993)).
As indicated by LaBoube (1986), the rotational stiffness of the panel-to-purlin connection
is primarily a function of the member thickness, sheet thickness, fastener type and
fastener location. For compressed glass fibre blanket insulation of initial thicknesses of
zero to six inches (152 mm), the rotational stiffness is not measurably affected (LaBoube,
1986). To ensure adequate rotational stiffness of the roof and wall systems designed in
accordance with the Standard, Clause 3.3.3.4 of the Standard explicitly states the
acceptable panel and fastener types. In Australia, cleats were used in the vacuum rig tests,
and these are specified clearly in Clause 3.3.3.4 of the Standard. Also in Australia, no
insulation was used between the purlin and sheeting.
Continuous beam tests were made in Australia on three equal spans under both wind uplift
and downwards loading and the R values were calculated from the failure loads assuming
a bending moment distribution based on double the second moment of area in the lapped
regions.
The provisions of Clause 3.3.3.4 of the Standard apply to beams for which the tension
flange is attached to sheeting by screw fasteners and the compression flange is unbraced
or has bridging which effectively prevents lateral and torsional deformation at support
points.
In the case of simply supported purlins without bridging, considerable twisting of the
purlins occurred which caused the screw heads to pull through the sheeting when
load-spreading (cyclone) washers were not used. The AISI R-factor values are
considerably lower than those in the Standard and can be used in the case when
load-spreading (cyclone) washers are not used.
The R-factor method cannot be used with clip fasteners of any form including sliding
clips unless testing validates the restraint provided by the clips is adequate to prevent
torsional deformation and allow the R factors in Clause 3.3.3.4 to be used.
For the limit states method, the use of the reduced nominal member moment capacity with
a capacity factor of φb equal to 0.90, Equation 3.3.3.4 of the Standard provides β values
varying from 2.5 to 2.7, which are satisfactory for the target value of 2.5. This analysis
was based on the load combination of (Wu − 0.8G) (Rousch and Hancock (1996a, 1996b)).
C3.3.4 Shear The shear strength of beam webs is governed by either yielding or
buckling, depending on the d1/t ratio and the mechanical properties of steel. For beam
webs having small d1/t ratios, the nominal shear capacity is governed by shear yielding
and can be calculated as follows:

. . . C3.3.4(1)
Accessed by BHP BILLITON on 27 Jun 2002

where
Aw = area of the beam web computed by d 1tw

τ y = yield point of steel in shear, which can be calculated by

COPYRIGHT
57 AS/NZS 4600 Supp1:1998

For beam webs having large d1/tw ratios, the nominal shear capacity is governed by elastic
shear buckling and can be calculated as follows:

. . . C3.3.4(2)

where
τ cr = critical shear buckling stress in the elastic range
kv = shear buckling coefficient
E = modulus of elasticity
ν = Poisson’s ratio
d1 = web depth
tw = web thickness
By using ν equal to 0.3, the shear strength (Vv) can be calculated as follows:

. . . C3.3.4(3)

For beam webs having moderate d1/t ratios, the nominal shear strength is based on
inelastic shear buckling and can be calculated as follows:

. . . C3.3.4(4)
The provisions in the Standard are applicable for the design of webs of beams and decks
either with or without transverse web stiffeners.
The nominal strength equations of Clause 3.3.4 of the Standard are similar to the nominal
shear strength equations given in the AISI LRFD Specification (AISI, 1991). The
acceptance of these nominal strength equations for cold-formed steel sections was
considered in the study summarized by LaBoube and Yu (1978a).
Previous editions of the ASD Specification (AISI, 1986) and AS 1538 — 1988 employed
three different factors of safety when evaluating the permissible shear strength of an
unreinforced web because it was intended to use the same permissible values for the AISI
and AISC specifications (i.e. 1.44 for yielding, 1.67 for inelastic buckling and 1.71 for
elastic buckling). For the limit states design method in the Standard, the φv factors used in
Clause 3.3.4 of the Standard were derived from the condition that the nominal capacities
for the limit states design method and the permissible stress method in AS 1538 — 1988
are the same, because the appropriate test data on shear were not available (Hsiao, Yu and
Galambos, 1988a; AISI, 1991).
Accessed by BHP BILLITON on 27 Jun 2002

C3.3.5 Combined bending and shear For cantilever beams and continuous beams,
high bending stresses often combine with high shear stresses at the supports. Such beam
webs should be safeguarded against buckling due to the combination of bending and shear
stresses.
For disjointed flat rectangular plates, the critical combination of bending and shear
stresses can be approximated by the following equation (Bleich, 1952):

. . . C3.3.5(1)

COPYRIGHT
AS/NZS 4600 Supp1:1998 58

where
fb = actual compressive bending stress
fcr = theoretical buckling stress in pure bending
τ = actual shear stress
τ cr = theoretical buckling stress in pure shear
Equation C3.3.5(1) was found to be conservative for beam webs with adequate transverse
stiffeners, for which a diagonal tension field action can be developed. Based on the
studies made by LaBoube and Yu (1978a), the following equation was developed for
beam webs with transverse stiffeners satisfying the requirements of Clause 2.7 of the
Standard:

. . . C3.3.5(2)

Equation C3.3.5(2) was added to the AISI Specification in 1980 and to AS 1538 in 1988.
The correlations between Equation C3.3.5(2) and the test results of beam webs having a
diagonal tension field action are shown in Figure C3.3.5.
For the limit states design method, the Equation in Clause 3.3.5 of the Standard for
combined bending and shear are based on Equation C3.3.5(1) and Equation C3.3.5(2) by
using the nominal section moment capacity and nominal shear capacity.
Accessed by BHP BILLITON on 27 Jun 2002

NOTE: Solid symbols represent test specimens without additional sheets on top and bottom flanges.

FIGURE C3.3.5 INTERACTION DIAGRAM FOR τ/τ max. AND f b/ f b max.

COPYRIGHT
59 AS/NZS 4600 Supp1:1998

C3.3.6 Bearing For cold-formed steel beams, transverse and shear stiffeners are not
frequently used. The webs of beams may cripple due to the high local intensity of the load
or reaction. Figure C3.3.6(1) shows the types of failure caused by web crippling
of unreinforced single webs (see Figure C3.3.6(1)(a)) and of I-beams (see
Figure C3.3.6(1)(b)).

FIGURE C3.3.6(1) WEB CRIPPLING OF COLD-FORMED STEEL BEAMS

In the past, the buckling problem of separate flat rectangular plates and web crippling
behaviour of cold-formed steel beam webs under locally distributed edge forces have been
studied by numerous investigators (Yu, 1991) and it was found that the theoretical
analysis of web crippling for cold-formed steel flexural members is rather complicated
because it involves the following:
(a) Non-uniform stress distribution under the applied load and adjacent portions of the
web.
(b) Elastic and inelastic stability of the web element.
(c) Local yielding in the immediate region of load application.
(d) Bending produced by eccentric load (or reaction) when it is applied on the bearing
flange at a distance beyond the curved transition of the web.
(e) Initial out-of-plane imperfection of plate elements.
(f) Various edge restraints provided by beam flanges and interaction between flange
and web elements.
(g) Inclined webs for decks and panels.
For these reasons, the present AISI design provisions for web crippling and the
AS/NZS 4600 design provisions for bearing are based on the extensive experimental
investigations conducted at Cornell University by Winter and Pian (Winter and Pian,
1946), and by Zetlin (Zetlin, 1955) in the 1940s and 1950s, and at the University of
Accessed by BHP BILLITON on 27 Jun 2002

Missouri-Rolla by Hetrakul and Yu (Hetrakul and Yu, 1978). In these experimental


investigations, the web crippling tests were carried out under the following four loading
conditions for beams having single unreinforced webs and I-beams:
(i) End one-flange (EOF) loading.
(ii) Interior one-flange (IOF) loading.
(iii) End two-flange (ETF) loading.
(iv) Interior two-flange (ITF) loading.
All loading conditions are shown in Figure C3.3.6(2). In Figures (a) and (b), the distances
between bearing plates were kept to no less than 1.5 times the web depth in order to avoid
the two-flange loading action.

COPYRIGHT
AS/NZS 4600 Supp1:1998 60

FIGURE C3.3.6(2) LOADING CONDITIONS FOR WEB CRIPPLING TESTS

Clause 3.3.6 of the Standard provides design equations to determine the web crippling
strength of flexural members having flat single webs (channels, Z-sections, hat sections,
tubular members, roof deck, floor deck and the like) and I-beams (made of two channels
connected back to back, by welding two angles to a channel, or by connecting three
channels). Different design equations are used for various loading conditions, as shown in
Figure C3.3.6(3), and Tables 3.3.6(1) and 3.3.6(2) in the Standard. These design equations
are based on experimental evidence (Winter, 1970; Hetrakul and Yu, 1978) and the
assumed distributions of loads or reactions into the web as shown in Figure C3.3.6(4).
The assumed distributions of loads or reactions into the web, as shown in
Figure C3.3.6(4), are independent of the flexural response of the beam. Due to flexure,
the point of bearing will vary relative to the plane of bearing resulting in non-uniform
bearing load distribution into the web. The value of R b will vary because of a transition
from the interior one-flange loading (see Figure C3.3.6(4)(b)) to the end one-flange
loading condition (see Figure C3.3.6(4)(a)). These discrete conditions represent the
experimental basis on which the design provisions were founded (Winter, 1970; Hetrakul
and Yu, 1978).
From Tables 3.2.6(1) and 3.3.6(2) in the Standard, it can be seen that the nominal
capacity for concentrated load or reaction of cold-formed steel beams depends on the
Accessed by BHP BILLITON on 27 Jun 2002

ratios of d1/tw, lb/tw, ri/tw, the web thickness (t(tw)), the yield stress ( fy), and the web
inclination angle (θ).
With regard to the limit states design approach, the use of φ w equal to 0.75 for single
unreinforced webs and φw equal to 0.80 for I-sections provide values of safety index
ranging from 2.4 to 3.8.
Recent research indicated that a Z-section having its end support flange bolted to the
section’s supporting member through two 12.7 mm diameter bolts will experience an
increase in end-one-flange web crippling capacity (Bhakta, LaBoube and Yu, 1992; Cain,
LaBoube and Yu, 1995). The increase in load-carrying capacity was shown to range from
27 to 55% for the sections under the limitations prescribed in the Standard. A lower
bound value of 30% increase is permitted in Clause 3.3.6 of the Standard.

COPYRIGHT
61 AS/NZS 4600 Supp1:1998
Accessed by BHP BILLITON on 27 Jun 2002

FIGURE C3.3.6(3) APPLICATION OF DESIGN EQUATIONS


GIVEN IN TABLES 3.3.6(1) AND 3.3.6(2) OF THE STANDARD

COPYRIGHT
AS/NZS 4600 Supp1:1998 62
Accessed by BHP BILLITON on 27 Jun 2002

FIGURE C3.3.6(4) ASSUMED DISTRIBUTION OF REACTION OR LOAD

COPYRIGHT
63 AS/NZS 4600 Supp1:1998

C3.3.7 Combined bending and bearing Clause 3.3.7 of the Standard contains
interactions for the combination of bending and bearing. Clauses 3.3.7(a) and 3.3.7(b) of
the Standard are based on the studies conducted at the University of Missouri-Rolla for
the effect of bending on the reduction of web crippling loads with the applicable capacity
factors used for bending and bearing (Hetrakul and Yu, 1978 and 1980; Yu, 1981 and
1991). Figures C3.3.7(1) and C3.3.7(2) show the correlations between the interactions and
test results. For embossed webs, bearing should be determined by tests in accordance with
Section 6 of the Standard.
The exception included in Clause 3.3.7 of the Standard for single unreinforced webs
applies to the interior supports of continuous spans using decks and beams, as shown in
Figure C3.3.7(3). Results of continuous beam tests of steel decks (Yu, 1981) and several
independent studies by manufacturers indicate that, for these types of members, the
post-buckling behaviour of webs at interior supports differs from the type of failure mode
occurring under concentrated loads on single span beams. This post-buckling strength
enables the member to redistribute the moments in continuous spans. For this reason, the
interaction in Clause 3.3.7(a) of the Standard is not applicable to the interaction between
bending and the reaction at interior supports of continuous spans. This exception applies
only to the members shown in Figure C3.3.7(3) and similar situations explicitly described
in Clause 3.3.7 of the Standard.
The exception should be interpreted to mean that the effects of combined bending and
bearing need not be checked for determining load-carrying capacity. Furthermore, the
positive bending resistance of the beam should be at least 90% of the negative bending
resistance in order to ensure the safety specified by Clause 3.3.7 of the Standard. Using
this procedure, serviceability loads may —
(a) produce slight deformations in the beam over the support;
(b) increase the actual compressive bending stresses over the support to as high as
0.8 fy; and
(c) result in additional bending deflection of up to 22% due to elastic moment
redistribution.
If load-carrying capacity is not the primary design concern because of the above
behaviour, the designer is urged to ignore the exception in Clause 3.3.7(a) of the
Standard.
With regard to Clause 3.3.7(b) of the Standard, previous tests indicate that when the d 1/tw
ratio of an I-beam web does not exceed and when , the bending
moment has little or no effect on the web crippling load (Yu, 1991). For this reason, the
permissible reaction or concentrated load can be determined by the equations given in
Clause 3.3.6 of the Standard without reduction for the presence of bending.
In the development of the limit states equations, a total of 551 tests were calibrated for
combined bending and bearing. Based on φw equal to 0.75 for single unreinforced webs
and φw equal to 0.80 for I-sections, the values of safety index vary from 2.5 to 3.3.
Accessed by BHP BILLITON on 27 Jun 2002

COPYRIGHT
AS/NZS 4600 Supp1:1998 64

FIGURE C3.3.7(1) GRAPHIC PRESENTATION FOR WEB CRIPPLING AND COMBINED


WEB CRIPPLING AND BENDING FOR SINGLE UNREINFORCED WEBS
Accessed by BHP BILLITON on 27 Jun 2002

FIGURE C3.3.7(2) INTERACTION BETWEEN WEB CRIPPLING AND BENDING FOR


I-BEAMS HAVING UNREINFORCED WEBS

COPYRIGHT
65 AS/NZS 4600 Supp1:1998

FIGURE C3.3.7(3) SECTIONS USED FOR EXCEPTION OF CLAUSE 3.3.7(a)


OF THE STANDARD

C3.4 CONCENTRICALLY LOADED COMPRESSION MEMBERS Depending on


the configuration of the cross-section, thickness of material, unbraced length, and end
restraint, axially loaded compression members should be designed for the following
ultimate limit state conditions:
(a) Yielding It is well known that a very short, compact column under an axial load
may fail by yielding. The yield load is determined by the following equation:
. . . C3.4(1)
where
Ag = gross area of the column
fy = yield stress of steel
(b) Overall column buckling (flexural buckling, torsional buckling or flexural-torsional
buckling)
(i) Flexural buckling of columns
(A) Elastic buckling stress A slender, axially loaded column may fail by
overall flexural buckling if the cross-section of the column is a
doubly-symmetric shape, closed shape (square or rectangular tube),
cylindrical shape, or point-symmetric shape. For singly-symmetric
Accessed by BHP BILLITON on 27 Jun 2002

shapes, flexural buckling is one of the possible failure modes. Wall


studs connected with sheeting material can also fail by flexural
buckling.
The elastic critical buckling load for a long column can be determined
by the following Euler formula:

. . . C3.4(2)

COPYRIGHT
AS/NZS 4600 Supp1:1998 66

where
Noc = column buckling load in the elastic range
E = modulus of elasticity
I = second moment of area
k = effective length factor
l = unbraced length
Accordingly, the elastic column buckling stress can be calculated as
follows:

. . . C3.4(3)

where
r = radius gyration of the full cross-section
le/r = the effective slenderness ratio
(B) Inelastic buckling stress When the elastic column buckling stress
calculated using Equation C3.4(3) exceeds the proportional limit (fpr),
the column will buckle in the inelastic range. Prior to 1996, the
following equation was used in the AISI Specification for calculating
the inelastic column buckling stress:

. . . C3.4(4)

It should be noted that because Equation C3.4(4) is based on the


assumption that fpr is equal to fy/2, it is applicable only for (foc)E greater
than or equal to fy/2.
By using λc as the column slenderness parameter instead of the
slenderness ratio (le/r), Equation C3.4(4) can be rewritten as follows:

. . . C3.4(5)

where

. . . C3.4(6)

Accordingly, Equation C3.4(5) is applicable only for λ c less than or


equal to .
Accessed by BHP BILLITON on 27 Jun 2002

In AS 1538 — 1988, the Perry curve (Ayrton and Perry (1986)) was
used to define the column strength since geometric imperfections were
included in the column design philosophy used in Australia. The Perry
curve gives a lower column strength than Equation 3.4(5) in the
Standard due to the inclusion of imperfections.

COPYRIGHT
67 AS/NZS 4600 Supp1:1998

(C) Design compressive axial force for locally stable columns If the
individual components of compression members have small b/t ratios,
local buckling will not occur before the compressive stress reaches the
column buckling stress and the yield stress of steel. Therefore, the
nominal axial strength can be determined as follows:
. . . C3.4(7)
where
Nc = nominal member capacity of the member in
compression
Ag = gross area of the column
= column buckling stress (elastic or inelastic as
appropriate)
(D) Design compressive axial force for locally unstable columns For
cold-formed steel compression members with large b/t ratios, local
buckling of individual component plates may occur before the applied
load reaches the nominal axial strength determined by
Equation C3.4(7). The interaction effect of the local and overall
column buckling may result in a reduction of the overall column
strength. From 1946 through 1986, the effect of local buckling on
column strength was considered in the AISI Specification and
AS 1538 — 1988 by using a form factor (Q) in the determination of the
permissible stress for the design of axially loaded compression
members (Winter, 1970; Yu, 1991). Even though the Q-factor method
was used successfully for the design of cold-formed steel compression
members, research work conducted at Cornell University and other
institutions has shown that this method is capable of improvement. On
the basis of the test results and analytical studies of DeWolf, Peköz,
Winter, and Mulligan (DeWolf, Peköz and Winter, 1974; Mulligan and
Peköz, 1984) and Peköz’s development of a unified approach for the
design of cold-formed steel members (Peköz, 1986b), the Q-factor
method was eliminated in the 1986 edition of the AISI Specification.
In order to reflect the effect of local buckling on the reduction of
column strength, the nominal axial strength is determined by the
critical column buckling stress and the effective area (Ae) instead of the
full sectional area. For a more in depth discussion of the background
for these provisions, see Peköz (1986b). Therefore, the nominal
member capacity of cold-formed steel compression members can be
determined by the following equation:
. . . C3.4(8)
Accessed by BHP BILLITON on 27 Jun 2002

where is the column buckling stress (elastic or inelastic as


appropriate).
An exception for Equation C3.4(7) is for C-shapes and Z-shapes, and
single-angle sections with unstiffened flanges. For these cases, the
nominal axial strength is also limited by the following capacity, which
is determined by the local buckling stress of the unstiffened element
and the area of the full cross-section as follows:

. . . C3.4(9)

COPYRIGHT
AS/NZS 4600 Supp1:1998 68

Equation C3.4(9) was included in Section C4(b) of the 1986 edition of


the AISI Specification when the unified design approach was adopted.
A recent study conducted by Rasmussen at the University of Sydney
(Rasmussen, 1994) indicated that the design provisions of
Section C4(b) of the 1986 AISI Specification lead to unnecessarily and
excessively conservative results. This conclusion was based on the
analytical studies carefully validated against test results as reported by
Rasmussen and Hancock (1992). Consequently, Section C4(b) of the
AISI Specification (Equation C-C4(9)) was deleted in 1996 and is not
included in the Standard.
In the AISI Specification, the design equations for calculating the
critical stress have been changed from those given by
Equations C3.4(3) and C3.4(4) to those used in the AISC LRFD
Specification (AISC, 1993). As given in Clause 3.4.1 of the Standard,
these design regulations are as follows:
. . . C3.4(10)

. . . C3.4(11)

where
fn = critical stress which depends on the value of
λc f y / ( fo c )E
( fo c )E = elastic flexural buckling stress calculated using
Equation C3.4(3).
Consequently, the equation for determining the nominal axial strength
can be written as follows:
. . . C3.4(12)
which is Equation 3.4.2(2) of the Standard with foc equal to ( fo c )E .
The reasons for changing the design equations from Equation C3.4(4)
to Equation C3.4(10) for inelastic buckling stress and from
Equation C3.4(3) to Equation C3.4(11) for elastic buckling stress are as
follows:
(1) The revised column design equations (Equations C3.4(10)
and C3.4(11) were shown to be more accurate by Peköz and
Sumer (1992). In this study, 299 test results on columns and
beam-columns were evaluated. The test specimens included
members with component elements in the post-local buckling
range as well as those that were locally stable. The test specimens
Accessed by BHP BILLITON on 27 Jun 2002

included members subject to flexural buckling as well as flexural-


torsional buckling.
(2) Because the revised column design equations represent the
maximum strength with due consideration given to initial
crookedness and can provide a better fit to test results, the
required factor of safety can be reduced. In addition, the revised
equations enable the use of a single capacity factor for all λc
values even though the nominal member capacity of columns
decreases as the slenderness increases because of initial out-of-
straightness.

COPYRIGHT
69 AS/NZS 4600 Supp1:1998

The design provisions included in the AISI Specification (AISI,


1986), the AISI-LRFD Specification (AISI, 1991), the combined
ASD/LRFD Specification (AISI, 1996) and the Standard are
shown in Figure C3.4(1).

FIGURE C3.4(1) COMPARISON BETWEEN THE CRITICAL BUCKLING STRESS


EQUATIONS

(E) Effective length factor (k) The effective length factor (k) accounts for
the influence of restraint against rotation and translation at the ends of
a column on its load-carrying capacity. For the simplest case, a column
with both ends hinged and braced against lateral translation, buckling
occurs in a single half-wave and the effective length (le) is equal to kl
being the length of this half-wave, is equal to the actual physical
length of the column (see Figure C3.4(2)); correspondingly, for this
case, k is equal to 1. This situation is approached if a given
compression member is part of a structure that is braced in such a
Accessed by BHP BILLITON on 27 Jun 2002

manner that no lateral translation (sidesway) of one end of the column


relative to the other can occur. This is so for columns or studs in a
structure with diagonal bracing, diaphragm bracing, shear-wall
construction or any other provision that prevents horizontal
displacement of the upper relative to the lower column ends. In these
situations, it is safe and only slightly, if at all, conservative to take k
equal to 1.

COPYRIGHT
AS/NZS 4600 Supp1:1998 70

FIGURE C3.4(2) OVERALL COLUMN BUCKLING

If translation is prevented and abutting members (including


foundations) at one or both ends of the member are rigidly connected
to the column in a manner that provides substantial restraint against
rotation, k values smaller than 1 are sometimes justified. Table C3.4
gives the theoretical k values for six idealized conditions in which joint
rotation and translation are either fully realized or non-existent.
Table C3.4 also includes the k values recommended by the Structural
Stability Research Council for design use (Galambos, 1988).
In trusses, the intersection of members provides rotational restraint to
the compression members at service loads. As the collapse load is
approached, the member stresses approach the yield stress, which
greatly reduces the restraint they can provide. For this reason, the k
value is usually taken as unity regardless of whether they are welded,
bolted, or connected by screws. However, when sheeting is attached
directly to the top flange of a continuous compression chord, recent
research (Harper, LaBoube and Yu, 1995) has shown that the k values
may be taken as 0.75 (AISI, 1995).
On the other hand, when no lateral bracing against sidesway is present,
such as in the portal frame shown in Figure C3.4(3), the structure
depends on its own bending stiffness for lateral stability. In this case,
if buckling of the columns were to occur, it invariably takes place by
Accessed by BHP BILLITON on 27 Jun 2002

the sidesway motion shown. This occurs at a lower load than the
columns would be able to carry if they were braced against sidesway
and the figure shows that the half-wave length into which the columns
buckle is longer than the actual column length. Hence, in this case, k is
larger than 1 and its value can be read from the graph shown in
Figure C3.4(4) (Winter, et al. 1948a and 1970). Since column bases are
rarely either actually hinged or completely fixed, k values between the
two curves should be estimated depending on actual base fixity.

COPYRIGHT
71 AS/NZS 4600 Supp1:1998

TABLE C3.4
EFFECTIVE LENGTH FACTORS (k) FOR CONCENTRICALLY LOADED
COMPRESSION MEMBERS

(a) (b) (c) (d) (e) (f)

Buckled shape of column


is shown by dashed line

Theoretical k value 0.5 0.7 1.0 1.0 2.0 2.0


Recommended k value
when ideal conditions are 0.65 0.80 1.2 1.0 2.10 2.0
approximated
Rotation fixed, translation fixed
Rotation free, translation fixed
End condition code
Rotation fixed, translation free
Rotation free, translation free
Accessed by BHP BILLITON on 27 Jun 2002

FIGURE C3.4(3) LATERALLY UNBRACED PORTAL FRAME

COPYRIGHT
AS/NZS 4600 Supp1:1998 72

FIGURE C3.4(4) EFFECTIVE LENGTH FACTOR (k) IN LATERALLY UNBRACED


PORTAL FRAMES

Figure C3.4(4) can also serve as a guide for estimating k for other
simple situations. For multi-bay or multi-storey frames, or both, simple
alignment charts for determining k are given in the AISC
Commentaries (AISC, 1989; 1993). For additional information on
frame stability and second order effects, see SSRC Guide to Stability
Design Criteria for Metal Structures (Galambos, 1988) or the
Commentaries to AS 4100 — 1998 or NZS 3404:1997. The Australian
Standard AS 4100 — 1998 and NZS 3404:1997 give graphs for
determining effective length factors for sway and non-sway frames.
If roof or floor slabs, anchored to shear walls or vertical plane bracing
systems, are counted upon to provide lateral support for individual
columns in a building system, their stiffness should be considered
when functioning as horizontal diaphragms (Winter, 1958).
(ii) Torsional buckling of columns As pointed out at the beginning of this
Clause, purely torsional buckling, i.e. failure by sudden twist without
concurrent bending, is also possible for certain thin-walled open shapes.
These are all point-symmetric shapes in which the shear centre and centroid
coincide, such as doubly-symmetric I-shapes, anti-symmetric Z-shapes and
such unusual sections as cruciforms, swastikas, and the like. Under
concentric load, torsional buckling of such shapes very rarely governs
design. This is so because such members of realistic slenderness will buckle
flexurally or by a combination of flexural and local buckling at loads smaller
Accessed by BHP BILLITON on 27 Jun 2002

than those which would produce torsional buckling. However, for relatively
short members of this type, carefully dimensioned to minimize local
buckling, such torsional buckling cannot be completely ruled out. If such
buckling is elastic, it occurs at the critical stress (f oz) calculated as follows
(Winter, 1970):

. . . C3.4(13)

COPYRIGHT
73 AS/NZS 4600 Supp1:1998

Equation C3.4(13) is the same as Equation 3.3.3.2(12) in the Standard —


where
G = shear modulus
J = St. Venant torsion constant of the cross-section
A = full cross-sectional area
ro1 = polar radius of gyration of the cross-section about the shear centre
Iw = torsional warping constant of the cross-section
lez = effective length for twisting
For inelastic buckling, the critical torsional buckling stress can also be
calculated using Equations C3.4(10) and C3.4(11) and using foz instead of foc.
(iii) Flexural-torsional buckling of columns Concentrically loaded columns can
buckle in the flexural buckling mode by bending about one of the principal
axes; or in the torsional buckling mode by twisting about the shear centre; or
in the flexural-torsional buckling mode by simultaneous bending and
twisting. For singly-symmetric shapes such as channels, hat sections, angles,
T-sections, and I-sections with unequal flanges, for which the shear centre
and centroid do not coincide, flexural-torsional buckling is one of the
possible buckling modes as shown in Figure C3.4(5). Non-symmetric
sections will always buckle in the flexural-torsional mode.
Accessed by BHP BILLITON on 27 Jun 2002

FIGURE C3.4(5) TORSIONAL-FLEXURAL BUCKLING OF A CHANNEL IN AXIAL


COMPRESSION

It should be emphasized that one needs to design for flexural-torsional


buckling only when it is physically possible for such buckling to occur. This
means that if a member is so connected to other parts of the structure, such
as wall sheeting, that it can only bend but cannot twist, it needs to be
designed for flexural buckling only. This may hold for the entire member or
for individual parts. For instance, a channel member in a wall or the chord of
a roof truss is easily connected to girts or purlins in a manner that prevents

COPYRIGHT
AS/NZS 4600 Supp1:1998 74

twisting at these connection points. In this case, flexural-torsional buckling


needs to be checked only for the unbraced lengths between such connections.
Likewise, a doubly-symmetric compression member can be made up by
connecting two spaced channels at intervals by batten plates. In this case,
each channel constitutes an ‘intermittently fastened component of a built-up
shape’. Here, the entire member, being doubly-symmetric, is not subject to
flexural-torsional buckling so that this mode needs to be checked only for the
individual component channels between batten connections (Winter, 1970).
The governing elastic flexural-torsional buckling load can be calculated using
the following equation (Chajes and Winter, 1965; Chajes, Fang and Winter,
1966; Yu, 1991):

. . . C3.4(14)

If both sides of this equation are divided by the cross-sectional area (A), one
obtains the equation for the elastic, flexural-torsional buckling stress (foc) as
follows:

. . . C3.4(15)

In Equation C3.4(15) as in all provisions which deal with flexural-torsional


buckling, the x-axis is the axis of symmetry; fox = π2E/(lex/rx)2 is the flexural
(Euler) buckling stress about the x-axis, foz is the torsional buckling stress
(see Equation C3.4(13)) and β = 1 − (xo/r o1)2.
It is worth noting that the flexural-torsional buckling stress is always lower
than the Euler stress (fox) for flexural buckling about the symmetry axis.
Hence, for these singly-symmetric sections, flexural buckling can only occur,
if at all, about the y-axis which is the principal axis perpendicular to the axis
of symmetry.
For inelastic buckling, the critical flexural-torsional buckling stress can also
be calculated by using Equations C3.14(10) and C3.14(11).
An inspection of Equation C3.14(15) will show that in order to calculate β
and foz, it is necessary to determine xo equal to the distance between the shear
centre and centroid, J equal to the St. Venant torsion constant and I w equal to
the warping constant, in addition to several other, more familiar cross-
sectional properties.
Because of these complexities, the calculation of the flexural-torsional
buckling stress cannot be made as simply as that for flexural buckling.
Equations for these properties are given in Paragraph E1 of Appendix E of
Accessed by BHP BILLITON on 27 Jun 2002

the Standard.
Any singly-symmetric shape can buckle either flexurally about the y-axis or
flexurally-torsionally, depending on its detailed dimensions. For instance, a
channel stud with narrow flanges and wide web will generally buckle
flexurally about the y-axis (axis parallel to web); in contrast, a channel stud
with wide flanges and a narrow web will generally fail in flexural-torsional
buckling.

COPYRIGHT
75 AS/NZS 4600 Supp1:1998

The above discussion refers to members subject to flexural-torsional


buckling, but made up of elements whose b/t ratios are small enough so that
no local buckling will occur. For shapes that are sufficiently thin, i.e. with
b/t ratios sufficiently large, local buckling can combine with flexural-
torsional buckling similar to the combination of local with flexural buckling.
For this case, the effect of local buckling on the flexural-torsional buckling
strength can also be handled by using the effective area (Ae) determined at
the stress (fn) for flexural-torsional buckling.
(c) Local buckling of individual elements (See Item (b)(i)(D).)
(d) Distortional buckling (See Clause C3.4.6.)
C3.4.2 Sections not subject to torsional or flexural-torsional buckling If
concentrically loaded compression members can buckle in the flexural buckling mode by
bending about one of the principal axes, the nominal flexural buckling strength of the
column should be determined in accordance with Clause 3.4.1 of the Standard. The elastic
flexural buckling stress is calculated using Equation 3.4.2 of the Standard, which is the
same as Equation C3.4(3). This provision is applicable to doubly-symmetric sections,
closed cross-sections and any other sections not subject to torsional or flexural-torsional
buckling.
C3.4.3 Doubly- or singly-symmetric sections subject to torsional or
flexural-torsional buckling As discussed in Clause C3.4, torsional buckling is one of
the possible buckling modes for doubly- and point-symmetric sections. For singly-
symmetric sections, flexural-torsional buckling is one of the possible buckling modes. The
other possible buckling mode is flexural buckling by bending about the y-axis
(i.e. assuming x-axis is the axis of symmetry).
For torsional buckling, the elastic buckling stress can be calculated using
Equation C3.4(13). For flexural-torsional buckling, Equation C3.4(15) can be used to
calculate the elastic buckling stress. The following simplified equation for elastic flexural-
torsional buckling stress is an alternative permitted by the Standard:

. . . C3.4(16)

Equation C3.4(16) is based on the following relationship given by Peköz and Winter
(1969a):

. . . C3.4(17)

or

. . . C3.4(18)
Accessed by BHP BILLITON on 27 Jun 2002

C3.4.4 Point-symmetric sections Point-symmetric sections either buckle flexurally or


torsionally, and so the lower buckling load is chosen.
C3.4.5 Non-symmetric sections For non-symmetric open shapes, the analysis for
flexural-torsional buckling becomes extremely tedious unless its need is sufficiently
frequent to warrant computerization. For one thing, instead of the quadratic equations, a
cubic equation has to be solved. For another, the calculation of the required section
properties, particularly Iw, becomes quite complex. Clause 3.4.5 of the Standard specifies
that calculation in accordance with this Clause shall be used or tests in accordance with
Section 6 shall be made when dealing with non-symmetric open shapes.

COPYRIGHT
AS/NZS 4600 Supp1:1998 76

C3.4.6 Singly-symmetric sections subject to distortional buckling Some


singly-symmetric sections such as the storage rack columns shown in Figure C3.4.6 are
particularly sensitive to distortional buckling in the modes shown. Distortional buckling in
this mode was investigated in detail by Hancock (1985) mainly for sections used in steel
storage racks, by Lau and Hancock (1987, 1988a, 1988b, 1990) and by Kwon and
Hancock (1992a, 1992b) for high-strength steel channel sections with intermediate
stiffeners. The elastic distortional buckling stress (fod) can be based on a rational elastic
buckling analysis such as that given in Papangelis and Hancock (1995), or in
Paragraphs D1 and D2 of Appendix D of the Standard, which are based on Lau and
Hancock (1987). The equation for the critical stress (fn) is developed in Hancock, Kwon
and Bernard (1994). It is normally not necessary to check simple lipped channels in
accordance with Clause 3.4.6, because they are adequately designed by Clause 2.4.3 of the
Standard for the distortional mode.

FIGURE C3.4.6 DISTORTIONAL BUCKLING MODES OF SINGLY SYMMETRIC


COMPRESSION MEMBERS

C3.4.7 Columns with one flange through-fastened to sheeting For axially loaded
C-sections or Z-sections having one flange attached to deck or sheeting, and the other
flange unbraced, e.g. a roof purlin or wall girt subjected to wind or seismic-generated
compression forces, the axial load capacity is less than a fully braced member, but greater
than an unbraced member. The partial restraint relative to weak axis buckling is a
function of the rotational stiffness provided by the panel-to-purlin connection.
Equation 3.4.7 in the Standard is used to calculate the weak axis capacity. This Equation
is not valid for sections attached to standing seam or clip-fastened roofs. The Equation
was developed by Glaser, Kaehler and Fisher (1994) and is also based on the work
Accessed by BHP BILLITON on 27 Jun 2002

contained in the reports of Hatch, Easterling and Murray (1990) and Simaan (1973).
A limitation on the maximum yield point of the C-section or Z-section is not given in the
Standard since Equation 3.4.7 of the Standard is based on elastic buckling criteria. A
limitation on minimum length is not contained in the Standard because Equation 3.4.7 is
conservative for spans less than 4.5 m.
As indicated in Clause 3.4.7 of the Standard, the strong axis axial load capacity is
determined assuming that the weak axis of the strut is braced, and so Clauses 3.5.1
and 3.5.2 of the Standard should be used.
The controlling axial capacity (weak or strong axis) is suitable for use in the combined
axial load and bending equations in Clause 3.5 of the Standard (Hatch, Easterling, and
Murray, 1990).

COPYRIGHT
77 AS/NZS 4600 Supp1:1998

C3.5 COMBINED AXIAL LOAD AND BENDING In the 1996 edition of the AISI
Specification, the design provisions for combined axial load and bending were expanded
to include expressions for the design of members subject to combined tensile axial load
and bending. These additional expressions are also included in the Standard.
C3.5.1 Combined axial compressive load and bending Cold-formed steel members
under a combination of compressive axial load and bending are usually referred to as
beam-columns. The bending may result from eccentric loading, transverse loads or applied
moments. Such members are often found in framed structures, trusses and exterior wall
studs. For the design of such members, interactions have been developed for locally stable
and unstable beam-columns on the basis of thorough comparisons with rigorous theory
and verified by the available test results (Peköz, 1986a; Peköz and Sumer, 1992).
The structural behaviour of beam-columns depends on the shape and dimensions of the
cross-section, the location of the applied eccentric load, the column length, the end
restraint and the condition of bracing.
When a beam-column is subject to an axial load (N*) and end moments (M*) as shown in
Figure C3.5.1, the combined axial and bending stress in compression is calculated using
the following equation, as long as the member remains straight:

. . . C3.5.1(1)

=
where
f* = combined stress in compression
f*a = axial compressive stress
f*b = bending stress in compression
N* = applied axial load
A = cross-sectional area
M* = applied bending moment
Z = elastic section modulus
It should be noted that in the design of such a beam-column by using the limiting stress
the combined stress should be limited to a certain stress (F), that is —

or

. . . C3.5.1(2)

If the left-hand term in Equation C3.5.1(2) is multiplied by the area (A) in the numerator
and demoninator, and the right-hand term is multiplied by the section modulus in the
Accessed by BHP BILLITON on 27 Jun 2002

numerator and denominator, then Equation C3.5.1(2) becomes —

. . . C3.5.1(3)

where
N* = applied axial load

COPYRIGHT
AS/NZS 4600 Supp1:1998 78

Ns = axial section capacity


= AF
M* = applied moment
= Zf*b
Ms = bending section capacity
= ZF
If the axial section capacity is reduced by the capacity factor (φ c) for compression and the
bending section capacity is reduced by the capacity factor (φb) for bending, then
Equation C3.5.1(3) becomes —

. . . C3.5.1(4)

Equation C3.5.1(4) is well-known, and has been adopted in some specifications and
Standards for the design of beam-columns. It can be used with reasonable accuracy for
short members and members subjected to a relative small axial load. It should be realized
that in practical applications, when end moments are applied to the member, it will be
bent as shown in Figure C3.5.1(b) due to the applied moment (M*), the secondary
moment resulting from the applied axial load (N*) and the deflection (δ) of the member.
The maximum bending moment at mid-length (point C) can be calculated as follows:

. . . C3.5.1(5)

where
M*max. = maximum bending moment at midlength
M* = applied end moments
αn = amplification factor
The amplification factor (αn) may be calculated as follows:

. . . C3.5.1(6)

where
Ne = elastic column buckling load (Euler load)

= π2 EI / leb
2

If the maximum bending moment (M*max.) is used to replace M*, the following equation
can be obtained from Equation C3.5.1(4) and Equation C3.5.1(6) as follows:
Accessed by BHP BILLITON on 27 Jun 2002

. . . C3.5.1(7)

For members that can buckle under axial compression, as described in Clause C3.4, N s is
replaced by N c and for members that can buckle laterally or distortionally, as described in
Clause C3.3, Ms is replaced by Mb, so that —

. . . C3.5.1(8)

COPYRIGHT
79 AS/NZS 4600 Supp1:1998

It has been found that Equation C3.5.1(8), developed for a member subjected to an axial
compressive load and equal end moments, can be used with reasonable accuracy for
braced members with unrestrained ends subjected to an axial load and a uniformly
distributed transverse load. However, it could be conservative for compression members
in unbraced frames (with sidesway), and for members bent in reverse curvature. For this
reason, the Equation C3.5.1(8) should be further modified by a coefficient (Cm) as given
in Equation C3.5.1(9) to account for the effect of end moments —

. . . C3.5.1(9)

Equation C3.5.1(9) forms the basis of Clause 3.5.1(a) of the Standard.


In Equation C3.5.1(9), C m can be determined for one of the three cases specified in
Clause 3.5.1 of the Standard. It can be calculated using Equation C3.5.1(10) for restrained
compression members braced against joint translation and not subject to transverse
loading (Case ii) as follows:

. . . C3.5.1(10)

where M1/M2 is the ratio of the smaller to the larger end moments. For Case (iii), Cm may
be approximated by using the value given in the AISC Commentaries for the applicable
condition of transverse loading and end restraint (AISC, 1989 and 1993).
When the maximum moment occurs at braced points, Equation C3.5.1(11)
(i.e. Clause 3.5.1(b) of the Standard) should be used to check the member at the braced
ends as follows:

. . . C3.5.1(11)

Furthermore, for the condition of small axial load, the influence of Cm/αn is usually small
and may be neglected. Therefore, when N*/φcNc ≤ 0.15, Equation C3.5.1(4) may be used
for the design of beam-columns.
If a second-order elastic analysis is performed in accordance with Appendix E of
AS 4100 — 1998 or NZS 3404:1997, then it is appropriate to use the second-order elastic
moments in place of the first order elastic moments (M*) in Equation C3.5.1(9), and the
value of Cm/αn may be taken as 1.0 since the effect of non-uniform moment and the
moment amplification are now included in the determination of the second order
moments.
C3.5.2 Combined axial tensile load and bending The design criteria included in
Accessed by BHP BILLITON on 27 Jun 2002

Clause 3.5.2 of the Standard are new requirements. These provisions apply to concurrent
bending and tensile axial load. If bending can occur without the presence of tensile axial
load, the member should also conform to the provisions of Clause 3.3 of the Standard.
Care should be taken not to overestimate the tensile load, as this would be unconservative.
Clause 3.5.2(a) of the Standard provides a design criterion to prevent failure of the
compression flange. Clause 3.5.2(b) of the Standard provides a design criterion to prevent
yielding of the tension flange of a member under combined tensile axial load and bending.

COPYRIGHT
AS/NZS 4600 Supp1:1998 80

FIGURE C3.5.1 BEAM-COLUMN SUBJECTED TO AXIAL LOADS AND END MOMENTS

C3.6 CYLINDRICAL TUBULAR MEMBERS Thin-walled cylindrical tubular


members are economic sections for compression and torsional members because of their
large ratio of radius of gyration to area, of having the same radius of gyration in all
directions and their large torsional rigidity. Like other cold-formed steel compression
members, cylindrical tubes should be designed to provide adequate safety not only against
overall column buckling but also against local buckling. It is well known that the classical
theory of local buckling of longitudinally compressed cylinders overestimates the actual
buckling strength and that inevitable imperfections and residual stresses reduce the actual
strength of compressed tubes radically below their theoretical value. For this reason, the
design provisions for local buckling have been based largely on test results.
Considering the post-buckling behaviour (local buckling stress) of the axially compressed
cylinder and the important effect of the initial imperfection, the design provisions
included in the Standard were originally based on Plantema’s graphic representation and
the additional results of cylindrical shell tests made by Wilson and Newmark at the
University of Illinois (Winter, 1970).
From the tests of compressed tubes, Plantema found that the ratio f ult/fy depends on the
parameter (E/fy) (t/D) in which t is the wall thickness, D is the mean diameter of the tubes
and fult is the ultimate stress or collapse stress. As shown in Figure C3.6(1), Line 1
corresponds to the collapse stress below the proportional limit, Line 2 corresponds to the
collapse stress between the proportional limit and the yield point, and Line 3 represents
the collapse stress occurring at the yield point. In the range of Line 3, local buckling will
not occur before yielding. In ranges 1 and 2, local buckling occurs before the yield point
is reached. The cylindrical tubes should be designed to safeguard against local buckling.
Based on a conservative approach, the Standard specifies that when the do/t ratio is
Accessed by BHP BILLITON on 27 Jun 2002

smaller than or equal to 0.112E/fy, the tubular member should be designed for yielding.
This provision is based on point A 1, for which (E/fy) (t/do) is equal to 8.93.
When 0.112E/fy < do/t < 0.441E/fy, the design of tubular members is based on the inelastic
local buckling criteria. For the purpose of developing a design equation for inelastic
buckling, point B1 was selected to represent the proportional limit. For point B 1, the
maximum stress of cylindrical tubes can be calculated as follows:

. . . C3.6(1)

COPYRIGHT
81 AS/NZS 4600 Supp1:1998

Using Line A1B1, the maximum stress of cylindrical tubes can be calculated as follows:

. . . C3.6(2)

When do/t ≥ 0.441 E/fy, the following equation represents Line 1 for the elastic local
buckling stress:

. . . C3.6(3)

FIGURE C3.6(1) CRITICAL STRESS OF CYLINDRICAL TUBES FOR LOCAL BUCKLING


Accessed by BHP BILLITON on 27 Jun 2002

The correlation between the available test data and Equations C3.6(1) and C3.6.(3) is
shown in Figure C3.6(2).
It should be noted that the design provisions of Clause 3.6 of the Standard are applicable
only for members having a ratio of outside diameter-to-wall thickness (d o/t) not greater
than 0.441 E/fy because the design of extremely thin tubes will be governed by elastic
local buckling resulting in an uneconomical design. In addition, cylindrical tubular
members with unusually large do/t ratios are very sensitive to geometric imperfections.

COPYRIGHT
AS/NZS 4600 Supp1:1998 82

FIGURE C3.6(2) CORRELATION BETWEEN TEST DATA AND CRITERIA FOR LOCAL
BUCKLING OF CYLINDRICAL TUBES UNDER AXIAL COMPRESSION

C3.6.2 Bending For non-slender cylinders in bending, the initiation of yielding does
not represent a failure condition as is generally assumed for axial loading. Failure is at the
plastic moment capacity which is at least 1.29 times the moment at first yielding. In
addition, the conditions for inelastic local buckling are not as severe as in axial
compression due to the stress gradient.
Equations 3.6.2(1), 3.6.2(2) and 3.6.2(3) in the Standard are based upon the work reported
by Sherman (1985) and an assumed minimum shape factor of 1.25. This slight reduction
in the inelastic range has been made to limit the maximum bending stress to 0.75f y, a
value typically used for solid sections in bending for the permissible stress method. The
reduction also brings the criteria closer to a lower bound for inelastic local buckling. A
small range of elastic local buckling has been included so that the upper d o/t limit of
0.441E/fy is the same as for axial compression.
All three equations in Clause 3.6.2 of the Standard for determining the nominal flexural
strength of cylindrical tubular members are shown in Figure C3.6.2. These equations have
been used in the AISI Specification since 1986 and are retained in the 1996 Specification
and the Standard. The capacity factor (φb) is the same as that used in Clause 3.3.2 of the
Standard for sectional bending strength.
Accessed by BHP BILLITON on 27 Jun 2002

COPYRIGHT
83 AS/NZS 4600 Supp1:1998

FIGURE C3.6.2 NOMINAL FLEXURAL STRENGTH OF


CYLINDRICAL TUBULAR MEMBERS

C3.6.3 Compression When cylindrical tubes are used as concentrically loaded


compression members, the nominal axial strength is determined by the same equation as
given in Clause 3.4 of the Standard, except that —
(a) the nominal buckling stress (foc) is determined only for flexural buckling; and
(b) the effective area (Ae) is calculated as follows:

. . . C3.6.3(1)
where

. . . C3.6.3(2)
Accessed by BHP BILLITON on 27 Jun 2002

. . . C3.6.3(3)

where A is the area of the unreduced cross-section. The capacity factor (φc) is the same as
that used in Clause 3.4 of the Standard for compression members.
Equation C3.6.3(3) is used for calculating the reduced area due to local buckling. It is
derived from Equation C3.6(2) for inelastic local buckling stress (Yu, 1991).
C3.6.4 Combined bending and compression The interactions given in Clause 3.5 of
the Standard can also be used for the design of cylindrical tubular members when these
members are subject to combined bending and compression.

COPYRIGHT
AS/NZS 4600 Supp1:1998 84

S E C T I O N C 4 S T R U C T U R A L A S S E M B L I E S

C4.1 BUILT-UP SECTIONS


C4.1.1 I-sections composed of two channels I-sections made by connecting two
channels back to back are often used as either compression or flexural members. Cases (b)
and (h) of Figure C1.3.2 and Cases (c) and (g) of Figure C1.3.3 show several built-up
I-sections, as follows:
(a) Compression members For the I-sections to be used as compression members, the
longitudinal spacing of connectors should not exceed the value of s max., calculated
using Equation 4.1.1(1) of the Standard. This prevents flexural buckling of the
individual channels about the axis parallel to the web at a load smaller than that at
which the entire I-section would buckle. This provision is based on the requirement
that the slenderness ratio of an individual channel between connectors (s max./rcy)
should not be greater than one-half of the pertinent slenderness ratio (l/r I) of the
entire I-section (Winter, 1970; Yu, 1991) and follows for one of the connectors
becoming loose or ineffective.
Even though Clause 4.1.1 of the Standard refers only to I-sections,
Equation 4.1.1(1) of the Standard can also be used for determining the maximum
spacing of welds for box-shaped compression members made by connecting two
channels tip to tip. In this case, rI is the smaller of the two radii of gyration of the
box-shaped section.
(b) Flexural members For the I-sections to be used as flexural members, the
longitudinal spacing of connectors is limited by Equation 4.1.1(2) of the Standard.
The first requirement is an arbitrarily selected limit to prevent any possible
excessive distortion of the top flange between connectors. The second is based on
the strength and arrangement of connectors and the intensity of the load acting on
the beam (Yu, 1991).
The maximum spacing of connectors required by Equation 4.1.1(2) of the Standard is
based on the fact that the shear centre of the channel is neither coincident with nor
located in the plane of the web; and that when a load (Q) is applied in the plane of the
web, it produces a twisting moment (Qm) about its shear centre, as shown in
Figure C4.1.1(1). The tensile force of the top connector (N*) can then be calculated from
the equality of the twisting moment (Qm) and the resisting moment (N*s g), as follows:

. . . C4.1.1(1)

. . . C4.1.1(2)
Accessed by BHP BILLITON on 27 Jun 2002

Considering that q is the intensity of the load and that s is the spacing of connectors as
shown in Figure C4.1.1(2), the applied load (Q) is equal to qs/2. The maximum spacing
(smax.) used in the Standard can easily be obtained by substituting the value of Q into
Equation C4.1.1(2). The determination of the load intensity (q) is based upon the type of
loading applied to the beam.
In addition to the required strength of connections, the spacing of connectors should not
be so great as to cause excessive distortion between connectors by separation along the
top flange. In view of the fact that channels are connected back to back and are
continuously in contact along the bottom flange, a maximum spacing of l/3 may be used.
Considering the possibility that one connection may be defective, a maximum spacing of
smax. equal to l/6 is also required in Equation 4.1.1(2) of the Standard.

COPYRIGHT
85 AS/NZS 4600 Supp1:1998

FIGURE C4.1.1(1) TENSILE FORCE DEVELOPED IN THE TOP CONNECTOR


FOR CHANNEL

FIGURE C4.1.1(2) SPACING OF CONNECTORS

C4.1.2 Cover plates, sheets or non-integral stiffeners in compression When


compression elements are joined to other parts of built-up members by intermittent
connections, these connectors should be closely spaced to develop the required strength of
the connected element. Figure C4.1.2 shows a box-shaped beam made by connecting a flat
sheet to an inverted hat section. If the connectors are appropriately placed, this flat sheet
will act as a stiffened compression element with a width (b) equal to the distance between
rows of connectors, and the sectional properties can be calculated accordingly. This is the
intent of the provisions in Clause 4.1.2 of the Standard.
Clause 4.1.2(a) of the Standard requires that the necessary shear strength be provided by
the same standard structural design procedure that is used in calculating flange
Accessed by BHP BILLITON on 27 Jun 2002

connections in bolted or welded plate girders or similar structures.


Clause 4.1.2(b) of the Standard ensures that the part of the flat sheet between two
adjacent connectors will not buckle as a column (see Figure C4.1.2) at a stress less than
1.67fc, where fc is the stress at service load in the connected compression element (Winter,
1970; Yu, 1991). The AISI requirement is based on the following Euler equation for
column buckling:

. . . C4.1.2(1)

COPYRIGHT
AS/NZS 4600 Supp1:1998 86

by substituting σcr equal to 1.67 f c, le equal to 0.6s and r equal to . This provision is
conservative because the length is taken as the centre distance instead of the clear distance
between connectors, and the effective length (le) is taken as 0.6s instead of 0.5s, which is
a theoretical value for a column with fixed end supports.
Clause 4.1.2(c) of the Standard ensures satisfactory spacing to make a row of connectors
act as a continuous line of stiffening for the flat sheet under most conditions (Winter,
1970; Yu, 1991).

FIGURE C4.1.2 SPACING OF CONNECTORS IN COMPOSITE SECTION

C4.2 MIXED SYSTEMS When cold-formed steel members are used in conjunction
with other construction materials, the design requirements of the other material standards
should be satisfied.

C4.3 LATERAL RESTRAINTS Bracing design requirements were expanded in the


1986 Specification to include a general statement regarding bracing for symmetrical
beams and columns and specific requirements for the design of roof systems subjected to
gravity load. These requirements are retained with some revisions of Clause 4.3.3.4 of the
Standard for the required number of braces.
C4.3.2 Symmetrical beams and columns There are no simple, generally accepted
techniques for determining the required strength and stiffness for discrete braces in steel
construction. Winter (1960) offered a partial solution and others have extended this
knowledge (Haussler, 1964; Haussler and Pahers, 1973; Lutz and Fisher, 1985; Salmon
and Johnson, 1990; Yura, 1993; SSRC, 1993). The design engineer is encouraged to seek
out the stated references to obtain guidance for the design of a brace or brace system.
In the Standard, the provisions of AS 4100 — 1998 or NZS 3404:1997 are used for
Clause 4.3.2 of the Standard. Reference should be made to Clause C5.4.3 of the
Commentary to AS 4100, or to Clause C5.4.3 of the Commentary to NZS 3404:1997. The
latter is augmented by the guidance given in HERA Report R4 — 92.
Accessed by BHP BILLITON on 27 Jun 2002

C4.3.3 Channel and Z-section beams Channel and Z-sections used as beams to
support transverse loads applied in the plane of the web may twist and deflect laterally
unless adequate lateral supports are provided. Clauses 4.3.3.2 and 4.3.3.3 of the Standard
deal with the bracing requirements when one flange of the beam is connected to deck or
sheeting material. Clause 4.3.3.4 of the Standard covers the requirements for spacing and
design of braces, when neither flange of the beam is connected to sheeting or is connected
to sheeting with concealed fasteners.
C4.3.3.2 One flange connected to sheeting and subjected to wind uplift Clause 4.3.3.2
of the Standard permits the bracing not to be connected to a stiff member but be capable
of preventing torsional deformation of the beam at the point of attachment, if the cleat
and screw-fastening requirements of Clause 3.3.3.4, Items (ix) to (xiv), of the Standard

COPYRIGHT
87 AS/NZS 4600 Supp1:1998

are satisfied. This provision is based on testing at the University of Sydney (Hancock,
Celeban and Healy (1993)) of typical Australian purlin-sheeting systems with screw-
fastened sheeting, cleats and side lap fasteners.
C4.3.3.3 Bracing for roof systems under gravity load The provisions of Clause 4.3.3.3
of the Standard are based on USA research where cleats are not used at support points.
The design rules for Z-purlin-supported roof systems are based on a first order, elastic
stiffness model (Murray and Elhouar, 1985). For the design of lateral bracing,
Equations 4.3.3.3(1) to 4.3.3.3(6) of the Standard can be used to determine the restraint
forces for single-span and multiple-span systems with braces at various locations. These
design equations are written in terms of the cross-sectional dimensions of the purlin,
number of purlin lines, number of spans, span length for multiple-span systems, and the
total load applied to the system. The accuracy of these design equations was verified by
Murray and Elhouar using their experimental results of six prototype and 33 quarter-scale
tests.
C4.3.3.4 Neither flange connected to sheeting or connected to sheeting with concealed
fasteners Where neither flange is connected to sheeting or where the flange is connected
to sheeting with concealed fasteners, the following should be considered:
(a) Bracing of channel beams If channels are used singly as beams, rather than being
paired to form I-sections, they should evidently be braced at intervals so as to
prevent them from rotating in the manner shown in Figure C4.3.3.4(1).
Figure C4.3.3.4(2), for simplicity, shows two channels braced at intervals against
each other. The situation is evidently much the same as in the composite I-section
shown in Figure C4.1.1(2), except that the role of the connectors is now played by
the braces. The difference is that the two channels are not in contact and that the
spacing of braces is generally considerably larger than the connector spacing. In
consequence, each channel may actually rotate very slightly between braces and this
will cause some additional stresses which superpose on the usual, simple bending
stresses. Bracing should be so arranged that —
(i) these additional stresses are small enough so that they will not reduce the
load-carrying capacity of the channel (as compared to what it would be in
the continuously braced condition); and
(ii) rotations should be kept small enough to be acceptable (of the order of 1 to 2
degrees).
In order to develop information on which to base appropriate bracing provisions,
different channel shapes were tested at Cornell University (Winter, 1970). Each of
these was tested with full, continuous bracing, without any bracing, and with
intermediate bracing at two different spacings. In addition to this experimental
work, an approximate method of analysis was developed and checked against the
test results. A condensed account of this is given by Winter, Lansing and McCalley
(1949). The reference indicates that the requirements are satisfied for most
distributions of beam load if between supports not less than three equidistant braces
Accessed by BHP BILLITON on 27 Jun 2002

are placed (i.e. at quarter-points of the span or closer). The exception is the case
where a large part of the total load of the beam is concentrated over a short portion
of the span. In this case, an additional brace is to be placed at such a load.
Correspondingly, previous editions of the AISI Specification (AISI, 1986; AISI,
1991) provided that the distance between braces shall not be greater than one-
quarter of the span. It also defined the conditions under which an additional brace
shall be placed at a load concentration.
For such braces to be effective, it is not only necessary that their spacing be
appropriately limited. In addition, their strength should suffice to provide the force
required to prevent the channel from rotating. It is, therefore, necessary also to
determine the forces that will act in braces, such as those forces shown in

COPYRIGHT
AS/NZS 4600 Supp1:1998 88

Figure C4.3.3.4(3). These forces are found if one considers that the action of a load
applied in the plane of the web (which causes a torque (Qm)) is equivalent to that
same load when applied at the shear centre (where it causes no torque) plus two
forces N* equal to Qm/d which, together, produce the same torque Qm. As shown in
Figure C4.3.3.4(4) and shown in some detail by Winter, Lansing and McCalley
(1949b), each half of the channel can then be regarded as a continuous beam loaded
by the horizontal forces and supported at the brace points. The horizontal brace
force is then, simply, the appropriate reaction of this continuous beam. The
provisions of Clause 4.3.3.4 of the Standard represent a simple and conservative
approximation for determining these reactions. These reactions are equal to the force
N*ib, which the brace is required to resist at each flange.
(b) Bracing of Z-section beams Most Z-sections are anti-symmetrical about the
vertical and horizontal centroidal axes, i.e. they are point-symmetrical. In view of
this, the centroid and the shear centre coincide and are located at the midpoint of
the web. A load applied in the plane of the web has, then, no lever arm about the
shear centre (m equal to 0) and does not tend to produce the kind of rotation a
similar load would produce on a channel. However, in Z-sections, the principal axes
are oblique to the web (see Figure C4.3.3.4(5)). A load applied in the plane of the
web, resolved in the direction of the two axes, produces deflections in each of them.
By projecting these deflections onto the horizontal and vertical planes, it is found
that a Z-beam loaded vertically in the plane of the web deflects not only vertically
but also horizontally. If such deflection is permitted to occur, then the loads,
moving sideways with the beam, are no longer in the same plane with the reactions
at the ends. In consequence, the loads produce a twisting moment about the line
connecting the reactions. In this manner, it is seen that a Z-beam, unbraced between
ends and loaded in the plane of the web, deflects laterally and also twists. Not only
are these deformations likely to interfere with a proper functioning of the beam, but
the additional stresses caused by them produce failure at a load considerably lower
than when the same beam is used fully braced.
In order to develop information on which to base appropriate bracing provisions,
tests were carried out on three different Z-shapes at Cornell University, unbraced as
well as with variously spaced intermediate braces. In addition, an approximate
method of analysis was developed and checked against the test results. An account
of this was given by Zetlin and Winter (1955). Briefly, it shows that intermittently
braced Z-beams can be analysed in much the same way as intermittently braced
channels. It is merely necessary, at the point of each actual vertical load (Q), to
apply a fictitious horizontal load N* equal to Q(Ix′y′/Ix′). One can then calculate the
vertical and horizontal deflections, and the corresponding stresses, in conventional
ways by utilizing the centroidal axes x′ and y′ (rather than x and y, as shown in
Figure C4.3.3.4(5)) except that certain modified section properties have to be used.
In this manner, it has been shown that as to the location of braces, the same
provisions that apply to channels also apply to Z-beams. Likewise, the forces in the
Accessed by BHP BILLITON on 27 Jun 2002

braces are again obtained as the reactions of continuous beams horizontally loaded
by fictitious loads N*ib.
(c) Spacing of braces During the period from 1956 through 1996, the AISI
Specification required that braces be attached both to the top and bottom flanges of
the beam, at the ends and at intervals not greater than one-quarter of the span
length, in such a manner as to prevent tipping at the ends and lateral deflection of
either flange in either direction at intermediate braces. The lateral buckling
equations provided in Clause 3.3.3.2 of the Standard can be used to predict the
moment capacity of the member. Recently, beam tests conducted by Ellifritt, Sputo
and Haynes (1992) have shown that for typical sections, a mid-span brace may
reduce service load horizontal deflections and rotations by as much as 80% when

COPYRIGHT
89 AS/NZS 4600 Supp1:1998

compared to a completely unbraced beam. However, the restraining effect of braces


may change the failure mode from lateral-torsional buckling to distortional buckling
of the flange and lip at a brace point. The natural tendency of the member under
vertical load is to twist and translate in such a manner as to relieve the compression
on the lip. When such movement is restrained by intermediate braces, the
compression on the stiffening lip is not relieved, and may increase. In this case,
distortional buckling may occur at loads lower than that predicted by the lateral
buckling equations given in Clause 3.3.3.2 of the Standard. Hence, equations for
distortional buckling were included in Clause 3.3.3.3 of the Standard.
The Standard permits omission of discrete braces when all loads and reactions on a beam
are transmitted through members that frame into the section in such a manner as to
effectively restrain the member against torsional rotation and lateral displacement.
The inclusion of purlins with sheeting connected with concealed fasteners in
Clause 4.3.3.4 of the Standard is based on testing at the University of Sydney (Hancock,
Celeban and Healy (1994)) where concealed fastened sheeting was found not to provide
adequate lateral restraint due to shear slippage between the sheets.

FIGURE C4.3.3.4(1) ROTATION OF CHANNEL BEAMS


Accessed by BHP BILLITON on 27 Jun 2002

FIGURE C4.3.3.4(2) TWO CHANNELS BRACED AT INTERVALS


AGAINST EACH OTHER

COPYRIGHT
AS/NZS 4600 Supp1:1998 90

FIGURE C4.3.3.4(3) LATERAL FORCES APPLIED TO CHANNEL

FIGURE C4.3.3.4(4) HALF OF CHANNEL TREATED AS A CONTINUOUS BEAM


LOADED BY HORIZONTAL FORCES
Accessed by BHP BILLITON on 27 Jun 2002

FIGURE C4.3.3.4(5) PRINCIPAL AXES OF Z-SECTION

COPYRIGHT
91 AS/NZS 4600 Supp1:1998

C4.4 WALL STUDS AND WALL STUD ASSEMBLIES It is well known that
column strength can be increased considerably by using adequate bracing, even though the
bracing is relatively flexible. This is particularly true for those sections generally used as
load-bearing wall studs which have large Ix/Iy ratios.
Cold-formed channels or box-type studs are generally used in walls with their webs
placed perpendicular to the wall surface. The walls may be made of different materials,
such as fibre board, pulp board, plywood or gypsum board. If the wall material is strong
enough and there is adequate attachment provided between wall material and studs for
lateral support of the studs, then the wall material can contribute to the structural
economy by increasing the useable strength of the studs substantially.
In order to determine the necessary requirements for adequate lateral support of the wall
studs, theoretical and experimental investigations were conducted in the 1940s by Green,
Winter, and Cuykendall (1947). The study included 102 tests on studs and 24 tests on a
variety of wall material. Based on the findings of this earlier investigation, specific AISI
provisions were developed for the design of wall studs.
In the 1970s, the structural behaviour of columns braced by steel diaphragms was a
special subject investigated at Cornell University and other institutions. The renewed
investigation of wall-braced studs has indicated that the bracing provided for studs by
steel panels is of the shear diaphragm type rather than the linear type that was considered
in the 1947 study. Simaan (1973) and Simaan and Pekoz (1976), which are summarized
by Yu (1991), contain procedures for calculating the strength of Channel and Z-section
wall studs that are braced by sheeting materials. The bracing action is due to both the
shear rigidity and the rotational restraint supplied by the sheeting material. The treatment
by Simaan (1973) and Simaan and Peköz (1976) is quite general and includes the case of
studs braced on one as well as on both flanges.
In Australia, it is not common to account for sheeting in the design of wall studs. Hence,
intermediate braces such as noggins (dwangs) are used. Design is then performed in
accordance with Section 3 of the Standard with appropriate effective lengths about the x, y
and z axes.
In New Zealand, bracing gypsum board is commonly used in conjunction with the means
of attachment so that stud flange restraint is provided. For assemblies, this bracing
gypsum board can be used with tension strap bracing to provide adequate bracing without
dwangs.
Accessed by BHP BILLITON on 27 Jun 2002

COPYRIGHT
AS/NZS 4600 Supp1:1998 92

S E C T I O N C 5 C O N N E C T I O N S

C5.1 GENERAL Welds, bolts, screws, rivets and other special devices such as
clinching, nailing and structural adhesives are generally used for cold-formed steel
connections (Brockenbrough, 1995). The 1996 edition of the AISI Specification contains
provisions in Chapter E for welded connections, bolted connections and screw
connections. The Standard includes the same provisions in Section 5 and also
blind-riveted connections. Design rules for clinching are not given at this stage, since
clinches are proprietary devices for which information on strength of connections should
be obtained from tests carried out by or for the user. The provisions in Section 6 of the
Standard are to be used in these tests.
The plans or specifications, or both, are to contain information and design requirement
data for the detailing of each connection, if each connection is not detailed on the
engineering design drawings.
For steel thickness less than 0.75 mm, the design value of the connection will generally
be that determined by the material thickness. However, for heavier steels, the shear or
tension value of the fastener (dependent on size) will usually govern. The fastener
strength in both shear and tension, should be 1.25 times that of the connection.
The design capacity for a connection can be determined as follows:
(a) Use the equations given in Clause 5.2 for welds, Clause 5.3.5 for bolts,
Clauses 5.4.2 and 5.4.3 for screws, and Clause 5.5.2 for rivets, where the calculated
nominal capacity is factored by the capacity [strength reduction] factor (φ) given in
Table 1.6 of the Standard.
(b) Determine the design capacity [strength] of a single-point connection by test as
follows:

. . . C5.2
where
φRu = design capacity [strength] by test of a single-point
connection
Rt = target test loads for the member of units to be tested
kt = variability factor given in Table 6.2.2 of the Standard
fu = minimum tensile strength of the connected material
given in Table 1.5 of the Standard
fu(average, tested) = average value of the tensile strength from the sample
tested
Accessed by BHP BILLITON on 27 Jun 2002

Consideration should be given to the durability of the connections and these should be
designed to functionally perform during the expected life of the structure. Compatibility is
important when dissimilar metals are combined or when fasteners are exposed to other
than dry internal environments.
The ratio fu/fu(average, tested) adjusts the test loads to allow for the fact that the material from
which the specimens are made may be of higher strength than the minimum specified.

COPYRIGHT
93 AS/NZS 4600 Supp1:1998

C5.2 WELDED CONNECTIONS Welds used for cold-formed steel construction may
be classified as arc welds and resistance welds. Arc welding is used for connecting cold-
formed steel members to each other as well as connecting such members to heavy, hot-
rolled steel framing (such as floor panels to beams of the steel frame). It is used in butt
welds, fillet welds, arc spot welds (puddle welds), arc seam welds and flare welds.
The design provisions contained in the Standard for arc welds are based primarily on
experimental evidence obtained from an extensive test program conducted at Cornell
University. The results of this program are reported by Peköz and McGuire (1979) and
summarized by Yu (1991). All possible failure modes are covered in the present
provisions, whereas the earlier provisions mainly dealt with shear failure.
For most of the connection tests reported by Peköz and McGuire (1979), the onset of
yielding was either poorly defined or followed closely by ultimate failure. Therefore, in
the provisions of Section 5 of the Standard, rupture rather than yielding is used as a more
reliable criterion of failure.
In addition, the Cornell research has provided the experimental basis for the AWS
Structural Welding Code for Sheet Steel (AWS, 1989). In most cases, the provisions of
the AWS code are in agreement with Section 5 of the Standard.
The welded connection tests, which served as the basis of the provisions given in
Clause 5.2 of the Standard, were conducted on sections with single and double sheets. The
largest total sheet thickness of the cover plates was approximately 3.81 mm. However,
within the Standard, the validity of the equations is limited to welded connections in
which the thickness of the thinnest connected part is 3 mm (2.5 mm for fillet welds) or
less. For welds in thicker material, the provisions of AS 4100 or NZS 3404 should be
used. These limitations were based on testing by Zhao and Hancock (1995a, 1995b).
The terms used in Section 5 of the Standard agree with the standard nomenclature given
in the AWS Welding Code for Sheet Steel (AWS, 1992).
C5.2.2 Butt welds The design equations for determining nominal capacity for butt
welds are taken from the AISC LRFD Specification (AISC, 1993) where the effective
throat thickness (te) is replaced by its equivalent, the design throat thickness (tt), as
specified in AS/NZS 1554.1:1995.
C5.2.3 Fillet welds For fillet welds on the lap joint specimens tested in the Cornell
research (Peköz and McGuire, 1979), the dimension of the leg on the sheet edge generally
was equal to the sheet thickness; the other leg often was two or three times longer. In
connections of this type, the fillet weld throat commonly is larger than the throat of a
conventional fillet weld of the same size. Usually, ultimate failure of fillet welded joints
has been found to occur by the tearing of the plate adjacent to the weld (see
Figure C5.2.3).
In most cases, the higher strength of the weld material prevents weld shear failure;
therefore, the provisions of the standard section are based on sheet tearing. For sections
thicker than 2.5 mm, the provisions of AS 4100 and NZS 3404, as appropriate, should be
Accessed by BHP BILLITON on 27 Jun 2002

used. This limit is based on the research of Zhao and Hancock (1995a, 1995b).
C5.2.4 Arc spot welds (puddle welds) Arc spot welds (puddle welds) used for
connecting thin sheets are similar to plug welds used for relatively thicker plates. The
difference between plug welds and arc spot welds is that the former are made with
prepunched holes, but for the latter no prepunched holes are required. Instead, a hole is
burned in the top sheet by the arc and then filled with weld metal to fuse it to the bottom
sheet or a framing member.

COPYRIGHT
AS/NZS 4600 Supp1:1998 94

FIGURE C5.2.3 FILLET WELD FAILURE MODES

C5.2.4.2 Shear and C5.2.4.3 Tearout The Cornell tests (Peköz and McGuire, 1979)
identified the following modes of failure for arc spot welds:
(a) Shear failure of welds in the fused area.
(b) Tearing of the sheet along the contour of the weld with the tearing spreading the
sheet at the leading edge of the weld.
(c) Sheet tearing combined with buckling near the trailing edge of the weld.
(d) Shearing of the sheet behind the weld.
Items (a), (b) and (c) are included in Clause 5.2.4.2 of the Standard and Item (d) is
included in Clause 5.2.4.3. It should be noted that many failures, particularly those of the
plate tearing type, may be preceded or accompanied by considerable inelastic out-of-plane
deformation of the type indicated in Figure C5.2.4.2. This form of behaviour is similar to
that observed in wide, pin-connected plates. Such behaviour should be avoided by closer
spacing of welds. When arc spot welds are used to connect two sheets to a framing
member as shown in Figure 5.2.4(1)(b), consideration should also be given to the possible
shear failure between thin sheets.
The thickness limitation of 3 mm is due to the range of the test program that served as the
basis of these provisions and the need to match with AS 4100 and NZS 3404. On sheets
below 0.711 mm thick, weld washers are required to avoid excessive burning of the sheets
and, therefore, inferior quality welds.
Accessed by BHP BILLITON on 27 Jun 2002

FIGURE C5.2.4.2 OUT-OF-PLANE DISTORTION OF WELDED CONNECTION

COPYRIGHT
95 AS/NZS 4600 Supp1:1998

C5.2.4.4 Tension For tensile capacity of arc spot welds, the design provisions in the
Standard are based on the tests reported by Fung (1978) and the study made by Albrecht
(1988). The provisions are limited to sheet failure with restrictive limitations on material
properties and sheet thickness. These design criteria were revised in the 1996 AISI
Specification because the recent tests conducted at the University of Missouri-Rolla
(LaBoube and Yu, 1991 and 1993) have shown that two potential limit states may occur.
The most common failure mode is that of sheet tearing around the perimeter of the weld.
This failure condition was found to be influenced by the sheet thickness, the average weld
diameter and the material tensile strength. In some cases, it was found that tensile failure
of the weld can occur. The strength of the weld was determined to be a function of the
cross-section of the fused area and tensile strength of the weld material. Tests (LaBoube
and Yu, 1991 and 1993) have also shown that when reinforced by a weld washer, thin
sheet weld connections can achieve the design strength calculated using
Equations E2.2.2-2 and E2.2.2-3 of the 1996 AISI Specification and using the thickness of
the thinner sheet. The provisions in the 1996 AISI Specification have not been included in
the Standard at this stage.
The equations given in the Standard were derived from the tests for which the applied
tension load imposed a concentric load on the weld, as would be the case, for example,
for the interior welds on a roof system subjected to wind uplift. Welds on the perimeter of
a roof or floor system would experience an eccentric tensile loading due to wind uplift.
Tests have shown that as much as a 50% reduction in nominal connection strength could
occur because of the eccentric load application (LaBoube and Yu, 1991 and 1993).
Eccentric conditions may also occur at connection laps shown in Figure C5.2.4.4.
At a lap connection between two deck sections as shown in Figure C5.2.4.4, the length of
the unstiffened flange and the extent of the encroachment of the weld into the unstiffened
flange have a measurable influence on the strength of the welded connection (LaBoube
and Yu, 1991). The 1996 AISI Specification recognizes the reduced capacity of this
connection detail by imposing a 30% reduction on the calculated nominal strength. This
requirement is not stated explicitly in the Standard but should be considered in accordance
with Clause 5.2.4.4 of the Standard.
Accessed by BHP BILLITON on 27 Jun 2002

FIGURE C5.2.4.4 INTERIOR WELD, EXTERIOR WELD AND LAP CONNECTION

C5.2.5 Arc seam welds The general behaviour of arc seam welds is similar to that of
arc spot welds. No simple shear failures of arc seam welds were observed in the Cornell
tests (Peköz and McGuire, 1979). Therefore, Equation 5.2.5.2(1) in the Standard, which
accounts for shear failure of welds, is adopted from the AWS welding provisions for sheet
steel (AWS, 1992).
Equation 5.2.5.2(2) in the Standard is intended to prevent failure by a combination of
tensile tearing plus shearing of the cover plates.

COPYRIGHT
AS/NZS 4600 Supp1:1998 96

C5.2.6 Flare welds The primary mode of failure in cold-formed steel sections welded
by flare welds, loaded transversely or longitudinally, also was found to be sheet tearing
along the contour of the weld (see Figure C5.2.6). The provisions of Clause 5.2.6 of the
Standard are intended to prevent shear tear failure. For sections thicker than 3 mm, the
provisions of AS 4100 and NZS 3404, as appropriate, should be used to check weld
failure.

FIGURE C5.2.6 FLARE GROOVE WELD FAILURE MODES

C5.2.7 Resistance welds The values of the nominal shear capacities given in
Table 5.2.7 of the Standard for outside sheets of 3.20 mm or less in thickness are based
on ‘Recommended Practice for Resistance Welding Coated Low-Carbon Steels’,
AWS C1.3-70, (Table 2.1 — Spot Welding Galvanized Low-Carbon Steel). The values of
the nominal shear capacities in Table 5.2.7 of the Standard for outside sheets thicker than
3.20 mm are based on ‘Recommended Practices for Resistance Welding’, AWS C1.1-66,
(Table 1.3 – Pulsation Welding Low-Carbon Steel) and apply to pulsation welding as well
as spot welding. They are applicable for all structural grades of low-carbon steel,
uncoated or galvanized with 275 grams/m 2 of sheet, or less, and are based on values
selected from Table 2.1 of AWS C1.3-70 and from Table 1.3 of AWS C1.1-66. Values for
intermediate thicknesses may be obtained by straight-line interpolation. The above values
may also be applied to medium carbon and low-alloy steels. Spot welds in such steels
give somewhat higher shear strengths than those upon which the above values are based;
however, they may require special welding conditions. In all cases, welding is to be
performed in accordance with AWS C1.3-70 and AWS C1.1-66 (AWS, 1966 and 1970).

C5.3 BOLTED CONNECTIONS The structural behaviour of bolted connections in


Accessed by BHP BILLITON on 27 Jun 2002

cold-formed steel construction is somewhat different from that in hot-rolled heavy


construction, mainly because of the thinness of the connected parts. Prior to 1980, the
provisions included in the AISI Specification and AS 1538 — 1974 for the design of
bolted connections were developed on the basis of the Cornell tests (Winter, 1956a,
1956b). These provisions were updated in 1980 (1988 in Australia) to reflect the results of
additional research performed in the United States (Yu, 1982) and to provide a better
coordination with the specifications of the Research Council on Structural Connections
(RCSC, 1980) and the AISC (1978). In AS 1538 — 1988, design provisions for the
maximum size of bolt holes were added.
The Standard gives special provisions for oversized and slotted holes. These provisions
differ for Australia and New Zealand, and are based on accepted practice in each country.

COPYRIGHT
97 AS/NZS 4600 Supp1:1998

The oversize and slotted holes described in Clauses 5.3.1(a) and (b) of the Standard for
Australia were used in the vacuum rig testing of purlins at the University of Sydney
(Hancock, Celeban and Healy (1993)).
The following provisions are taken into consideration:
(a) Scope Previous studies and practical experiences have indicated that the structural
behaviour of bolted connections used for joining relatively thick cold-formed steel
members is similar to that for connecting hot-rolled shapes and built-up members.
The criteria specified in Clause 5.3 of the Standard are applicable only to cold-
formed steel members or elements less than 3 mm in thickness. For materials greater
than or equal to 3 mm, reference is made to AS 4100 or NZS 3404, as appropriate.
Because of a lack of appropriate test data and the use of numerous surface
conditions, the Standard does not provide design criteria for slip-critical (also called
friction-type) connections. When such connections are used with cold-formed
members where the thickness of the thinnest connected part is less than 3 mm, it is
recommended that tests be conducted to confirm their design capacity. The test data
should verify that the specified design capacity for the connection provides a safety
against initial slip at least equal to that implied by the AS 4100 and NZS 3404
provisions. In addition, the reliability when calculating the ultimate capacity should
be greater than or equal to that implied by the Standard for bearing-type
connections.
These provisions apply only when there are no gaps between plies. The designer
should recognize that the connection of a rectangular tubular member by means of
bolt(s) through such members may have less strength than if no gap existed.
Structural performance of connections containing unavoidable gaps between plies
requires tests in accordance with Section 6.
(b) Materials Clause 1.5.3.1 of the Standard describes the different types of fasteners
that are normally used for cold-formed steel construction. During recent years, other
types of fasteners, with or without special washers, have been widely used as
connectors in steel structures using cold-formed steel members. The design of these
fasteners should be determined by tests in accordance with Section 6 of the
Standard.
(c) Bolt installation Bolted connections in cold-formed steel structures use either
mild- or high-strength steel bolts, and are designed as a bearing-type connection.
Bolt pretensioning is not required for attaining the design capacity because the
ultimate strength of a bolted connection is independent of the level of bolt preload.
It may be required to provide twist restraint to supporting members; however,
installation should ensure that the bolted assembly will not come apart during
service. For examples where this applies, see Clause 5.4.2 of NZS 3404. Experience
has shown that bolts installed do not loosen or back off, if tightened to a snug tight
condition under normal building conditions and are not subject to vibration or
fatigue.
Accessed by BHP BILLITON on 27 Jun 2002

Bolts in slip-critical connections should be tightened in a manner that assures the


development of the fastener tension forces required by the Research Council on
Structural Connections (1985 and 1988) for the particular size and type of bolts.
Turn-of-nut rotations specified in Section 15 of AS 4100 or NZS 3404 may not be
applicable because such rotations are based on larger grip lengths than are
encountered in usual cold-formed construction. Reduced turn-of-nut values would
have to be established for the actual combination of grip and bolt. A similar test
program (RCSC, 1985 and 1988) could establish a cut-off value for calibrated
wrenches.

COPYRIGHT
AS/NZS 4600 Supp1:1998 98

(d) Hole sizes In Table 5.3.1 of the Standard, the maximum size of holes for bolts
having diameters not less than 12 mm is based on Table 1 of the Research Council
on Structural Connections (1985 and 1988), except that for the oversized hole
diameter, a slightly larger hole diameter is permitted.
For bolts having diameters less than 12 mm, the diameter of a standard hole is the
diameter of bolt plus 1 mm.
Short-slotted holes are usually treated in the same manner as oversized holes.
Washers or backup plates should be used over oversized or short-slotted holes in an
outer ply unless suitable performance is demonstrated by tests.
The inclusion of the design information in Table 5.3.1 of the Standard for oversized
and slotted holes is because such holes are sometimes used in Australian practice to
meet dimensional tolerances during erection. However, when using oversized holes
care should be exercised by the designer to ensure that excessive deformation due to
slip will not occur at working loads. Excessive deformations, which can occur in the
direction of the slots, may be prevented by requiring bolt pretensioning, as specified
in Clause 5.4.2 of NZS 3404 for restraint provision.
C5.3.2 Tearout The provisions for minimum spacing and edge distance were revised
in the 1980 Specification and AS 1538 — 1988 to include additional design requirements
for bolted connections with standard, oversized and slotted holes. The minimum edge
distance of each individual connected part (emin.) is determined by using the tensile
strength of steel (fu) and the thickness of connected part. In accordance with the different
ranges of the fu/fy ratio, two different capacity factors are used for determining the
required minimum edge distance. These design provisions are based on the following
basic equation established from the test results:

. . . C5.3.2

where
e = the required minimum edge distance to prevent shear failure of the connected
part
Vf = force transmitted by one bolt
t = thickness of the thinnest connected part. For design purposes, a capacity
factor of 0.70 was used for fu/fy greater than or equal to 1.08, and 0.60 for
steel with fu/fy less than 1.08, in accordance with the degree of correlation
between Equation C5.3.2 and the test data. As a result, whenever fu/fy is less
than or equal to 1.08, the requirement is the same as the AISC specification.
In addition, several requirements were added to the AISI Specification in
1980 and AS 1538 — 1988 concerning the following:
(a) The minimum distance between centres of holes, as required for
installation of bolts.
Accessed by BHP BILLITON on 27 Jun 2002

(b) The clear distance between edges of two adjacent holes.


(c) The minimum distance between the edge of the hole and the end of the
member.
The same design provisions were retained in the 1986 AISI Specification and are also
used in the Standard, except that the limiting f u/fy ratio has been reduced from 1.15 to 1.08
for consistency with Clause 1.5.1.5 of the Standard. The test data used for the
development of Equation 5.3.2 of the Standard are documented by Winter (1956a and
1956b) and Yu (1982, 1985, and 1991).

COPYRIGHT
99 AS/NZS 4600 Supp1:1998

C5.3.3 Net section tension In the Standard, the nominal tensile capacity (Nf) on the
net section of connected parts is based on the loads determined in accordance with
Clauses 3.2 and 5.3.3 of the Standard, whichever is smaller. In the use of the equations
provided in Clause 5.3.3 of the Standard, the following design features should be noted:
(a) The provisions are applicable only to the thinnest connected part less than 3 mm in
thickness. For materials thicker than 3 mm, the design tensile force is determined in
accordance with AS 4100 or NZS 3404, as applicable.
(b) The nominal tensile strength (Nf) on the net section of a connected member is
determined by the tensile strength of the connected part (f u) and the ratios rf and
df/s f.
(c) Different equations are used for bolted connections with and without washers
(Chong and Matlock, 1974).
(d) The tensile capacity on the net section of a connected member is based on the type
of joint, either a single shear lap joint or a double shear butt joint.
C5.3.4 Bearing The available test data has shown that the bearing strength of bolted
connections depends on the following:
(a) The tensile strength and thickness of the connected parts.
(b) Joints with single shear or double shear conditions.
(c) The fu/fy ratio.
(d) The use of washers (Winter, 1956a and 1956b; Yu, 1982 and 1991; Chong and
Matlock, 1974).
The nominal bearing capacities between the connected parts for different conditions are
given in Tables 5.3.4.1 and 5.3.4.2 of the Standard. The capacity factors are provided in
the tables for different types of joints and f u/fy ratios. It should be noted that in the 1996
edition of the AISI Specification and the Standard, the limiting value of f u/fy used in
Tables 5.3.4.1 and 5.3.4.2 of the Standard was changed from 1.15 to 1.08 in order to be
consistent with Clause 1.5.1.5 of the Standard.
C5.3.5 Bolts The provisions for bolts in shear, tension and combined shear and tension
are the same as those for AS 4100 or NZS 3404. Reference should be made to
Clause C9.3.2 of the Commentaries to these Standards.

C5.4 SCREWED CONNECTIONS


C5.4.1 General The Standard covers cases where the loads applied to the connection
are either predominantly shear or normal tension (head pull). It is not intended to cover
cases where the connection will experience moments, combined loading or significant
secondary forces such as prying. For these cases, testing in accordance with Section 6 of
the Standard should be used. Testing can also be used where more accurate shear and
normal tension capacities are required.
Accessed by BHP BILLITON on 27 Jun 2002

Testing is particularly useful where —


(a) the thickness of cold-formed high strength G550 steel sections is less than 0.90 mm;
and
(b) the fu/fy ratios are 1.0 for 0.40 mm BMT to 1.08 for 0.90 mm BMT.
It is recommended that at least two screws should be used to connect individual
components. This provides redundancy against poor installation and limits lap shear
connection and distortion of flat unformed members, such as straps. The screws should be
installed in accordance with the manufacturer’s recommendations. Fine-threaded screws
perform better in thick material, where several threads will engage. Conversely,
coarse-threaded screws usually perform better in thin materials, especially where the
material thickness fits neatly between two threads.

COPYRIGHT
AS/NZS 4600 Supp1:1998 100

For connections of steels with low ductility, e.g. Grade 550 less than 0.9 mm thick, the
tensile strength (fu) used in Equations 5.4.2.3(1) to 5.4.3(2) of the Standard should be
taken as the lesser of 75% of the specified minimum tensile strength or 450 MPa as
specified in Clause 1.5.1.5 of the Standard. This reduction in fu does not apply if the
design capacity is determined by test in accordance with Clause 6.2 of the Standard.
It is intended that this reduction will provide a safety factor to prevent tensile failure. To
ensure ductility, it is considered preferable to provide some yielding at the connection,
although section buckling should occur before connection failure. Lighter sections usually
produce a more flexible structure which, although strong, can be subject to cyclic flexing
due to e.g. wind forces, which a heavier structure would absorb and resist without flexing.
The Standard applies only to screws with nominal diameters between 3 mm and 7 mm.
This reflects the range of screws used in the tests on which the predictive equations have
been based. Table C5.4.1 gives the nominal diameter of screws between 3 mm and 7 mm.

TABLE C5.4.1
NOMINAL DIAMETER FOR COMMON
SIZE SCREWS
Nominal diameter (df)
Size designation mm
(see Figure C5.4(1))
No. 5 3.2
No. 6 3.5
No. 7 3.8
No. 8 4.2
No. 10 4.8
No. 12 5.4
No. 14 6.3

FIGURE C5.4(1) NOMINAL DIAMETER FOR SCREWS

C5.4.2 Screwed connections in shear To ensure equal distribution of forces between


the screws in a connection, it is important to limit the maximum spacing between the
screws. This is particularly important for the outermost screws. The AISI Specification
gives no guidance on this, but the European Recommendations (ECCS 1983) specifies the
following, when the applied force is parallel to the rows of screws:
Accessed by BHP BILLITON on 27 Jun 2002

(a) Where the distance between the outermost screws is less than 15df, the force may be
distributed uniformly over the screws.
(b) Where the distance between the outermost screws is 65d f, the force in the
connection should be limited to 75% of the sum of the design strengths of the
screwed fastenings.
(c) For distances between 15d f and 65df, linear interpolation should be used.
Screwed connections loaded in shear can fail in one mode or in a combination of several
modes. These modes are screw shear, edge tearing, tilting and subsequent pull-out of the
screw, and bearing of the joined materials.

COPYRIGHT
101 AS/NZS 4600 Supp1:1998

Tilting of the screw followed by threads tearing out of the lower sheet reduces the
connection shear capacity from that of the typical connection bearing strength
(see Figure C5.4.2.3(1)).
The provisions in Clause 5.4.2.3 of the Standard focus on the tilting and bearing failure
modes. Two cases are given depending on the ratio of thicknesses of the connected
members. Normally, the head of the screw will be in contact with the thinner material as
shown in Figure C5.4.2.3(2). However, when both members are of the same thickness, or
when the thicker member is in contact with the screw head, tilting is also to be
considered, as shown in Figure C5.4.2.3(3).
It is necessary to determine the lower bearing capacity of the two members, based on the
product of their respective thicknesses and tensile strengths.

FIGURE C5.4.2.3(1) COMPARISON OF TILTING AND BEARING

FIGURE C5.4.2.3(2) DESIGN EQUATIONS FOR t 2/ t 1 ≥ 2.5


Accessed by BHP BILLITON on 27 Jun 2002

FIGURE C5.4.2.3(3) DESIGN EQUATIONS FOR t 2/ t 1 ≤ 1.0

COPYRIGHT
AS/NZS 4600 Supp1:1998 102

C5.4.2.4 Screws in shear To prevent the connection failing in a brittle manner, the
design shear capacity of the screw itself should be 1.25 times the design capacity due to
tilting and bearing. In general, the shear capacity of the screw itself will be approximately
0.6 times the axial tensile strength of the screw.
The shear values given by the manufacturer are not relevant where t2 is less than or equal
to 1.6 mm, where t2 is the thickness of the material not in contact with the head of the
screw.
C5.4.3 Screwed connections in tension
C5.4.3.1 Pull-out and pull-over (pull-through) Clause 5.4.3.1 of the Standard applies
to static loading conditions. For pull-over (pull-through) behaviour, the tensile strength
may be affected by repeated loading, such as cyclonic wind conditions in Australia and
high wind areas in New Zealand, such as wind regions I, V and VII specified in
NZS 4203. The AISI Specification gives no guidance on this, while the Eurocode
recommends using a cyclic loading factor of 0.5 for the calculated static design capacity.
Alternatively, testing to cyclic loading in accordance with AS 4040.3 may be considered.
The thickness of the washer, including where it is integrated into the head of the screw,
needs to be at least 1.3 mm or formed to an equivalent strength to withstand the bending
forces with little or no deformation. Washers having diameters larger than 12.5 mm can
be used. However, in this case the value of dw used in Equation 5.4.3(2) of the Standard is
limited to 12.5 mm maximum. Alternatively, tests in accordance with Section 6 of the
Standard may be considered if the full design capacity of larger washers is desired.
Design capacities for connections where the members are not in contact at the point of
fastening, such as crest fixing of cladding, have not been included as the design capacity
of the connection is dependent on the type of profile used (Mahendran 1994).
For non-cyclonic areas of Australia and protected inland areas of New Zealand, and for
the design of crest-fixed profile sheeting using No. 14 screws and 0.35−0.48 mm G550
steel without washers, the nominal pull-over (pull-through) capacity (N ov) can be
approximated as follows:

. . . C5.4.3.1
where
k2 = 1 for static loads (non-cyclonic areas)
k3 = 0.54 for corrugated sheeting
= 0.89 for wide pan trapezoidal sheeting
= 0.79 for narrow pan trapezoidal sheeting
t1 = thickness of the sheet in contact with the screw head
fu1 = tensile strength of the sheet in contact with the screw head
For non-drilling screws, the diameter of the hole of the member in contact with the screw
Accessed by BHP BILLITON on 27 Jun 2002

head should not exceed that recommended in AS B194. The minimum withdrawal axial
force for screws specified in AS 3566 is not relevant where t 2 is less than 1.6 mm, where
t2 is the thickness of the material not in contact with the head of the screw.
C5.4.3.2 Screws in tension To prevent the connections failing in a brittle manner, the
tensile design capacity [strength] of the screw should be 1.25 times the design capacity
due to pull-out and pull-over (pull-through). The maximum tensile capacities [strengths]
of self-drilling screws, as specified in AS 3566, are given in Table C5.4.3.2. The values
given in Table C5.4.3.2 are for the screws only and not the connection. The thickness and
grade of steel will generally determine the design value of the connection.

COPYRIGHT
103 AS/NZS 4600 Supp1:1998

TABLE C5.4.3.2
MINIMUM AXIAL TENSILE CAPACITY [STRENGTH]
OF SELF-DRILLING SCREWS

Maximum axial tensile capacity [strength]


Size designation kN
Type ASD Type BSD Type CSD
No. 6 4.35 4.35 5.33
No. 8 6.35 6.35 8.46
No. 10 7.50 8.60 10.01
No. 12 11.34 11.63 14.44
No. 14 14.95 16.15 18.90

The values given in Table C5.4.3.2 are not the design tensile capacity [strength] of the
screw connecting plies of thin gauge steel. The screw types are described in AS 3566.

C5.5 BLIND RIVETED CONNECTIONS Clause 5.5 of the Standard is based on


Eurocode 3, Part 1.3 (Toma et al. 1993) in which only shear is covered.
At least two rivets should be used to connect individual components. This provides
redundancy against poor installation and limits lap shear connection and distortion of flat
unformed members, such as straps. The rivets should be installed in accordance with the
manufacturer’s recommendations. Where sheets of different thicknesses are joined, it is
recommended that the preformed head be placed against the thinnest sheet.
To limit distortion and ensure equal distribution of forces, the maximum spacing between
rivets should be limited to 20 times the nominal diameter of the rivets.
For connections of steels with low ductility, e.g. Grade 550 less than 0.9 mm thick, the
tensile strength (fu1) used in Equations 5.5.2.3(1) to 5.5.2.3(5) of the Standard should be
taken as the lesser of 75% of the specified minimum tensile strength or 450 MPa,
whichever is the lesser, as specified in Clause 1.5.1.5 of the Standard. This reduction in f u
does not apply if the design capacity is determined by test in accordance with Clause 6.2
of the Standard.
C5.5.2.4 Rivets in shear To prevent the connection failing in a brittle manner, the
design shear capacity of the rivet itself should be 1.25 times the shear capacity due to
tilting and bearing. The minimum shear and axial tensile capacities [strengths] of blind
rivets, as specified in the Industrial Fasteners Institute Standard F114 (IFI 1988), are
given in Table C5.5.2.4.
The values given in Table C5.5.2.4 are not the design shear or tensile capacity [strength]
of the rivet connecting plies of thin gauge material.
Similar design considerations that are required for screws are required for blind rivets.
Accessed by BHP BILLITON on 27 Jun 2002

Blind rivets are manufactured in various shapes, sizes and materials but are all subject to
pull-over failure in tension because the head of the rivet is much smaller than a screw
head. The practical size for rivets used for structural purposes is, therefore, limited to
4.0 mm, 4.8 mm and 6.3 mm. To provide the shear and tension capacities quoted by the
manufacturer, it is essential to use the correct size drill (4.1 mm, 4.9 mm and 6.4 mm)
and the correct length of rivet for the materials being fastened. It is usual to use more
rivets than screws because the rivet head is not normally washered and consequently
lesser values are obtained with rivets. Their spacing should be greater than 3df and less
than 20df, to minimize distortion.
Galvanized steel rivets are plated only and are not considered durable except in a
protected environment.

COPYRIGHT
AS/NZS 4600 Supp1:1998 104

TABLE C5.5.2.4
MINIMUM SHEAR AND AXIAL TENSILE CAPACITY [STRENGTH]
OF BREAK MANDREL BLIND RIVETS

Nominal Minimum shear capacity Minimum axial tensile capacity


Nominal
head [strength] [strength]
diameter
diameter (kN) (kN)
(mm) (mm) A1 St M S/S A1 St M S/S
4.0 8.0 1.3 1.7 2.5 3.0 1.62 2.17 3.24 3.8
4.8 9.5 1.6 2.5 3.7 4.4 2.3 3.15 4.63 5.5
6.3 12.7 3.2 4.63 6.48 7.8 4.26 5.74 0.56 9.72

where
A1 = aluminium alloy, 5154, 5056, 5754 with carbon steel mandrel
St = low carbon steel with carbon steel mandrel
M = nickel-copper alloy (Monel) with carbon steel mandrel
S/S = 300 series stainless steel with carbon or stainless steel mandrel

C5.6 RUPTURE
C5.6.1 Shear rupture Connection tests conducted by Birkemoe and Gilmor (1978)
have shown that on coped beams a tearing failure mode as shown in Figure C5.6.1(1) can
occur along the perimeter of the holes. The provisions in Clause 5.6.1 of the Standard for
shear rupture are adopted from the AISC Specification (AISC, 1978). These provisions are
considered to yield a conservative estimate of the rupture capacity at the coped end of a
beam by neglecting the contribution of the tensile area. For additional design information
on tension rupture strength and block shear rupture strength of connections (see
Figures C5.6.1(1) and C5.6.1(2)), refer to the AISC Specifications (AISC, 1989 and
1993).
Accessed by BHP BILLITON on 27 Jun 2002

FIGURE C5.6.1(1) FAILURE MODES FOR BLOCK SHEAR RUPTURE

COPYRIGHT
105 AS/NZS 4600 Supp1:1998

FIGURE 5.6.1(2) BLOCK SHEAR RUPTURE IN TENSION

C5.6.3 Block shear rupture Block shear is a limit state in which the resistance is
determined by the sum of the shear capacity on a failure path(s) and the tensile strength
on a perpendicular segment. A comprehensive test program does not exist regarding block
shear for cold-formed steel members. However, a limited unpublished study conducted at
the University of Missouri-Rolla indicates that the AISC LRFD equations may be applied
to cold-formed steel members.
Accessed by BHP BILLITON on 27 Jun 2002

COPYRIGHT
AS/NZS 4600 Supp1:1998 106

S E C T I O N C 6 T E S T I N G

C6.1 TESTING FOR DETERMINING MATERIAL PROPERTIES


C6.1.1 Testing of unformed steel The Clause applies to steel classified as ‘other steel’
in Clause 1.5.1.2 of the Standard, when used in formed sections for which the strength
increase from cold-forming is calculated from the unformed properties in accordance with
Clause 1.5.1.3 of the Standard.
C6.1.2 Compression testing Clause 6.1.2 is an exact copy from AS 1538 — 1988. This
test is required for the application of Clauses 1.5.1.3 and 6.1.3 of the Standard.
C6.1.3 Testing of full sections Clause 6.1.3 is an exact copy from AS 1538 — 1988.
C6.1.4 Testing of flat coupons of formed members Two procedures are provided
depending on the purpose of the test. If the purpose of the test is to assess the strength
increase resulting from cold-forming for the application of Clause 1.5.1.3 of the Standard,
then Clause 6.1.4.1 of the Standard is applicable. Alternatively, if the purpose of the test
is to determine the design properties directly from flat coupons of formed members, then
Clause 6.1.4.2 of the Standard is applicable.
C6.1.5 Testing for determining section properties The Clause applies to the
determination of section properties for sections whose properties may be difficult to
calculate, such as those with single-point fasteners, e.g. clinches and the like.
C6.1.6 Testing of single-point fastener connections New standard tests for single-
point fastener connections are given in Appendix G of the Standard. The purpose of the
tests is to allow comparison of performance of different types of single-point fasteners
and they were used to derive the predictive equations for the performance of screwed
connection (see Clauses 5.4.2 and 5.4.3 of the Standard). The results of these tests are not
necessarily applicable to connections that have more complex geometry or use multiple
fasteners. In these situations, prototype testing in accordance with Clause 6.2 of the
Standard is recommended.

C6.2 TESTING FOR ASSESSMENT OR VERIFICATION


C6.2.1 General The Clause provides an alternative method for assessment or
verification in place of calculation. This method is uniform across all Australian/New
Zealand structural Standards and is different to the equivalent AISI Specification. Both
methods, however, produce similar outcomes.
The test method of the Clause is generally known as prototype testing, i.e. the testing of a
small number of samples of a product for the assessment or verification of a much larger
population of a nominally identical product.
The method is material independent and is, therefore, suitable for use with composite
products. It is most advantageous for products that are produced in a large number of
Accessed by BHP BILLITON on 27 Jun 2002

identical units with low variation in structural characteristics but difficult to assess by
calculation (see Clause 6.2.2.3 of the Standard).
At present, only assessment of static behaviour (for strength, or serviceability, or both) is
included in the Standard. Procedures for assessing cyclic performance under cyclonic
loading (for cyclonic areas of Australia) and seismic loading (for New Zealand) can be
obtained from the following publications:
(a) For cyclonic or high wind assessment TR440 (Department of Construction 1978),
TR5 (James Cook University, 1980).
(b) For seismic assessment NZS 4203, Volume 2, Appendix C4A or BRANZ P21.

COPYRIGHT
107 AS/NZS 4600 Supp1:1998

C6.2.2 Static tests for strength or serviceability


C6.2.2.1 Test specimens The requirement that ‘test specimens shall be nominally
identical to the class of units for which structural verification is required’ is applicable
only to the component under test. It is permissible to temporarily strengthen those parts of
the structure that are not under test. However, if such strengthening is carried out, care
should be taken to ensure that the component under test receives the correct loading
without artificial restraint or other form of strengthening that will not exist in the real
structure. For example, if trusses are tested to verify the performance of the truss apex
connection, then only the apex connection (including the members that connect to it) has
to be ‘identical’ to production units. Outside the joint area, the members may be
strengthened in order to produce the failure at the connection. On the other hand, if
trusses are tested to verify the behaviour of the trusses, then the whole test truss should be
identical to the production units.
C6.2.2.2 Test loads The determination of the factor (k t), given in Table 6.2.2 of the
Standard, is designed so that the probability that any production unit will not have the
required performance is 50%, with a 75% confidence. Interpolation is permitted in the use
of Table 6.2.2. However, it is not realistic to expect that the coefficient of structural
characteristics can be determined to 2.5% accuracy (see Clause C6.2.2.3).
C6.2.2.3 Coefficient of variation of structural characteristics If the samples under test
are sufficiently representative of the total population, then the coefficient of variation of
structural characteristics can be obtained directly from the test data. However, in most
practical cases, the samples are not likely to be representative of the total population of
the product. Therefore, the assumed coefficient of variation of the structural
characteristics is usually larger than that obtained from the test data. For cold-formed steel
components and assemblies, the AISI Specification gives the following guidance on the
coefficient of variation for material (km) and fabrication (kf):
(a) km = 0.06 − 0.10.
(b) kf = 0.05 − 0.15.
The following values for coefficients of variation of structural characteristics should be
used unless demonstrated otherwise:
(i) Strength of members . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10%.
(ii) Strength of connections and assemblies . . . . . . . . . . . . . . . . . . . . . . . . . . . 15%.
(iii) Stiffness of members . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5%.
(iv) Stiffness of connections and assemblies . . . . . . . . . . . . . . . . . . . . . . . . . . . 10%.
C6.2.2.4 Test requirements The method by which the loading should be applied to the
unit that is to be tested, and the positions at which deflections should be measured, can
only be decided with special reference to the particular structure or element and to the
particular loading conditions that are to be investigated.
The test loading should be applied and resisted in a manner which reasonably
Accessed by BHP BILLITON on 27 Jun 2002

approximates the actual service conditions. Although, in general, prototype testing is


likely to involve loading in a vertical plane, additional out-of-plane loading of a structure
or element to simulate other load effects may be required. Horizontal support to the unit
as a whole or to individual members of the unit should also represent as closely as
possible actual service conditions.
Any eccentricities not inherent in the design of the structure or element, or not resulting
from typical loading in service, should be avoided at points of loading and reaction, and
care should also be taken to ensure that no inadvertent restraints are present. Where it is
clear that the method of test involves a significant or appreciable divergence from service
conditions, either in loading or restraint, due allowance should be made in determining the
design strength or stiffness from the test results.

COPYRIGHT
AS/NZS 4600 Supp1:1998 108

All likely combinations of permanent loads and imposed loads of shorter duration,
including those due to wind and, where applicable, those due to impact, should be taken
into account when determining the worst loading conditions.
A load-deflection curve should be plotted during each test on each unit. Such a curve will
serve not only as a check against observational errors, but also to indicate any
irregularities in the behaviour under load of the structure or element. It is desirable that a
minimum of six points, not including the zero load point, be obtained to define the shape
of the load-deflection curve if the latter is predominantly linear, and a minimum of
10 points if the curve is significantly non-linear.
It is usual practice to give the test specimen a few cycles of preliminary loading and to
apply a small amount of pre-load, such as 20%, at the beginning of a test to obtain more
reliable deformation readings.
C6.2.2.5 Criteria for acceptance The following criteria of acceptance should be
considered:
(a) Acceptance for static strength The acceptance criterion is based on the minimum
result of the tests. Abnormally low results can be discarded provided that the cause
of the abnormality is identified (e.g. testing machine not functioning properly,
incorrect test set up and the like) and judged irrelevant to the purpose of the test.
(b) Acceptance for serviceability The 95% recovery limit on deformation is to ensure
that no extensive yielding of the product will occur under service loading.
Consideration should be given to the extent of yielding provided by the ductility of the
connection and to ensure that this distortion does not exceed serviceability requirements
suitable for the structure. Deflections, even within elastic limits, may determine the
suitable capacity of a connection.
C6.2.2.6 Test report By its very nature, prototype test procedure cannot be
standardized. It is, therefore, important that proper adequate test reports be made
available.
C6.2.2.7 Design capacity of specific products and assemblies There is no need to apply
the capacity factor to the result of the Clause because it is included in the factor k t.
Accessed by BHP BILLITON on 27 Jun 2002

COPYRIGHT
109 AS/NZS 4600 Supp1:1998

APPENDIX A
REFERENCED DOCUMENTS

1 Albrecht, R.E. (1988), Developments and Future Needs in Welding Cold-formed


Steel, Proceedings Ninth International Specialty Conference on Cold-formed Steel
Structures, University of Missouri-Rolla, Rolla, MO, November 1988.
2 Allen, D.E. and Murray, T.M. (1993), Designing Criterion for Vibrations Due to
Walking, Engineering Journal, AISC, Fourth Quarter, 1993.
3 American Institute of Steel Construction (1978), Specification for the Design,
Fabrication and Erection of Structural Steel for Buildings, Chicago, IL,
November 1978.
4 American Institute of Steel Construction (1986), Load and Resistance Factor Design
Specification For Structural Steel Buildings, Chicago, IL, 1986.
5 American Institute of Steel Construction (1989), Specification for Structural Steel
Buildings — Allowable Stress Design and Plastic Design, Chicago, IL, 1989.
6 American Institute of Steel Construction (1993), Load and Resistance Factor Design
Specification for Structural Steel Buildings, Chicago, IL, December 1993.
7 American Iron and Steel Institute (1946), Specification for the Design of Light
Gauge Steel Structural Members, New York, NY, 1946.
8 American Iron and Steel Institute (1968), Specification for the Design of Light
Gauge Structural Members, New York, 1968.
9 American Iron and Steel Institute (1983), Cold-formed Steel Design Manual, (Part I,
Specification, 1980 Edition, Part II, Commentary, Part II, Supplementary
Information, Part IV, Illustrative Examples, Part V, Charts and Tables),
Washington D.C., 1983.
10 American Iron and Steel Institute (1986), Cold-formed Steel Design Manual, (Part I,
Specification, 1986 Edition with the 1989 Addendum, Part II, Commentary, 1986
Edition with the 1989 Addendum, Part III, Supplementary Information, Part IV,
Illustrative Examples, Part V, Charts and Tables, Part VI, Computer Aids, Part VII,
Test Procedures), Washington D.C., 1986.
11 American Iron and Steel Institute (1991), LRFD Cold-formed Steel Design Manual,
(Part I, Specification, Part II, Commentary, Part III, Supplementary Information,
Part IV, Illustrative Examples, Part V, Charts and Tables, Part VI, Computer Aids,
Part VII, Test Procedures), Washington D.C., 1991.
12 American Iron and Steel Institute (1995), Design Guide for Cold-formed Steel
Accessed by BHP BILLITON on 27 Jun 2002

Trusses, Publication RG-95-18, Washington D.C., 1995.


13 American Iron and Steel Institute (1996), Specification for the Design of Cold-
formed Steel Structural Members, Publication CF 96-1, Washington D.C., 1996.
14 American Society of Civil Engineers, ASCE (1995), Minimum Design Loads for
Buildings and Other Structures, ASCE Standard 7-95, 1995.
15 American Welding Society (1966), Recommended Practice for Resistance Welding,
AWS C1.1-66, Miami, FL, 1966.
16 American Welding Society (1970), Recommended Practice for Resistance Welding
Coated Low Carbon Steels, AWS C1.3-70, (Reaffirmed 1987), Miami, FL, 1970.

COPYRIGHT
AS/NZS 4600 Supp1:1998 110

17 American Welding Society (1989), Structural Welding Code — Sheet Steel,


ANSI/AWS D1.3-89, Miami, FL, 1989.
18 Ayrton, W.E. and Perry, J. (1986), On Struts, The Engineer, Vol. 62, 1986.
19 Bernard, E.S., Bridge, R.Q. and Hancock, G.J. (1996), Flange Curling in Profiled
Steel Decks, Thin-walled Structures, Vol. 25, No.1, 1996.
20 Bhakta, B.H., LaBoube, R.A. and Yu, W.W. (1992), The Effect of Flange Restraint
on Web Crippling Strength, Final Report, Civil Engineering Study 92-1, University
of Missouri-Rolla, Rolla, MO, March 1992.
21 Birkemoe, P.C. and Gilmor, M.I. (1978), Behaviour of Bearing — Critical Double-
angle Beam Connections, Engineering Journal, AISC, Fourth Quarter, 1978.
22 Bleich, F. (1952), Buckling Strength of Metal Structures, McGraw-Hill Book Co.,
New York, NY, 1952.
23 British Standards Institution (1992), British Standard: Structural Use of Steelwork
in Building, Part 5: Code of Practice for Design of Cold-formed Sections,
BS 5950: Part 5, CF92-2, 1992.
24 Brockenbrough (1995), Fastening of Cold-formed Steel Framing, AISI,
Washington D.C., September 1995.
25 Bulson, P.S. (1969), The Stability of Flat Plates, American Elsevier Publishing
Company, New York, NY, 1969.
26 Cain, D.E., LaBoube, R.A. and Yu, W.W. (1995), The Effect of Flange Restraint on
Web Crippling Strength of Cold-formed Steel Z- and I-sections, Final Report, Civil
Engineering Study 95-2, University of Missouri-Rolla, Rolla, MO, May 1995.
27 Chajes, A., Britvec, S.J. and Winter, G. (1963), Effects of Cold-straining on
Structural Steels, Journal of the Structural Division, ASCE, Vol. 89, No. ST2,
February 1963.
28 Chajes, A. and Winter, G. (1965), Torsional-Flexural Buckling of Thin-walled
Members, Journal of the Structural Division, ASCE, Vol. 91, No. ST4,
August 1965.
29 Chajes, A., Fang, P.J. and Winter, G. (1966), Torsional-Flexural Buckling, Elastic
and Inelastic, of Cold-formed Thin-walled Columns, Engineering Research Bulletin,
No. 66-1, Cornell University, 1966.
30 Chong, K.P. and Matlock, R.B. (1974), Light Gauge Steel Bolted Connections
without Washers, Journal of the Structural Division, ASCE, Vol. 101, No. ST7,
July 1974.
31 Cohen, J.M. and Peköz, T.B. (1987), Local Buckling Behaviour of Plate Elements,
Research Report, Cornell University, 1987.
32 Department of Construction (1978), Guidelines for the Testing and Evaluation of
Accessed by BHP BILLITON on 27 Jun 2002

Products for Cyclonic-prone Areas, Technical Record 440, February 1978.


33 Desmond, T.P., Peköz, T.B. and Winter, G. (1981a), Edge Stiffeners for
Thin-walled Members, Journal of the Structural Division, ASCE, Vol. 107,
No. ST2, February 1981.
34 Desmond, T.P., Peköz, T.B. and Winter, G. (1981b), Intermediate Stiffeners for
Thin-walled Members, Journal of the Structural Division, ASCE, Vol. 107,
No. ST4, April 1981.
35 DeWolf, J.T., Peköz, T.B. and Winter, G. (1974), Local and Overall Buckling of
Cold-formed Steel Members, Journal of the Structural Division, ASCE, Vol. 100,
October 1974.

COPYRIGHT
111 AS/NZS 4600 Supp1:1998

36 Dhalla, A.K., Errera, S.J. and Winter, G. (1971), Connections in Thin Low-ductility
Steels, Journal of the Structural Division, ASCE, Vol. 97, No. ST10, October 1971.
37 Dhalla, A.K. and Winter, G. (1974a), Steel Ductility Measurements, Journal of the
Structural Division, ASCE, Vol. 100, No. ST2, February 1974.
38 Dhalla, A.K. and Winter, G. (1974b), Suggested Steel Ductility Requirements,
Journal of the Structural Division, ASCE, Vol. 100, No. ST2, February 1974.
39 Douty, R.T. (1962), A Design Approach to the Strength of Laterally Unbraced
Compression Flanges, Bulletin No. 37, Cornell University, Ithaca, NY, 1962.
40 Ellifritt, D.S., Sputo, T. and Haynes, J. (1992), Flexural Capacity of Discretely
Braced C’s and Z’s, Proceedings Eleventh International Specialty Conference on
Cold-formed Steel Structures, St. Louis, MO, 1992.
41 Ellingwood, B., Galambos, T.V., MacGregor, J.G. and Cornell, C.A. (1980),
Development of a Probability Based Load Criterion for American National Standard
A58: Building Code Requirements for Minimum Design Loads in Buildings and
Other Structures, U.S. Department of Commerce, National Bureau of Standards,
NBS Special Publication 577, June 1980.
42 Ellingwood, B., MacGregor, J.G., Galambos, T.V. and Cornell, C.A. (1982),
Probability Based Load Criteria: Load Factors and Load Combinations, Journal of
the Structural Division, ASCE, Vol. 108, No. ST5, May 1982.
43 Ellingwood, B. (1989), Serviceability Guidelines for Steel Structures, Engineering
Journal, AISC, First Quarter, 1989.
44 European Committee for Standardisation (1996), Eurocode 3: Design of Steel
Structures, Part 1.3: General Rules — Supplementary Rules for Cold Formed Thin
Gauge Members and Sheeting, Brussels, European Prestandard, ENV 1993-1-3,
1996.
45 European Convention for Constructional Steelwork (1983), European
Recommendations for the Design and Testing of Connections in Steel Sheeting and
Sections, Publication No. 21, European Convention for Constructional Steelwork,
1983.
46 European Convention for Constructional Steelwork (1987), European
Recommendations for the Design of Light Gauge Steel Members, First Edition,
Brussels, Belgium, 1987.
47 Fisher, J.M. and West, M.A (1990), Serviceability Design Considerations for Low-
rise Buildings, Steel Design Guide Series, AISC, 1990.
48 Fung, C. (1978), Strength of Arc-Spot Welds in Sheet Steel Construction, Final
Report to Canadian Steel Industries Construction Council (CSICC), Westeel-Rosco
Limited, Canada, 1978.
49 Galambos, T.V. (1963), Inelastic Buckling of Beams, Journal of the Structural
Accessed by BHP BILLITON on 27 Jun 2002

Division, ASCE, Vol. 89, No. ST5, October 1963.


50 Galambos, T.V., Ellingwood, B., MacGregor, J.G. and Cornell, C.A. (1982),
Probability Based Load Criteria: Assessment of Current Design Practice, Journal of
the Structural Division, ASCE, Vol. 108, No. ST5, May 1982.
51 Galambos, T.V. (Editor) (1988), Guide to Stability Design Criteria for Metal
Structures, Fourth Edition, John Wiley and Sons, New York, NY, 1988.
52 Glaser, N.J., Kaehler, R.C. and Fisher, J.M. (1994), Axial Load Capacity of Sheeted
C and Z Members, Proceedings Twelfth International Specialty Conference on Cold-
formed Steel Structures, St. Louis, MO, 1994.

COPYRIGHT
AS/NZS 4600 Supp1:1998 112

53 Green, G.G., Winter, G. and Cuykendall, T.R. (1947), Light Gauge Steel Columns
in Wall-braced Panels, Bulletin, No. 35/2, Cornell University Engineering
Experimental Station, 1947.
54 Hancock, G.J. (1985), Distortional Buckling of Steel Storage Rack Columns,
Journal of Structural Engineering, ASCE, Vol. 111, No. 12, 1985.
55 Hancock, G.J., Celeban, M. and Healy, C. (1993), Behaviour of Purlins with Screw-
fastened Sheeting under Wind Uplift and Downwards Loading, Civil Engineering
Transactions, Institute of Engineers, Australia, Vol. CE35, No. 3, 1993.
56 Hancock, G.J. (1998), Design of Cold-formed Steel Structures, 3rd edition, Sydney,
Australian Institute of Steel Construction, 1998.
57 Hancock, G.J., Celeban, M. and Healy, C. (1994), Tests of Purlins with Concealed
Fixed Sheeting, Proceedings 12th International Specialty Conference on Cold-
formed Structures, St. Louis, MO, 1994.
58 Hancock, G.J., Kwon, Y.B. and Bernard, E.S. (1994), Strength Design Curves for
Thin-walled Sections Undergoing Distortional Buckling, Journal of Constructional
Steel Research, Vol. 31, 1994.
59 Hancock, G.J., Rogers, C.A. and Schuster, R.M. (1996), Design of Thin-walled
Beams for Distortional Buckling, Proceedings, 13th International Specialty
Conference on Cold-formed Steel Structures, St. Louis, MO, 1996.
60 Hancock, G.J. (1997), Design for Distortional Buckling of Flexural Members, Thin-
walled Structures, Vol. 27, No. 1, 1997.
61 Harper, M.M., LaBoube, R.A. and Yu, W.W. (1995), Behaviour of Cold-formed
Steel Roof Trusses, Summary Report, Civil Engineering Study 95-3, University of
Missouri-Rolla, Rolla, MO, May 1995.
62 Hatch, J., Easterling, W.S. and Murray, T.M. (1990), Strength Evaluation of
Strut-purlins, Proceedings of the Tenth International Specialty Conference on
Cold-formed Steel Structures, University of Missouri-Rolla, October 1990.
63 Haussler, R.W. (1964), Strength of Elastically Stabilized Beams, Journal of
Structural Division, ASCE, Vol. 90, No. ST3, June 1964; also ASCE Transactions,
Vol. 130, 1965.
64 Haussler, R.W. and Pahers, R.F. (1973), Connection Strength in Thin Metal Roof
Structures, Proceedings Second Specialty Conference on Cold-formed Steel
Structures, St. Louis, MO, 1973.
65 Haussler, R.W. (1988), Theory of Cold-formed Steel Purlin/Girt Flexure,
Proceedings Ninth International Specialty Conference on Cold-formed Steel
Structures, St. Louis, MO, 1988.
66 HERA report R4-92.
67 Hetrakul, N. and Yu, W.W. (1978), Structural Behaviour of Beam Webs Subjected
Accessed by BHP BILLITON on 27 Jun 2002

to Web Crippling and a Combination of Web Crippling and Bending, Final Report,
Civil Engineering Study 78-4, University of Missouri-Rolla, Rolla, MO, June 1978.
68 Hetrakul, N. and Yu, W.W. (1980), Cold-formed Steel I-beams Subjected to
Combined Bending and Web Crippling, Thin-walled Structures — Recent Technical
Advances and Trends in Design, Research and Construction, Rhodes, J. and
A.C. Walker (Eds.), Granada Publishing Limited, London, 1980.
69 Hsiao, L.E., Yu, W.W. and Galambos, T.V. (1988), Load and Resistance Factor
Design of Cold-formed Steel: Calibration of the AISI Design Provisions, Ninth
Progress Report, Civil Engineering Study 88-2, University of Missouri-Rolla, Rolla,
MO, February 1988.

COPYRIGHT
113 AS/NZS 4600 Supp1:1998

70 Hsiao, L.E. (1989), Reliability Based Criteria for Cold-formed Steel Members,
Thesis presented to the University of Missouri-Rolla, Rolla, Missouri, in partial
fulfilment of the requirements for the Degree of Doctor of Philosophy, 1989.
71 Hsiao, L.E., Yu, W.W. and Galambos, T.V. (1990), AISI LRFD Method for Cold-
formed Steel Structural Member, Journal of Structural Engineering, ASCE,
Vol. 116, No. 2, February 1990.
72 IFI (1988), Fastener Standards, Industrial Fasteners Institute Standard F114, 1988.
73 James Cook University (1980), Recommendations for the Testing of Roofs and Walls
to Resist High Wind Forces, Tech. Rep. No. 5, Cyclonic Testing Station, June 1980.
74 Johnston, N. and Hancock, G.J. (1993), Calibration of the AISI R-Factor Design
Approach for Purlins using Australian Test Data, Engineering Structures, Vol. 16,
No. 5, 1993.
75 Kalyanaraman, V., Peköz, T. and Winter, G. (1977), Unstiffened Compression
Elements, Journal of the Structural Division, ASCE, Vol. 103, No. ST9,
September 1977.
76 Kalyanaraman, V. and Peköz, T. (1978), Analytical Study of Unstiffened Elements,
Journal of the Structural Division, ASCE, Vol. 104, No. ST9, September 1978.
77 Karren, K.W. (1967), Corner Properties of Cold-formed Steel Shapes, Journal of the
Structural Division, ASCE, Vol. 93, No. ST1, February, 1967.
78 Karren, K.W. and Winter, G. (1967), Effects of Cold-work on Light Gauge Steel
Members, Journal of the Structural Division, ASCE, Vol. 93, No. ST1,
February 1967.
79 Kavanagh, K.T. and Ellifritt, D.S. (1993), Bracing of Cold-formed Channels Not
Attached to Deck or Sheeting, Is Your Building Suitably Braced?, Structural
Stability Research Council, April 1993.
80 Kavanagh, K.T. and Ellifritt, D.S. (1994), Design Strength of Cold-formed Channels
in Bending and Torsion, Journal of Structural Engineering, ASCE, Vol. 120, No. 5,
May 1994.
81 Kian, T. and Peköz, T.B. (1994), Evaluation of Industry-type Bracing Details for
Wall Stud Assemblies, Final Report, submitted to American Iron and Steel Institute,
Cornell University, January 1994.
82 Kirby, P.A. and Nethercot, D.A. (1979), Design for Structural Stability, John Wiley
and Sons, Inc., New York, NY, 1979.
83 Kwon, Y.B. and Hancock, G.J. (1992a), Design of Channels against Distortional
Buckling, 11th International Specialty Conference on Cold-formed Steel Structures,
St. Louis, MO, 1992.
84 Kwon, Y.B. and Hancock, G.J. (1992b), Strength Tests of Cold-formed Channel
Sections undergoing Local and Distortional Buckling, Journal of Structural
Accessed by BHP BILLITON on 27 Jun 2002

Engineering, ASCE, Vol. 117, No. 2, 1992.


85 LaBoube, R.A. and Yu, W.W. (1978a), Structural Behaviour of Beam Webs
Subjected Primarily to Shear Stress, Final Report, Civil Engineering Study 78-2,
University of Missouri-Rolla, Rolla, MO, June 1978.
86 LaBoube, R.A. and Yu, W.W. (1978b), Structural Behaviour of Beam Webs
Subjected to a Combination of Bending and Shear, Final Report, Civil Engineering
Study 78-3, University of Missouri-Rolla, Rolla, MO, June 1978.
87 LaBoube, R.A. and Yu, W.W. (1982), Bending Strength of Webs of Cold-formed
Steel Beams, Journal of the Structural Division, ASCE, Vol. 108, No. ST7,
July 1982.

COPYRIGHT
AS/NZS 4600 Supp1:1998 114

88 LaBoube, R.A. (1983), Laterally Unsupported Purlins Subjected to Uplift, Final


Report, Metal Building Manufacturers Association, 1983.
89 LaBoube, R.A. (1986), Roof Panel to Purlin Connection: Rotational Restraint
Factor, Proceedings IABSE Colloquium on Thin-walled Metal Structures in
Buildings, Stockholm, Sweden, 1986.
90 LaBoube, R.A., Golovin, M., Montague, D.J., Perry, D.C. and Wilson, L.L. (1988),
Behaviour of Continuous Span Purlin System, Proceedings Ninth International
Specialty Conference on Cold-formed Steel Structures, St. Louis, MO, 1988.
91 LaBoube, R.A. and Yu, W.W. (1991), Tensile Strength of Welded Connections,
Final Report, Civil Engineering Study 91-3, University of Missouri-Rolla, Rolla,
MO, June 1991.
92 LaBoube, R.A. and Yu, W.W. (1993), Behaviour of Arc Spot Weld Connections in
Tension, Journal of Structural Engineering, ASCE, Vol. 119, No. 7, July 1993.
93 Lau, S.C.W. and Hancock, G.J. (1987), Distortional Buckling Formulas for Channel
Columns, Journal of Structural Engineering, ASCE, Vol. 113, No. 5, May 1987.
94 Lau, S.C.W. and Hancock, G.J. (1988a), Distortional Buckling of Tests of Cold-
formed Channel Sections, 9th International Specialty Conference on Cold-formed
Steel Structures, St. Louis, MO, 1998.
95 Lau, S.C.W. and Hancock, G.J. (1988b), Strength Tests and Design Methods for
Cold-formed Channel Columns Undergoing Distortional Buckling, Research Report
R579, School of Civil and Mining Engineering, University of Sydney, 1988.
96 Lau, S.C.W. and Hancock, G.J. (1990), Inelastic Buckling of Channel Columns in
the Distortional Mode, Thin-walled Structures, Vol. 10, 1990.
97 Leicester, R.H., Pham, L. and Kleeman, P.W. (1985), Use of Reliability Concepts in
the Conversion of Codes to Limit States Design, Civil Engineering Transactions,
Institute of Engineers, Australia, Vol. CE27, No. 1, 1985.
98 Lutz, L.A. and Fisher, J.M. (1985), A Unified Approach for Stability Bracing
Requirements, Engineering Journal, AISC, 4th Quarter, Vol. 22, No. 4, 1985.
99 Macadam, J.N., Brockenbrough, R.L., LaBoube, R.A., Peköz, T. and Schneider, E.J.
(1988), Low-strain-hardening Ductile-steel Cold-formed Members, Proceedings of
the Ninth International Specialty Conference on Cold-formed Steel Structures,
University of Missouri-Rolla, Rolla, MO, November 1988.
100 Mahendran, M. (1994), Crest-fixed Profiled Steel Roof Claddings, Engineering
Structures, Vol. 16, No. 5, 1994.
101 Mulligan, G.P. and Peköz, T.B. (1984), Locally Buckled Thin-walled Columns,
Journal of the Structural Division, ASCE, Vol. 110, No. ST11, November 1984.
102 Murray, T.M. and Elhouar, S. (1985), Stability Requirements of Z-purlin Supported
Conventional Metal Building Roof Systems, Annual Technical Session Proceedings,
Accessed by BHP BILLITON on 27 Jun 2002

Structural Stability Research Council, 1985.


103 Murray, T.M. (1991), Building Floor Vibrations, Engineering Journal, AISC, Third
Quarter, 1991.
104 Nguyen, P. and Yu, W.W. (1978a), Structural Behaviour of Transversely Reinforced
Beam Webs, Final Report, Civil Engineering Study 78-5, University of Missouri-
Rolla, Rolla, MO, July 1978.
105 Nguyen, P. and Yu, W.W. (1978b), Structural Behaviour of Longitudinally
Reinforced Beam Webs, Final Report, Civil Engineering Study 78-6, University of
Missouri-Rolla, Rolla, MO, July 1978.

COPYRIGHT
115 AS/NZS 4600 Supp1:1998

106 Ortiz-Colberg, R. and Peköz, T.B. (1981), Load Carrying Capacity of Perforated
Cold-formed Steel Columns, Research Report, No. 81-12, Cornell University,
Ithaca, NY, 1981.
107 Papangelis, J.P. and Hancock, G.J. (1995), Computer Analysis of Thin-walled
Structural Members, Computers and Structures, Vol. 56, No. 1, 1995.
108 Peköz, T.B. and Winter, G. (1969), Torsional-flexural Buckling of Thin-walled
Sections under Eccentric Load, Journal of the Structural Division, ASCE, Vol. 95,
No. ST5, May 1969.
109 Peköz, T.B. and McGuire, W. (1979), Welding of Sheet Steel, Report SG-79-2,
American Iron and Steel Institute, January 1979.
110 Peköz, T.B. and Soroushian, P. (1981), Behaviour of C- and Z- Purlins under Uplift,
Report No. 81-2, Cornell University, Ithaca, NY, 1981.
111 Peköz, T.B. and Soroushian, P. (1982), Behaviour of C- and Z- Purlins under Wind
Uplift, Proceedings of the Sixth International Specialty Conference on Cold-formed
Steel Structures, University of Missouri-Rolla, Rolla, MO, November 1982.
112 Peköz, T.B. (1986a), Combined Axial Load and Bending in Cold-formed Steel
Members, Thin-walled Metal Structures in Buildings, IABSE Colloquium,
Stockholm, Sweden, 1986.
113 Peköz, T.B. (1986b), Development of a Unified Approach to the Design of Cold-
formed Steel Members, Report SG-86-4, American Iron and Steel Institute, 1986.
114 Peköz, T.B. (1986c), Developments of a Unified Approach to the Design of Cold-
formed Steel Members, Proceedings Eighth International Specialty Conference on
Cold-formed Steel Structures, St. Louis, MO, 1986.
115 Peköz, T.B. (1990), Design of Cold-formed Steel Screw Connections, Proceedings
Tenth International Specialty Conference on Cold-formed Steel Structures,
St. Louis, MO, 1990.
116 Peköz, T.B. and Sumer, O. (1992), Design Provisions for Cold-formed Steel
Columns and Beam-columns, Final Report, submitted to American Iron and Steel
Institute, Cornell University, September 1992.
117 Pham, L. (1985), Load Combinations and Probabilistic Load Models for Limit State
Codes, Civil Engineering Transactions, Institute of Engineers, Australia, Vol. CE27,
No. 1, 1985.
118 Pi, Y.L. and Trahair, N.S. (1997), Lateral-distortional Buckling of Hollow Flange
Beams, Journal of Structural Engineering, ASCE, Vol. 123, 1997.
119 Rang, T.N., Galambos, T.V., Yu, W.W. and Ravindra, M.K. (1978), Load and
Resistance Factor Design of Cold-formed Steel Structural Members, Proceedings of
the Fourth International Specialty Conference on Cold-formed Steel Structures,
University of Missouri-Rolla, Rolla, MO, June 1978.
Accessed by BHP BILLITON on 27 Jun 2002

120 Rang, T.N., Galambos, T.V. and Yu, W.W. (1979a), Load and Resistance Factor
Design of Cold-formed Steel: Study of Design Formats and Safety Index Combined
with Calibration of the AISI Formulas for Cold Work and Effective Design Width,
First Progress Report, Civil Engineering Study 79-1, University of Missouri-Rolla,
Rolla, MO, January 1979.
121 Rang, T.N., Galambos, T.V. and Yu, W.W. (1979b), Load and Resistance Factor
Design of Cold-formed Steel: Statistical Analysis of Mechanical Properties and
Thickness of Material Combined with Calibration of the AISI Design Provisions on
Unstiffened Compression Elements and Connections, Second Progress Report, Civil
Engineering Study 79-2, University of Missouri-Rolla, Rolla, MO, January 1979.

COPYRIGHT
AS/NZS 4600 Supp1:1998 116

122 Rang, T.N., Galambos, T.V. and Yu, W.W. (1979c), Load and Resistance Factor
Design of Cold-formed Steel: Calibration of the Design Provisions on Connections
and Axially Loaded Compression Members, Third Progress Report, Civil
Engineering Study 79-3, University of Missouri-Rolla, Rolla, MO, January 1979.
123 Rang, T.N., Galambos, T.V. and Yu, W.W. (1979d), Load and Resistance Factor
Design of Cold-formed Steel: Calibration of the Design Provisions on Laterally
Unbraced Beams and Beam-Columns, Fourth Progress Report, Civil Engineering
Study 79-4, University of Missouri-Rolla, Rolla, MO, January 1979.
124 Rasmussen, K.J.R. and Hancock, G.J. (1992), Nonlinear Analyses of Thin-walled
Channel Section Columns, Thin-walled Structures, Vol. 13, Nos. 1-2, 1992.
125 Rasmussen, K.J.R. (1994), Design of Thin-walled Columns with Unstiffened
Flanges, Engineering Structures, (G.J. Hancock, Guest Editor), Vol. 16,
No. 5, 1994.
126 Ravindra, M.K. and Galambos, T.V. (1978), Load and Resistance Factor Design for
Steel, Journal of the Structural Division, ASCE, Vol. 104, No. ST9,
September 1978.
127 Reck, H.P., Peköz, T. and Winter, G. (1975), Inelastic Strength of Cold-formed
Steel Beams, Journal of the Structural Division, ASCE, Vol. 101, No. ST11,
November 1975.
128 Research Council on Structural Connections (1980), Specification for Structural
Joints Using ASTM A325 or A490 Bolts, 1980.
129 Research Council on Structural Connections (1985), Allowable Stress Design
Specification for Structural Joints Using ASTM A325 or A490 Bolts, 1985.
130 Research Council on Structural Connections (1988), Load and Resistance Factor
Design Specification for Structural Joints Using ASTM A325 or A490 Bolts, 1988.
131 Roark, R.J. (1965), Formulas for Stress and Strain, Fourth Edition, McGraw-Hill
Book Company, New York, NY, 1965.
132 Rousch, C.J. and Hancock, G.J. (1996a), Determination of Purlin R-Factors Using a
Non-Linear Analysis, Research Report R722, School of Civil and Mining
Engineering, University of Sydney, 1996.
133 Rousch, C.J. and Hancock, G.J. (1996b), Determinations of Purlin R-Factors Using
a Non-Linear Analysis, Proceedings 13th International Specialty Conference on
Cold-formed Steel Structures, St. Louis, 1996.
134 Salmon, C.G. and Johnson, J.E. (1990), Steel Structures: Design and Behaviour,
Third Edition, Harper & Row, New York, NY, 1990.
135 Santaputra, C., Parks, M.B. and Yu, W.W. (1989), Web Crippling Strength of Cold-
formed Steel Beams, Journal of Structural Engineering, ASCE, Vol. 115, No. 10,
October 1989.
Accessed by BHP BILLITON on 27 Jun 2002

136 Serrette, R.L. and Peköz, T.B. (1992), Local and Distortional Buckling of Thin-
walled Beams, Proceedings Eleventh International Specialty Conference on Cold-
formed Steel Structures, St. Louis, MO, 1992.
137 Serrette, R.L. and Peköz, T.B. (1994), Flexural Capacity of Continuous Span
Standing Seam Panels: Gravity Load, Proceedings Twelfth International Specialty
Conference on Cold-formed Steel Structures, St. Louis, MO, 1994.
138 Serrette, R.L. and Peköz, T.B. (1995), Behaviour of Standing Seam Panels,
Proceedings Third International Conference on Steel and Aluminium Structures,
Bogazici University, Istanbul, Turkey, May 1995.

COPYRIGHT
117 AS/NZS 4600 Supp1:1998

139 Sherman, D.R. (1976), Tentative Criteria for Structural Applications of Steel Tubing
and Pipe, American Iron and Steel Institute, Washington D.C., 1976.
140 Sherman, D.R. (1985), Bending Equations for Circular Tubes, Annual Technical
Session, SSRC 1985.
141 Simaan, A. (1973), Buckling of Diaphragm-braced Columns of Unsymmetrical
Sections and Applications to Wall Studs Design, Report No. 353, Cornell
University, Ithaca, NY, 1973.
142 Simaan, A. and Peköz, T. (1976), Diaphragm-braced Members and Design of Wall
Studs, Journal of the Structural Division, ASCE, Vol. 102, ST1, January 1976.
143 Standards Association of Australia (1968), SAA Steel Structures Code,
AS CA1—1968.
144 Standards Association of Australia (1974), SAA Cold-formed Steel Structures Code,
AS 1538—1974.
145 Standards Association of Australia (1988), Cold-formed Steel Structures,
AS 1538—1988.
146 Standards Australia (1988), Cold-formed Steel Structures Code, AS 1538—1988.
147 Standards Australia (1988), Screws—Self-drilling—For the building and
construction industries, AS 3566—1988.
148 Standards Australia (1989a), Minimum design loads on structures, Part 1: Dead and
live loads, Sydney, AS 1170.1 — 1989.
149 Standards Australia (1989b), Minimum design loads on structures, Part 2: Wind
loads, Sydney, AS 1170.2 — 1989.
150 Standards Australia (1990a), Minimum design loads on structures, Part 3: Snow
loads, Sydney, AS 1170.3 — 1990.
151 Standards Australia (1990b), Steel structures, Sydney, AS 4100 — 1990.
152 Standards Australia (1990c), Methods for tensile testing of metals, Sydney,
AS 1391 — 1991.
153 Standards Australia (1992), Methods of testing sheet roof and wall cladding,
Method 3: Resistance to wind pressures for cyclone regions, AS 4040.3—1992.
154 Standards Australia (1993a), Domestic metal framing, Sydney, AS 3623 — 1993.
155 Standards Australia (1993b), Minimum design loads on structures, Part 4:
Earthquake loads, Sydney, AS 1170.4 — 1993.
156 Standards Australia/Standards New Zealand (1995), Structural steel welding,
Part 1: Welding of steel structures, AS/NZS 1554.1:1995.
157 Standards Australia/Standards New Zealand (1996), Cold-formed steel structures,
AS/NZS 4600:1996.
Accessed by BHP BILLITON on 27 Jun 2002

158 Standards New Zealand (1992), Code of practice for general structural design and
design loads for buildings, NZS 4203:1992.
159 Structural Stability Research Council (1993), Is Your Structure Suitably Braced?,
Lehigh University, Bethlehem, PA, April 1993.
160 Supornsilaphachai, B., Galambos, T.V. and Yu, W.W. (1979), Load and Resistance
Factor Design of Cold-formed Steel: Calibration of the Design Provisions on Beam
Webs, Fifth Progress Report, Civil Engineering Study 79-5, University of Missouri-
Rolla, Rolla, MO, September 1979.

COPYRIGHT
AS/NZS 4600 Supp1:1998 118

161 Trahair, N.S., Flexural-Torsional Buckling of Structures, London, E. & F. N. Spon,


1993.
162 Toma, A., Sedlacek, G. and Weynand, K. (1993), Connections in Cold-formed Steel,
Thin-walled Structures, Vol. 16, Nos. 1–4, 1993.
163 Von Karman, T., Sechler, E.E. and Donnell, L.H. (1932), The Strength of Thin
Plates in Compression, Transactions, ASME, Vol. 54, 1932.
164 Weng, C.C. and Peköz, T.B. (1986), Subultimate Behaviour of Uniformly
Compressed Stiffened Plate Elements, Research Report, Cornell University, Ithaca,
NY, 1986.
165 Winter, G. (1940), Stress Distribution in and Equivalent Width of Flanges of Wide,
Thin-walled Steel Beams, Technical Note No. 784, National Advisory Committee
for Aeronautics, Washington D.C., 1940.
166 Winter, G. (1943), Lateral Stability of Unsymmetrical I-beams and Trusses,
Transactions, ASCE, Vol. 198, 1943.
167 Winter, G. (1944), Strength of Slender Beams, Transactions, ASCE, Vol. 109, 1944.
168 Winter, G. and Pian, R.H.J. (1946), Crushing Strength of Thin Steel Webs, Cornell
Bulletin 35, pt. 1, April 1946.
169 Winter, G., Hsu, P.T., Koo, B. and Loh, M.H. (1948a), Buckling of Trusses and
Rigid Bulletin No. 36, Cornell University, Ithaca, NY, 1948.
170 Winter, G. (1948b), Performance of Thin Steel Compression Flanges, Preliminary
Publication, 3rd Congress of the International Association of Bridge and Structural
Engineers, Liege, Belgium.
171 Winter, G., Lansing, W. and McCalley, R.B. Jr. (1949), Performance of Laterally
Loaded Channel Beams, Research, Engineering Structures Supplement. (Colston
Papers, Vol. II), 1949.
172 Winter, G. (1956a), Light Gauge Steel Connections with High-strength, High-
torqued Bolts, Publications, IABSE, Vol. 16, 1956.
173 Winter, G. (1956b), Tests on Bolted Connections in Light Gauge Steel, Journal of
the Structural Division, ASCE, Vol. 82, No. ST2, February 1956.
174 Winter, G. (1958), Lateral Bracing of Columns and Beams, Journal of the
Structural Division, ASCE, Vol. 84, No. ST2, March 1958.
175 Winter, G. (1959a), Development of Cold-formed, Light Gauge Steel Structures,
AISI Regional Technical Papers, October 1, 1959.
176 Winter, G. (1959b), Cold-formed, Light Gauge Steel Construction, Journal of the
Structural Division, ASCE, Vol. 85, No. ST9, November 1959.
177 Winter, G. (1960), Lateral Bracing of Columns and Beams, Transactions, ASCE,
Vol. 125, 1960.
Accessed by BHP BILLITON on 27 Jun 2002

178 Winter, G. and Uribe, J. (1968), Effects of Cold-work on Cold-formed Steel


Members, Thin-walled Steel Structures — Their Design and Use in Buildings,
K.C. Rockey and H.V. Hill (Eds.), Gordon and Breach Science Publishers, United
Kingdom, 1968.
179 Winter, G. (1970), Commentary on the 1968 Edition of the Specification for the
Design of Cold-formed Steel Structural Members, American Iron and Steel Institute,
New York, NY, 1970.
180 Yener, M. and Peköz, T.B. (1985a), Partial Stress Redistribution in Cold-formed
Steel, Journal of Structural Engineering, ASCE, Vol. 111, No. 6, June 1985.

COPYRIGHT
119 AS/NZS 4600 Supp1:1998

181 Yener, M. and Peköz, T.B. (1985b), Partial Moment Redistribution in Cold-formed
Steel, Journal of Structural Engineering, ASCE, Vol. 111, No. 6, June 1985.
182 Yu, W.W. and Davis, C.S. (1973a), Cold-formed Steel Members with Perforated
Elements, Journal of the Structural Division, ASCE, Vol. 99, No. ST10,
October 1973.
183 Yu, W.W. Liu, V.A. and McKinney, W.M. (1973b), Structural Behaviour of Thick
Cold-formed Steel Members, Journal of the Structural Division, ASCE, Vol. 100,
No. ST1, January 1974.
184 Yu, W.W., Liu, V.A. and McKinney, W.M. (1974), Structural Behaviour of Thick
Cold-formed Steel Members, Journal of the Structural Division, ASCE, Vol. 100,
No. ST1, January 1974.
185 Yu, W.W. (1981), Web Crippling and Combined Web Crippling and Bending of
Steel Decks, Civil Engineering Study 81-2, University of Missouri-Rolla, Rolla,
MO, April 1981.
186 Yu, W.W. (1982), AISI Design Criteria for Bolted Connections, Proceedings Sixth
International Specialty Conference on Cold-formed Steel Structures, St. Louis, MO,
1982.
187 Yu, W.W. (1985), Cold-formed Steel Design, Wiley-Interscience, New York,
NY, 1985.
188 Yu, W.W. (1991), Cold-formed Steel Design, 2nd Edition, Wiley-Interscience, New
York, NY, 1991.
189 Yura, J.A. (1993), Fundamentals of Beam Bracing, Is Your Structure Suitably
Braced?, Structural Stability Research Council, April 1993.
190 Zetlin, L. (1955), Elastic Instability of Flat Plates Subjected to Partial Edge Loads,
Journal of the Structural Division, ASCE, Vol. 81, September 1955.
191 Zetlin, L. and Winter, G. (1955), Unsymmetrical Bending of Beams with and
without Lateral Bracing, Journal of the Structural Division, ASCE, Vol. 81, 1955.
192 Zhao, X.L. and Hancock, G.H. (1995a), Butt Welds and Transverse Fillet Welds in
Thin Cold-formed RHS Members, Journal of Structural Engineering, ASCE,
Vol. 121, No. 11, November 1995.
193 Zhao, X.L. and Hancock, G.H. (1995b), Longitudinal Fillet Welds in Thin
Cold-formed RHS Members, Journal of Structural Engineering, ASCE, Vol. 121,
No. 11, November 1995.
Accessed by BHP BILLITON on 27 Jun 2002

COPYRIGHT
This document has expired. To access the current document, please go to your on-l
Accessed by BHP BILLITON on 27 Jun 2002

You might also like