AISI S100-07 Commentary
AISI S100-07 Commentary
AISI S100-07 Commentary
AISI STANDARD
Commentary on
2001 EDITION
The material contained herein has been developed by a joint effort of the American Iron
and Steel Institute Committee on Specifications, the Canadian Standards Association Technical
Committee on Cold Formed Steel Structural Members (S136), and Camara Nacional de la
Industria del Hierro y del Acero (CANACERO) in Mexico. The organizations and the
Committees have made a diligent effort to present accurate, reliable, and useful information on
cold-formed steel design. The Committees acknowledge and are grateful for the contributions
of the numerous researchers, engineers, and others who have contributed to the body of
knowledge on the subject. Specific references are included in the Commentary on the
Specification.
With anticipated improvements in understanding of the behavior of cold-formed steel and
the continuing development of new technology, this material may eventually become dated. It
is anticipated that future editions of this specification will update this material as new
information becomes available, but this cannot be guaranteed.
The materials set forth herein are for general information only. They are not a substitute
for competent professional advice. Application of this information to a specific project should
be reviewed by a registered professional engineer. Indeed, in most jurisdictions, such review is
required by law. Anyone making use of the information set forth herein does so at their own
risk and assumes any and all resulting liability arising therefrom.
PREFACE
This document provides a commentary on the 2007 edition of the North American
Specification for the Design of Cold-Formed Steel Structural Members. This Commentary should be
used in combination with the 2008 edition of the AISI Cold-Formed Steel Design Manual.
The purpose of the Commentary includes: (a) to provide a record of the reasoning behind,
and justification for the various provisions of the North American Specification by cross-
referencing the published supporting research data and to discuss the changes make in the
current Specification; (b) to offer a brief but coherent presentation of the characteristics and
performance of cold-formed steel structures to structural engineers and other interested
individuals; (c) to furnish the background material for a study of cold-formed steel design
methods to educators and students; and (d) to provide the needed information to those who
will be responsible for future revisions of the Specification. The readers who wish to have more
complete information, or who may have questions which are not answered by the abbreviated
presentation of this Commentary, should refer to the original research publications.
Consistent with the Specification, the Commentary contains a main document, Chapters A
through G, and Appendices 1 and 2, and Appendices A and B. A symbol !A,B is used in the
main document to point out that additional discussions are provided in the corresponding
country specific provisions in Appendices A and/or B.
The assistance and close cooperation of the North American Specification Committee
under the Chairmanship of Professor Reinhold M. Schuster and the AISI Committee on
Specifications under the Chairmanship of Mr. Roger L. Brockenbrough and the Vice
Chairmanship of Mr. Jay W. Larson are gratefully acknowledged. Special thanks are extended
to Professor Wei-Wen Yu for revising the draft of this Commentary. The Institute is very grateful
to members of the Editorial Subcommittee and all members of the AISI Committee on
Specifications for their careful review of the document and their valuable comments and
suggestions. The background materials provided by various subcommittees are appreciated.
TABLE OF COTENTS
COMMENTARY ON THE 2007 EDITION OF THE NORTH AMERICAN
SPECIFICATION FOR THE DESIGN OF COLD-FORMED
STEEL STRUCTURAL MEMBERS
PREFACE..................................................................................................................................................iii
INTRODUCTION.........................................................................................................................................1
A. GENERAL PROVISIONS......................................................................................................................2
A1 Scope, Applicability, and Definitions ......................................................................................... 2
A1.1 Scope ................................................................................................................................. 2
A1.2 Applicability........................................................................................................................... 3
A1.3 Definitions ............................................................................................................................ 4
A1.4 Units of Symbols and Terms.............................................................................................. 8
A2 Material ............................................................................................................................................ 9
A2.1 Applicable Steels ................................................................................................................. 9
A2.2 Other Steels ........................................................................................................................ 10
A2.3 Ductility .............................................................................................................................. 10
A2.4 Delivered Minimum Thickness ....................................................................................... 12
A3 Loads .............................................................................................................................................. 13
A4 Allowable Strength Design ......................................................................................................... 13
A4.1 Design Basis ....................................................................................................................... 13
A4.1.1 ASD Requirements ............................................................................................. 13
A4.1.2 Load Combinations for ASD............................................................................. 13
A5 Load and Resistance Factor Design ........................................................................................... 13
A5.1 Design Basis ....................................................................................................................... 13
A5.1.1 LRFD Requirements ........................................................................................... 14
A5.1.2 Load Factors and Load Combinations for LRFD ........................................... 19
A6 Limit States Design....................................................................................................................... 19
A6.1 Design Basis ....................................................................................................................... 19
A6.1.1 LSD Requirements.............................................................................................. 20
A6.1.2 Load Factors and Load Combinations for LSD.............................................. 20
A7 Yield Stress and Strength Increase from Cold Work of Forming .......................................... 21
A7.1 Yield Stress ......................................................................................................................... 21
A7.2 Strength Increase from Cold Work of Forming ............................................................ 22
A8 Serviceability ................................................................................................................................. 26
A9 Referenced Documents ................................................................................................................ 26
B. ELEMENTS ....................................................................................................................................... 27
B1 Dimensional Limits and Considerations................................................................................... 27
B1.1 Flange Flat-Width-to-Thickness Considerations .......................................................... 27
B1.2 Maximum Web Depth-to-Thickness Ratios................................................................... 29
B2 Effective Widths of Stiffened Elements ..................................................................................... 30
B2.1 Uniformly Compressed Stiffened Elements .................................................................. 33
B2.2 Uniformly Compressed Stiffened Elements with Circular or Non-Circular Holes . 35
iv July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
B2.3 Webs and other Stiffened Elements under Stress Gradient ........................................ 36
B2.4 C-Section Webs with Holes under Stress Gradient ...................................................... 37
B3 Effective Widths of Unstiffened Elements ................................................................................ 38
B3.1 Uniformly Compressed Unstiffened Elements ............................................................. 40
B3.2 Unstiffened Elements and Edge Stiffeners with Stress Gradient ............................... 40
B4 Effective Width of Uniformly Compressed Elements with a Simple Lip Edge Stiffener... 42
B5 Effective Widths of Stiffened Elements with Single or Multiple Intermediate Stiffeners or
Edge Stiffened Elements with Intermediate Stiffener(s) ......................................................... 43
B5.1 Effective Width of Uniformly Compressed Stiffened Elements with Single or
Multiple Intermediate Stiffeners ..................................................................................... 43
B5.2 Edge Stiffened Elements with Intermediate Stiffener(s).............................................. 44
C. MEMBERS ....................................................................................................................................... 46
C1 Properties of Sections ................................................................................................................... 46
C2 Tension Members ......................................................................................................................... 46
C3 Flexural Members ......................................................................................................................... 46
C3.1 Bending ............................................................................................................................... 47
C3.1.1 Nominal Section Strength [Resistance] ........................................................... 47
C3.1.2 Lateral-Torsional Buckling Strength [Resistance].......................................... 50
C3.1.2.1 Lateral-Torsional Buckling Strength [Resistance] for Open Cross
Section Members............................................................................... 50
C3.1.2.2 Lateral-Torsional Buckling Strength [Resistance] for Closed Box
Members ............................................................................................ 57
C3.1.3 Flexural Strength [Resistance] of Closed Cylindrical Tubular Members ... 58
C3.1.4 Distortional Buckling Strength [Resistance] ................................................... 59
C3.2 Shear ............................................................................................................................... 63
C3.2.1 Shear Strength [Resistance] of Webs without Holes ..................................... 63
C3.2.2 Shear Strength [Resistance] of C-Section Webs With Holes......................... 64
C3.3 Combined Bending and Shear......................................................................................... 65
C3.3.1 ASD Method........................................................................................................ 66
C3.3.2 LRFD and LSD Methods.................................................................................... 66
C3.4 Web Crippling ................................................................................................................... 66
C3.4.1 Web Crippling Strength [Resistance] of Webs without Holes ..................... 66
C3.4.2 Web Crippling Strength [Resistance] of C-Section Webs with Holes ......... 72
C3.5 Combined Bending and Web Crippling ........................................................................ 73
C3.5.1 ASD Method........................................................................................................ 73
C3.5.2 LRFD and LSD Methods.................................................................................... 74
C3.6 Combined Bending and Torsional Loading .................................................................. 74
C3.7 Stiffeners ............................................................................................................................. 75
C3.7.1 Bearing Stiffeners................................................................................................ 75
C3.7.2 Bearing Stiffeners in C-Section Flexural Members ........................................ 76
C3.7.3 Shear Stiffeners ................................................................................................... 76
C3.7.4 Non-Conforming Stiffeners............................................................................... 76
C4 Concentrically Loaded Compression Members ....................................................................... 76
C4.1 Nominal Strength for Yielding, Flexural, Flexural-Torsional and
Torsional Buckling ............................................................................................................ 77
C4.1.1 Sections Not Subject to Torsional or Flexural-Torsional Buckling .............. 87
July 2007 v
Table of Contents
vi July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
REFERENCES....................................................................................................................................... 139
July 2007 ix
Commentary on the 2007 North American Cold-Formed Steel Specification
INTRODUCTION
Cold-formed steel members have been used economically for building construction and
other applications (Winter, 1959a, 1959b; Yu, 2000). 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 0.0147 in. (0.373 mm) to about 1/4 in. (6.35 mm). Steel plates
and bars as thick as 1 in. (25.4 mm) can be cold-formed successfully into structural shapes.
In general, cold-formed steel structural members can offer several advantages for building
construction (Winter, 1970; Yu, 2000): (1) light members can be manufactured for relatively light
loads and/or short spans, (2) unusual sectional configurations can be produced economically
by cold-forming operations and consequently favorable strength-to-weight ratios can be
obtained, (3) 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, and (4) 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 North America such steel members were not widely used in buildings until the
publication of the first edition of the American Iron and Steel Institute (AISI) 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 Canada, the Canadian Standards Association
(CSA) published its first edition of Design of Light Gauge Steel Structural Members in 1963 based
on the 1962 edition of the AISI Specification. Subsequent editions were published in 1974, 1984,
1989 and 1994. The Canadian Standard for Cold-Formed Steel Structural Members (CSA, 1994) was
based on the Limit States Design (LSD) method.
In Mexico, cold-formed steel structural members have also been designed on the basis of
AISI Specifications. The 1962 edition of the AISI Design Manual (AISI, 1962) was translated to
Spanish in 1965 (Camara, 1965).
The first edition of the unified North American Specification (AISI, 2001) was prepared and
issued in 2001. It was applicable to the United States, Canada, and Mexico for the design of
cold-formed steel structural members. The 2001 edition of the Specification was developed on
the basis of the 1996 AISI Specification with the 1999 Supplement (AISI, 1996, 1999), the 1994 CSA
Standard (CSA, 1994), and subsequent developments. In 2001, the term “Allowable Stress
Design” was renamed to “Allowable Strength Design” to clarify the nature of this design
method. In the North American Specification, the ASD and LRFD methods are used in the United
States and Mexico, while the LSD method is used in Canada. The Supplement to the 2001 edition
of the North American Specification was published in 2004 (AISI, 2004b), in which the new Direct
Strength Method was added in the Specification as Appendix 1. Following the successful use of
the first North American Specification for seven years, it was revised and expanded in 2007 on the
basis of the results of continued research and new developments (AISI, 2007a). This updated
July 2007 1
Chapter A, General Provisions
edition of the Specification includes the new Appendix 2 for the Second-Order Analysis of
structural systems. Additionally, Appendix A has been expanded to be applicable to Mexico
and, consequently, Appendix C has been deleted.
In addition to the issuance of the design specification, AISI also published the first edition of
the Design Manual in 1949 (AISI, 1949). This allowable stress design manual was revised later in
1956, 1961, 1962, 1968, 1977, 1983, and 1986. In 1991, the LRFD Design Manual was published for
using the load and resistance factor design criteria. The AISI 1996 Cold-Formed Design Manual
was prepared for the combined AISI ASD and LRFD Specifications. For using the 2001 edition of
the North American Specification, AISI published the 2002 edition of the Cold-Formed Steel Design
Manual (AISI, 2002). In 2008, the new Design Manual (AISI, 2008) will be published by AISI
based on the 2007 edition of the North American Specification.
During the period from 1958 through 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. From 1983, the format used for the AISI Commentary
has been changed in that the same section numbers are used in the Commentary as in the
Specification. The Commentary on the 1996 AISI Specification was prepared by Professor Wei-Wen
Yu of the University of Missouri-Rolla (Yu, 1996). The 2001 edition of the Commentary (AISI,
2001) was based on the Commentary on the 1996 AISI Specification. The current edition of the
Commentary (AISI, 2007b) was updated for the 2007 edition of the North American Specification
with extensive additions and revisions. It contains Chapters A through G, Appendices 1 and 2,
and Appendices A and B, where commentary on provisions that are only applicable to a specific
country is included in the corresponding Appendix.
As in previous editions of the Commentary, this document contains a brief presentation of
the characteristics and the performance of cold-formed steel members, connections and
assemblies. In addition, it provides a record of the reasoning behind, and the justification for,
various provisions of the specification. A cross-reference is provided between various design
provisions and the published research data.
In this Commentary, the individual sections, equations, figures, and tables are identified by
the same notation as in the Specification and the material is presented in the same sequence.
Bracketed terms used in the Commentary are equivalent terms that apply particularly to the LSD
method in Canada.
The Specification and Commentary are intended for use by design professionals with
demonstrated engineering competence in their fields.
A. GENERAL PROVISIONS
A1 Scope, Applicability, and Definitions
A1.1 Scope
The cross-sectional configurations, manufacturing processes and fabrication practices of
cold-formed steel structural members differ in several respects from that 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 (1) absence
of the residual stresses caused by uneven cooling due to hot-rolling, (2) lack of corner fillets,
(3) presence of increased yield stress with decreased proportional limit and ductility resulting
from cold-forming, (4) presence of cold-reducing stresses when cold-rolled steel stock has not
2 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
been finally annealed, (5) prevalence of elements having large width-to-thickness ratios, (6)
rounded corners, and (7) stress-strain curves can be either sharp-yielding type or gradual-
yielding type.
The Specification is applicable only to cold-formed sections not more than 1 inch (25.4 mm)
in thickness. Research conducted at the University of Missouri-Rolla (Yu, Liu, and McKinney,
1973b and 1974) has verified the applicability of the specification’s provisions for such cases.
In view of the fact that most of the design provisions have been developed on the basis of
the experimental work subject to static loading, the Specification 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.
!A
A1.2 Applicability
The Specification (AISI, 2007a) is limited to the design of steel structural members cold-
formed from carbon or low-alloy sheet, strip, plate or bar. The design can be made by using
either the Allowable Strength Design (ASD) method or the Load and Resistance Factor
Design (LRFD) method for the United States and Mexico. Only the Limit States Design (LSD)
method is permitted in Canada.
In this Commentary, the bracketed terms are equivalent terms that apply particularly to
LSD. A symbol !x is used to point out that additional provisions are provided in the country
specific appendices as indicated by the letter, x.
Because of the diverse forms in which cold-formed steel structural members can be used,
it is not possible to cover all design configurations by the design rules presented in the
Specification. For those special cases where the available strength [factored resistance]¾
and/or stiffness cannot be so determined, it can be established either by (a) testing and
evaluation in accord with the provisions of Chapter F, or (b) rational engineering analysis.
Prior to 2001, the only option in such cases was testing. However, since 2001, in recognition of
the fact that this was not always practical or necessary, the rational engineering analysis
option was added. It is essential that such analysis be based on theory that is appropriate for
the situation, any available test data that is relevant, and sound engineering judgment. Safety
and resistance factors are provided for ease of use, but these factors should not be used if
applicable safety factors or resistance factors in the main Specification are more conservative,
where the main Specification refers to Chapters A through G, Appendices A and B, and
Appendix 2. These provisions must not be used to circumvent the intent of the Specification.
Where the provisions of Chapters B through G of the Specification and Appendices A and B
apply, those provisions must be used and cannot be avoided by testing or rational analysis.
In 2004, Appendix 1, Design of Cold-Formed Steel Structural Members Using the Direct
Strength Method, was introduced (AISI, 2004b). The Appendix provides an alternative design
procedure for several Sections of Chapters C. The Direct Strength Method detailed in
Appendix 1 requires (1) determination of the elastic buckling behavior of the member, and
then provides (2) a series of nominal strength [resistance] curves for predicting the member
strength based on the elastic buckling behavior. The procedure does not require effective
width calculations, nor iteration, and instead uses gross properties and the elastic buckling
behavior of the cross-section to predict the strength. The applicability of the provided
provisions is detailed in the General Provisions of Appendix 1.
July 2007 3
Chapter A, General Provisions
In 2007, Appendix 2, Second-Order Analysis, was added in the Specification (AISI, 2007a).
The provisions of this Appendix are based on the studies conducted by Sarawit and Pekoz at
Cornell University with due considerations given to flexural-torsional buckling, semi-rigid
joints, and local instabilities. The second-order analysis was found to be more accurate than
the effect length approach.
A1.3 Definitions
Many of the definitions in Specification Section A1.3 for ASD, LRFD and LSD are self-
explanatory. Only those which are not self-explanatory are briefly discussed below.
General Terms
4 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
w1 w2 w3
t
b1
1/2be 1/2be
t
b2 N.A.
(1)
Multiple Stiffened Hat-Section
w w
b1
1/2 be 1/2 be t
b2 N.A.
(2)
Multiple Stiffened Inverted "U"-Type Section
July 2007 5
Chapter A, General Provisions
w w w w
b2 b2 1/2 b 1/2 b
b1 b2 b1 b1
ds b1 ds b1 ds b1
t t
b1 b1
b2 N.A. t b2 N.A. t
(4) (5)
Box-Type Section Inverted
"U"-Type Section
Flexural Members, Such as Beams (Top Flange in Compression)
w1
b1 b2 w1
1/2 b1 1/2 b1
ds
1/2 b2
1/2 b2
w2 w2
t 1/2 b2
t 1/2 b2
(7)
(6) Box-Type Section
Lipped Channel
w1 w1 w1
b2 b2
b1 b1 b1 b2
ds ds
1/2 b2
w2
t
t 1/2 b2
(9)
(8) Lipped Angle
I-Section Made of Two Lipped
Channels Back-to-Back
Compression Members, Such as Columns
6 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
w w w w
b b b b
b1 b1 b1
h
b2 b2 b2
h h h
b
w w
w w w w
b1 b1 b1 b1
w
w w w
b1
July 2007 7
Chapter A, General Provisions
In the North American Specification, the terminologies for Limit States Design (LSD) are
given in brackets parallel to those for load and resistance factor design (LRFD). The inclusion
of LSD terminology is intended to help engineers who are familiar with LSD better
understand the Specification.
It should be noted that the design concept used for the LRFD and the LSD methods is the
same, except that the load factors, load combinations, assumed dead-to-live ratios, and target
reliability indexes are slightly different. In most cases, same nominal strength [nominal
resistance] equations are used for ASD, LRFD, and LSD approaches.
Table C-A1.4-1
Conversion Table
To Convert To Multiply by
in. mm 25.4
mm in. 0.03937
Length
ft m 0.30480
m ft 3.28084
in2 mm2 645.160
mm2 in2 0.00155
Area
ft2 m2 0.09290
m2 ft2 10.7639
kip kN 4.448
kip kg 453.5
lb N 4.448
lb kg 0.4535
Force
kN kip 0.2248
kN kg 101.96
kg kip 0.0022
kg N 9.808
ksi MPa 6.895
ksi kg/cm2 70.30
MPa ksi 0.145
Stress
MPa kg/cm2 10.196
kg/cm2 ksi 0.0142
kg/cm2 MPa 0.0981
8 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
A2 Material
A2.1 Applicable Steels
The American Society for Testing and Materials (ASTM) is the basic source of steel
designations for use with the Specification. Section A2.1 contains the complete list of ASTM
Standards for steels that are accepted by the Specification. Dates of issue are included in
Section A9. Other standards that are applicable to a specific country are listed in the
corresponding Appendix.
In the AISI 1996 Specification, the ASTM A446 Standard was replaced by the ASTM
A653/A653M Standard. At the same time, the ASTM A283/A283M Standard, High-Strength,
Low-Alloy Steel (HSLAS) Grades 70 (480) and 80 (550) of ASTM A653/A653M and ASTM
A715 were added.
In 2001, the ASTM A1008/A1008M and ASTM A1011/A1011M Standards replaced the
ASTM A570/A570M, ASTM A607, ASTM A611, and ASTM A715 Standards. ASTM
A1003/A1003M was added to the list of Specification Section A2.1.
In 2007, the ASTM A1039 Standard was added to the list of Specification Section A2.1. For
all grades of steel, ASTM A1039 complies with the minimum required Fu/Fy ratio of 1.08.
Thicknesses equal to or greater than 0.064 in. (1.6 mm) also meet the minimum elongation
requirements of Specification Section A2.3.1 and no reduction in the specified minimum yield
stress is required. However, steel thicknesses less than 0.064 in. (1.6 mm) with yield stresses
greater than 55 ksi (380 MPa) do not meet the requirements of Specification Section A2.3.1 and
are subject to the limitations of Specification Section A2.3.2.
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 sizable
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 2-inch (51 mm) gage 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 ASTM Standards, the yield stresses of steels range from 24 to 80 ksi (165 to
550 MPa or 1690 to 5620 kg/cm2) and the tensile strengths vary from 42 to 100 ksi (290 to 690
MPa or 2950 to 7030 kg/cm2). The tensile-to-yield ratios are no less than 1.13, and the
elongations are no less than 10 percent. Exceptions are ASTM A653/A653M SS Grade 80
(550); specific thicknesses of ASTM A1039/A1039M 55 (380), 60 (410), 70 (480), and 80 (550),
ASTM A1008/A1008M SS Grade 80 (550); and ASTM A792/A792M SS Grade 80 (550) steels
with a specified minimum yield stress of 80 ksi (550 MPa or 5620 kg/cm2), a specified
minimum tensile strength of 82 ksi (565 MPa or 5770 kg/cm2), and with no stipulated
minimum elongation in 2 inches (51 mm). These low ductility steels permit only limited
amounts of cold forming, require fairly large corner radii, and have other limits on their
applicability for structural framing members. Nevertheless, they have been used successfully
for specific applications, such as decks and panels with large corner radii and little, if any,
stress concentrations. The conditions for use of these SS Grade 80 (550) steels are outlined in
Specification Section A2.3.2.
For ASTM A1003/A1003M steel, even though the minimum tensile strength is not
specified in the ASTM Standard for each of Types H and L Steels, the footnote of Table 2 of
the Standard states that for Type H steels the ratio of tensile strength to yield stress shall not
be less than 1.08. Thus, a conservative value of Fu = 1.08 Fy can be used for the design of
July 2007 9
Chapter A, General Provisions
cold-formed steel members using Type H steels. Based on the same Standard, a conservative
value of Fu = Fy can be used for the design of purlins and girts using Type L steels. In 2004,
the Specification listing of ASTM A1003/A1003M steel was revised to list only the grades
designated Type H, because this is the only grade that satisfies the criterion for unrestricted
usage. Grades designated Type L can still be used but are subject to the restrictions of
Specification Section A2.3.1.
10 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
July 2007 11
Chapter A, General Provisions
be used to calculate the web crippling strength [resistance] of deck panels, the Specification is
adopting a conservative approach in Specification Section C3.4. The lesser of 0.75 Fsy and 60
ksi (414 MPa or 4220 kg/cm2) is used to determine both the web crippling strength
[resistance] and the shear strength [resistance] for the low ductility steels. This is consistent
with the previous edition of the AISI Specification.
Another UMR study (Koka, Yu, and LaBoube, 1997) confirmed that for the connection
design using SS Grade 80 (550) of A653/A653M steel, the tensile strength used in design
should be taken as 75 percent of the specified minimum tensile strength or 62 ksi (427 MPa or
4360 kg/cm2), whichever is less. It should be noted that the current design provisions are
limited only to the design of members and connections subjected to static loading without the
considerations of fatigue strength.
Load tests are permitted, but not for the purpose of using higher loads than can be
calculated under Specification Chapters B through G.
For the calculation of the strength [resistance] of concentrically loaded compression
members with a closed box section, Specification Exception 2 was added on the basis of a
study at University of Sydney (Yang, Hancock, 2002). For short members where Fn = Fy in
Specification Section C4, the study shows that the limit of the yield stress used in the design
can be 90 percent of the specified minimum yield stress Fsy for low ductility steels. Tests were
performed on box-sections composed of G550 steel of AS1397 which is similar to ASTM A792
Grade 80. The box-section is formed by connecting the lips of two hat sections.
Further, for calculating the strength [resistance] of concentrically loaded long
compression members, Specification Equations A2.3.2-3 and A2.3.2-4, based on the University
of Sydney research findings (Yang, Hancock and Rasmussen, 2002), were added in the
Specification Section A2.3.2 in Exception 2 when determining the nominal axial strength
[nominal axial resistance] according to Specification Section C4.1.1. The reduction factor Rr
specified in Specification Equation A2.3.2-4 is to be applied to the radius of gyration r and
allows for the interaction of local and flexural (Euler) buckling of thin high strength low
ductility steel sections. The reduction factor is a function of the length varying from 0.65 at
KL = 0 to 1.0 at KL = 1.1L0, where L0 is the length at which the local buckling stress equals
the flexural buckling stress.
12 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
The responsibility of meeting this requirement for a cold-formed product is clearly that of
the manufacturer of the product, not the steel producer.
In 2004, the country specific section, Specification Section A2.4a, was deleted from
Appendix B.
A3 Loads
Comments on loads and load combinations for different countries are provided in the
corresponding Appendices of this Commentary. A,B
!
A4 Allowable Strength Design
A4.1 Design Basis
The Allowable Strength Design method has been featured in AISI specifications beginning
with the 1946 edition. It is included in the Specification along with the LRFD and the LSD
methods for use in the United States, Mexico, and Canada since the 2001 edition.
July 2007 13
Chapter A, General Provisions
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., postbuckling 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 in (Winter, 1970; Pekoz, 1986b;
and Yu, 2000), and a continuing research effort provides further improvement in
understanding them.
Two types of limit states are considered in the load and resistance factor design method.
They are: (1) the limit state of the strength required to resist the extreme loads during the
intended life of the structure, and (2) the limit state of the ability of the structure to perform
its intended function during its life. These two limit states are usually referred to as the limit
state of strength and limit state of serviceability. Like the ASD method, the LRFD method
focuses on the limit state of strength in Specification Section A5.1.1 and the limit state of
serviceability in Specification Section A8.
14 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
Probability Density
Qm Rm
In(R/Q) m
βσ In(R/Q)
In(R/Q)
July 2007 15
Chapter A, General Provisions
changing the value of β (Figure C-A5.1.1-2), since βσln(R/Q) = ln(R/Q)m, from which
approximately
ln ( R m / Q m )
β= (C-A5.1.1-2)
VR2 + VQ
2
16 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
July 2007 17
Chapter A, General Provisions
specifications such as the ANSI/AISC S360 for hot-rolled steel, the AISI Specification for
cold-formed steel, the ACI Code for reinforced concrete members, etc. The studies for hot-
rolled steel are summarized by Ravindra and Galambos (1978), where also many further
papers are referenced which contain additional data. 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, 1988a; Rang, Galambos, and Yu,
1979a, 1979b, 1979c, and 1979d; Supornsilaphachai, Galambos, and Yu, 1979), where both
the basic research data as well as the β’s inherent in the AISI Specification are presented in
great detail. The β’s computed in the above referenced publications were developed with
slightly different load statistics than those of this Commentary, but the essential conclusions
remain the same.
The entire set of data for hot-rolled steel and cold-formed steel designs, as well as data
for reinforced concrete, aluminum, laminated timber, and masonry walls was re-analyzed
by Ellingwood, Galambos, MacGregor, and Cornell (Ellingwood et al., 1980; Galambos et
al., 1982; Ellingwood et al., 1982) using (a) updated load statistics and (b) 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 give, on the average, essentially the same
design by the LRFD method as is obtained by current design for all materials of
construction. These target reliabilities βo for use in LRFD are:
Basic case: Gravity loading, βo = 3.0
For connections: βo = 4.5
For wind loading: βo = 2.5
These target reliability indices are the ones inherent in the load factors recommended
in the ASCE 7-98 Load Standard (ASCE, 1998).
For simply supported, braced cold-formed steel beams with stiffened flanges, which
were designed according to the allowable strength design method in the current
Specification or to any previous version of the AISI Specification, it was shown that for the
representative dead-to-live load ratio of 1/5 the reliability index β = 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 = 2.5 is recommended as a lower limit
in the United States. The resistance factors φ were selected such that βo = 2.5 is essentially
the lower bound of the actual β’s for members. In order to assure that failure of a structure
is not initiated in the connections, a higher target reliability of βo = 3.5 is recommended for
joints and fasteners in the United States. These two targets of 2.5 and 3.5 for members and
connections, respectively, are somewhat lower than those recommended by the ASCE 7-98
(i.e., 3.0 and 4.5, respectively), but they are essentially the same targets as are the basis for
the AISC LRFD Specification (AISC, 1999). For wind loading, the same ASCE target value
18 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
of βo = 2.5 is used for connections in the US LRFD method. For flexural members such as
individual purlins, girts, panels, and roof decks subjected to the combination of dead and
wind loads, the target βo value used in the United States is reduced to 1.5. With this
reduced target reliability index, the design based on the US LRFD method is comparable to
the US allowable strength design method.
(c) Resistance Factors
The following portions of this Commentary present the background for the resistance
factors φ which are recommended for various members and connections in Chapters B
through E (AISI, 1996). These φ factors are determined in conformance with the ASCE 7
load factors to provide approximately a target βo of 2.5 for members and 3.5 for
connections, respectively, for a typical load combination 1.2D+1.6L. For practical reasons, it
is desirable to have relatively few different resistance factors, and so the actual values of β
will differ from the derived targets. This means that
φRn = c(1.2D+1.6L) = (1.2D/L+1.6)cL (C-A5.1.1-11)
where c is the deterministic influence coefficient translating load intensities to load effects.
By assuming D/L = 1/5, Equations C-A5.1.1-11 and C-A5.1.1-9 can be rewritten as
follows:
Rn = 1.84(cL/φ) (C-A5.1.1-12)
Qm = (1.05D/L+1)cL = 1.21cL (C-A5.1.1-13)
Therefore,
Rm/Qm =(1.521/φ)(Rm/Rn) (C-A5.1.1-14)
The φ factor can be computed from Equation C-A5.1.1-15 on the basis of Equations C-
A5.1.1-2, C-A5.1.1-4 and C-A5.1.1-14 (Hsiao, Yu and Galambos, 1988b, AISI 1996):
φ = 1.521 (PmMmFm)exp(-βo VR 2 + VQ 2 ) (C-A5.1.1-15)
in which, βo is the target reliability index. Other symbols were defined previously.
By knowing the φ factor, the corresponding safety factor, Ω, for allowable strength
design can be computed for the load combination 1.2D+1.6L as follows:
Ω = (1.2D/L + 1.6)/[φ(D/L + 1)] (C-A5.1.1-16)
where D/L is the dead-to-live load ratio for the given condition.
July 2007 19
Chapter A, General Provisions
limit states have been established through experience in practice or in the laboratory, and
they have been thoroughly investigated through analytical and experimental research.
Two types of limit states are considered in the Limit States Design method. They are: (1)
the limit state of the strength required to resist the extreme loads during the intended life of
the structure, and (2) the limit state of the ability of the structure to perform its intended
function during its life. These two limit states are usually referred to as the limit state of
strength and limit state of serviceability. The LSD method focuses on the limit state of
strength in Specification Section A6.1.1 and the limit state of serviceability in Specification
Section A8.
20 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
Commentary.
Inelastic
Strain hardening
range
Fu
Elastic
range
Stress, σ
Fy
tan-1 E
E= ε
σ
Strain, ε
(a)
Fu
tan-1 E t
Stress, σ
Et = dσ
ft
f pr dε
tan-1 E
E= ε
σ
(b) Strain, ε
The strength [resistance] 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 (Figure C-A7.1-1). The measured values of E on the basis of the standard methods
usually range from 29,000 to 30,000 ksi (200 to 207 GPa or 2.0x106 to 2.1x106 kg/cm2). A value
July 2007 21
Chapter A, General Provisions
of 29,500 ksi (203 GPa or 2.07x106 kg/cm2) is used in the Specification for design purposes.
The tangent modulus is defined by the slope of the stress-strain curve at any stress level, as
shown in Figure C-A7.1-1(b).
For sharp-yielding steels, Et = E up to the yield point, but with gradual-yielding steels, Et
= 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 Specification have been written for gradual-yielding
steels whose proportional limit is not lower than about 70 percent of the specified minimum
yield stress.
Determination of proportional limits for information purposes can be done simply by
using the offset method shown in Figure C-A7.1-2(b) with the distance “om” equal to 0.0001
length/length (0.01 percent offset) and calling the stress R where “mn” intersects the stress-
strain curve at “r”, the proportional limit.
A n
n
R R R
r r
Yield Point
Stress
Stress
Stress
(a) Showing Yield Point (b) Showing Yield Point or (c) Determination of Yield
Corresponding with Yield Strength by the Strength by Extension
Top of Knee. Offset Method. (Also Used Under Load Method.
for Proportional Limit)
22 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
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-aging, as illustrated in Figure C-A7.2-2 (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-aging. It is interesting to note
that the yield stresses 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 aging.
Cornell research also revealed that the effects of cold work on the mechanical properties
of corners usually depend on (1) the type of steel, (2) the type of stress (compression or
tension), (3) the direction of stress with respect to the direction of cold work (transverse or
longitudinal), (4) the Fu/Fy ratio, (5) the inside radius-to-thickness ratio (R/t), and (6) the
amount of cold work. Among the above items, the Fu/Fy and R/t ratios are the most
important factors to affect the change in mechanical properties of formed sections. Virgin
material with a large Fu/Fy ratio possesses a large potential for strain hardening.
Consequently as the Fu/Fy ratio increases, the effect of cold work on the increase in the yield
stress of steel increases. Small inside radius-to-thickness ratios, R/t, correspond to a large
degree of cold work in a corner, and therefore, for a given material, the smaller the R/t ratio,
the larger the increase in yield stress.
Investigating the influence of cold work, Karren derived the following equations for the
ratio of corner yield stress-to-virgin yield stress (Karren, 1967):
Fy c Bc
= (C-A7.2-1)
Fyv (R / t ) m
where
2
F ⎛F ⎞
Bc = 3.69 uv − 0.819⎜ uv ⎟ − 1.79
Fyv ⎜ Fyv ⎟
⎝ ⎠
and
F
m = 0.192 uv − 0.068
Fyv
Fyc = corner yield stress
Fyv = virgin yield stress
Fuv = virgin ultimate tensile strength
R = 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:
Fya = CFyc + (1 - C)Fyf (C-A7.2-2)
where
Fya = full-section tensile yield stress
Fyc = average tensile yield stress of corners = BcFyv/(R/t)m
Fyf = average tensile yield stress of flats
C = ratio of corner area to total cross-sectional area. For flexural members having
July 2007 23
Chapter A, General Provisions
70 Virgin
ultimate
strength
65
K
1"
A
4
60
1"
Stress, ksi
B J
4
3"
116
C 55
5" 1"
H
16 4
D E F G Virgin
yield
str.
50
1"
16
45
5" 1" 3" 1" 3" 1" 3" 1" 5"
16 4 16 4 16 4 16 4 16
3"
2 16
40
A B C D E F G H J K L
Yield Strength
Ultimate Strength
(a)
75
70
65
60
3.68" 55
0.90" 0.90" 50
0.25" 0.25" 0.25" 0.25" Virgin
A B G H 50 ultimate
C C strength
50 Virgin
D F yield
0.25" C EC strength
35
30
0.25" A BCDC E C F CG H
Yield Strength
Ultimate Strength
(b)
24 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
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 stress of the section, Fya, to be determined by (1) full section tensile tests, (2)
stub column tests, or (3) computed in accordance with Equation C-A7.2-2. However, such a
strength increase is limited only to relatively compact sections designed according to
Specification Section C2 (tension members), Section C3.1 (bending strength excluding the use
of inelastic reserve capacity), Section C4 (concentrically loaded compression members),
Section C5 (combined axial load and bending), Section D4 (cold-formed steel light-frame
construction), and Section D6.1 (purlins, girts and other members). Design Example of the
2008 Cold-Formed Steel Design Manual (AISI, 2008) demonstrates the use of strength increase
from cold work of forming for a channel section to be used as a beam.
In some cases, when evaluating the effective width of the web, the reduction factor ρ
according to Specification Section B2.3 may be less than unity but the sum of b1 and b2 of
Figure B2.3-1 of the Specification may be such that the web is fully effective, and cold work of
forming may be used. This situation only arises when the web width to flange width ratio,
ho/bo, is less than or equal to 4.
Increase in Fu
Strain aging
D
A Strain aging
Stress
C
Increase
in Fy
A Ductility after
Strain strain aging
hardening
B
C
Strain
Ductility after strain
hardening
Virgin ductility
Figure C-A7.2-2 Effect of Strain Hardening and Strain Aging on
Stress-Strain Characteristics
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 resistance factors φ:
(Fy)m = 1.10Fy; Mm = 1.10; Vfy = VM =0.10
(Fya)m=1.10Fya; Mm = 1.10; VFya = VM =0.11
(Fu)m = 1.10Fu; Mm = 1.10; VFu = VM =0.08
Fm = 1.00; VF = 0.05
In the above expressions, m refers to mean value, V represents coefficient of variation, M
and F are, respectively, the ratios of the actual-to-the nominal material property and cross-
sectional property; and Fy, Fya, and Fu are, respectively, the specified minimum yield stress,
July 2007 25
Chapter A, General Provisions
the average yield stress including the effect of cold forming, and the 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.
A8 Serviceability
Serviceability limit states are conditions under which a structure can no longer perform its
intended functions. Safety and strength [resistance] 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:
1. 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.
2. Excessive vibrations which may cause occupant discomfort of equipment malfunctions.
3. Deterioration over time which may include corrosion or appearance considerations.
When checking serviceability, the designer should consider appropriate service loads, the
response of the structure, and the reaction of building occupants.
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 analyzed assuming linear
elastic behavior. However, members that accumulate residual deformations under service loads
may require consideration of this long-term behavior.
Serviceability limits depend on the function of the structure and on the perceptions of the
observer. In contrast to the strength [resistance] limit states, it is not possible to specify general
serviceability limits that are applicable to all structures. The Specification 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), AISC (2005) and ATC (1999).
A9 Referenced Documents
Other specifications and standards to which the Specification makes references to have been
listed and updated in Specification Section A9 to provide the effective dates of these standards at
the time of approval of this Specification.
Additional references which the designer may use for related information are listed at the
end of the Commentary.
26 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
B. ELEMENTS
In cold-formed steel construction, individual elements of steel structural members are thin
and the width-to-thickness ratios are large as compared with hot-rolled steel shapes. These thin
elements may buckle locally at a stress level lower than the yield stress of steel when they are
subjected to compression in flexural bending, axial compression, shear, or bearing. Figure C-B-1
illustrates some local buckling patterns of certain beams and columns (Yu, 2000).
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 the failure by
local instability with due consideration given to the postbuckling strength of structural
components. Chapter B of the Specification 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, elements with edge stiffeners or
intermediate stiffeners, and beam webs. The design provisions are provided for the use of
stiffeners in Specification Section C3.7 for flexural members.
Compression Compression
flange flange
(a)
A A
Section A-A
(b)
July 2007 27
Chapter B, Elements
The limitation to a maximum w/t of 60 for the compression flanges having one
longitudinal edge connected to a web and the other edge is stiffened by a simple lip is
based on the fact that if the w/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. This is the reason why the w/t ratio is limited to
60 for stiffened compression elements having one longitudinal edge connected to a web
or flange element and the other is stiffened by a simple lip.
The limitation to w/t = 90 for compression flanges with any other kind of stiffeners
indicates that thinner flanges with large w/t ratios are quite flexible and liable to be
damaged in transport, handling and erection. The same is true for the limitation to w/t =
500 for stiffened compression elements with both longitudinal edges connected to other
stiffened elements and for the limitation to w/t = 60 for unstiffened compression
elements. The provision specifically states that wider flanges are not unsafe, but that
when the w/t ratio of unstiffened flanges exceeds 30 and the w/t ratio of stiffened flanges
exceeds 250, it is likely to develop noticeable deformation at the full design strength
[resistance], without affecting the ability of the member to develop required strength
[resistance]. In both cases the maximum w/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 such cases where the limits are exceeded, tests in accordance with Specification
Chapter F are required.
(b) Flange Curling
In beams which have unusually wide and thin, but stable flanges, (i.e., primarily tension
flanges with large w/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, center
portions of flanges of box or hat beams) tend to deflect toward the neutral axis. An
approximate, analytical treatment of this problem was given by Winter (1948b). Equation
B1.1-1 of the Specification permits one to compute the maximum permissible flange width,
wf, for a given amount of flange curling, cf.
It should be noted that Section B1.1(b) does not stipulate the amount of curling which
can be regarded as tolerable, but an amount of curling in the order of 5 percent 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 important, the out-of-plane distortion should be closely controlled in practice.
Example of the AISI Cold-Formed Steel Design Manual (AISI, 2008) illustrates the design
consideration for flange curling.
(c) Shear Lag Effects - Short Spans Supporting Concentrated Loads
For the 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 the width of the flange, similar to that in stiffened compression
28 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
elements (see Section B2 of the Commentary), 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 wf 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 Section B1.1(c) 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 Specification Table B1.1(c)
are taken directly from Curve A of Figure 4 of Winter (1940).
It will be noted that according to Specification Section B1.1(c), the use of a reduced width
for stable, wide flanges is required only for concentrated load as shown in Figure C-B1.1-
1. For uniform load it is seen from Curve B of the figure that the width reduction due to
shear lag for any unrealistically large span-width ratios is so small as to be practically
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 according to Specification Section B1.1(c). For
design purpose, see Example of the AISI Design Manual (AISI, 2008).
1.0
B
Actual width
0.9
AISI design
0.8 criteria
A
0.7
For concentrated load
0.6
0.5
0 10 20 30
L
wf
July 2007 29
Chapter B, Elements
Specification. Because the definition for “h” 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 h/t ratio may appear to be more
liberal. An unpublished study by LaBoube concluded that the present definition for h had
negligible influence on the web strength [resistance].
a
b
c
d
30 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
Table C-B2-1
Values of Plate Buckling Coefficients
s.s.
(a) s.s. s.s. Compression 4.0
s.s.
Fixed
(b) s.s. s.s. Compression 6.97
Fixed
s.s.
(c) s.s. s.s. Compression 0.425
Free
Fixed
(d) s.s. s.s. Compression 1.277
Free
Fixed
(e) s.s. s.s. Compression 5.42
s.s.
s.s.
(f) s.s. s.s. Shear 5.34
s.s.
Fixed
(g) Fixed Fixed Shear 8.98
Fixed
s.s.
(h) s.s. s.s. Bending 23.9
s.s.
Fixed
(i) Fixed Fixed Bending 41.8
Fixed
July 2007 31
Chapter B, Elements
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
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 behavior of the model is as follows:
(a) there is no collapse by unrestrained deflections, as in unsupported columns, and (b) the
various struts will deflect unequal amounts, those nearest the supported edges being held
almost straight by the ties, those nearest the center being able to deflect most.
In consequence of (a), the model will not collapse and fail when its buckling stress (Equation
C-B2-1) is reached; in contrast to columns it will merely develop slight deflections but will
continue to carry increasing load. In consequence of (b), the struts (strips of the plate) closest to
the center, which deflect most, “get away from the load,” and hardly participate in carrying any
further load increases. These center strips may in fact, even transfer part of their pre-buckling
load to their neighbors. 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 on Figure C-B2-3, the stresses being largest at the edges and smallest in the
center. With further increase in load this non-uniformity increases further, as also shown on
Figure C-B2-3. 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 fmax reaches the yield stress Fy.
c
W
32 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
This postbuckling strength [resistance] 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 postbuckling
strength [resistance] can be found in the series of photographs on Figure 7 of Winter (1959b).
The model of Figure C-B2-2 is representative of the behavior of a compression element
supported along both longitudinal edges, as the flange in Figure C-B2-1. In fact, such elements
buckle into approximately square waves.
In order to utilize the postbuckling strength [resistance] of the stiffened compression
element for design purposes, the AISI Specification has used the effective design width approach to
determine the sectional properties since 1946. In Section B2 of the present Specification, design
equations for computing the effective widths are provided for the following four cases: (1)
uniformly compressed stiffened elements, (2) uniformly compressed stiffened elements with
circular or noncircular holes, (3) webs and other stiffened elements with stress gradient, (4)
unstiffened elements with uniform or gradient stress, and (5) C-section webs with holes under
stress gradient. The background information on various design requirements is discussed in
subsequent sections.
fmax
b/2 b/2
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-gage, cold-
formed steel sections at Cornell University, the following equation was developed by
Winter in 1946 for determining the effective width b for stiffened compression elements
simply supported along both longitudinal edges:
July 2007 33
Chapter B, Elements
E ⎡ ⎛ t ⎞ E ⎤
b = 1.9 t ⎢1 − 0.475⎜ ⎟ ⎥ (C-B2.1-1)
fmax ⎢⎣ ⎝ w ⎠ fmax ⎥⎦
The above equation can be written in terms of the ratio of Fcr/fmax as follows:
b Fcr ⎛ Fcr ⎞
= ⎜ 1 − 0.25 ⎟ (C-B2.1-2)
w fmax ⎜ f ⎟
⎝ max ⎠
where Fcr is the critical elastic buckling stress of a plate, and is expressed in Equation C-
B2-1.
Thus, the effective width expression (e.g., C-B2.1-1) provides a prediction of the nominal
strength [resistance] based only on the critical elastic buckling stress and the applied
stress of the plate. During the period from 1946 to 1968, the AISI design provision for the
determination of the effective design width was based on Equation C-B2.1-1. A long-time
accumulated experience has indicated that a more realistic equation, as shown below may
be used for the determination of the effective width b (Winter, 1970):
E ⎡ t E ⎤
b = 1.9 t ⎢1 − 0.415( ) ⎥ (C-B2.1-3)
fmax ⎢⎣ w fmax ⎥⎦
The correlation between the test data on stiffened compression elements and Equation
C-B2.1-3 is illustrated by Yu (2000).
It should be noted that Equation C-B2.1-3 may also be rewritten in terms of the Fcr/fmax
ratio as follows:
b Fcr ⎛ Fcr ⎞
= ⎜ 1 − 0.22 ⎟ (C-B2.1-4)
w fmax ⎜ fmax ⎟⎠
⎝
Therefore, the effective width, b, can be determined as
b = ρw (C-B2.1-5)
where ρ = reduction factor
= (1 − 0.22 / fmax / Fcr ) / fmax / Fcr = (1 − 0.22 / λ )/λ ≤ 1 (C-B2.1-6)
In Equation C-B2.1-6, λ is a slenderness factor determined below.
λ = fmax / Fcr (C-B2.1-7)
Figure C-B2.1-1 shows the relationship between ρ and λ. It can be seen that when
λ ≤ 0.673, ρ = 1.0.
Based on Equations C-B2.1-5 through C-B2.1-7 and the unified approach proposed by
Pekoz (1986b and 1986c), the 1986 edition of the AISI Specification adopted the
nondimensional format in Section B2.1 for determining the effective design width, b, for
uniformly compressed stiffened elements. The same design equations were used in the
1996 edition of the AISI Specification and was retained in this edition of the North American
Specification. For design examples, see Part I of the AISI Design Manual (AISI, 2008).
(b) Effective Width for Serviceability Determination
The effective design width equations discussed above for strength [resistance]
determination can also be used to obtain a conservative effective width, bd, for
serviceability determination. It is included in Section B2.1(b) of the Specification as
Procedure I.
34 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
1.0
0.9
0.8
0.7
Eq. C-B2.1-6
0.6
ρ = (1 - 0.22/ λ )/ λ < 1
ρ 0.5
0.4
0.3
0.2
0.1
0
0 0.673 1 2 3 4 5 6 7 8
λ
Figure C-B2.1-1 Reduction Factor, ρ, vs. Slenderness Factor, λ
July 2007 35
Chapter B, Elements
Although Figure B2.2-1 in the Specification shows a hole centered within the flat width, w,
holes not centered within w are allowed. For such a case, the unstiffened strip, c, and
resulting effective width, b, must be calculated separately for the strips on each side of the
hole. For sections with perforations, which do not meet these limits, the effective area, Ae, can
be determined by stub column tests.
The geometric limitations (w/t, etc.) and hole size for the circular and non-circular hole
provisions in Specification Section B2.2 are not consistent with one another. This anomaly in
the limitations is due to the differing scopes of the test programs that serve as the basis for
these effective width equations. Ongoing research on perforations will provide a consistent,
compatible design methodology in the future. The provisions for non-circular holes generally
give a more conservative prediction of the effective width than the provisions for circular
holes, as long as dh/w < 0.4.
36 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
(CSA, 1994) Standard, there are only minor changes for members with ho/bo > 4, but an
increase in strength [resistance] will be experienced when ho/bo ≤ 4.
It should be noted that in the North American Specification, the stress ratio ψ is defined as
an absolute value. As a result, some signs for ψ have been changed in Specification Equations
B2.3-2, B2.3-3, B2.3-6 and B2.3-7 as compared with the 1996 edition of the AISI Specification
(AISI, 1996).
July 2007 37
Chapter B, Elements
Lh
dh
dh
38 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
63.3/ Fy 144/ Fy
Inelastic Elastic
Yielding Buckling Buckling
Fy
A
B w/ t = 25
Stress
Based on postbuckling strength
C
E
D
f cr
0 10 20 30 40 50 60
w
t
July 2007 39
Chapter B, Elements
1.2
1.0
A
0.8 B
σ
0.6
Fy C
0.4
0
63.3 144
0 50 100 150 200 250
w
Fy
t
40 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
1.25
ψ= 1 Compression
+
Tension
1.00
b
Unwelded plate tests
Welded plate tests } Uniform
compression
Unwelded plate tests ψ= 0
ψ= Welded plate tests ψ= 0
1 ψ= 1 Compression at
0.75 Compression ψ=
ψ=
0.7
5
Unwelded plate tests
Welded plate tests
}
unsupported
ψ= 1 edge
ψ= 1 0.5
b ψ=
---- + 0.2
w 5
b
0.50
ψ= 0 Compression
+
f f
0.25 ψ= ---2- b ψ= -----2
f1 f1 + f1
f2 + f1 Compression
f2
b Compression Tension b
0
0 0.5 1.0 1.5 2.0 2.5 3.0 3.5
Slenderness (λ)
elements with stress gradient proposed by Bambach and Rasmussen (2002a, 2002b and
2002c), based on an extensive experimental investigation of unstiffened plates tested as
isolated elements in combined compression and bending. The effective width, b, (measured
from the supported edge) of unstiffened elements with stress gradient causing compression
at both longitudinal edges, is calculated using the Winter equation. For unstiffened elements
with stress gradients causing tension at one longitudinal edge and compression at the other
longitudinal edge, modified Winter equations are specified when tension exists at either the
supported or the unsupported edges. The effective width equations apply to any unstiffened
element under stress gradient, and are not restricted to particular cross-sections. Figure C-
B3.2-1 demonstrates how the effective width of an unstiffened element increases as the stress
at the supported edge changes from compression to tension. As shown in the figure, the
effective width curve is independent of the stress ratio, ψ, when both edges are in
compression. In this case, the effect of stress ratio is accounted for by the plate buckling
coefficient, k, which varies with stress ratio and affects the slenderness, λ. When the
supported edge is in tension and the unsupported edge is in compression, both the effective
width curve and the plate buckling coefficient depend on the stress ratio, as per Equations
B3.2-4 and B3.2-5 of the Specification.
Equations are provided for k, determined from the stress ratio, ψ, applied to the full
element width such that iteration is not required, and k will usually be higher than 0.43. The
equations for k are theoretical solutions for long plates assuming simple support along the
longitudinal edge. A more accurate determination of k by accounting for interaction between
adjoining elements is permitted for plain channels in minor axis bending (causing
compression at the unsupported edge of the unstiffened element), based on research of plain
channels in compression and bending by Yiu and Pekoz (2001).
The effective width is located adjacent to the supported edge for all stress ratios,
including those producing tension at the unsupported edge. Research has found (Bambach
July 2007 41
Chapter B, Elements
and Rasmussen 2002a) that for the unsupported edge to be effective, tension must be applied
over at least half of the width of the element starting at the unsupported edge. For less
tension, the unsupported edge will buckle and the effective part of the element is located
adjacent to the supported edge. Further, when tension is applied over half of the element or
more starting at the unsupported edge, the compressed part of the element will remain
effective for elements with w/t ratios less than the limits set out in Section B1.1 of the
Specification.
The method for serviceability determination is based on the method used for stiffened
elements with stress gradient in Section B2.3(b) of the Specification.
B4 Effective Width of Uniformly Compressed Elements with a Simple Lip Edge Stiffener
An edge stiffener is used to provide continuous support along a longitudinal edge of the
compression flange to improve the buckling stress. In most cases, the edge stiffener takes the
form of a simple lip. Other types of edge stiffeners can be beneficial and are also used for cold-
formed steel members, but are not covered in Specification Section B4.
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. As far as the design provisions are concerned, the 1980 and earlier editions of the
AISI Specification included the requirements for the minimum moment of inertia of stiffeners to
provide sufficient rigidity. When the size of the actual stiffener does not satisfy the required
moment of inertia, the load-carrying capacity of the beam had to be determined either on the
basis of a flat element disregarding the stiffener or through tests.
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 Section B4 of the 1986 AISI Specification were based on the investigations on adequately
stiffened and partially stiffened elements conducted by Desmond, Pekoz and Winter (1981a),
with additional research work of Pekoz and Cohen (Pekoz, 1986b). These design provisions
were developed on the basis of the critical buckling criterion and the postbuckling strength
[resistance] criterion.
Specification Section B4 recognizes that the necessary stiffener rigidity depends upon the
slenderness (w/t) of the plate element being stiffened. 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 (Pekoz, 1986b).
In the 1996 edition of the AISI Specification (AISI, 1996), the design equations for buckling
coefficient were changed for further clarity. The requirement of 140° ≥ θ ≥ 40° for the
applicability of these provisions was decided on an intuitive basis. For design examples, see
Part I of the Cold-Formed Steel Manual (AISI, 2008).
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 Specification Section B4.
The 1996 Commentary provided a warning to the user that lip lengths with a d/t ratio greater
than 14 may give unconservative results. Examination of available experimental data on both
flexural members (Rogers and Schuster, 1996, Schafer and Pekoz, 1999) and compression
members (Schafer, 2000) with edge stiffeners indicates that the Specification does not have an
inherent problem for members with large d/t ratios. Existing experimental data covers d/t
42 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
July 2007 43
Chapter B, Elements
element having closely spaced intermediate stiffeners, (4) the effective width of sub-element
with w/t > 60, and (5) the reduced area of stiffeners. In the Canadian Standard, a different
design equation was used to determine the equivalent thickness.
Plate Sub-element
In 2001, Specification Section B5.1 was revised to reflect recent research findings for
flexural members with multiple intermediate stiffeners in the compression flange (Papazian
et al. 1994, Schafer and Peköz 1998, Acharya and Schuster 1998). The method is based on
determining the plate buckling coefficient for the two competing modes of buckling: local
buckling, in which the stiffener does not move; and distortional buckling in which the
stiffener buckles with the entire plate. See Figure C-B5.1-1. Experimental research shows that
the distortional mode is prevalent for members with multiple intermediate stiffeners.
The reduction factor, ρ, is applied to the entire element (gross area of the
element/thickness) instead of only the flat portions. Reducing the entire element to an
effective width, which ignores the geometry of the stiffeners, for effective section property
calculation allows distortional buckling to be treated consistently with the rest of the
Specification, rather than as an “effective area” or other method. The resulting effective width
must act at the centroid of the original element including the stiffeners. This insures that the
neutral axis location for the member is unaffected by the use of the simple effective width,
which replaces the more complicated geometry of the element with multiple intermediate
stiffeners. One possible result of this approach is that the calculated effective width (be) may
be greater than bo. This may occur when ρ is near 1, and is due to the fact that be includes
contributions from the stiffener area and bo does not. As long as the calculated be is placed at
the centroid of the entire element the use of be>bo is correct.
44 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
However, an edge stiffened element does not have the same web rotational restraint as a
stiffened element, therefore the constant R of Specification Section B5.1 is conservatively
limited to be less than or equal to 1.0.
If the edge stiffened element is partially effective (bo/t > 0.328S and Is < Ia and thus k < 4,
per the rules of Specification Section B4) then the intermediate stiffener(s) should be ignored
and the provisions of Specification Section B4 followed. Elastic buckling analysis of the
distortional mode for an edge stiffened element with intermediate stiffener(s) indicates that
the effect of intermediate stiffener(s) on the distortional buckling stress is ±10 percent for
practical intermediate and edge stiffener sizes.
When applying Specification Section B5.2 for effective width determination of edge
stiffened elements with intermediate stiffeners, the effective width of the intermediately
stiffened flange, be, is replaced by an equivalent flat section (as shown in Specification Figure
B5.1-2). The edge stiffener should not be used in determining the centroid location of the
equivalent flat effective width, be, for the intermediately stiffened flange.
Stub compression testing performed in 2003 demonstrates the adequacy of this approach
(Yang and Hancock, 2003).
Distortional Buckling of
the Edge Stiffened Element
Figure C-B5.2-1 Buckling Modes in an Edge Stiffened Element with
Intermediate Stiffeners
July 2007 45
Chapter C, Members
C. MEMBERS
This Chapter provides the design requirements for (a) tension members, (b) flexural
members, (c) concentrically loaded compression members, and (d) members subjected to
combined axial load and bending.
In 2007, the following design provisions were moved from Specification Chapter C,
Members, to Section D6, Metal Roof and Wall Systems: (1) Flexural Members Having One
Flange Through-Fastened to Deck or Sheathing, (2) Flexural Members Having One Flange
Fastened to a Standing Seam Roof System, (3) Compression Members Having One Flange
Through-Fastened to Deck or Sheathing, and (4) Strength [Resistance] of Standing Seam Panel
System. For closed cylindrical tubular members the design provisions have been moved to new
Section C3.1.3 for flexural members and new Section C4.1.5 for compression members.
In general, a common nominal strength [resistance] equation is provided in the Specification
for a given limit state with a required safety factor (Ω) for Allowable Strength Design (ASD) and
a resistance factor (φ) for Load and Resistance Factor design (LRFD) or Limit State Design
(LSD). Design provisions that are applicable to a specific country are provided in the
corresponding Appendix.
C1 Properties of Sections
The geometric properties of a member (i.e., area, moment of inertia, section modulus, radius
of gyration, etc.) 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, both gross and net sections are employed when
computing the nominal tensile strength [resistance] of the axially loaded tension members.
For flexural members and axially loaded compression members, both full and effective
dimensions are used to compute sectional properties. The full dimensions are used when
calculating the critical load or moment, while the effective dimensions, evaluated at the stress
corresponding to the critical load or moment, are used to calculate the nominal strength
[resistance]. For serviceability consideration, the effective dimension should be determined for
the compressive stress in the element corresponding to the service load. Pekoz (1986a and
1986b) discussed this concept in more detail.
Section 3 of Part I of the AISI Design Manual (AISI, 2008) deals with the calculation of
sectional properties for C-sections, Z-sections, angles, hat sections, and decks.
C2 Tension Members
The design provisions of this section are given in Section C2 of the Appendices. The
discussion for this section is provided in the Commentary on the corresponding Appendix.
C3 Flexural Members
!A,B
For the design of cold-formed steel flexural members, consideration should be given to
several design features: (a) bending strength [resistance] and serviceability, (b) shear strength
[resistance] of webs and combined bending and shear, (c) web crippling strength [resistance]
and combined bending and web crippling, and (d) bracing requirements. For some cases,
46 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
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 Specification Sections
C3, and D6.1 and D6.2, while the requirements for lateral and stability bracing are given in
Specification Sections D3 and D6.3. The treatments for flange curling and shear lag were
discussed in Section B1.1(b) and (c) of the Commentary, respectively.
Example problems are given in Part II of the AISI Cold-Formed Steel Design Manual (AISI,
2008) for the design of flexural members.
C3.1 Bending
Bending strengths [resistances] of flexural members are differentiated according to
whether or not the member is laterally braced. If such members are laterally supported, then
they are proportioned according to the nominal section strength [resistance] (Specification
Section C3.1.1). Since the distortional buckling has an intermediate buckling half wavelength,
the distortional buckling still needs to be considered even for braced members. See the Direct
Strength Method Design Guide (AISI, 2006) for detailed discussion and design examples. If
they are laterally unbraced, then the limit state is lateral-torsional buckling (Specification
Section C3.1.2). For C- or Z-sections with the tension flange attached to deck or sheathing and
with compression flange laterally unbraced, the bending capacity is less than that of a fully
braced member but greater than that of an unbraced member (Specification Section D6.1.1).
For C- or Z-sections supporting a standing seam roof system under gravity or uplift loads,
the bending capacity is greater than that of an unbraced member and may be equal to that of
a fully braced member (Specification Section D6.1.2). Similarly, for standing seam roof
systems, design provisions are provided in Specification Section D6.2.1 for evaluating the
bending strength of the system based on tests. The governing nominal bending strength
[resistance] is the smallest of the values determined from the applicable conditions.
July 2007 47
Chapter C, Members
Fy Fy Fy
Fy Fy Fy
(a)
Fy Fy Fy
(b)
Fy Fy Fy
Neutral
Neutral axis Neutral
axis axis
48 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
stress is known. The 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 successive approximation.
For determining the design flexural strength [factored resistance], φbMn, by using the
LRFD approach, slightly different resistance factors are used for the sections with
stiffened or partially stiffened compression flanges and the sections with unstiffened
compression flanges. These φb values were derived from the test results and a dead-to-
live load ratio of 1/5. They provide the β values from 2.53 to 4.05 (AISI, 1991; Hsiao, Yu
and Galambos, 1988a).
(b) Procedure II - 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 steel shapes have large width-to-thickness ratios
that are considerably in excess of the limits required by plastic design.
In the 1970s and early 1980s, research work on the inelastic strength of cold-formed
steel beams was carried out by Reck, Pekoz, Winter, and Yener at Cornell University
(Reck, Pekoz and Winter, 1975; Yener and Pekoz, 1985a, 1985b). These studies showed
that the inelastic reserve strength [resistance] 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 [resistance] can be utilized to achieve more economical design of
such members.
In order to utilize the available inelastic reserve strength [resistance] 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. The same provisions are
retained in the 2001 and the 2007 editions of the Specification. According to Procedure II
of Section C3.1.1(b) of the Specification, the nominal section strength [resistance], Mn, of
those beams satisfying certain specific limitations can be determined on the basis of the
inelastic reserve capacity with a limit of 1.25My, where My is the effective yield
moment. The ratio of Mn/My represents the inelastic reserve strength [resistance] of a
beam cross section.
The nominal moment Mn is the maximum bending capacity of the beam by
considering the inelastic reserve strength [resistance] through partial plastification of
the cross section. The inelastic stress distribution in the cross section depends on the
maximum strain in the compression flange, εcu. Based on the Cornell research work on
hat sections having stiffened compression flanges (Reck, Pekoz and Winter, 1975), the
AISI design provision limits the maximum compression strain to be Cyεy, where Cy is a
compression strain factor determined by using the equations provided in Specification
Section C3.1.1(b) (i) as shown in Figure C-C3.1.1-2.
On the basis of the maximum compression strain εcu allowed in the Specification, the
neutral axis can be located by using Equation C-C3.1.1-2 and the nominal moment Mn
can be determined by using Equation C-C3.1.1-3:
∫ σdA = 0 (C-C3.1.1-2)
∫ σydA = Mn (C-C3.1.1-3)
where σ is the stress in the cross section.
July 2007 49
Chapter C, Members
3
w/ t - λ 1
Cy = 3 - 2
λ2 - λ1
2
ε
Cy = εcu
y
0
0 λ1 λ2
1.11/ Fy /E 1.28/ Fy /E
w
t
50 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
π ⎛ π 2 EC w ⎞⎟
σ cr = EI y GJ⎜ 1 + (C-C3.1.2.1-1)
LS f ⎜ GJL 2 ⎟
⎝ ⎠
For other than simply supported end conditions, Equation C-C3.1.2.1-1 can be
generalized as given in Equation C-C3.1.2.1-1a (Galambos, 1998):
π ⎡ π 2 EC w ⎤
σ cr = EI y GJ ⎢1 + ⎥ (C-C3.1.2.1-1a)
(K y L y )S f ⎢⎣ GJ(K t L t ) 2 ⎥⎦
In the above equation, Ky and Kt are effective length factors and Ly and Lt are
unbraced lengths for bending about the y-axis and for twisting, respectively, E is the
modulus of elasticity, G is the shear modulus, Sf is the elastic section modulus of the full
unreduced section relative to the extreme compression fiber, Iy is the moment of inertia
about the y-axis, Cw is the torsional warping constant, J is the Saint-Venant torsion
constant, and L is the unbraced length.
For equal-flanged I-members with simply supported end conditions both laterally
and torsionally, Equation C-C3.1.2.1-2 can be used to calculate the elastic critical
buckling stress (Winter, 1947a; Yu, 2000):
2
⎛ Iy ⎞ ⎛
π2 E JI y ⎞⎛ L ⎞ 2
σ cr = ⎜ ⎟ +⎜ ⎟⎜ ⎟ (C-C3.1.2.1-2)
2(L/d)2 ⎜⎝ 2I x ⎟⎠ ⎜⎝ 2(1 + µ )I x 2 ⎟⎠⎝ πd ⎠
In Equation C-C3.1.2.1-2, the first term under the square root represents the lateral
bending rigidity of the member, and the second term represents the Saint-Venant
torsional rigidity. For thin-walled cold-formed steel sections, the first term usually
exceeds the second term by a considerable margin.
For simply supported I-members with unequal flanges, the following equation has
been derived by Winter for the lateral-torsional buckling stress (Winter, 1943):
⎛ ⎞
π 2 Ed ⎜ 4GJL2 ⎟
σ cr = ⎜ I yc - I yt + I y 1 + ⎟ (C-C3.1.2.1-3)
2L2 S f ⎜ π 2 I y Ed 2 ⎟
⎝ ⎠
where Iyc and Iyt are the moments of inertia of the compression and tension portions
of the full section, respectively, about the centroidal axis parallel to the web. Other
symbols were defined previously. For equal-flange sections, Iyc = Iyt = Iy/2, Equations
C-C3.1.2.1-2 and C-C3.1.2.1-3 are identical.
For other than simply supported end conditions, Equation C-C3.1.2.1-3 can be
generalized as given in Equation C-C3.1.2.1-3a:
⎛ ⎞
π 2 Ed ⎜ 4GJ(K t L t ) 2 ⎟
σ cr = ⎜ I yc - I yt + I y 1 + ⎟ (C-C3.1.2.1-3a)
2(K y L y ) 2 S f ⎜ π 2 I y Ed 2 ⎟
⎝ ⎠
In Equation C-C3.1.2.1-3a, the second term under the square root represents the
Saint-Venant torsional rigidity, which can be neglected without any loss in economy.
Therefore, Equation C-C3.1.2.1-3a can be simplified as shown in Equation C-C3.1.2.1-4
by considering Iy = Iyc + Iyt and neglecting the term 4GJ(KtLt)2/(π2IyEd2):
July 2007 51
Chapter C, Members
π 2 EdI yc
σ cr = (C-C3.1.2.1-4)
(K y L y ) 2 S f
Equation C-C3.1.2.1-4 was derived on the basis of a uniform bending moment and is
conservative for other cases. For this reason σcr is modified by multiplying the right
hand side by a bending coefficient Cb, to account for non-uniform bending and the
symbol Fe is used for σcr, i.e.,
C b π 2 EdI yc
Fe = (C-C3.1.2.1-5)
(K y L y ) 2 S f
where Cb is the bending coefficient, which can conservatively be taken as unity, or
calculated from
Cb = 1.75 + 1.05 (M1/M2) + 0.3 (M1/M2)2 ≤ 2.3 (C-C3.1.2.1-6)
in which M1 is the smaller and M2 the larger bending moment at the ends of the
unbraced length.
The above Equation was used in the 1968, 1980, 1986, and 1991 editions of the AISI
Specification. Because it is valid only for straight-line moment diagrams, Equation C-
C3.1.2.1-6 was replaced by the following equation for Cb in the 1996 edition of the AISI
Specification and is retained in this edition of the Specification:
12.5M max
Cb = (C-C3.1.2.1-7)
2.5M max + 3M A + 4M B + 3M C
where
Mmax = absolute value of maximum moment in the unbraced segment
MA = absolute value of moment at quarter point of unbraced segment
MB = absolute value of moment at centerline of unbraced segment
M1 M1 2
2.5 C b = 1.75 + 1.05 + 0.3 < 2.3
M2 M2
2.0
1.5
Cb
0.5 MA MB MC
M2 M1
M1
M2
52 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
July 2007 53
Chapter C, Members
Fc
10
F
9 y I- and C-Sections
Fy
Z-Sections
0.56F y
0
0 Lu Lu Unbraced Length, L
54 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
π 2 EC w
C2 = (C-C3.1.2.1-14)
(K t ) 2
(b) for I-, C- or Z-sections bent about the centroidal axis perpendicular to the web, the
following equations may be used in lieu of (a) (AISI, 1996):
For doubly-symmetric I-sections and singly-symmetric C-sections:
0. 5
1 ⎡⎢ 0.36C b π EdI yc ⎤⎥
2
Lu = (C-C3.1.2.1-15)
Ky ⎢ Fy S f ⎥
⎣ ⎦
For point-symmetric Z-sections:
0. 5
1 ⎡⎢ 0.18C b π EdI yc ⎤⎥
2
Lu = (C-C3.1.2.1-16)
Ky ⎢ Fy S f ⎥
⎣ ⎦
For members with unbraced length, L ≤ Lu, or elastic lateral-torsional buckling
stress, Fe ≥ 2.78Fy, the flexural strength [moment resistance] is determined in accordance
with C3.1.1(a).
The above discussion dealt only with the lateral-torsional buckling strength
[resistance] of locally stable beams. For locally unstable beams, the interaction of the
local buckling of the compression elements and overall lateral-torsional buckling of
members may result in a reduction of the lateral-torsional buckling strength [resistance]
of the member. The effect of local buckling on the critical moment is considered in
Section C3.1.2.1 of the Specification by using the elastic section modulus Sc based on an
effective section. i.e.,
Mn = FcSc (C-C3.1.2.1-17)
where
July 2007 55
Chapter C, Members
The problems discussed above dealt with the type of lateral-torsional buckling of I-
members, C-sections, and Z-shaped sections for which the entire cross section rotates
and deflects in the lateral direction as a unit. But this is not the case for U-shaped beams
and the combined sheet-stiffener sections as shown in Figure C-C3.1.2.1-3. For this case,
when the section is loaded in such a manner that the brims and 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 C-
56 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
July 2007 57
Chapter C, Members
provides for other than simply supported end conditions. Detailed discussions are
provided in Section C3.1.2.1 of the Commentary.
1.6
1.4
Spec. Eq. C3.1.3-2
1.25
1.2 Spec. Eq. C3.1.3-3
1.0
0.6
0.4
0.2 D/t=
0.441E/Fy D/t=0.318E/Fy D/t=0.0714E/Fy
0
0 2 4 6 8 10 12 14 16
E t
Fy D
58 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
All three equations for determining the nominal flexural strength [moment resistance]
of closed cylindrical tubular members are shown in Figure C-C3.1.3-1. These equations
have been used in the AISI Specification since 1986 and are retained in this Specification. In
1999, the limiting D/t ratios for Specification Equations C3.1.3-2 and C3.1.3-3 have been
revised to provide an appropriate continuity. The safety factor Ωb and the resistance factor
φb are the same as that used in Specification Section C3.1.1 for sectional bending strength.
July 2007 59
Chapter C, Members
bound estimate of kφ. The member lateral deformation may be removed from the
measured lateral deformation to provide a more accurate estimate of kφ.
Testing on 8 in. and 9.5 in. (203 and 241 mm) deep Z-sections with a thickness between
0.069 in. (1.75 mm) and 0.118 in. (3.00 mm), through-fastened 12 in. (205 mm) o.c., to a
36 in. (914 mm) wide, 1 in. (25.4 mm) and 1.5 in. (38.1 mm) high steel panels, with up to
6 in. (152 mm) of blanket insulation between the panel and the Z-section, results in a kφ
between 0.15 to 0.44 kip-in./rad./in. (0.667 to 1.96 kN-mm/rad./mm) (MRI 1981).
Additional testing on C- and Z-sections with pairs of through-fasteners provides
considerably higher rotational stiffness: for 6 and 8 in. (152 and 203 mm) deep C-sections
with a thickness between 0.054 and 0.097 in. (1.27 and 2.46 mm), fastened with pairs of
fasteners on each side of a 1.25 in. (31.8 mm) high steel panel flute at 12 in. (305 mm) o.c.,
kφ is 0.4 kip-in./rad./in. (1.78 kN-mm/rad./mm); and for 8.5 in. (216 mm) deep Z-sections
with a thickness between 0.070 in. and 0.120 in. (1.78 mm to 3.05 mm), fastened with pairs
of fasteners on each side of 1.25 in. (31.8 mm) high steel panel flute at 12 in. (305 mm) o.c.,
kφ is 0.8 kip-in./rad./in. (3.56 kN-mm/rad./mm) (Yu and Schafer 2003, Yu 2005).
Examples of rational engineering analysis to estimate the rotational stiffness are
provided in the Direct Strength Method Design Guide (AISI 2006). For a flexural member,
kφ can be approximated as:
kφ ≈ EI/(W/2) (C-C3.1.4-1)
where E is the modulus of the attached material, I is the moment of inertia of the
engaged attachment, and W is the member spacing. The primary complication in such a
method is determining how much of the attachment (decking, sheathing, etc.) is engaged
when the flange attempts to deform. For the Z-sections tested in Yu (2005) experimental kφ
is 0.8 kip-in./rad./in. (3.56 kN-mm/rad./mm). Using an estimate of EI/(W/2) the rational
engineering values are kφ of 9 kip-in./rad/in. (40.0 kN-mm/rad./mm) if the entire panel,
flutes and all, are engaged; kφ of 1.2 kip-in./rad/in. (5.34 kN-mm/rad./mm) if only the
corrugated bottom panel, but not the flutes, is engaged; and kφ of 0.003 kip-in./rad./in.
(0.0133 kN-mm/rad./mm) if plate bending of the t = 0.019 in. (0.483 mm) panel occurs. The
observed panel engagement is between the last two estimates, and assuming the
corrugated bottom pan, but not the 1.25 in. (31.8 mm) high flutes is engaged is reasonable.
For members with wood sheathing attached, little experimental information is
available. The problem has been studied numerically using the same paired fastener detail
as in Yu’s (2005) and Yu and Schafer (2003) tests but replacing the steel panel with a
simulated wood member, thickness = 0.5 in. (12.7 mm), E = 1000 ksi (6900 MPa), and µ =
0.3. The calculated kφ is 5.1 kip-in./rad./in. (22.7 kN-mm/rad./mm) for 6 and 8 in. (152 to
203 mm) deep C-sections with a thickness between 0.054 and 0.097 in. (1.37 and 2.46 mm);
and kφ is 4.1 kip-in./rad./in. (18.2 kN-mm/rad./mm) for 8.5 in. (216 mm) deep Z-sections
with thickness between 0.070 and 0.120 in. (1.78 mm and 3.05 mm). From calculations
assuming a fully engaged 1/2 in. (12.7 mm) thick wood sheet on top of C- or Z-section
members spaced 12 in. (305 mm) apart, kφ is predicted to be 1.7 kip-in./rad./in. (7.56 kN-
mm/rad./mm). Thus, use of EI/(W/2) provides a reasonably conservative approximation,
with I calculated assuming the full engagement of wood sheet.
The presence of moment gradient can also increase the distortional buckling moment
(or equivalently stress, Fd). However, this increase is lessened if the moment gradient
occurs over a longer length. Thus, in determining the influence of moment gradient (β) the
60 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
ratio of the end moments, M1/M2, and the ratio of the critical distortional buckling length
to the unbraced length, L/Lm, should both be accounted for. Yu (2005) performed elastic
buckling analysis with shell finite element models of C- and Z-sections under different
moment gradients to examine this problem. Significant scatter exists in the results,
therefore a lower bound prediction (Specification Equation C3.1.4-11) for the increase was
selected.
M y =107.53kip-in.
1
M cr / M y
0.5
0
0 1 2 3
10 10 10 10
half-wavelength (in.)
Figure C-C3.1.4-1 Rational Elastic Buckling Analysis of a Z-Section under Restrained Bending
Showing Local, Distortional, and Lateral-Torsional Buckling Modes
(a) Simplified Provision for Unrestrained C- and Z-Sections with Simple Lip Stiffeners
The provision of Specification Section C3.1.4(a) provides a conservative approximation
to the distortional buckling length, Lcr, and stress, Fd, for C- and Z-sections with simple lip
stiffeners bent about an axis perpendicular to the web. The provision ignores any
rotational restraint, which would restrain distortional buckling. The expressions were
specifically derived as a conservative simplification to those provided in Specification
Sections C3.1.4(b) and (c).
(b) For C- and Z-Sections or any Open Section with a Stiffened Compression Flange Extending to
One Side of the Web where the Stiffener is either a Simple Lip or a Complex Edge Stiffener
The provisions of Specification Section C3.1.4(b) provide a general method for
calculation of the distortional buckling stress, Fd, for any open section with an edge
stiffened compression flange, including complex edge stiffeners. The provisions of
Specification Section C3.1.4(b) also provide a more refined answer for any C- and Z-section
including those meeting the criteria of C3.1.4(a). The expressions employed here are
derived in Schafer and Peköz (1999) and verified for complex stiffeners in Schafer et al.
(2006). The equations used for the distortional buckling stress, Fd, in AS/NZS 4600 are also
similar to those in Specification Section C3.1.4 (b), except that when the web is very slender
and is restrained by the flange, AS/NZS 4600 uses a simpler, conservative treatment. Since
the provided expressions can be complicated, solutions for the geometric properties of C-
and Z-sections based on centerline dimensions are provided in Table C-C3.1.4(b)-1.
July 2007 61
Chapter C, Members
1.2
Eq. C3.1.4-2
Distortional buckling tests
1.1
0.9
Mtes t/My
0.8
0.7
0.6
0.5
0.4
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
0 .5
(My /Mc r )
62 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
Table C-C3.1.4(b)-1
Geometric flange properties for C- and Z-sections
b b
θ d θ
h h
A f = (b + d )t A f = (b + d )t
J f = 1 3 bt 3 + 1 3 dt 3 J f = 1 3 bt 3 + 1 3 dt 3
I xf =
(
t t 2 b 2 + 4 bd 3 + t 2 bd + d 4 ) I xf =
( )
t t 2 b 2 + 4 bd 3 − 4 bd 3 cos 2 (θ ) + t 2 bd + d 4 − d 4 cos 2 (θ)
12(b + d ) 12(b + d )
I yf =
(
t b 4 + 4db 3 ) I yf =
(
t b 4 + 4db 3 + 6d 2 b 2 cos(θ) + 4d 3 b cos 2 (θ ) + d 4 cos 2 (θ ) )
12(b + d ) 12(b + d )
tb 2 d 2 tbd 2 sin (θ )(b + d cos(θ ))
I xyf = I xyf =
4(b + d ) 4(b + d )
C wf = 0 C wf = 0
b2 b 2 − d 2 cos(θ )
xo = xo =
2(b + d ) 2(b + d )
hx =
(
− b 2 + 2db ) hx =
(
− b 2 + 2db + d 2 cos(θ) )
2 (b + d ) 2(b + d )
− d2 − d 2 sin (θ )
h y = yo = h y = yo =
2 (b + d ) 2 (b + d )
C3.2 Shear
C3.2.1 Shear Strength [Resistance] of Webs without Holes
The shear strength [resistance] of beam webs is governed by either yielding or
buckling, depending on the h/t ratio and the mechanical properties of steel. For beam
webs having small h/t ratios, the nominal shear strength [resistance] is governed by shear
yielding, i.e.,
Vn = A w τ y = Α w Fy / 3 ≈ 0.60Fy ht (C-C3.2.1-1)
in which Aw is the area of the beam web computed by (ht), and τy is the yield stress of steel
in shear, which can be computed by Fy / 3 .
For beam webs having large h/t ratios, the nominal shear strength [resistance] is
governed by elastic shear buckling (Yu, 2000), i.e.,
k v π 2 EA w
Vn = A w τ cr = (C-C3.2.1-2)
12(1 − µ 2 )( h/t ) 2
in which τcr is the critical shear buckling stress in the elastic range, kv is the shear buckling
coefficient, E is the modulus of elasticity, µ is the Poisson’s ratio, h is the web depth, and t
is the web thickness. By using µ = 0.3, the shear strength [resistance], Vn, can be
determined as follows:
July 2007 63
Chapter C, Members
Vn = 0.904Ek v t 3 / h (C-C3.2.1-3)
For beam webs having moderate h/t ratios, the nominal shear strength [resistance] is
based on inelastic shear buckling (Yu, 2000), i.e.,
Vn = 0.64 t 2 k v Fy E (C-C3.2.1-4)
The Specification provisions are applicable for the design of webs of beams and decks
either with or without transverse web stiffeners.
The nominal strength [resistance] equations of Section C3.2.1 of the Specification are
similar to the nominal shear strength [resistance] equations given in the AISI LRFD
Specification (AISI, 1991). The acceptance of these nominal strength [resistance] equations
for cold-formed steel sections has been considered in the study summarized by LaBoube
and Yu (1978a).
Previous editions of the AISI ASD Specification (AISI, 1986) used three different safety
factors when evaluating the allowable shear strength [resistance] of an unreinforced web
because it was intended to use the same nominal strength [resistance] equations for the
AISI and AISC Specifications. To simplify the design of shear using only one safety factor
for ASD and one resistance factor for LRFD, Craig (Craig, 1999) carried out a calibration
using the data by LaBoube and Yu (LaBoube, 1978a). Based on this work, the constant used
in Specification Equation C3.2.1-3 was reduced from 0.64 to 0.60. In addition, the ASD
safety factor for yielding, elastic and inelastic buckling is now taken as 1.60, with a
corresponding resistance factor of 0.95 for LRFD and 0.80 for LSD.
64 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
1.2
A B
1.0
Eq. C-C3.3-3
0.8
C
Eq. C-C3.3-1
t h = 120
t max 0.6 t
h = 150
t
0.4
h = 200
t
Note: Shaded symbols represent test
0.2 specimens without additional sheets
on top and bottom flanges.
0 D
0 0.2 0.4 0.6 0.8 1.0 1.2
fb
fb max
July 2007 65
Chapter C, Members
The above equation was added to the AISI Specification in 1980. The correlations between
Equation C-C3.3-3 and the test results of beam webs having a diagonal tension field action
are shown in Figure C-C3.3-1.
(a) (b)
In the past, the buckling problem of plates and the web crippling behavior of cold-
formed steel members under locally distributed edge loading have been studied by
numerous investigators (Yu, 2000). It has been found that the theoretical analysis of web
crippling for cold-formed steel flexural members is rather complicated because it involves
the following factors: (1) nonuniform stress distribution under the applied load and
66 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
adjacent portions of the web, (2) elastic and inelastic stability of the web element, (3) local
yielding in the immediate region of load application, (4) 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, (5) initial out-of-plane imperfection of plate elements, (6) various
edge restraints provided by beam flanges and interaction between flange and web
elements, and (7) inclined webs for decks and panels.
For these reasons, the present AISI design provision for web crippling is based on the
extensive experimental investigations conducted at Cornell University by Winter and Pian
(1946) and Zetlin (1955a); at the University of Missouri-Rolla by Hetrakul and Yu (1978 and
1979), Yu (1981), Santaputra (1986), Santaputra, Parks and Yu (1989), Bhakta, LaBoube and
Yu (1992), Langan, Yu and LaBoube (1994), Cain, LaBoube and Yu (1995) and Wu, Yu and
LaBoube (1997); at the University of Waterloo by Wing (1981), Wing and Schuster (1982),
Prabakaran (1993), Gerges (1997), Gerges and Schuster (1998), Prabakaran and Schuster
(1998), Beshara (1999), and Beshara and Schuster (2000 and 2000a); and at the University of
Sydney by Young and Hancock (1998). 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, single hat sections and multi-web deck
sections:
1. End one-flange (EOF) loading
2. Interior one-flange (IOF) loading
3. End two-flange (ETF) loading
4. Interior two-flange (ITF) loading
All loading conditions are illustrated in Figure C-C3.4.1-2. 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. Application of the various load cases is shown in
Region
of failure
Region Region
of failure of failure
(a) (b)
(c) (d)
July 2007 67
Chapter C, Members
(a)
Lo
End One-Flange
Loading
> 1.5h
(b)
Figure C-C3.4.1-3 and the assumed reaction or load distributions are illustrated in Figure
C-C3.4.1-4.
In the 1996 edition of the AISI Specification, and in previous editions, different web
crippling equations were used for the various loading conditions stated above. These
equations were based on experimental evidence (Winter, 1970; Hetrakul and Yu, 1978) and
the assumed distributions of loads or reactions acting on the web as shown in Figure C-
C3.4.1-4. The equations were also based on the type of section geometry, i.e., shapes having
single webs and I-sections (made of two channels connected back to back, by welding two
angles to a channel, or by connecting three channels). C-and Z-sections, single hat sections
and multi-web deck sections were considered in the single web member category. I-
sections made of two channels connected back to back by a line of connectors near each
flange or similar sections that provide a high degree of restraint against rotation of the web
were treated separately. In addition, different equations were used for sections with
stiffened or partially stiffened flanges and sections with unstiffened flanges.
68 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
>1.5h >1.5h
<1.5h
<1.5h
July 2007 69
Chapter C, Members
Prabakaran (1993) and Prabakaran and Schuster (1998) developed one consistent
unified web crippling equation with variable coefficients (Specification Equation C3.4.1-1).
These coefficients accommodate one or two flange loading for both end and interior
loading conditions of various section geometries. Beshara (1999) extended the work of
Prabakaran and Schuster (1998) by developing new web crippling coefficients using the
available data as summarized by Beshara and Schuster (2000). The web crippling
coefficients are summarized in Tables C3.4.1-1 to C3.4.1-5 of the Specification and the
parametric limitations given are based on the experimental data that was used in the
development of the web crippling coefficients. From Specification Equation C3.4.1-1, it can
be seen that the nominal web crippling strength of cold-formed steel members depends on
an overall web crippling coefficient, C, the web thickness, t, the yield stress, Fy, the web
inclination angle, θ, the inside bend radius coefficient, CR, the inside bend radius ratio, R/t,
the bearing length coefficient, CN, the bearing length ratio, N/t, the web slenderness
coefficient, Ch, and the web slenderness ratio, h/t.
This new equation is presented in a normalized format and is non-dimensional,
allowing for any consistent system of measurement to be used. Consideration was given to
whether or not the test specimens were fastened to the bearing plate/support during
testing. It was discovered that some of the test specimens in the literature were not
fastened to the bearing plate/support during testing, which can make a considerable
difference in the web crippling capacity of certain sections and loading conditions.
Therefore, it was decided to separate the data on the basis of members being fastened to
the bearing plate/support and those not being fastened to the bearing plate/support. The
fastened to the bearing plate/support data in the literature were primarily based on
specimens being bolted to the bearing plate/support, hence, a few control tests were
carried out by Schuster, the results of which are contained in (Beshara 1999), using self-
drilling screws to establish the web crippling integrity in comparison to the bolted data.
Based on these tests, the specimens with self-drilling screws performed equally well in
comparison to the specimens with bolts. Fastened to the bearing plate/support in practice
can be achieved by either using bolts, self-drilling/self-tapping screws or by welding.
What is important is that the flange elements are restrained from rotating at the location of
load application. In fact, in most cases, the flanges are frequently completely restrained
against rotation by some type of sheathing material that is attached to the flanges.
The data was further separated in the Specification based on section type, as follows.
1) Built-up sections (Table C3.4.1-1);
2) Single web channel and C-sections (Table C3.4.1-2);
3) Single web Z-sections (Table C3.4.1-3);
4) Single hat sections (Table C3.4.1-4); and
5) Multi-web deck sections (Table C3.4.1-5).
In the case of unfastened built-up members such as I-sections (not fastened to the
bearing plate/support), the available data was for specimens that were fastened together
with a row of fasteners near each flange line of the member (Winter and Pian 1946) and
Hetrakul and Yu (1978) as shown in Figure C-C3.4.1-5(a). For the fastened built-up
member data of I-sections (fastened to the bearing plate/support), the specimens were
fastened together with two rows of fasteners located symmetrically near the centerline
length of the member, as shown in Figure C-C3.4.1-5(b) (Bhakta, LaBoube and Yu, 1992).
70 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
Calibrations were carried out by Beshara and Schuster (2000) in accordance with
Supornsilaphachai, Galambos and Yu (1979) to establish the safety factors, Ω, and the
resistance factors, φ, for each web crippling case. Based on these calibrations, different
safety factors and corresponding resistance factors are presented in the web crippling
coefficient tables for the particular load case and section type. In 2005, the safety and the
resistance factors for built-up and single hat sections with interior one-flange loading case
have been revised based on a more consistent calibration. For the fastened built-up
sections, the factors were revised from 1.65 to 1.75 (for ASD), 0.90 to 0.85 (for LRFD) and
0.80 to 0.75 (for LSD). For the fastened single hat section, the factors were revised from
1.90 to 1.80 (for ASD) and 0.80 to 0.85 (for LRFD). For the unfastened single hat sections,
the factors were revised from 1.70 to 1.80 (for ASD), 0.90 to 0.80 (for LRFD) and 0.75 to 0.70
(for LSD). Also in 2005 the coefficients for built-up sections were revised to remove
inconsistencies between unfastened and fastened conditions and to better reflect the
calibration for the safety factor and the resistance factors. Also, a minimum bearing length
of 3/4 in. (19 mm) was introduced based on the data used in the development of the web
crippling coefficients. For fastened to support single web C- and Z-section members under
interior two-flange loading or reaction, the distance from the edge of bearing to the end of
the member (Figure C-C3.4.1-2(d)) must be extended at least 2.5h. This requirement is
necessary because a total of 5h specimen length was used for the test setup shown in
Figure C-C3.4.1-2(d) (Beshara, 1999). The 2.5h length is conservatively taken from the edge
of bearing rather than the centerline of bearing.
The assumed distributions of loads or reactions acting on the web of a member, as
shown in Figure C-C3.4.1-4, are independent of the flexural response of the member. Due
to the flexural action, the point of bearing will vary relative to the plane of bearing,
resulting in a non-uniform bearing load distribution on the web. The value of Pn will vary
because of a transition from the interior one-flange loading (Figure C3.4.1-4(b)) to the end
one-flange loading (Figure C3.4.1-4(a)) condition. These discrete conditions represent the
experimental basis on which the design provisions were founded (Winter, 1970; Hetrakul
and Yu, 1978). Based on additional updated calibrations, the resistance factor for Canada
LSD for the unfastened interior one-flange loading (IOF) case in Table C3.4.1-4 was
changed from 0.75 to 0.70 in 2004.
t t
(a) Winter and Pian 1946 (b) Bhakta, LaBoube and Yu 1992
Hetrakul and Yu 1978
Figure C-C3.4.1-5 Typical Bolt Pattern for I-Section Test Specimens
The research indicates that a Z-section having its end support flange bolted to the
section’s supporting member through two 1/2-in. (12.7 mm) diameter bolts will experience
an increase in end-one-flange web crippling capacity (Bhakta, LaBoube and Yu, 1992; Cain,
July 2007 71
Chapter C, Members
LaBoube and Yu, 1995). The increase in load-carrying capacity was shown to range from 27
to 55 percent for the sections under the limitations prescribed in the Specification. A lower
bound value of 30 percent increase was permitted in Specification Section C3.4 of the 1996
Specification. This is now incorporated under “Fastened to Support” condition.
In 2005, the R/t limit in Table C3.4.1-3 regarding Interior-one-flange loading for
fastened Z-sections was changed from 5 to 5.5 to achieve consistency with Specification
Equations C3.5.1-3 and C3.5.2-3 which stipulate a limit of R/t = 5.5.
For two nested Z-sections, the 1996 AISI Specification permitted the use of a slightly
different safety factor and resistance factor for the interior one flange loading condition.
This is no longer required since the new web crippling approach now takes this into
account in Table C3.4.1-3 of the Specification under the category of “Fastened to Support”
for the interior one flange loading case.
The previous web crippling coefficients in Table C3.4.1-5 for end one flange loading
(EOF) of multi-web deck sections in the design provisions (AISI 2001) were based on
limited data. This data was based on specimens that were not fastened to the support
during testing, hence, the previous coefficients for this case were also being used
conservatively for the case of fastened to the support. Based on extensive testing, web
crippling coefficients were developed by James A. Wallace (2003) for both the unfastened
and fastened case of EOF loading. Calibrations were also carried out to establish the
respective safety factors and resistance factors.
In 2004, the definitions of “one-flange loading” and “two-flange loading” were revised
according to the test setup, specimen lengths, development of web crippling coefficients,
and calibration of safety factors and resistance factors. In Figures C-C3.4.1-3 and C-C3.4.1-
4 of the Commentary, the distances from the edge of bearing to the end of the member were
revised to be consistent with the Specification.
Specification Equation C3.4.1-2 for single web C- and Z-sections with an overhang or
overhangs is based on a study of the behavior and resultant failure loads from an end-one-
flange loading investigation performed at the University of Missouri-Rolla (Holesapple
and LaBoube, 2002). This Equation is applicable within the limits of the investigation. The
UMR test results indicated that in some situations with overhangs, the interior-one-flange
load capacity may not be achieved and the interior-one-flange loading condition was
therefore removed from Figures C-C3.4.1-3 and C-C3.4.1-4.
72 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
July 2007 73
Chapter C, Members
(a) Decks
Deck or cladding
<10"
(b) Beams Deck, cladding
or braces
Figure C-C3.5-1 Sections Used for Exception Clause of Specification Section C3.5
74 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
supporting a through fastened roof panel that will prevent movement in the plane of the roof
panel. It is important that the designer ensure that torsion is adequately constrained when
evaluating a specific situation.
In situations where torsional loading cannot be avoided, discrete bracing will reduce the
torsional effects. Torsional bracing at the third points of the span would be adequate for
most light construction applications. The bracing should be designed to prevent torsional
twisting at the braced points.
Specification Section C3.6 provides design criteria for a member that is subjected to
torsional loading. The provision uses a reduction factor to reduce the nominal moment
strength as determined by Specification Section C3.1.1(a) This reduction factor requires
calculation of both the usual bending stresses and the torsional warping stresses at critical
points on the cross-section. The largest combination of these is the denominator of the
reduction factor while the bending stress alone at this same point is the numerator. The
member is then selected based on bending alone with the effect of torsion accounted for by
the reduction in the nominal moment capacity.
The largest combination of compression stresses on the cross section may occur at the
junction of the web and flange or at the junction of the edge of flange and flange stiffener.
The second condition has the more severe effect on reducing the moment capacity of the
member. This can occur when the applied load is positioned off the member away from both
the web and the shear center. This is shown from the test results reported in the referenced
paper by Bogdan M. Put and others (Put et al., 1999). This is not a practical situation for
structural assemblies, however this location of the critical compression stresses would occur
at the position of a torsional brace located at mid-span of a member. To allow for the more
favorable situation, the provisions of Specification Section C3.6 permit the moment capacity to
be increased by 15% when the critical combination of compressive stresses occurs at the
junction of the flange and web. This is also supported by tests on channels conducted by
Winter in 1950 (Winter et al., 1950), which indicated that an overstress of 15% did not
significantly affect the carrying capacity.
The provisions of this Section need not be used in combination with the bending
provisions in Specification Sections D6.1.1 and D6.1.2 since these provisions are based on tests
in which torsional effects were present.
C3.7 Stiffeners
C3.7.1 Bearing Stiffeners
Design requirements for attached bearing stiffeners (previously called transverse
stiffeners) and for shear stiffeners were added in the 1980 AISI Specification and were
unchanged in the 1986 Specification. The same design equations are retained in Section
C3.7 of the current Specification. The term “transverse stiffener” was renamed to “bearing
stiffeners” in 2004. The nominal strength [resistance] equation given in Item (a) of
Specification Section C3.7.1 serves to prevent end crushing of the bearing stiffeners, while
the nominal strength [resistance] equation given in Item (b) is to prevent column-type
buckling of the web-stiffeners. The equations for computing the effective areas (Ab and Ac)
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 bearing stiffeners were evaluated
July 2007 75
Chapter C, Members
by Hsiao, Yu and Galambos (1988a). A total of 61 tests were examined. The resistance
factor of 0.85 used for the LRFD method was selected on the basis of the statistical data.
The corresponding reliability indices vary from 3.32 to 3.41.
In 1999, the upper limit of w/ts ratio for the unstiffened elements of cold-formed steel
bearing stiffeners was revised from 0.37 E Fys to 0.42 E Fys for the reason that the
former was calculated based on the allowable strength design approach, while the latter is
based on the effective area approach. The revision provided the same basis for the
stiffened and unstiffened elements of cold-formed steel bearing stiffeners.
76 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
flexural-torsional buckling), (3) local buckling of individual elements, and (4) distortional
buckling. The first three limit states are discussed in Section C4.1 and ditortional buckling limit
state is discussed in Section C4.2. For the design tables and example problems on columns, see
Parts I and III of the AISI Cold-Formed Steel Design Manual (AISI, 2008).
July 2007 77
Chapter C, Members
78 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
Pn = AeFcr (C-C4.1-8)
where Fcr is either elastic buckling stress or inelastic buckling stress whichever is
applicable, and Ae is the effective area at Fcr.
In the 1986 edition of the AISI Specification, the nominal axial strength [resistance] for C-
1
or
AISI 1996 and Eq. C-C4-3
Fn 2001 Specifications
0.4
Fy
Eq. C-C4-11
0.2
0 0.5 1 1.5 2
lc
0.6
Based on the AISI 1986
Specification and Variable F.S.
0.5
0.1
0 0.5 1 1.5 2
lc
July 2007 79
Chapter C, Members
0.8
Based on the AISI 1991
0.6
Pn
Py
Based on the AISI 1996 and
0.4 2001 Specifications
0.2
0 0.5 1 1.5 2
lc
KL = L
P
Figure C-C4.1-4 Overall Column Buckling
and Z-sections and single angle sections was limited by Equation C-C4.1-9, which is
determined by the local buckling stress of the unstiffened element and the area of the full
cross-section:
Aπ 2 E
Pn = (C-C4.1-9)
25.7( w/t ) 2
This equation was deleted since the 1996 edition of the Specification based on a study
80 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
July 2007 81
Chapter C, Members
[compressive resistances] used for the 1991, 1996, 2001 and the 2007 LRFD design
provisions are shown in Figure C-C4.1-3. The curve for the LSD provisions would be
the same as the curve for LRFD.
(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 KL, being the length of this half-wave, is equal
to the actual physical length of the column (Figure C-C4.1-4); correspondingly, for this
case, K = 1. This situation is approached if a given compression member is part of a
structure which is braced in such a 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 which 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 = 1.
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 which provides
substantial restraint against rotation, K-values smaller than 1 (one) are sometimes
justified. Table C-C4.1-1 provides the theoretical K values for six idealized conditions in
which joint rotation and translation are either fully realized or nonexistent. The same
table also includes the K values recommended by the Structural Stability Research
Council for design use (Galambos, 1998).
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
Table C-C4.1-1
Effective Length Factors K for Concentrically Loaded
Compression Members
(a) (b) (c) (d) (e) (f)
82 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
approach the yield stress which greatly reduces the restraint they can provide. For this
reason K value is usually taken as unity regardless of whether they are welded, bolted, or
connected by screws. However, when sheathing is attached directly to the top flange of a
continuous compression chord, 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 of Figure C-C4.1-5, the structure depends on its own bending stiffness for
lateral stability. In this case, when failure occurs by buckling of the columns, it invariably
takes place by the sidesway motion shown. This occurs at a lower load than the columns
would be able to carry if they where braced against sidesway and the figure shows that
KL P P
5.0
4.0
Hinged
3.0 base
(I/L) beam
(I/L) column
2.0
Fixed
1.0 base
0
1.0 2.0 3.0 4.0
K
Figure C-C4.1-6 Effective Length Factor K in Laterally
Unbraced Portal Frames
July 2007 83
Chapter C, Members
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 (one) and its value can be read from the
graph of Figure C-C4.1-6 (Winter et al., 1948a and Winter, 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.
Figure C-C4.1-6 can also serve as a guide for estimating K for other simple situations.
For multi-bay and/or multi-story frames, simple alignment charts for determining K are
given in the AISC Commentaries (AISC, 1989; 1999; 2005). For additional information on
frame stability and second order effects, see SSRC Guide to Stability Design Criteria for
Metal Structures (Galambos, 1998) and the AISC Specifications and Commentaries.
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 must be considered when functioning as horizontal diaphragms (Winter,
1958a).
C. Torsional Buckling of Columns
It was pointed out at the beginning of this section that purely torsional buckling, i.e.,
failure by sudden twist without concurrent bending, is also possible for certain cold-
formed open shapes. These are all point-symmetric shapes (in which shear center 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 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 σt calculated as follows (Winter, 1970):
1 ⎡ π 2 EC w ⎤
σt = ⎢ GJ + ⎥ (C-C4.1-13)
Aro2 ⎣⎢ ( K t L t ) 2 ⎦⎥
The above equation is the same as Specification Equation C3.1.2.1-9, in which A is the full
cross-sectional area, ro is the polar radius of gyration of the cross section about the shear
center, G is the shear modulus, J is Saint-Venant torsion constant of the cross section, E is
the modulus of elasticity, Cw is the torsional warping constant of the cross section, and Kt
Lt is the effective length for twisting.
For inelastic buckling, the critical torsional buckling stress can also be calculated
according to Equation C-C4.1-10 by using σt as Fe in the calculation of λc.
D. Flexural-Torsional Buckling of Columns
As discussed previously, 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 center; 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
center and centroid do not coincide, flexural-torsional buckling is one of the possible
buckling modes as shown in Figure C-C4.1-7. Unsymmetric sections will always buckle in
the flexural-torsional mode.
84 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
P
Shear Centroid
Center
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 sheathing 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 which prevents
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 of a column can be found from the
following equation, (Chajes and Winter, 1965; Chajes, Fang and Winter, 1966; Yu, 2000):
1 ⎡
Pn = (Px + Pz ) − (Px + Pz )2 − 4βPx Pz ⎤⎥ (C-C4.1-14)
2β ⎢⎣ ⎦
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 Fe as follows:
1 ⎡
Fe = ⎢ (σ ex + σ t ) − (σ ex + σ t )2 − 4βσ ex σ t ⎤⎥ (C-C4.1-15)
2β ⎣ ⎦
For this equation, as in all provisions which deal with flexural-torsional buckling, the x-
axis is the axis of symmetry; σex = π2E/(KxLx/rx)2 is the flexural Euler buckling stress
about the x-axis, σt is the torsional buckling stress (Equation C-C4.1-13) and β=1-(xo/ro)2.
It is worth noting that the flexural-torsional buckling stress is always lower than the Euler
July 2007 85
Chapter C, Members
stress σex 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 Equation C-C4.1-10.
An inspection of Equation C-C4.1-15 will show that in order to calculate β and σt, it is
necessary to determine xo = distance between shear center and centroid, J = Saint-Venant
torsion constant, and Cw = 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 simple as that for flexural buckling. Formulas
for typical C-, Z-sections, angle and hat sections are provided in Part I of the Design
Manual (AISI, 2008).
For one thing, any singly-symmetric shape can buckle either flexurally about the y-axis
or flexural-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. If flexural-torsional buckling is indicated, the
information and design aids in Parts I and VII of the AISI Design Manual (AISI, 2008)
facilitate and expedite the necessary calculations.
The above discussion refers to members subject to flexural-torsional buckling, but made
up of elements whose w/t ratios are small enough so that no local buckling will occur.
For shapes which are sufficiently thin, i.e., with w/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.
E. Additional Design Consideration for Angles
During the development of a unified approach to the design of cold-formed steel
members, Pekoz realized the possibility of a reduction in column strength due to initial
sweep (out-of-straightness) of angle sections. Based on an evaluation of the available test
results, an initial out-of-straightness of L/1000 was recommended by Pekoz for the
design of concentrically loaded compression angle members and beam-columns in the
1986 edition of the AISI Specification. Those requirements were retained in Sections C4.1,
C5.2.1, and C5.2.2 of the 1996 edition of the Specification. A study conducted at the
University of Sydney (Popovic, Hancock, and Rasmussen, 1999) indicated that for the
design of singly-symmetric unstiffened angles sections under the axial compression load,
the required additional moment about the minor principal axis due to initial sweep
should only be applied to those angle sections, for which the effective area at stress Fy is
less than the full, unreduced cross-sectional area. Consequently, clarifications have been
made in Sections C5.2.1 and C5.2.2 of the 2001 edition of the AISI Specification to reflect
the research findings.
F. Slenderness Ratios
The slenderness ratio, KL/r, of all compression members preferably should not exceed
200, except that during construction only, KL/r should not exceed 300. In 1999, the above
86 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
July 2007 87
Chapter C, Members
or
1 1 1
= + (C-C4.1-18)
Fe σ ex σ t
Research at the University of Sydney (Popovic, Hancock, and Rasmussen, 1999) has
shown that singly-symmetric unstiffened cold-formed steel angles, which have a fully
effective cross-section under yield stress, do not fail in a flexural-torsional mode and can be
designed based on flexural buckling alone as specified in Specification Section C4.1.1. There
is also no need to include a load eccentricity for these sections when using Specification
Section C5.2.1 or Section C5.2.2 as explained in Item E of Section C4.1.
88 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
important effect of the initial imperfection, the design provisions included in the AISI
Specification 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 Fult/Fy depends on
the parameter (E/Fy)(t/D), in which t is the wall thickness, D is the mean diameter of the
tube, and Fult is the ultimate stress or collapse stress. As shown in Figure C-C4.1-8, 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 stress, and line 3 represents the
collapse stress occurring at yield stress. In the range of line 3, local buckling will not occur
before yielding. In ranges 1 and 2, local buckling occurs before the yield stress is reached.
The cylindrical tubes should be designed to safeguard against local buckling.
Elastic
buckling Inelastic buckling Yielding
A1 3
1.0
2
0.6
F ult
Fy
0.4 1 Eq. (C-C4.1-21)
0.2
For D/t = 0.441E/Fy For D/t = 0.112E/Fy
0
2.27 8.93
0 2 4 6 8 10 12
E t
Fy D
Based on a conservative approach, the Specification specifies that when the D/t ratio is
smaller than or equal to 0.112E/Fy, the tubular member shall be designed for yielding. This
provision is based on point A1, for which (E/Fy)(t/D) = 8.93.
When 0.112E/Fy < D/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 B1,
⎛ E ⎞⎛ t ⎞ Fult
⎜ ⎟⎜ ⎟ = 2.27 , = 0.75 (C-C4.1-19)
⎜ Fy ⎟⎝ D ⎠ Fy
⎝ ⎠
Using line A1B1, the maximum stress of cylindrical tubes can be represented by
July 2007 89
Chapter C, Members
Fult ⎛ E ⎞⎛ t ⎞
= 0.037 ⎜ ⎟⎜ ⎟ + 0.667 (C-C4.1-20)
Fy ⎜ Fy ⎟⎝ D ⎠
⎝ ⎠
When D/t ≥ 0.441E/Fy, the following equation represents Line 1 for elastic local
buckling stress:
Fult ⎛ E ⎞⎛ t ⎞
= 0.328⎜ ⎟⎜ ⎟ (C-C4.1-21)
Fy ⎜ Fy⎟⎝ D ⎠
⎝ ⎠
The correlations between the available test data and Equations C-C4.1-20 and C-C4.121
are shown in Figure C-C4.1-9. The definition of symbol “D” was changed from “mean
diameter” to “outside diameter” in the 1986 AISI Specification in order to be consistent with
the general practice.
1.2
A1
1.0
0.8
B1
Fult
0.6 Eq. (C-C4.1-20)
Fy
0.4
University of Illinois
Lehigh University
Eq. (C-C4.1-21)
University of Tokyo
0.2 University of Alberta
0
0 2 4 6 8 10 12 14 16 18
E t
Fy D
Figure C-C4.1-9 Correlation between Test Data and AISI Criteria for Local Buckling of
Cylindrical Tubes under Axial Compression
90 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
where
R = Fy / 2Fe (C-C4.1-23)
⎡ 0.037 ⎤
Ao = ⎢ + 0.667 ⎥A ≤ Α (C-C4.1-24)
⎢⎣ DFy / tE ⎥⎦
A = area of the unreduced cross section.
Equation C-C4.1-24 is used for computing the reduced area due to local buckling. It is
derived from Equation C-C4.1-20 for inelastic local buckling stress (Yu, 2000).
In 1999, the coefficient, R, was limited to one (1.0) so that the effective area, Ae, will
always be less than or equal to the unreduced cross sectional area, A. To simplify the
equations, R = Fy/(2Fe) is used rather than R = Fy /( 2 Fe ) as in the previous edition of the
AISI Specification. The equation for the effective area is simplified to Ae = Ao + R(A - Ao) as
given in Equation C4.1.5-1 of the North American Specification.
0.5
Z-section with lips (AISI 2002 Ex. I-10)
0.45
0.4
P y =45.23kips
0.35
Flexural
0.3
Pcr / Py
0.25
0.15
0.1
0.05
0
0 1 2 3
10 10 10 10
half-wavelength (in.)
July 2007 91
Chapter C, Members
Section C4.2 Distortional Buckling Strength provisions need not be applied since distortional
buckling is inherently included as a limit state in the Section D6.1.3 strength prediction
equations.
Determination of the nominal strength in distortional buckling (Specification Equation
C4.2-2) was validated by testing. Equation C4.2-2 was originally developed for the Direct
Strength Method of Appendix 1 of the Specification. Calibration of the safety and resistance
factors for Specification Equation C4.2-2 is provided in the commentary to Appendix 1. In
addition, the Australian/New Zealand Specification (AS/NZS 4600) has used an expression
of similar form to Specification Equation C4.2-2, but yielding slightly less conservative
strength predictions than Equation C4.2-2, since 1996.
Distortional buckling is unlikely to control the strength of a column if (a) the web is
slender and triggers local buckling far in advance of distortional buckling, as is the case for
many common C-sections, (b) edge stiffeners are sufficiently stiff and thus stabilize the flange
(as is often the case for C-sections, but typically not for Z-sections due to the use of sloping lip
stiffeners), (c) unbraced lengths are long and flexural or flexural-torsional buckling strength
limits the capacity, or (d) adequate rotational restraint is provided to the flanges from
attachments (panels, sheathing, etc.).
The primary difficulty in calculating the strength in distortional buckling is to efficiently
estimate the elastic distortional buckling stress, Fd. Recognizing the complexity of this
calculation this section of the Specification provides three alternatives: Specification Section
C4.2(a) provides a conservative prediction for unrestrained C- and Z-sections, Section C4.2(b)
provides a more comprehensive method for C- and Z-Section members and any open section
with a single web and flanges of the same dimension, and Section C4.2(c) offers the option to
use rational elastic buckling analysis. See the Appendix 1 commentary for further discussion.
The equations of Section C4.2(a) assume the compression flange is unrestrained; however, the
methods of Sections C4.2(b) and (c) allow for a rotational restraint, kφ, to be included to
account for attachments which restrict flange rotation. Additional guidance on kφ is provided
in the Commentary Section C3.1.4.
(a) Simplified Provision for Unrestrained C- and Z-sections with Simple Lip Stiffeners
The provision of Specification Section C4.2(a) provides a conservative approximation to
the distortional buckling stress, Fd, for C- and Z-sections with simple lip stiffeners. The
expressions were specifically derived as a conservative simplification to those provided in
Sections C4.2(b) and (c). For many common sections the provisions of Section C4.2(a) may
be used to show that distortional buckling of the column will not control the capacity.
(b) For C- and Z-Sections or Hat Sections or Any Open Section with Stiffened Flanges of Equal
Dimension where the Stiffener is either a Simple Lip or a Complex Edge Stiffener
The provisions of Specification Section C4.2(b) provide a general method for calculation
of the distortional buckling stress, Fd, for any open section with equal edge stiffened
compression flanges, including those with complex edge stiffeners. The provisions of
Specification Section C4.2(b) also provide a more refined answer for any C- and Z-section
including those meeting the criteria of Section C4.2(a). The expressions employed here are
derived in Schafer (2002) and verified for complex stiffeners in Schafer et al. (2006). The
equations used for the distortional buckling stress, Fd, in AS/NZS 4600 are also similar to
those in Specification Section C4.2(b), except that when the web is very slender and is
restrained by the flange, AS/NZS 4600 uses a simpler, conservative treatment. Since the
92 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
provided expressions can be complicated, solutions for the geometric properties of C- and
Z-sections based on centerline dimensions are provided in Table C-C3.1.4(b)-1.
(c) Rational Elastic Buckling Analysis
Rational elastic buckling analysis consists of any method following the principles of
mechanics to arrive at an accurate prediction of the elastic distortional buckling stress. It
is important to note that this is a rational elastic buckling analysis and not simply an
arbitrary rational method to determine strength. A variety of rational computational and
analytical methods can provide the elastic buckling moment with a high degree of
accuracy. Complete details are provided in Section 1.1.2 of the Commentary to Appendix
1 of the Specification. The safety and resistance factors of this section have been shown to
apply to a wide variety of cross-sections undergoing distortional buckling (via the
methods of Appendix 1). As long as the member falls within the geometric limits of main
Specification Section B1.1 the same safety and resistance factors have been assumed to
apply.
July 2007 93
Chapter C, Members
94 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
P P
M B M
Lb C
M A M
P P
(a) (b)
fa + fb ≤ F
or
fa fb
+ ≤ 1.0 (C-C5.2.1-2)
F F
As specified in Sections C3.1, D6.1 and C4 of the Specification, the safety factor Ωc for
the design of compression members is different from the safety factor Ωb for beam design.
Therefore Equation C-C5.2.1-2 may be modified as follows:
fa fb
+ ≤ 1.0 (C-C5.2.1-3)
Fa Fb
where
Fa = allowable stress for the design of compression members
Fb = allowable stress for the design of beams
If the strength ratio is used instead of the stress ratio, Equation C-C5.2.1-3 can be
rewritten as follows:
P M
+ ≤ 1.0 (C-C5.2.1-4)
Pa M a
where
P = applied axial load = Afa
Pa = allowable axial load = AFa
M = applied moment = Sfb
Ma = allowable moment = SFb
According to Equation C-A4.1.1-1,
P
Pa = n
Ωc
Mn
Ma =
Ωb
In the above equations, Pn and Ωc are given in Specification Sections C4 and D6.1, while
Mn and Ωb are specified in Specification Sections C3.1 and D6.1. Substituting the above
July 2007 95
Chapter C, Members
96 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
M1
C m = 0.6 − 0.4 (C-C5.2.1-11)
M2
where M1/M2 is the ratio of smaller to the larger end moments. For Case 3, 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, 1999, and 2005).
Figure C-C5.2-2 illustrates the interaction relation. In order to simplify the illustration,
bending about only one axis is considered in Figure C-C5.2-2 and the safety factors, Ωc and
Ωb, are taken as unity. The ordinate is the compressive axial load on the member and the
abscissa is the bending moment. When the moment is zero, the limiting axial load is Pn
determined in accordance with Specification Section C4, which is based on column buckling
and local buckling. When the axial load is zero, the limiting moment, Mn, is determined in
accordance with Specification Sections C3 and D6.1 and is the lowest of the effective yield
moment, the moment based on inelastic reserve capacity (if applicable) or the moment
based on lateral-torsional buckling. The interaction relation cannot exceed either of these
limits.
When Specification Equation C5.2.1-1 is plotted in Figure C-C5.2-2, the axial load limit is
Pn and the moment limit is Mn/Cm, which will exceed Mn when Cm < 1. Therefore,
Specification Equation C5.2.1-2 is used as a mathematical stratagem to limit the moment to
Mn and match the rigorous solution at low axial loads. The interaction limit is the lower of
the two equations as shown by hash marks. Specification Equation C5.2.1-2 is a linear
relation between the nominal axial yield strength Pno = FyAe and Mn, and does not
represent a failure state over its whole range. If Specification Equation C5.2.1-2 uses the
moment capacity based only on yield or local buckling, Mno = FySeff, it would be
represented by the dashed line, which could exceed an Mn limit based on lateral-torsional
buckling. Clearly, load combinations in the shaded region would be unconservative. If Mn
is determined by Mno, the relation in Figure C-C5.2-2 still apply. If Cm/α ≥ 1, Specification
Equation C5.2.1-1 controls.
P
Pno
0.15Pn
Specification Eq. C5.2.1-3
M
Mn Mno Mn C m
July 2007 97
Chapter C, Members
For low axial loads, Specification Equation C5.2.1-3 may be used. This is a conservative
simplification of the interaction relation defined by Specification Equations C5.2.1-1 and
C5.2.1-2.
In 2001, a requirement of each individual ratio in Specification Equations C5.2.1-1 to
C5.2.1-3 not exceeding unity was added to avoid situations of the load ΩcP exceeding the
Euler buckling load PE, which leads to amplification factor Φ (given in Equation C-C5.2.1-
8) negative.
For the design of angle sections using the ASD method, the required additional
bending moment of PL/1000 about the minor principal axis is discussed in Item E of
Section C4 of the Commentary.
98 July 2007
Commentary on the 2007 North American Cold-Formed Steel Specification
Ts
S.C.
g
m
Ts
July 2007 99
Chapter D, Structural Assemblies and Systems
g s
s
Note that in this expression, the overall slenderness ratio, (KL/r)o, is computed about the
same axis as the modified slenderness ratio, (KL/r)m. Further, the modified slenderness ratio,
(KL/r)m, replaces KL/r in the Specification Section C4 for both flexural and flexural-torsional
buckling.
This modified slenderness approach is used in other steel standards, including the AISC
(AISC, 1999 and 2005), CSA S136 (CSA S136, 1994), and CAN/CSA S16.1 (CAN/CSA S16.1-
94, 1994).
To prevent the flexural buckling of the individual shapes between intermediate
connectors, the intermediate fastener spacing, a, is limited such that a/ri does not exceed one
half the governing slenderness ratio of the built-up member (i.e. a/ri ≤ 0.5(KL/r)o). This
intermediate fastener spacing requirement is consistent with the previous edition of the AISI
Specification with the one half factor included to account for any one of the connectors
becoming loose or ineffective. Note that the previous edition of S136 (S136, 1994) had no limit
on fastener spacing.
The importance of preventing shear slip in the end connection is addressed by the
prescriptive requirements in Specification Section D1.2(2) adopted from the AISC (AISC, 1999)
and CAN/CSA S16.1 (CAN/CSA S16.1-94, 1994). These provisions were added to the North
American Specification since 2001.
The intermediate fastener(s) or weld(s) at any longitudinal member tie location is
required, as a group, to transmit a force equal to 2.5 percent of the nominal axial strength
[resistance] of the built-up member. A longitudinal member tie is defined as a location of
interconnection of the two members in contact. In the 2001 edition of the Specification, a 2.5
percent total force determined in accordance with appropriate load combinations was used
for design of the intermediate fastener(s) or weld(s). This requirement was adopted from
CSA S136-94. In 2004, the requirement has been changed to be a function of the nominal axial
strength. This change is to ensure that the nominal axial strength [resistance] of the built-up
member is valid and is not compromised by the strength [resistance] of the member
interconnections.
Note that the provision in Specification Section D1.2 has been substantially taken from
research in hot-rolled built-up members connected with bolts or welds. These hot-rolled
provisions have been extended to include other fastener types common in cold-formed steel
construction (such as screws) provided they meet the 2.5 percent requirement for shear
strength [resistance] and the conservative spacing requirement a/ri ≤ 0.5(KL/r)o.
Section D1.3(b) of the Specification ensures that the part of the flat sheet between two
adjacent connectors will not buckle as a column (see Figure C-D1.3-1) at a stress less than
1.67fc, where fc is the stress at service load in the connected compression element (Winter,
1970; Yu, 2000). The AISI requirement is based on the following Euler equation for column
buckling:
π2E
σ cr =
( KL/r ) 2
by substituting σcr = 1.67fc, K = 0.6, L = s, and r = t/ 12 . This provision is conservative
because the length is taken as the center distance instead of the clear distance between
connectors, and the coefficient K is taken as 0.6 instead of 0.5, which is theoretical value for a
column with fixed end supports.
Section D1.3(c) 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, 2000).
D2 Mixed Systems
When cold-formed steel members are used in conjunction with other construction materials,
the design requirements of the other material specifications also must be satisfied.
distance between braces shall not be greater than one-quarter of the span; it also
defined the conditions under which an additional brace should 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 [resistance] should suffice to provide
the force required to prevent the C-section from rotating. It is, therefore, necessary also
to determine the forces that will act in braces, such as those forces shown in Figure C-
D3.2.1-3. 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 center (where it causes no torque) plus two forces P = Qm/d
which, together, produce the same torque Qm. As is sketched in Figure C-D3.2.1-4, 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 Specification Section
D3.2.1 provide expressions for determining bracing forces PL1 and PL2, which the
braces are required to resist at each flange.
(b) Bracing of Z-Section Beams
Most Z-sections are anti-symmetrical about the vertical and horizontal centroidal
Q
Q Q P = Qm
d
m
Q
S.C.
dm
V
S.C.
V
P = Qm
V d
Figure C-D3.2.1-3 Lateral Forces Applied to C-Section
Q
Figure C-D3.2.1-1 Rotation of C-Section Beams
axes, i.e. they are point-symmetrical. In view of this, the centroid and the shear center
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 center (m = 0) and does not tend to
produce the kind of rotation a similar load would produce on a C-section. However, in
Z-sections the principal axes are oblique to the web (Figure C-D3.2.1-5). A load applied
in the plane of the web, resolved in the direction of the two axes, produces deflections
along 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 obtain information for developing appropriate bracing provisions, tests
have been carried out on three different Z-sections at Cornell University, unbraced as
well as with variously spaced intermediate braces. In addition, an approximate method
of analysis has been developed and checked against the test results. An account of this
was given by Zetlin and Winter (1955b). Briefly, it is shown that intermittently braced
Z-beams can be analyzed in much the same way as intermittently braced C-beams. It is
merely necessary, at the point of each actual vertical load Q, to apply a fictitious
horizontal load Q(Ixy/Ix) or Q[Ixy/(2Ix)] to each flange. One can then compute the
vertical and horizontal deflections, and the corresponding stresses, in conventional
ways by utilizing the convenient axes x and y (rather than 1 and 2, Figure C-D3.2.1-5),
except that certain modified section properties have to be used. To control the lateral
deflection, brace forces, P, must statically balance the fictitious force.
+y
2
Q
P=QIxy/(2Ix)
Fictitious
load
1
-x +x
1
Fictitious
load
P=QIxy/(2Ix)
-y
2
In this manner it has been shown that as to location of braces the same provisions that
apply to C-sections are also adequate for Z-sections. Likewise, the forces in the braces
are again obtained as the reactions of continuous beams horizontally loaded by
fictitious loads P. It should, however, be noted that the direction of the bracing forces
in Z-beams is different from the direction in C-beams. In the Z-beam, the bracing
forces are acting in the same direction, as shown in Fig. C-D3.2.1-5 in order to constrain
bending of the section about the axis x-x in Figure C-D3.2.1-5. The directions of the
bracing forces in the C-beam flanges are in the opposite direction as shown in Figure C-
D3.2.1-3 in order to resist the torsion caused by the applied load. In the previous
edition of the Specification, the magnitude of the Z-beam bracing force was shown as P
= Q(Ixy/Ix) on each flange. In 2001, this force was corrected to P = Q[Ixy/(2Ix)].
Py P
θ
ys y ys y
esx Px
PL1
m esy
Py
S.C. x, xs
C. x, x sd Mz Px
PL2
x- and y-direction, respectively; Mz = torsional moment about the shear center; and PL1
= bracing force applied to the flange located in the quadrant with both positive x and y
axes, and PL2 = bracing force applied on the other flange. Positive PL1 and PL2 indicate
that a restraint is required to prevent the movement of the corresponding flange in the
negative x-direction.
For a special case where design load, Q, is through the web, as shown in Figure C-
D3.2.1-3, Py = -Q, Px = 0; esx = m, esy = d/2, and from Equation C-D3.2.1-3, Mz = -Qm.
Therefore
PL1 = -Qm/d (C-D3.2.1-4)
PL2 = Qm/d (C-D3.2.1-5)
In which, m = distance from centerline of web to the shear center.
For the Z-section member as shown in Figure C-D3.2.1-7, bracing forces, PL1 and PL2,
are given in Equations C-D3.2.1-6 and C-D3.2.1-7.
I xy P M
PL 1 = Py ( )− x + z (C-D3.2.1-6)
2I x 2 d
I xy P M
PL 2 = Py ( )− x − z (C-D3.2.1-7)
2I x 2 d
where Ix, Ixy = unreduced moment of inertia and product of inertia; respectively. Other
variables are defined under the discussion for C-section members.
Py P
y, ys θ y, ys
esx PL1
Px
esy Py x, xs
x, x s Px
Mz
d PL2
Assuming that a gravity load, P, acts at 1/3 of the top flange width, bf, and the Z-
Section member rests on a sloped roof with an angle of θ, Px = -Psinθ; Py = -Pcosθ; esx =
bf/3; esy = d/2 and Mz = Psinθ(d/2) - Pcosθ(bf/3). Substituting the above expressions
into equations C-D3.2.1-6 and C-D3.2.1-7 results in
I xy Pb f cos θ
PL 1 = −P cos θ( ) + P sin θ −
2I x 3d
I xy Pb f cos θ
PL 2 = −P cos θ( )+
2I x 3d
In considering the distribution of loads and the braces along the member length, it is
required that the resistance at each brace location along the member length be greater
than or equal to the design load within a distance of 0.5a on each side of the brace for
distributed loads. For concentrated loads, the resistance at each brace location should
be greater than or equal to the concentrated design load within a distance 0.3a each
side of the brace, plus 1.4(1-l/a) times each design load located farther than 0.3a but not
farther than 1.0a from the brace. In the above, a is the distance between centerline of
braces along the member length and l is the distance from concentrated load to the
brace to be considered.
(d) 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-torsional buckling equations provided in Specification Section
C3.1.2.1 can be used to predict the moment capacity of the member. 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 percent when 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,
local distortional buckling may occur at loads lower than that predicted by the lateral-
torsional buckling equations of Specification Section C3.1.2.1.
Research (Ellifritt, Sputo and Haynes, 1992) has also shown that the lateral-torsional
buckling equations of Specification Section C3.1.2.1 predict loads, which are
conservative for cases where one mid-span brace is used but may be unconservative
where more than one intermediate brace is used. Based on such research findings,
Section D3.2.1 of the Specification was revised in 1996 to eliminate the requirement of
quarter-point bracing. It is suggested that, minimally, a mid-span brace be used for C-
section and Z-section beams to control lateral deflection and rotation at service loads.
The lateral-torsional buckling strength [resistance] of an open cross section member
should be determined by Specification Section C3.1.2.1 using the distance between
centerlines of braces “a” as the unbraced length of the member “L” in all design
equations. In any case, the user is permitted to perform tests, in accordance with
Specification Section F1, as an alternative, or use a rigorous analysis, which accounts for
biaxial bending and torsion.
Section D3.2.1 of the Specification provides the lateral forces for which these discrete
braces must be designed.
The Specification 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.
Frequently, this occurs in the end walls of metal buildings.
In 2007, the title of this section was changed to clarify that it is and was formerly
anticipated that the C- and Z-sections covered by these provisions would be
supporting sheathing and be loaded as a result of providing this support function. The
revised title reflects that the supported sheathing is not contributing to the strength and
stiffness of these members by virtue of the nature of its connection to the C- and Z-
sections.
Cold-formed I-, C-, Z-, 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 fiberboard, 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 usable strength [resistance] 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 behavior 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, which was considered in the 1947 study.
Simaan (1973) and Simaan and Pekoz (1976), which are summarized by Yu (2000), contain
procedures for computing the strength [resistance] of C- and Z-section wall studs that are
braced by sheathing materials. The bracing action is due to both the shear rigidity and the
rotational restraint supplied by the sheathing material. The treatment by Simaan (1973) and
Simaan and Pekoz (1976) is quite general and includes the case of studs braced on one as well
as on both flanges. However, the provisions of Section D4 of the 1980 AISI Specification dealt
only with the simplest case of identical sheathing material on both sides of the stud. For
simplicity, only the restraint due to the shear rigidity of the sheathing material was
considered.
The 1989 Addendum to the AISI 1986 Specification included the design limitations from
the Commentary and introduced stub column tests and/or rational analysis for the design of
studs with perforations (Davis and Yu, 1972; Rack Manufacturers Institute, 1990).
In 1996, the design provisions were revised to permit (a) all steel design and (b) sheathing
braced design of wall studs with either solid or perforated webs. For sheathing braced
design, in order to be effective, sheathing must retain its design strength [resistance] and
integrity for the expected service life of the wall. Of particular concern is the use of gypsum
sheathing in a moist environment.
In 2004 the sheathing braced design provisions were removed from the Specification and a
requirement added that sheathing braced design be based on appropriate theory, tests, or
rational engineering analysis that can be found in AISI (2004); Green, Winter, and
Cuykendall (1947); Simaan (1973); and Simaan and Pekoz (1976).
In 2007, in addition to the revisions of the Sepecification Section D4 as discussed in
Section D4 of this Commentary, the provisions for non-circular holes were moved from
Specification Section D4.1 to Section B2.2 on Uniformly Compressed Stiffened Elements with
Circular or Non-Circular Holes. Within the limitations stated for the size and spacing of
perforations and section depth, the provisions were deemed appropriate for members with
uniformly compressed stiffened elements, not just wall studs.
The Steel Deck Institute (1987) and the Department of Army (1992) have consistently
recommended a safety factor of two to limit “out of plane buckling” of diaphragms. Out of
plane buckling is related to panel profile, while the other diaphragm limit state is connection
related. The remainder of the Specification requires different safety and resistance factors for the
two limit states and larger safety factors for connection controlled states. The safety and
resistance factor for panel buckling were changed and the limit state being considered was
clarified relative to the previous edition. The prescribed factors for out of plane panel buckling
are constants for all loading types.
The Specification allows mechanical fasteners other than screws. The diaphragm shear value
using any fastener must not be based on a safety factor less than the individual fastener shear
strength safety factor unless: 1) sufficient data exists to establish a system effect, 2) an analytical
method is established from the tests, and 3) test limits are stated.
degrees or greater with the plane of the compression flange and braces to the compression
flange located at third points or more frequently, member capacities may be increased over
those without discrete braces.
For the LRFD method, the use of the reduced nominal flexural strength [resistance]
(Specification Equation D6.1.1-1) with a resistance factor of φb = 0.90 provides the β values
varying from 1.5 to 1.60 which are satisfactory for the target value of 1.5. This analysis was
based on the load combination of 1.17 W - 0.9D using a reduction factor of 0.9 applied to
the load factor for the nominal wind load, where W and D are nominal wind and dead
loads, respectively (Hsiao, Yu and Galambos, 1988a; AISI, 1991).
In 2007 the panel depth was reduced from 1-1/4 inch (32 mm) to 1-1/8 inch (29 mm).
This reduction in depth was justified because the behavior during full-scale tests indicated
that the panel deformation was restricted to a relatively small area around the screw
attachment of the panel to the purlin. Also, tests by LaBoube (1986) demonstrated that the
panel depth did not influence the rotational stiffness of the panel to purlin attachment.
Prior to the 2001 edition, the Specification specifically limited the applicability of these
provisions to continuous purlin systems in which any given span length did not vary from
any other span length by more than 20 percent. This limitation was included in recognition
of the fact that the research was based on systems with equal bay spacing. In 2007, the
Specification was revised to permit purlin systems with adjacent span lengths varying more
than 20 percent to use the reduction factor, R, for the simply supported condition. The
revision allows a row of continuous purlins to be treated with a continuous beam
condition R-factor in some bays and a simple span beam condition R-factor in others. The
20 percent span variation rule is a local effect and as such, only variation in adjacent spans
is relevant.
D6.1.2 Flexural Members Having One Flange Fastened to a Standing Seam Roof System
The design provision of this section is only applicable to the United States and Mexico.
The discussion for this section is provided in the Commentary on Appendix A. !A
D6.1.3 Compression Members Having One Flange Through-Fastened to Deck or Sheathing
For axially loaded C- or Z- sections having one flange attached to deck or sheathing
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.
Specification Equation D6.1.3-1 is used to calculate the weak axis capacity. This equation is
not valid for sections attached to standing seam roofs. The equation was developed by
Glaser, Kaehler and Fisher (1994) and is also based on the work contained in the reports of
Hatch, Easterling and Murray (1990) and Simaan (1973).
A limitation on the maximum yield stress of the C- or Z- section is not given in the
Specification since Specification Equation D6.1.3-1 is based on elastic buckling criteria. A
limitation on minimum length is not contained in the Specification because Equation
D6.1.3-1 is conservative for spans less than 15 feet. The gross area, A, has been used rather
than the effective area, Ae, because the ultimate axial stress is generally not large enough to
result in a significant reduction in the effective area for common cross section geometries.
As indicated in the Specification, the strong axis axial load capacity is determined
assuming that the weak axis of the strut is braced.
The controlling axial capacity (weak or strong axis) is suitable for usage in the
combined axial load and bending equations in Section C5 of the Specification (Hatch,
Easterling, and Murray, 1990).
[resistance] and 1.3 on serviceability. A buckling or crease does not have the same
consequences as a failure of a clip. In the latter case, the roof panel itself may become
detached and expose the contents of a building to the elements of the environment.
Further, Galambos (1988a) recommended a value of 2.0 for the target reliability index, βo,
when slight damage is expected and a value of 2.5 when moderate damage is expected.
The resulting ratio is 1.25.
In Specification Section D6.2.1, a target reliability index of 2.5 is used for connection
limits. It is used because the consequences of a panel fastener failure (βo = 2.5) are not
nearly so severe as the consequences of a primary frame connection failure (βo = 3.5). The
intermittent nature of wind load as compared to the relatively long duration of snow load
further justifies the use of βo = 2.5 for panel anchors. In Specification Section D6.2.1, the
coefficient of variation of the material factor, VM, is recommended to be 0.08 for failure
limited by anchor or connection failure, and 0.10 for limits caused by flexural or other
modes of failure. Specification Section D6.2.1 also eliminates the limit on coefficient of
variation of the test results, Vp, because consistent test results often lead to Vp values lower
than the 6.5 percent value set in Specification Section F1. The elimination of the limit will be
beneficial when test results are consistent.
The value for the number of tests for fasteners is set as the number of anchors tested
with the same tributary area as the anchor that failed. This is consistent with design
practice where anchors are checked using a load calculated based on tributary area. Actual
anchor loads are not calculated from a stiffness analysis of the panel in ordinary design
practice.
lateral stiffness of each of the components. The method is more computationally intensive
but allows for analysis of more complex bracing configurations such as supports plus third
points lateral anchorage and supports plus third points torsional braces.
A method to predict lateral anchorage forces using the finite element method is
presented in Seek and Murray (2004). The model uses shell finite elements to model the Z-
sections and sheathing in the roof system. The model accurately represents Z-section
behavior and is capable of handling configurations other than lateral anchorage applied at
the top flange. However, the computational complexity limits the size of the roof system
that can be modeled by this method.
Rational analysis may also be performed using the elastic stiffness model developed by
Sears and Murray (2007) upon which the provisions of Specification Section D6.3.1 are
based. The model uses frame finite elements to represent the Z-sections and a truss system
to represent the diaphragm. The model is computationally efficient allowing for analysis
of large systems.
D6.3.2 Alternate Lateral and Stability Bracing for Purlin Roof Systems
Tests (Shadravan and Ramseyer, 2007) have shown that C- and Z-sections can reach the
capacity determined by Specification Section C3.1 through the application of torsional
braces along the span of the member. Torsional braces applied between pairs of purlins
prevent twist of the section at a discrete location. The moments developed due to the
torsional brace can be resolved by forces in the plane of the web of each section and do not
require external anchorage at the location of the brace. The vertical forces should,
however, be accounted for when determining the applied load on the section.
Torsional braces should be applied at or near each flange of the Z- or C-section to
prevent deformation of the web of the section and insure the effectiveness of the brace.
When twist of the section is thus prevented, a section may deflect laterally and retain its
strength [resistance]. Second order moments can be resisted by the rotational restraints.
Therefore, a more liberal lateral deflection of L/180 between the supports is permitted for a
C- or Z- section with torsional braces. Anchorage is required at the frame line to prevent
excessive deformation at the support location that undermines the strength [resistance] of
the section. A lateral displacement limit therefore is imposed along the frame lines to
insure that adequate restraint along the frame lines is provided.
In this edition of the Specification, the ASD, LRFD and LSD design provisions for welded and
bolted connections were based on the 1996 edition of the AISI Specification with some revisions
and additions which will be discussed in subsequent sections.
E2 Welded Connections
Welds used for cold-formed steel construction may be classified as fusion welds (or arc
welds) and resistance welds. Fusion 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 groove welds, arc spot welds, arc seam
welds, fillet welds, and flare groove welds.
The design provisions contained in this Specification section for fusion welds have been
based primarily on experimental evidence obtained from an extensive test program conducted
at Cornell University. The results of this program are reported by Pekoz and McGuire (1979)
and summarized by Yu (2000). All possible failure modes are covered in the Specification since
1996, whereas the earlier Specification mainly dealt with shear failure.
For most of the connection tests reported by Pekoz and McGuire (1979), the onset of
yielding was either poorly defined or followed closely by failure. Therefore, in the provisions of
this section, rupture rather than yielding is used as a more reliable criterion of failure.
The welded connection tests, which served as the basis of the provisions given in
Specification Sections E2.1 through E2.5, were conducted on sections with single and double
sheets. See Specification Figures E2.2-1 and E2.2-2. The largest total sheet thickness of the cover
plates was approximately 0.15 inch (3.81 mm). However, within this Specification, the validity of
the equations was extended to welded connections in which the thickness of the thinnest
connected part is 0.18 inch (4.57 mm) or less. For arc spot welds, the maximum thickness of a
single sheet (Specification Figure E2.2.1.2-1) and the combined thickness of double sheets
(Specification Figure E2.2.1.2-2) are set at 0.15 inch (3.81 mm).
In 2001, the safety factors and resistance factors in this section were modified for consistency
based on the research work by Tangorra, Schuster, and LaBoube (2001).
For design tables and example problems on welded connections, see Part IV of the Design
Manual (AISI, 2008).
See Appendix A or B for additional commentary.
! A,B
E2.2.1 Shear
E2.2.1.1 Minimum Edge Distance
The edge distance requirements provided in the Specification Section E2.2.1.1 are to
ensure the connection provides the sufficient strength for preventing shear failure of
connected part in the direction of stress. Compared with previous editions of the AISI
Specification, the limiting Fu/Fsy ratio was revised to be consistent with Specification
Section A2.3.1.
The thickness limitation of 0.15 inch (3.81 mm) is due to the range of the test program
that served as the basis of these provisions. On sheets below 0.028 inch (0.711 mm) thick,
weld washers are required to avoid excessive burning of the sheets and, therefore,
inferior quality welds.
In the AISI 1996 Specification, Equation E2.2-1 was revised to be consistent with the
research report (Pekoz and McGuire, 1979).
In 2001, the equation used for determining da for multiple sheets was revised to be
(d-t).
E2.2.2 Tension
For tensile capacity of arc spot welds, the design provisions in the 1989 Addendum
were based on the tests reported by Fung (1978) and the study made by Albrecht (1988).
Those provisions were limited to sheet failure with restrictive limitations on material
properties and sheet thickness. These design criteria were revised in 1996 because the 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
[resistance] of the weld was determined to be a function of the cross-section of the fused
area and tensile strength of the weld material. Based on analysis by LaBoube (LaBoube,
2001), the nominal strength [resistance] equation was changed in 2001 to reflect the
ductility of the sheet, Fu/Fy, and the sheet thickness, the average weld diameter, and the
material tensile strength.
The multiple safety factors and resistance factors recognize the behavior of a panel
system with many connections versus the behavior of a member connection and the
potential for a catastrophic failure in each application. In Specification Section E2.2.2 a
target reliability index of 3.0 for the United States and Mexico and 3.5 for Canada is used
for the panel connection limit, whereas a target reliability index of 3.5 for the United States
and Mexico and 4 for Canada is used for the other connection limit. Precedence for the use
of a smaller target reliability index for systems was established in Section D6.2.1 of the
Specification.
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 [resistance]
given by Specification Equation E2.2.2-2 using the thickness of the thinner sheet.
The equations given in the Specification 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 percent reduction in nominal
connection strength [resistance] could occur because of the eccentric load application
(LaBoube and Yu, 1991 and 1993). Eccentric conditions may also occur at connection laps
depicted by Figure C-E2.2-2.
At a lap connection between two deck sections as shown in Figure C-E2.2-2, 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 [resistance] of the welded
connection (LaBoube and Yu, 1991). The Specification recognizes the reduced capacity of
this connection detail by imposing a 30 percent reduction on the calculated nominal
strength [resistance].
Interior Weld Exterior Weld
Lap Connection Subjected to Subjected to
Concentric Load Eccentric Load
Beam
Prequalified fillet welds are given in AWS D1.3-98 (AWS, 1998) or other equivalent weld
standards.
A-A
the surface. The vertical leg of the weld can either be greater, Figure E2.5-6, or less, Figure
E2.5-7, than the radius of outside bend surface. The definition of the horizontal leg of the
weld in each case is slightly different as indicated. No change was needed in the Specification
requirements from previous editions except in the definitions of the effective throat for use in
Specification Equation E2.5-4.
In 2001, the Specification was revised to require that weld strength be checked when the
plate thickness is greater than 0.10 in. (2.54 mm) based on the research by Zhao and Hancock
(1995).
E2.7 Rupture in Net Section of Members other than Flat Sheets (Shear Lag)
Shear lag has a debilitating effect on the nominal tensile strength [resistance] of a cross
section. The AISI Specification addresses the shear lag effect on tension members other than
flat sheets in welded connections. The AISC Specification’s design approach has been
adopted.
When computing U for combinations of longitudinal and transverse welds, L is taken as
the length of the longitudinal weld because the transverse weld does little to minimize shear
lag. For angle or channel sections, the distance, x , from shear plane to centroid of the cross
section is defined in Figure C-E2.7.
L L
x
x
E3 Bolted Connections
The structural behavior of bolted connections in 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 for the design of
bolted connections were developed on the basis of the Cornell tests (Winter, 1956a, 1956b).
These provisions were updated in 1980 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 AISC (1978). In 1986, design
provisions for maximum size of bolt holes and the allowable tension stress for bolts were added
in the AISI Specification (AISI, 1986). In the 1996 edition of the AISI Specification, minor changes
of the safety factors were made for computing the allowable and design tensile and shear
strengths [resistances] of bolts. The allowable tension stress for the bolts subject to the
combination of shear and tension was determined by the equations provided in Specification
Table E3.4-2 with the applicable safety factor.
(a) Scope
Previous studies and practical experiences have indicated that the structural behavior 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 AISI Specification criteria
are applicable only to cold-formed steel members or elements less than 3/16 inch (4.76 mm)
in thickness. For materials not less than 3/16 inch (4.76 mm), reference is made to the
specifications or standards stipulated in Section E3a of Appendix A or B.
Because of lack of appropriate test data and the use of numerous surface conditions, this
Specification 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 3/16 inch (4.76 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 sufficient safety against initial slip at
least equal to that implied by the provisions of the specifications or standards listed in
Section E3a of Appendix A or B. In addition, the safety against ultimate capacity should be
at least equal to that implied by this Specification for bearing-type connections.
The Specification 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 [resistance] than if no gap existed.
Structural performance of connections containing unavoidable gaps between plies would
require tests in accordance with Specification Section F1.
(b) Materials
This section lists five different types of fasteners which are normally used for cold-
formed steel construction. In view of the fact that A325 and A490 bolts are available only for
diameters of 1/2 inch (12.7 mm) and larger, A449 and A354 Grade BD bolts should be used
as an equivalent of A325 and A490 bolts, respectively, whenever smaller bolts (less than 1/2
inch (12.7 mm) in diameter) are required.
During recent years, other types of fasteners, with or without special washers, have been
widely used in steel structures using cold-formed steel members. The design of these
fasteners should be determined by tests in accordance with Chapter F of this Specification.
(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
because the ultimate strength of a bolted connection is independent of the level of bolt
preload. Installation must ensure that the bolted assembly will not come apart during
service. Experience has shown that bolts installed to a snug tight condition do not loosen or
“back-off” under normal building conditions and are not subject to vibration or fatigue.
Bolts in slip-critical connections, however, must be tightened in a manner which assures
the development of the fastener tension forces required by the Research Council on
Structural Connections (1985 and 2000) for the particular size and type of bolts. Turn-of-nut
rotations specified by the Research Council on Structural Connections may not be applicable
because such rotations are based on larger grip lengths than are encountered in usual cold-
formed construction. Reduced turn-of-the-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. Direct tension indicators (ASTM F959),
whose published clamping forces are independent of grip, can be used for tightening slip-
critical connections.
(d) Hole Sizes
Design information for oversized and slotted holes is included in the Appendices
because such holes are often used in practice to meet dimensional tolerances during
erection. A,B
!
E3.1 Shear, Spacing and Edge Distance
The design provisions of this section are given in Section E3.1 of Appendix A. The
discussion for this section is provided in the Commentary on the corresponding Appendix.
!A
E3.2 Rupture in Net Section (Shear Lag)
The design provisions of this section are given in Section E3.2 of Appendix A. The
discussion for this section is provided in the Commentary on the corresponding Appendix.
!A
E3.3 Bearing
Previous bolted connection tests have shown that the bearing strength [resistance] of
bolted connections depends on (1) the tensile strength Fu of the connected parts, (2) the
thickness of connected parts, (3) the diameter of bolt, (4) joints with single shear and double
shear conditions, (5) the Fu/Fy ratio, and (6) the use of washers (Winter, 1956a and 1956b;
Chong and Matlock, 1974; Yu, 1982 and 2000). These design parameters were used in the
1996 and earlier editions of the AISI Specification for determining the bearing strength
[resistance] between bolt and connected parts (AISI, 1996).
In the Canadian Standard (CSA, 1994), the d/t ratio was also used in the design equation
for determining the bearing strength [resistance] of bolted connections.
In this edition of the Specification, the design format and tables for determining the
bearing strength [resistance] without consideration of bolt hole deformation were revised in
2001 on the basis of the research work conducted at the University of Sydney (Rogers and
Hancock, 1998) and at the University of Waterloo (Wallace, Schuster, and LaBoube, 2001a and
2001b).
E4 Screw Connections
Results of over 3500 tests worldwide were analyzed to formulate screw connection
provisions (Pekoz, 1990). European Recommendations (1987) and British Standards (1992) were
considered and modified as appropriate. Since the provisions apply to many different screw
connections and fastener details, a greater degree of conservatism is implied than is otherwise
typical within this Specification. These provisions are intended for use when a sufficient number
of test results is not available for the particular application. A higher degree of accuracy can be
obtained by testing any particular connection geometry (AISI, 1992).
Over 450 elemental connection tests and eight diaphragm tests were conducted in which
compressible fiberglass insulation, typical of that used in metal building roof systems (MBMA,
2002), was placed between the two pieces of steel (between steel sheet samples in the elemental
connection tests and between the deck and purlin in the diaphragm tests) (Lease and Easterling,
2006a, 2006b). The results indicate that the equations in Section E4 of the Specification are valid
for applications that incorporate 6-3/8 in. (162 mm) or less of compressible fiberglass insulation.
Screw connection tests used to formulate the provisions included single fastener specimens
as well as multiple fastener specimens. However, it is recommended that at least two screws
should be used to connect individual elements. This provides redundancy against under-
torquing, over-torquing, etc., and limits lap shear connection distortion of flat unformed
members such as straps.
Proper installation of screws is important to achieve satisfactory performance. Power tools
with adjustable torque controls and driving depth limitations are usually used.
For the convenience of designers, Table C-E4-1 gives the correlation between the common
number designation and the nominal diameter for screws. See Figure C-E4-1 for the
0 0.060 1.52
1 0.073 1.85
2 0.086 2.18
3 0.099 2.51
4 0.112 2.84
5 0.125 3.18
6 0.138 3.51
7 0.151 3.84
8 0.164 4.17
10 0.190 4.83
12 0.216 5.49
1/4 0.250 6.35
E4.3 Shear
E4.3.1 Connection Shear Limited by Tilting and Bearing
Screw connections loaded in shear can fail in one mode or in 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.
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 (Figure C-
E4.3-1).
These provisions are focused on the tilting and bearing failure modes. Two cases are
given depending on the ratio of thicknesses of the connected members. Normally, the head
Tilting g
arin
Be
Pns
Spec. Eq. E4.3.1-3
t2
t1 tilting N/A
bearing Pns = 2.7 t1dFu1 or
t2 bearing Pns = 2.7 t2dFu2
of the screw will be in contact with the thinner material as shown in Figure C-E4.3-2.
However, when both members are the same thickness, or when the thicker member is in
contact with the screw head, tilting must also be considered as shown in Figure C-E4.3-3.
It is necessary to determine the lower bearing capacity of the two members based on
the product of their respective thicknesses and tensile strengths.
E4.4 Tension
Screw connections loaded in tension can fail either by pulling out of the screw from the
plate (pull-out) or pulling of material over the screw head and the washer, if a washer is
present, (pull-over) or by tensile fracture of the screw. The serviceability concerns of gross
distortion are not covered by the equations given in Specification Section E4.4.
Diameter and rigidity of the fastener head assembly as well as sheet thickness and tensile
strength have a significant effect on the pull-over failure load of a connection.
There are a variety of washers and head styles in use. Washers must be at least 0.050 inch
(1.27 mm) thick to withstand bending forces with little or no deformation.
E4.4.1 Pull-Out
For the limit state of pull-out, Specification Equation E4.4.1-1 was derived on the basis of
the modified European Recommendations and the results of a large number of tests. The
statistic data on pull-out design considerations were presented by Pekoz (1990).
E4.4.2 Pull-Over
For the limit state of pull-over, Specification Equation E4.4.2-1 was derived on the basis
of the modified British Standard and the results of a series of tests as reported by Pekoz
(1990). In 2007, a rational allowance was included to cover the contribution of steel washers
beneath screw heads. For the special case of screws with domed washers, that is washers
that are not solid or do not seat flatly against the sheet metal in contact with the washer,
the calculated nominal pull-over strength [resistance] should not exceed 1.5t1d'wFu1 with
d'w = 5/8 in. (16 mm). The 5/8 in. (16 mm) limit does not apply to solid steel washers in
full contact with the sheet metal. In accordance with Specification Section E4, testing is
allowed as an alternative method to determine fastener capacity. To use test data in design,
the tested material should be consistent with the design. When a polygon shaped washer is
used and capacity is determined using Specification Equation E4.4.2-1, the washer should
have rounded corners to prevent premature tearing.
attached panels, the 30 percent reduction associated with welds at sidelaps need not be
applied when evaluating the strength of sidelap screw connections at supports or sheet to
sheet. The reduction is due to transverse prying or peeling. It is acceptable to apply the 50
percent reduction at panel ends due to longitudinal prying.
E5 Rupture
The design provisions of this section are given in Section E5 of the Appendices. The
discussion of this section is provided in the Commentary on the corresponding Appendix.
!A,B
E6 Connections to other Materials
E6.1 Bearing
The design provisions for the nominal bearing strength [resistance] on the other materials
should be derived from appropriate material specifications.
E6.2 Tension
This Section is included in the Specification to raise the awareness of the design engineer
regarding tension on fasteners and the connected parts.
E6.3 Shear
This Section is included in the Specification to raise the awareness of the design engineer
regarding the transfer of shear forces from steel components to adjacent components of other
materials.
F1.1 Load and Resistance Factor Design and Limit States Design
The determination of load-carrying capacity of the tested elements, assemblies,
connections, or members is based on the same procedures used to calibrate the LRFD design
criteria, for which the φ factor can be computed from Equation C-A5.1.1-15. The correction
factor CP is used in Specification Equation F1.1-2 for determining the φ factor to account for
the influence due to a small number of tests (Pekoz and Hall, 1988b and Tsai, 1992). It should
be noted that when the number of tests is large enough, the effect of the correction factor is
negligible. In the 1996 edition of the AISI Specification, Equation F1.1-3 was revised because
the old formula for CP could be unconservative for combinations of a high VP and a small
sample size (Tsai, 1992). This revision enables the reduction of the minimum number of tests
from four to three identical specimens. Consequently, the ± 10 percent deviation limit was
relaxed to ± 15 percent. The use of CP with a minimum VP reduces the need for this
restriction. In Specification Equation F1.1-3, a numerical value of CP = 5.7 was found for n = 3
by comparison with a two-parameter method developed by Tsai (1992). It is based on the
given value of VQ and other statistics listed in Specification Table F1, assuming that VP will be
no larger than about 0.20. The requirements of Specification Section F1.1(a) for n = 3 help to
ensure this.
The 6.5 percent minimum value of VP, when used in Specification Equation F1.1-2 for the
case of three tests, produces safety factors similar to those of the 1986 edition of the AISI ASD
Specification, i.e. approximately 2.0 for members and 2.5 for connections. The LRFD
calibration reported by Hsiao, Yu and Galambos (1988a) indicates that VP is almost always
greater than 0.065 for common cold-formed steel components, and can sometimes reach
values of 0.20 or more. The minimum value for VP helps to prevent potential unconservatism
compared to values of VP implied in LRFD design criteria.
In evaluating the coefficient of variation VP from test data, care must be taken to use the
coefficient of variation for a sample. This can be calculated as follows:
s2
VP =
Rm
where
s2 = sample variance of all test results
1 n
= ∑ (R i − R m )2
n − 1 i =1
Rm = mean of all test results
Ri = test result i of n total results
Alternatively, VP can be calculated as the sample standard deviation of n ratios Ri/Rm.
For beams having tension flange through-fastened to deck or sheathing and with
compression flange laterally unbraced (subject to wind uplift), the calibration is based on a
load combination of 1.17W-0.9D with D/W = 0.1 (see Section D6.1.1 of this Commentary for
detailed discussion).
The statistical data needed for the determination of the resistance factor are listed in
Specification Table F1. The data listed for screw connections were added in 1996 on the basis
of the study of bolted connections reported by Rang, Galambos, and Yu (1979b). The same
statistical data of Mm, VM, Fm, and VF have been used by Pekoz in the development of the
design criteria for screw connections (Pekoz, 1990).
In 1999, two entries were added to Table F1, one for "Structural Members Not Listed
Above" and the other for "Connections Not Listed Above". It was considered necessary to
include these values for members and connections not covered by one of the existing
classifications. The statistical values were taken as the most conservative values in the
existing table.
In 2004, the statistic data VM for screw bearing strength was revised from 0.10 to 0.08.
This revision is based on the tensile strength statistic data provided in the UMR research
report (Rang, Galambos, and Yu, 1979b). In addition, Vf was revised from 0.10 to 0.05 to
reflect the tolerance of the cross-sectional area of the screw.
In 2007, additional entries were made to Table F1 to provide statistical data for all limit
states included within the Specification for the standard connection types. The entry
"Connections Not Listed Above" is intended to provide statistical data for connections other
than welded, bolted, or screwed.
Also in 2007 the specification more clearly defined the appropriate material properties
that are to be used when evaluating test results by specifying that supplier provided
properties are not to be used.
Since design is in accordance with the Specification, all that is needed is that the tested
specimen or assembly demonstrates a strength [resistance] not less than the applicable nominal
resistance, Rn.
L1
L2
x
b/2
(a)
x x
x
y
(b)
Research by Barsom et al. (1980) and Klippstein (1988, 1985, 1981, 1980) developed fatigue
information on the behavior of sheet and plate steel weldments and mechanical connections.
Although research indicates that the values of Fy and Fu do not influence fatigue behavior, the
Specification provisions are based on tests using ASTM A715 (Grade 80), ASTM A607 Grade 60,
and SAE 1008 (Fy = 30 ksi). Using regression analysis, mean fatigue life curves (S-N curves)
with the corresponding standard deviation were developed. The fatigue resistance S-N curve
has been expressed as an exponential relationship between stress range and life cycle (Fisher et
al, 1970). The general relationship is often plotted as a linear log-log function, Eq. C-G1.
REFERENCES
Acharya, V.V. and R.M. Schuster (1998), “Bending Tests of Hat Section with Multiple
Longitudinal Stiffeners,” Proceedings of the Fourteenth International Specialty Conference on
Cold-Formed Steel Structures, University of Missouri-Rolla, Rolla, MO, October, 1998.
Albrecht, R. E. (1988), “Developments and Future Needs in Welding Cold-Formed
Steel,” Proceedings of the Ninth International Specialty Conference on Cold-Formed Steel
Structures, University of Missouri-Rolla, Rolla, MO, November 1988.
Allen, D. E. and T. M. Murray (1993), “Designing Criterion for Vibrations Due to
Walking,” Engineering Journal, AISC, Fourth Quarter, 1993.
American Institute of Steel Construction (1978), Specification for the Design, Fabrication and
Erection of Structural Steel for Buildings, Chicago, IL, November 1978.
American Institute of Steel Construction (1986), Load and Resistance Factor Design
Specification For Structural Steel Buildings, Chicago, IL, 1986.
American Institute of Steel Construction (1989), Specification for Structural Steel Buildings -
Allowable Stress Design and Plastic Design, Chicago, IL, 1989.
American Institute of Steel Construction (1993), Load and Resistance Factor Design
Specification for Structural Steel Buildings, Chicago, IL, December 1993.
American Institute of Steel Construction (1997a), Steel Design Guide Series 9: Torsional
Analysis of Structural Steel Members, Chicago, IL, 1997.
American Institute of Steel Construction (1997b), AISC/CISC Steel Design Guide Series 11:
Floor Vibration Due to Human Activity, Chicago, IL, 1997.
American Institute of Steel Construction (1999), Load and Resistance Factor Design
Specification for Structural Steel Buildings, Chicago, IL, 1999.
American Institute of Steel Construction (2005), Specification for Structural Steel Buildings,
Chicago, IL, 2005.
American Iron and Steel Institute (1946), Specification for the Design of Light Gage Steel
Structural Members, New York, NY, 1946.
American Iron and Steel Institute (1949), Light Gage Steel Design Manual, New York, NY,
1949.
American Iron and Steel Institute (1956), Light Gage Cold-Formed Steel Design Manual,
(Part I - Specification, Part II - Supplementary Information, Part III - Illustrative
Examples, Part IV - Charts and Tables of Structural Properties, and Appendix), New
York, NY, 1956.
American Iron and Steel Institute (1960), Specification for the Design of Light Gage Cold-
Formed Steel Structural Members, New York, NY, 1960.
American Iron and Steel Institute (1961), Light Gage Cold-Formed Steel Design Manual,
(Part I - Specification, Part II - Supplementary Information, Part III - Illustrative
Examples, Part IV - Charts and Tables of Structural Properties, and Appendix), New
York, NY, 1961.
American Iron and Steel Institute (1962), Light Gage Cold-Formed Steel Design Manual,
(Part I - Specification, Part II - Supplementary Information, Part III - Illustrative
Examples, Part IV - Charts and Tables of Structural Properties, Appendix, and
Commentary on the 1962 Edition of the Specification by George Winter), New York, NY,
1962.
American Iron and Steel Institute (1967), Design of Light Gage Steel Diaphragms, First
Edition, New York, NY, 1967.
American Iron and Steel Institute (1968), Specification for the Design of Cold-Formed Steel
Structural Members, New York, NY, 1968.
American Iron and Steel Institute (1977), Cold-Formed Steel Design Manual, (Part I -
Specification, 1968 Edition; Part II - Commentary by George Winter, 1970 Edition; Part
IV -Illustrative Examples, 1972 Edition, March 1977; and Part V - Charts and Tables, 1977
Edition), Washington, D.C, 1977.
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.
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.
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, DC, 1991.
American Iron and Steel Institute (1992), “Test Methods for Mechanically Fastened Cold-
Formed Steel Connections,” Research Report CF92-2, Washington, DC, 1992.
American Iron and Steel Institute (1995), “Design Guide for Cold-Formed Steel Trusses,”
Publication RG-95-18, Washington, DC, 1995.
American Iron and Steel Institute (1996), Cold-Formed Steel Design Manual, Washington,
D. C., 1996.
American Iron and Steel Institute (1999), Specification for the Design of Cold-Formed Steel
Structural Members with Commentary, 1996 Edition, Supplement No.1, Washington, DC,
1999.
American Iron and Steel Institute (2001), North American Specification for the Design of
Cold-Formed Steel Structural Members with Commentary, Washington, DC, 2001.
American Iron and Steel Institute (2002), Cold-Formed Steel Design Manual, Washington,
DC, 2002.
American Iron and Steel Institute (2004a), Standard for Cold-Formed Steel Framing – Wall
Stud Design, Washington, DC, 2004.
American Iron and Steel Institute (2004b), Supplement 2004 to the North American
Specification for the Design of Cold-Formed Steel Structural Members, 2001 Edition,
Washington, DC, 2004.
American Iron and Steel Institute (2005), “Test Procedure for Determining a Strength
Value for a Roof Panel-to-Purlin-to-Anchorage Device Connection”, S912-05.
American Iron and Steel Institute (2006), Direct Strength Method (DSM) Design Guide,
Design Guide 06-1, Washington, DC, 2006.
American Iron and Steel Institute (2007a), North American Specification for the Design of
Cold-Formed Steel Structural Members, Washington, DC, 2007.
American Iron and Steel Institute (2007b), Commentary on North American Specification for
the Design of Cold-Formed Steel Structural Members, Washington, DC, 2007.
American Iron and Steel Institute (2008), Cold-Formed Steel Design Manual, Washington,
DC, 2008.
American Society for Testing and Materials (1997), “Standard Methods and Definitions
for Mechanical Testing of Steel Products,” ASTM 370, 1997.
American Society for Testing and Materials (1995), “Standard Test Method for Structural
Performance of Sheet Metal Roof and Siding Systems by Uniform Static Air Pressure
Difference,” E 1592-95, 1995.
American Society of Civil Engineers (1991), Specification for the Design and Construction of
Composite Slabs and Commentary on Specifications for the Design and Construction of
Composite Slabs, ANSI/ASCE 3-91, 1991.
American Society of Civil Engineers (1998), Minimum Design Loads for Buildings and Other
Structures, ASCE Standard 7-98, 1998.
American Society of Civil Engineers (2005), Minimum Design Loads for Buildings and Other
Structures, ASCE/SEI 7-05, Reston, VA, 2005.
American Welding Society (1966), Recommended Practice for Resistance Welding, AWS
C1.1-66, Miami, FL, 1966.
American Welding Society (1970), Recommended Practice for Resistance Welding Coated Low
Carbon Steels, AWS C1.3-70, (Reaffirmed 1987), Miami, FL, 1970.
American Welding Society (1996), Structural Welding Code - Steel, ANSI/AWS D1.1-96,
Miami, FL, 1996.
American Welding Society (1998), Structural Welding Code - Sheet Steel, ANSI/AWS D1.3-
98, Miami, FL, 1998.
American Welding Society (2000), Recommended Practices for Resistance Welding,
ANSI/AWS C1.1/C1.1M-2000, Miami, FL, 2000.
Applied Technology Council (1999), ATC Design Guide 1: Minimizing Floor Vibration,
Redwood City, California, 1999.
AS/NZS (1996). AS/NZS 4600: 1996 Cold-Formed Steel Structures. Standards Australia
and the Australian Institute of Steel Construction.
Bambach, M. R., and K. J. R. Rasmussen (2002a), "Tests on Unstiffened Elements under
Combined Bending and Compression," Research Report R818, Department of Civil
Engineering, University of Sydney, Australia, May 2002.
Bambach, M.R. and K.J.R. Rasmussen (2002b), “Elastic and Plastic Effective Width
Equations for Unstiffened Elements,” Research Report R819, Department of Civil
Engineering, University of Sydney, Australia, 2002.
Bambach, M.R. and K.J.R. Rasmussen (2002c), “Design Methods for Thin-Walled
Sections Containing Unstiffened Elements,” Research Report R820, Department of Civil
Engineering, University of Sydney, Australia, 2002.
Barsom, J. M., K. H. Klippstein, and A. K. Shoemaker (1980), “Fatigue Behavior of Sheet
Steels for Automotive Applications,” Research Report SG 80-2, American Iron and Steel
Institute, Washington, DC, 1980.
Beshara, B. (1999), “Web Crippling of Cold-Formed Steel Members,” M.A.Sc. Thesis,
University of Waterloo, Waterloo, Canada, 1999.
Beshara, B. and R.M. Schuster (2000), “Web Crippling Data and Calibrations of Cold-
Formed Steel Members,” Final Report, University of Waterloo, Waterloo, Canada, 2000.
Beshara, B. and R.M. Schuster (2000a), “Web Crippling of Cold Formed C- and Z-
Sections,” Proceedings of the Fifteenth International Specialty Conference on Cold-Formed Steel
Structures, University of Missouri-Rolla, Rolla, MO, October, 2000.
Bhakta, B.H., R.A. LaBoube, and W.W. Yu (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.
Birkemoe, P. C. and M. I. Gilmor (1978), Behavior of Bearing - Critical Double-Angle
Beam Connections, Engineering Journal, AISC, Fourth Quarter, 1978.
Bleich, F. (1952), Buckling Strength of Metal Structures, McGraw-Hill Book Co., New York,
NY, 1952.
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.
Brockenbrough, R. L. (1995), Fastening of Cold-Formed Steel Framing, American Iron and
Steel Institute, Washington, DC, September 1995.
Bryant, M.R. and Murray, T.M. (2001) “Investigation of Inflection Points as Brace Points
in Multi-Span Purlin Roof Systems” Report No. CE/VPI-ST 99/08, Virginia Polytechnic
Institute and State University, Blackburg, VA 2001.
Bulson, P. S. (1969), The Stability of Flat Plates, American Elsevier Publishing Company,
New York, NY, 1969.
Cain, D.E., R.A. LaBoube and W.W. Yu (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.
Camara Nacional de la Industria del Hierro y del Acero (1965), Manual de Diseno de
Secciones Estructurales de Acero Formadas en Frio de Calibre Ligero, Mexico, 1965.
Canadian Standards Association (1994a), Limit States Design of Steel Structures,
CAN/CSA-S16.1-94, Rexdale, Ontario, Canada, 1994.
Canadian Standards Association (1994b), Cold Formed Steel Structural Members, S136-94,
Rexdale, Ontario, Canada, 1994.
Canadian Standards Association (1995), Commentary on CSA Standard S136-94, Cold
Formed Steel Structural Members, S136.1-95, Rexdale, Ontario, Canada, 1995.
Eiler, M. R., R. A. LaBoube, and W.W. Yu (1997), “Behavior of Web Elements with
Openings Subjected to Linearly Varying Shear,” Final Report, Civil Engineering Series
97-5, Cold-Formed Steel Series, Department of Civil Engineering, University of
Missouri-Rolla, Rolla, MO, 1997.
Elhouar, S., and T.M. Murray (1985) “Adequacy of Proposed AISI Effective Width
Specification Provisions for Z- and C-Purlin Design.” Fears Structural Engineering
Laboratory, FSEL/MBMA 85-04, University of Oklahoma, Norman, Oklahoma, 1985.
Ellifritt, D. S. (1977), “The Mysterious 1/3 Stress Increase,” Engineering Journal, AISC,
Fourth Quarter, 1977.
Ellifritt, D. S., T. Sputo and J. Haynes (1992), “Flexural Capacity of Discretely Braced C’s
and Z’s,” Proceedings of the Eleventh International Specialty Conference on Cold-Formed Steel
Structures, University of Missouri-Rolla, Rolla, MO, October 1992.
Ellifritt, D. S., R. L. Glover, J. D. Hren (1998) “A Simplified Model for Distortional
Buckling of Channels and Zees in Flexure,” Proceedings of the Fourteenth International
Specialty Conference on Cold-Formed Steel Structures, University of Missouri-Rolla, Rolla,
MO, October, 1998.
Ellingwood, B., T. V. Galambos, J. G. MacGregor, and C. A. Cornell (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.
Ellingwood, B., J. G. MacGregor, T. V. Galambos, and C. A. Cornell (1982), “Probability
Based Load Criteria: Load Factors and Load Combinations,” Journal of the Structural
Division, ASCE, Vol. 108, No. ST5, May 1982.
Ellingwood, B. (1989), “Serviceability Guidelines for Steel Structures,” Engineering
Journal, AISC, First Quarter, 1989.
European Convention for Constructional Steelwork (1977), “European
Recommendations for the Stressed Skin Design of Steel Structures,” ECCS-XVII-77-1E,
CONSTRADO, London, March 1977.
European Convention for Constructional Steelwork (1987), “European
Recommendations for the Design of Light Gage Steel Members,” First Edition, Brussels,
Belgium, 1987.
Fisher, J. M. and M.A. West (1990), Serviceability Design Considerations for Low-Rise
Buildings, Steel Design Guide Series, AISC, 1990.
Fisher, J. M., (1996), “Uplift Capacity of Simple Span Cee and Zee Members with
Through - Fastened Roof Panels,” Final Report MBMA 95-01, Metal Building
Manufacturers Association, 1996.
Fisher, J.W., K. H. Frank, M. A. Hirt and B.M. McNamee (1970), “Effect of Weldments on
the Fatigue Strength of Steel Beams,” National Cooperative Highway Research Program
Report 102, Highway Research Board, Washington, DC, 1970.
Fisher, J. W., G. L. Kulak, and I. F.C. Smith (1998), “A Fatigue Primer for Structural
Engineers,” National Steel Bridge Alliance, 1998.
Fox, S.R. (2002), “Bearing Stiffeners in Cold Formed Steel C-Sections”, Ph.D. Thesis,
Department of Civil Engineering, University of Waterloo, Waterloo, Ontario, Canada,
2002.
Fox, S.R. and Schuster, R.M. (2002), “Bearing Stiffeners in Cold-Formed Steel C-
Sections.” Final Report, American Iron and Steel Institute, Washington, DC, 2002.
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.
Galambos, T. V. (1963), “Inelastic Buckling of Beams,”Journal of the Structural Division,
ASCE, Vol. 89, No. ST5, October 1963.
Galambos, T. V., B. Ellingwood, J. G. MacGregor, and C. A. Cornell (1982), “Probability
Based Load Criteria: Assessment of Current Design Practice,” Journal of the Structural
Division, ASCE, Vol. 108, No. ST5, May 1982.
Galambos, T. V. (Editor) (1988a), Guide to Stability Design Criteria for Metal Structures,
Fourth Edition, John Wiley and Sons, New York, NY, 1988.
Galambos, T. V. (1988b), “Reliability of Structural Steel Systems, “ Report No. 88-06,
American Iron and Steel Institute, Washington, D.C., 1988.
Galambos, T. V. (1998), Guide to Stability Design Criteria for Metal Structures, Fifth Edition,
John Wiley & Sons, Inc., 1998.
Gerges,R.R. (1997), “Web Crippling of Single Web Cold-Formed Steel Members
Subjected to End One-Flange Loading,” M.A.Sc. Thesis, University of Waterloo,
Waterloo, Canada, 1997.
Gerges,R.R. and R.M. Schuster (1998), “Web Crippling of Members Using High-Strength
Steels,” Proceedings of the Fourteenth International Specialty Conference on Cold-Formed Steel
Structures, University of Missouri-Rolla, October 1998.
Glaser, N.J., R. C. Kaehler and J. M. Fisher (1994), “Axial Load Capacity of Sheeted C
and Z Members,” Proceedings of the Twelfth International Specialty Conference on Cold-
Formed Steel Structures, University of Missouri-Rolla, October 1994.
Green, G. G., G. Winter and T. R. Cuykendall (1947), “Light Gage Steel Columns in Wall-
Braced Panels,” Bulletin, No. 35/2, Cornell University Engineering Experimental Station,
1947.
Green, P.S., T. Sputo, and V. Urala (2004). “Bracing Strength and Stiffness Requirements
for Axially Loaded Lipped Cee Studs.” Proceedings of the Seventeenth International
Specialty Conference on Cold-Formed Stee Structuresl, University of Missouri-Rolla, 2004.
Hancock, G. J., Y. B. Kwon and E. S. Bernard (1994), “Strength Design Curves for Thin-
Walled Sections Undergoing Distortional Buckling,” Journal of Constructional Steel
Research, Vol. 31, 1994.
Hancock, G. J. (1995), “Design for Distortional Buckling of Flexural Members,”
Proceedings, Third International Conference on Steel and Aluminum Structures, Istanbul,
Turkey, May 1995.
Hancock, G.J. (1997), “Design for Distortional Buckling of Flexural Members,” Thin-
Walled Structures, Vol. 27, No.1, 1997.
Hancock, G.J., C. A. Rogers, and R.M. Schuster (1996). “Comparison of the Distortional
Buckling Method for Flexural Members with Tests.” Proceedings of the Thirteenth
International Specialty Conference on Cold-Formed Steel Structures, University of Missouri-
Rolla, MO, 1996.
Hardash, S. G., and R. Bjorhovde (1985), “New Design Criteria for Gusset Plates in
Tension,” AISC Engineering Journal, Vol. 22, No. 2, 2nd Quarter, 1985.
Harper, M.M., R.A. LaBoube and W. W. Yu (1995), “Behavior of Cold-Formed Steel Roof
Trusses,” Summary Report, Civil Engineering Study 95-3, University of Missouri-Rolla,
Rolla, MO, May 1995.
Harris, P. S. and R. A. LaBoube (1985), “Understanding the Engineering Safety Factor in
Building Design,” Plant Engineering, August 1985.
Hatch, J., W. S. Easterling and T. M. Murray (1990), “Strength Evaluation of Strut-
Purlins,” Proceedings of the Tenth International Specialty Conference on Cold-Formed Steel
Structures, University of Missouri-Rolla, October 1990.
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.
Haussler, R. W. and R. F. Pahers (1973), “Connection Strength in Thin Metal Roof
Structures,” Proceedings of the Second Specialty Conference on Cold-Formed Steel Structures,
University of Missouri-Rolla, Rolla, MO, October 1973.
Haussler, R. W. (1988), “Theory of Cold-Formed Steel Purlin/Girt Flexure,” Proceedings
of the Ninth International Specialty Conference on Cold-Formed Steel Structures, University of
Missouri-Rolla, Rolla, MO, November 1988.
Hetrakul, N. and W. W. Yu (1978), “Structural Behavior of Beam Webs Subjected 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.
Hetrakul, N. and W. W. Yu (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.
Hill, H. N. (1954), “Lateral Buckling of Channels and Z-Beams,” Transactions, ASCE, Vol.
119, 1954.
Holcomb, B. D., R. A. LaBoube and W. W. Yu (1995), “Tensile and Bearing Capacities of
Bolted Connections,” Second Summary Report, Civil Engineering Study 95-1, University
of Missouri-Rolla, Rolla, Mo., May 1995.
Holesapple, M.W. and R.A. LaBoube (2002), “Overhang Effects on End-One-Flange Web
Crippling Capacity of Cold-Formed Steel Members,” Final Report, Civil Engineering
Study 02-1, Cold-Formed Steel Series, University of Missouri-Rolla, MO, 2002.
Hsiao, L. E., W. W. Yu and T. V. Galambos (1988a), “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.
Hsiao, L. E., W. W. Yu and T. V. Galambos (1988b), “Load and Resistance Factor Design
of Cold-Formed Steel: Comparative Study of Design Methods for Cold-Formed Steel,”
Eleventh Progress Report, Civil Engineering Study 88-4, University of Missouri-Rolla,
Rolla, MO, February 1988.
Hsiao, L. E. (1989), “Reliability Based Criteria for Cold-Formed Steel Members,” thesis
presented to the University of Missouri-Rolla, Rolla, Missouri, in partial fulfillment of
the requirements for the Degree of Doctor of Philosophy, 1989.
Hsiao, L. E., W. W. Yu, and T. V. Galambos (1990), “AISI LRFD Method for Cold-
Formed Steel Structural Members,” Journal of Structural Engineering, ASCE, Vol. 116, No.
2, February 1990.
Johnston, B. G. (Editor) (1976), Guide to Stability Design Criteria for Metal Structures, Third
Edition, John Wiley and Sons, New York, NY, 1976.
Joint Departments of the Army, Navy, Air Force, USA (1992), Chapter 13, Seismic Design
for Buildings, TM 5-809-10/NAVFACP-355/AFM 88-3, Washington, DC, 20 October
1992.
Kalyanaraman, V., T. Pekoz, and G. Winter (1977), “Unstiffened Compression
Elements,” Journal of the Structural Division, ASCE, Vol. 103, No. ST9, September 1977.
Kalyanaraman, V., and T. Pekoz (1978), “Analytical Study of Unstiffened Elements,”
Journal of the Structural Division, ASCE, Vol. 104, No. ST9, September 1978.
Karren, K. W. (1967), “Corner Properties of Cold-Formed Steel Shapes,” Journal of the
Structural Division, ASCE, Vol. 93, No. ST1, February, 1967.
Karren, K. W. and G. Winter (1967), “Effects of Cold-Work on Light Gage Steel
Members,” Journal of the Structural Division, ASCE, Vol. 93, No. ST1, February 1967.
Kavanagh, K. T. and D. S. Ellifritt (1993), “Bracing of Cold-Formed Channels Not
Attached to Deck or Sheeting,” Is Your Building Suitably Braced?, Structural Stability
Research Council, April 1993.
Kavanagh, K. T. and D. S. Ellifritt (1994), “Design Strength of Cold-Formed Channels in
Bending and Torsion,” Journal of Structural Engineering, ASCE, Vol. 120, No. 5, May 1994.
Kian, T. and T. B. Pekoz (1994), “Evaluation of Industry-Type Bracing Details for Wall
Stud Assemblies,” Final Report, submitted to American Iron and Steel Institute, Cornell
University, January 1994.
Kirby, P. A. and D. A. Nethercot (1979), Design for Structural Stability, John Wiley and
Sons, Inc., New York, NY, 1979.
Klippstein, K. H. (1980), “Fatigue Behavior of Sheet Steel Fabrication Details,”
Proceedings of the Fifth International Specialty Conference on Cold-Formed Steel Structures,
University of Missouri-Rolla, Rolla, MO, November, 1980.
Klippstein, K. H. (1981), “Fatigue Behavior of Steel-Sheet Fabrication Details,” SAE
Technical Paper Series 810436, International Congress and Exposition, Detroit, MI.
Klippstein, K. H. (1985), “Fatigue of Fabricated Steel-Sheet Details - Phase II,” SAE
Technical Paper Series 850366, International Congress and Exposition, Detroit, MI.
Klippstein, K. H. (1988), “Fatigue Design Curves for Structural Fabrication Details Made
of Sheet and Plate Steel,” unpublished AISI research report.
Miller, T. H. and T. Pekoz (1994), “Unstiffened Strip Approach for Perforated Wall
Studs,” Journal of the Structural Engineering, ASCE, Vol. 120, No. 2, February 1994.
Moreyra, M.E. (1993). “The Behavior of Cold-Formed Lipped Channels under Bending,”
M.S. Thesis, Cornell University, Ithaca, New York, 1993.
Mulligan, G. P. and T. B. Pekoz (1984), “Locally Buckled Thin-Walled Columns,” Journal
of the Structural Division, ASCE, Vol. 110, No. ST11, November 1984.
Murray, T. M. and S. Elhouar (1985), “Stability Requirements of Z-Purlin Supported
Conventional Metal Building Roof Systems,” Annual Technical Session Proceedings,
Structural Stability Research Council, 1985.
Murray, T. M. (1991), “Building Floor Vibrations,” Engineering Journal, AISC, Third
Quarter, 1991.
Nguyen, P. and W. W. Yu (1978a), “Structural Behavior of Transversely Reinforced
Beam Webs,” Final Report, Civil Engineering Study 78-5, University of Missouri-Rolla,
Rolla, MO, July 1978.
Nguyen, P. and W. W. Yu, (1978b), “Structural Behavior of Longitudinally Reinforced
Beam Webs,” Final Report, Civil Engineering Study 78-6, University of Missouri-Rolla,
Rolla, MO, July 1978.
Ortiz-Colberg, R. and T. B. Pekoz (1981), “Load Carrying Capacity of Perforated Cold-
Formed Steel Columns,” Research Report No. 81-12, Cornell University, Ithaca, NY, 1981.
Pan, L.C., and W. W. Yu (1988), "High Strength Steel Members with Unstiffened
Compression Elements," Proceedings of the Ninth International Specialty Conference on Cold-
Formed Steel Structures, University of Missouri-Rolla, MO, November, 1988.
Papazian, R.P., R.M. Schuster and M. Sommerstein (1994), “Multiple Stiffened Deck
Profiles,” Proceedings of the Twelfth International Specialty Conference on Cold-Formed Steel
Structures, University of Missouri-Rolla, Rolla, MO, October, 1994.
Pekoz, T. B. and G. Winter (1969a), “Torsional-Flexural Buckling of Thin-Walled Sections
Under Eccentric Load,” Journal of the Structural Division, ASCE, Vol. 95, No. ST5, May
1969.
Pekoz, T. B. and N. Celebi (1969b), “Torsional-Flexural Buckling of Thin-Walled Sections
under Eccentric Load,” Engineering Research Bulletin 69-1, Cornell University, 1969.
Pekoz, T. B. and W. McGuire (1979), “Welding of Sheet Steel,” Report SG-79-2, American
Iron and Steel Institute, January 1979.
Pekoz, T. B. and P. Soroushian (1981), “Behavior of C- and Z- Purlins Under Uplift,”
Report No. 81-2, Cornell University, Ithaca, NY, 1981.
Pekoz, T. B. and P. Soroushian (1982), “Behavior of C- and Z- Purlins Under Wind
Uplift,” Proceedings of the Sixth International Specialty Conference on Cold-Formed Steel
Structures, University of Missouri-Rolla, Rolla, MO, November 1982.
Pekoz, T. B. (1986a), “Combined Axial Load and Bending in Cold-Formed Steel
Members,” Thin-Walled Metal Structures in Buildings, IABSE Colloquium, Stockholm,
Sweden, 1986.
Pekoz, 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.
Rang, T. N., T. V. Galambos, and W. W. Yu (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.
Rasmussen, K. J. R. and G. J. Hancock (1992), “Nonlinear Analyses of Thin-Walled
Channel Section Columns,” Thin Walled Structures (J. Rhodes and K.P. Chong, Eds.), Vol.
13, Nos. 1-2, Elsevier Applied Science, Tarrytown, NY, 1992.
Rasmussen, K. J. R. (1994), “Design of Thin-Walled Columns with Unstiffened Flanges,”
Engineering Structures, (G. J. Hancock, Guest Editor), Vol. 16, No. 5, Butterworth-
Heinmann Ltd., London, July 1994.
Ravindra, M. K. and T. V. Galambos (1978), “Load and Resistance Factor Design for
Steel,” Journal of the Structural Division, ASCE, Vol. 104, No. ST9, September 1978.
Reck, H. P., T. Pekoz, and G. Winter (1975), “Inelastic Strength of Cold-Formed Steel
Beams” Journal of the Structural Division, ASCE, Vol. 101, No. ST11, November 1975.
Research Council on Structural Connections (1980), Specification for Structural Joints Using
ASTM A325 or A490 Bolts, 1980.
Research Council on Structural Connections (1985), Allowable Stress Design Specification
for Structural Joints Using ASTM A325 or A490 Bolts, 1985.
Research Council on Structural Connections (2000), Specification for Structural Joints Using
ASTM A325 or A490 Bolts, 2000.
Research Council on Structural Connections (2004), Specification for Structural Joints Using
ASTM A325 or A490 Bolts, 2004.
Rivard, P. and T.M. Murray (1986), “Anchorage Forces in Two Purlin Line Standing
Seam Z-Purlin Supported Roof Systems,” Research Report, University of Oklahoma,
Norman, OK, December 1986.
Roark, R. J. (1965), Formulas for Stress and Strain, Fourth Edition, McGraw-Hill Book
Company, New York, NY, 1965.
Rogers, C.A., and R. M. Schuster (1995) “Interaction Buckling of Flange, Edge Stiffener
and Web of C-Sections in Bending.” Research Into Cold Formed Steel, Final Report of
CSSBI/IRAP Project, Department of Civil Engineering, University of Waterloo, Waterloo,
Ontario, 1995.
Rogers, C. and R.M. Schuster (1996), “Cold-Formed Steel Flat Width Ratio Limits, d/t,
and di/w,” Proceedings of the Thirteenth International Specialty Conference on Cold-Formed
Steel Structures, University of Missouri-Rolla, Rolla, MO, October, 1996.
Rogers, C. A., and G. J. Hancock (1998), “Bolted Connection Tests of Thin G550 and
G300 Sheet Steels,” Journal of Structural Engineering, ASCE, Vol. 124, No. 7, 1998.
Salmon, C. G., and J.E. Johnson (1990), Steel Structures: Design and Behavior, Third
Edition, Harper & Row, New York, NY, 1990.
Santaputra, C. (1986), “Web Crippling of High Strength Cold-Formed Steel Beams,”
Ph.D. Thesis, University of Missouri-Rolla, Rolla, MO, 1986.
Santaputra, C., M. B. Parks, and W. W. Yu (1989), “Web Crippling Strength of Cold-
Formed Steel Beams,” Journal of Structural Engineering, ASCE, Vol. 115, No. 10, October
1989.
Schafer, B.W. and T. Pekoz (1998), “Cold-Formed Steel Members with Multiple
Longitudinal Intermediate Stiffeners in the Compression Flange,” Journal of Structural
Engineering, ASCE, Vol. 124, No.10, October, 1998.
Schafer, B.W., and T. Peköz (1999), “Laterally Braced Cold-Formed Steel Flexural
Members with Edge Stiffened Flanges.” Journal of Structural Engineering. ASCE, Vol. 125,
No. 2, February 1999.
Schafer, B.W. (2000), “Distortional Buckling of Cold-Formed Steel Columns,” Final
Report, sponsored by the American Iron and Steel Institute, Washington, D.C., 2000.
Schafer, B.W. (2002), “Local, Distortional, and Euler Buckling in Thin-Walled Columns,”
Journal of Structural Engineering, ASCE, Vol. 128, No.3, March 2002.
Schafer, B.W., Sarawit, A., Peköz, T. (2006). “Complex Edge Stiffeners for Thin-Walled
Members.” Journal of Structural Engineering, ASCE, Vol. 132, No. 2, February, 2006.
Schardt, R. W., Schrade (1982), “Kaltprofil-Pfetten.” Institut Für Statik, Technische
Hochschule Darmstadt, Bericht Nr. 1, Darmstadt, 1982.
Schuster, R.M. (1992). “Testing of Perforated C-Stud Sections in Bending.” Report for the
Canadian Sheet Steel Building Institute, University of Waterloo, Waterloo Ontario, 1992.
Schuster, R. M., C. A. Rogers, and A. Celli (1995), “Research into Cold-Formed Steel
Perforated C-Sections in Shear,” Progress Report No. 1 of Phase I of CSSBI/IRAP
Project, Department of Civil Engineering, University of Waterloo, Waterloo, Ontario,
Canada, 1995.
Sears, J. M. and T. M. Murray (2007), “Proposed Method for the Prediction of Lateral
Restraint Forces in Metal Building Roof Systems,” Annual Stability Conference Proceedings,
Structural Stability Research Council, 2007.
Seek, M. W. and T. M. Murray (2004). “Computer Modeling of Sloped Z-Purlin
Supported Roof Systems to Predict Lateral Restraint Force Requirements.” Conference
Proceedings, Seventeenth International Specialty Conference on Cold-Formed Steel Structures.
Department of Civil Engineering, University of Missouri-Rolla, Rolla, Missouri.
Seek, M. W. and T. M. Murray (2006). “Component Stiffness Method to Predict Lateral
Restraint Forces in End Restrained Single Span Z-Section Supported Roof Systems with
One Flange Attached to Sheathing.” Proceedings of the Nineteenth International Specialty
Conference on Cold-Formed Steel Structures. University of Missouri-Rolla. Rolla, MO,
2006.
Seek, M.W. and T.M. Murray (2007) “Lateral Brace Forces in Single Span Z-Section Roof
Systems with Interior Restraints Using the Component Stiffness Method.” Annual
Stability Conference Proceedings, Structural Stability Research Council, 2007.
Serrette, R. L. and T. B. Pekoz (1992), “Local and Distortional Buckling of Thin-Walled
Beams,” Proceedings of the Eleventh International Specialty Conference on Cold-Formed Steel
Structures, University of Missouri-Rolla, Rolla, MO, October 1992.
Serrette, R. L. and T. B. Pekoz (1994), “Flexural Capacity of Continuous Span Standing
Seam Panels: Gravity Load,” Proceedings of the Twelfth International Specialty Conference on
Cold-Formed Steel Structures, University of Missouri-Rolla, Rolla, MO, October 1994.
Serrette, R. L. and T. B. Pekoz (1995), “Behavior of Standing Seam Panels,” Proceedings of
the Third International Conference on Steel and Aluminum Structures, Bogazici University,
Istanbul, Turkey, May 1995.
Shadravan, S. and C. Ramseyer (2007), “Bending Capacity of Steel Purlins with Torsional
Bracing Using the Base Test, “ Annual Stability Conference Proceedings, Structural Stability
Research Council, 2007.
Shan, M. Y., R. A. LaBoube, and W. W. Yu (1994), "Behavior of Web Elements with
Openings Subjected to Bending, Shear and the Combination of Bending and Shear,"
Final Report, Civil Engineering Series 94-2, Cold-Formed Steel Series, Department of
Civil Engineering, University of Missouri-Rolla
Sherman, D. R. (1976), “Tentative Criteria for Structural Applications of Steel Tubing
and Pipe,” American Iron and Steel Institute, Washington, D. C., 1976.
Sherman, D. R. (1985), “Bending Equations for Circular Tubes,” Annual Technical Session
Proceedings, Structural Stability Research Council, 1985.
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.
Simaan, A. and T. Pekoz (1976), “Diaphragm-Braced Members and Design of Wall
Studs,” Journal of the Structural Division, ASCE, Vol. 102, ST1, January 1976.
Sputo, T., and K. Beery (2006). “Accumulation of Bracing Strength and Stiffness
Demand in Cold-Formed Steel Stud Walls.” Proceedings of the Eighteenth International
Specialty Conference on Cold-Formed Steel Structures, University of Missouri-Rolla, 2006.
Steel Deck Institute, Inc. (1981), Steel Deck Institute Diaphragm Design Manual, First
Edition, Canton, OH, 1981.
Steel Deck Institute, Inc. (1987), Steel Deck Institute Diaphragm Design Manual, Canton,
OH, 1987.
Steel Deck Institute, Inc. (2004), Steel Deck Institute Diaphragm Design Manual, Third
Edition, Fox River Grove, IL, 2004.
Steel Deck Institute, Inc. (2006), Design Manual for Composite Decks, Form Decks, Roof
Decks, Cellular Deck Floor Systems with Electrical Distribution, SDI Publication No. 30, 2006.
Stolarczyk, J. A., J. M. Fisher, A. Ghorbanpoor (2002), “Axial Strength of Purlins
Attached to Standing Seam Roof Panels,” Proceedings of the Sixteenth International
Specialty Conference on Cold-Formed Steel Structures, University of Missouri-Rolla, MO,
October 2002.
Structural Stability Resarch Council (1993), Is Your Structure Suitably Braced?, Lehigh
University, Bethlehem, PA, April 1993.
Supornsilaphachai, B., T. V. Galambos, and W. W. Yu (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.
Supornsilaphachai, B. (1980), “Load and Resistance Factor Design of Cold-Formed Steel
Structural Members,” thesis presented to the University of Missouri-Rolla, Missouri, in
partial fulfillment of the requirements for the Degree of Doctor of Philosophy, 1980.
Surry, D., R. R. Sinno, B. Nail, T.C.E. Ho, S. Farquhar, and G. A. Kopp (2007),
“Structurally-Effective Static Wind Loads for Roof Panels,” Journal of the Structural
Engineering, ASCE, Vol. 133, No. 6, June 2007.
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.
Yu, W. W. (1982), “AISI Design Criteria for Bolted Connections,” Proceedings of the Sixth
International Specialty Conference on Cold-Formed Steel Structures, University at Missouri-
Rolla, Rolla, MO, November 1982.
Yu, W. W. (1985), Cold-Formed Steel Design, Wiley-Interscience, New York, NY, 1985.
Yu, W.W.(1996), Commentary on the 1996 Edition of the Specification for the Design of Cold-
Formed Steel Structural Members, American Iron and Steel Institute, Washington, D.C.,
1996.
Yu, W. W. (2000), Cold-Formed Steel Design, Third Edition, John Wiley & Sons, New York,
NY, 2000.
Yura, J.A. (1993), “Fundamentals of Beam Bracing,” Is Your Structure Suitably Braced?
Structural Stability Research Council, April 1993.
Zetlin, L. (1955a), “Elastic Instability of Flat Plates Subjected to Partial Edge Loads,”
Journal of the Structural Division, ASCE, Vol. 81, September 1955.
Zetlin, L. and G. Winter (1955b), “Unsymmetrical Bending of Beams with and without
Lateral Bracing,” Journal of the Structural Division, ASCE, Vol. 81, 1955.
Zhao, X.L. and G.J. Hancock (1995), “Butt Welds and Transverse Fillet Welds in Thin
Cold-Formed RHS Members,” Journal of Structural Engineering, ASCE, Vol. 121, No.11,
November, 1995.
Zwick, K. and R. A. LaBoube (2002), “Self-Drilling Screw Connections Subject to
Combined Shear and Tension”, Center for Cold-Formed Steel Structures, University of
Missouri-Rolla, 2002.
Commentary on Appendix 1
Structural Members
2007 EDITION
Commentary on the 2007 North American Cold-Formed Steel Specification
Note:
¿
The North American Specification for the Design of Cold-Formed Steel Structural Members, Chapters A
through G and Appendices A and B and Appendix 2, is herein referred to as the main Specification.
original Rack Upright category were relaxed to match those found for C-section beams with
complex stiffeners (Schafer, et al., 2006).
It is intended that as more cross-sections are verified for use in the Direct Strength
Method, these tables and sections will be augmented. Companies with proprietary sections
may wish to perform their own testing and follow Chapter F of the main Specification to
justify the use of lower Ω and higher φ factors for a particular cross-section. Alternatively,
member geometries that are not pre-qualified may still use the method of Appendix 1, but
with the increased Ω and reduced φ factors consistent with any rational analysis method as
prescribed in A1.2 of the main Specification.
1.5
AISI (2002) Ex. I−8
My=126.55kip−in.
1
Mcr / My
Lateral−torsional
0
0 1 2 3
10 10 10 10
half−wavelength (in.)
0.4
0.35
Py=48.42kips
0.3
Flexural
0.25
Pcr / Py
0.2
0.1
0.05
0
0 1 2 3
10 10 10 10
half−wavelength (in.)
(a) 9CS2.5x059 of AISI 2002 Cold-Formed Steel Design Manual Example I-8
1.5
My=107.53kip−in.
1
Mcr / My
Lateral−torsional
0
0 1 2 3
10 10 10 10
half−wavelength (in)
0.5
0.4
Py=45.23kips
0.35
0.3
Flexural
Pcr / Py
0.25
Distortional P cr/Py=0.29
0.2
Local Pcr/Py=0.16
0.15
0.1
0.05
0
0 1 2 3
10 10 10 10
half−wavelength (in.)
(b) 8ZS2.25x059 of AISI 2002 Cold-Formed Steel Design Manual Example I-10
1.5
My=2.12kip−in.
1
Mcr / My
Distortional M cr/My=1.03
Lateral−torsional
0.5
0
0 1 2 3
10 10 10 10
half−wavelength (in.)
3.5
My=2.12kip−in.
2.5
2
Mcr / My
Distortional M cr/My=2.18
1.5
Lateral−torsional
0.5
0
0 1 2 3
10 10 10 10
half−wavelength (in.)
(c) 2LU2x060 of AISI 2002 Cold-Formed Steel Design Manual Example I-12
5
Mcr / My
Lateral−torsional
3
Local Mcr/My=3.46
0
0 1 2 3
10 10 10 10
half−wavelength (in.)
3.5
3
Pcr / Py
2.5
Flexural−torsional
2 Local Pcr/Py=2.66
1.5
0.5
0
0 1 2 3
10 10 10 10
half−wavelength (in.)
(d) 3HU4.5x135 of AISI 2002 Cold-Formed Steel Design Manual Example I-13
In the finite strip method, members are loaded with a reference stress distribution: pure
compression for finding Pcr, and pure bending for finding Mcr (see Figure C-1.1.2-1).
instead.
Half-wavelength
Distortional buckling occurs at a half-wavelength intermediate to local and global
buckling modes, as shown in the figures given in C-1.1.2-1. The half-wavelength is
typically several times larger than the largest characteristic dimension of the member. The
half-wavelength is highly dependent on both the loading and the geometry.
Mode shape
Distortional buckling involves both translation and rotation at the fold line of a
member. Distortional buckling involves distortion of one portion of the cross-section and
predominately rigid response of a second portion. For instance, the edge stiffened flanges
of the lipped cee and zee are primarily responding as one rigid piece while the web is
distorting.
Discussion
Distortional buckling may be indistinct (without a minimum) even when local buckling
and long half-wavelength (global) buckling are clear. The lipped cee and zee in bending
show this basic behavior. For some members distortional buckling may not occur.
Bracing can be effective in retarding distortional buckling and boosting the strength
[resistance] of a member. Continuous bracing may be modeled by adding a continuous
spring in a finite strip model. For discrete bracing of distortional buckling, when the
unbraced length is less than the critical distortional half-wavelength, best current practice
is to use the buckling load (or moment) at the unbraced length. The key consideration for
distortional bracing is limiting the rotation at the compression flange/web juncture.
Global bucking modes for columns include: flexural, torsional and flexural-torsional
buckling. For beams bent about their strong-axis, lateral-torsional buckling is the global
buckling mode of interest.
Half-wavelength
Global (or “Euler”) buckling modes: flexural, torsional, or flexural-torsional for
columns, lateral-torsional for beams, occur as the minimum mode at long half-
wavelengths.
Mode Shape
Global buckling modes involve translation (flexure) and/or rotation (torsion) of the
entire cross-section. No distortion exists in any of the elements in the long half-wavelength
buckling modes.
Discussion
Flexural and distortional buckling may interact at relatively long half-wavelengths
making it difficult to determine long column modes at certain intermediate to long lengths.
When long column end conditions are not simply supported, or when they are dissimilar
for flexure and torsion, higher modes are needed for determining the appropriate buckling
load. By examining higher modes in a finite strip analysis, distinct flexural and flexural-
torsional modes may be identified. Based on the boundary conditions, the effective length,
KL, for a given mode can be determined. With KL known, then Pcre (or Mcre) for that mode
may be read directly from the finite strip at a half-wavelength of KL by using the curve
corresponding to the appropriate mode. For beams, Cb of the main Specification may be
employed to account for the moment gradient. Mixed flexural and torsional boundary
conditions may not be directly treated. Alternatively, traditional manual solutions may be
used for global buckling modes with different bracing conditions.
Distortional Buckling
Distortional buckling of members with edge stiffened flanges may also be predicted by
manual solutions. Unfortunately, the complicated interaction that occurs between the edge
stiffened flange and the web leads to cumbersome and lengthy formulas.
For columns,
Pcrd = Agfcrd (C-1.1.2-3)
Ag = gross area of the member
fcrd = distortional buckling stress (see below)
For beams,
Mcrd = Sffcrd (C-1.1.2-4)
Sf = gross section modulus to the extreme compression fiber
fcrd = distortional buckling stress at the extreme compression fiber. Solutions and
design aids for fcrd are available for beams (Hancock et al., 1996; Hancock, 1997;
Schafer and Peköz, 1999) and for columns (Lau and Hancock, 1987; Schafer 2002).
Design aids for flanges with unusual edge stiffeners (e.g., Bambach et al., 1998)
or flexural members with a longitudinal stiffener in the web (Schafer, 1997) are
also available. See the Commentary on the Main Specification Sections C3.1.4 and
C4.2 for additional information.
Global Buckling
Global buckling of members is calculated in the main Specification. Therefore, for both
beams and columns, extensive closed-form expressions are already available and may be
used for manual calculation. See the Commentary to main Specification Sections C4 and C3 for
additional details.
For columns,
Pcre = Agfcre (C-1.1.2-5)
Ag = gross area of the member
fcre = minimum of the elastic critical flexural, torsional, or flexural-torsional buckling
stress. fcre is equal to Fe of Section C4 of the main Specification. The hand methods
presented in Specification Sections C4.1.1 through C4.1.4 provide all necessary
formula. Note, Section C4.1.4 specifically addresses the long-standing practice
that Fe (or fcre) may be calculated by rational analysis. Rational analysis hand
solutions to long column buckling are available - see the Commentary for main
Specification Section C4.1.4 as well as Yu (2000) or Hancock et al. (2001). The hand
calculations may be quite lengthy, particular if member properties xo and Cw are
unknown.
For beams,
Mcre = Sffcre (C-1.1.2-6)
Sf = gross section modulus to the extreme compression fiber
fcre = elastic critical lateral-torsional buckling stress. fcre is equal to Fe of main
Specification Section C3.1.2.1 for open cross-section members and C3.1.2.2 for
closed cross-section members. Hand solutions are well established for doubly-
and singly-symmetric sections, but not so for point symmetric sections (zees). Fe
1.2 MEMBERS
1.2.1 Column Design
Commentary Section C4 provides a complete discussion on the behavior of cold-formed steel
columns as it relates to the main Specification. This commentary addresses the specific issues
raised by the use of the Direct Strength Method of Appendix 1 for the design of cold-formed
steel columns. The thin-walled nature of cold-formed columns complicates behavior and
design. Elastic buckling analysis reveals at least three buckling modes: local, distortional, and
Euler (flexural, torsional, or flexural-torsional) that must be considered in design. Therefore, in
addition to usual considerations for steel columns: material non-linearity (e.g., yielding),
imperfections, and residual stresses, the individual role and potential for interaction of buckling
modes must also be considered. The Direct Strength Method of this Appendix emerged through
the combination of more refined methods for local and distortional buckling prediction,
improved understanding of the post-buckling strength and imperfection sensitivity in
distortional failures, and the relatively large amount of available experimental data.
Fully effective or compact columns are generally well predicted by conventional column
curves (AISC, 2001; Galambos, 1998, etc.). Therefore, the long column strength, Pne, follows the
same practice as the main Specification and uses the AISC (2001) curves for strength prediction.
The main Specification provides the long column strength in terms of a stress, Fn (Equations
C4.1-2 and C4.1-3). In the Direct Strength Method this is converted from a stress to a strength by
multiplying the gross area, Ag, resulting in the formulas for Pne given in Appendix 1.
In the main Specification, column strength is calculated by multiplying the nominal column
buckling stress, Fn, by the effective area, Ae, calculated at Fn. This accounts for local buckling
reductions in the actual column strength (i.e., local-global interaction). In the Direct Strength
Method, this calculation is broken into two parts: the long column strength without any
reduction for local buckling (Pne) and the long column strength considering local-global
interaction (Pnl).
The strength curves for local and distortional buckling of a fully braced column are
presented in Figure C-1.2.1-1. The curves are presented as a function of slenderness, which in
this case refers to slenderness in the local or distortional mode, as opposed to traditional long
column slenderness. Inelastic and post-buckling regimes are observed for both local and
distortional buckling modes. The magnitude of the post-buckling reserve for the distortional
buckling mode is less than the local buckling mode, as may be observed by the location of the
strength curves in relation to the critical elastic buckling curve.
1.0
Local: Eq. 1.2.1-6
Pn
Py 0.5
0.4
Pnl P Pcrl 0.4
= 1-0.15 crl
Py Py Py
0.6 0.6
Pnd P Pcrd
= 1-0.25 crd
Py Py Py
Pnd Pcr
=
Py Py
0
0 1 2 3 4
Slenderness = Py /Pcr
The development and calibration of the Direct Strength provisions for columns are reported
in Schafer (2000, 2002). The reliability of the column provisions was determined using the test
data of Appendix Section 1.1.1.1 and the provisions of Chapter F of the main Specification. Based
on a target reliability, β, of 2.5, a resistance factor, φ, of 0.84 was calculated for all the
investigated columns. Based on this information the safety and resistance factors of Appendix
Section 1.2.1 were determined for the pre-qualified members. For the United States and Mexico
φ = 0.85 was selected; while for Canada φ = 0.80 since a slightly higher reliability, β, of 3.0 is
employed. The safety factor, Ω, was back calculated from φ at an assumed dead to live load
ratio of 1 to 5. Since the range of pre-qualified members is relatively large, extensions of the
Direct Strength Method to geometries outside the pre-qualified set is allowed. Given the
uncertain nature of this extension, increased safety factors and reduced resistance factors are
applied in that case, per the rational analysis provisions of Section A1.2(b) of the main
Specification.
The provisions of Appendix 1, applied to the columns of Section 1.1.1.1, are summarized in
Figure C-1.2.1-2 below. The controlling strength is either by Appendix 1 Section 1.2.1.2, which
considers local buckling interaction with long column buckling, or by Section 1.2.1.3, which
considers the distortional mode alone. The controlling strength (minimum predicted of the two
modes) is highlighted for the examined members by the choice of marker. Overall performance
of the method can be judged by examination of Figure C-1.2.1-2. Scatter exists throughout the
data set, but the trends in strength are clearly shown, and further, the scatter (variance) is
similar to that of the main Specification.
1.5
Local: Eq. 1.2.1-6
⎛ Ptest ⎞ local
⎜ ⎟
⎜ P ⎟ 1
⎝ y ⎠d distortional
or
⎛ Ptest ⎞
⎜⎜ ⎟⎟
⎝ Pne ⎠ l
0.5
0
0 1 2 3 4 5 6 7 8
λ d = Py Pcrd or λ l = Pne Pcrl
Figure C-1.2.1-2 Direct Strength Method for Concentrically Loaded Pin-Ended Columns
the success of the Australian/New Zealand code (see Hancock et al., 2001 for discussion and
Hancock et al. 1994 for further details) the distortional buckling strength is limited to Py
instead of Pne. This presumes that distortional buckling failures are independent of long-
column behavior, i.e., little if any distortional-global interaction exists. See Section 1.1.2 for
information on rational analysis methods for calculation of Pcrd.
1.0
Local: Eq. 1.2.2-6
Mn
M y 0.5
0.4
Mnl Mcrl Mcrl 0.4
= 1-0.15
My My My
0.5
Mnd Mcrd Mcrd 0.5
= 1-0.22
My My My
Mnd Mcr
=
My M y
0
0 1 2 3 4
Slenderness = M y/Mcr
The lateral-torsional buckling strength, Mne, follows the same practice as the main
Specification. The main Specification provides the lateral-torsional buckling strength in terms of a
stress, Fc (Equations C3.1.2.1-2, -3, -4 and -5). In the Direct Strength Method, this is converted
from a stress to a moment by multiplying by the gross section modulus, Sf, resulting in the
formulas for Mne given in Appendix 1.
In the main Specification, for beams that are not fully braced and locally unstable, beam
strength is calculated by multiplying the predicted stress for failure in lateral-buckling, Fc, by
the effective section modulus, Sc, determined at stress Fc. This accounts for local buckling
reductions in the lateral-torsional buckling strength (i.e., local-global interaction). In the Direct
Strength Method, this calculation is broken into two parts: the lateral-torsional buckling
strength without any reduction for local buckling (Mne) and the strength considering local-
global interaction (Mnl).
The strength curves for local and distortional buckling of a fully braced beam are presented
in Figure C-1.2.2-1 and compared to the critical elastic buckling curve. While the strength in
both the local and distortional modes exhibit both an inelastic regime and a post-buckling
regime, the post-buckling reserve for the local mode is predicted to be greater than that of the
distortional mode.
The reliability of the beam provisions was determined using the test data of Section 1.1.1.2
and the provisions of Chapter F of the main Specification. Based on a target reliability, β, of 2.5, a
resistance factor, φ, of 0.90 was calculated for all the investigated beams. Based on this
information the safety and resistance factors of Appendix Section 1.2.2 were determined for the
pre-qualified members. For the United States and Mexico φ = 0.90; while for Canada φ = 0.85
because Canada employs a slightly higher reliability, β, of 3.0. The safety factor, Ω, is back
calculated from φ at an assumed dead to live load ratio of 1 to 5. Since the range of pre-qualified
members is relatively large, extensions of the Direct Strength Method to geometries outside the
pre-qualified set is allowed. However, given the uncertain nature of this extension, increased
safety factors and reduced resistance factors are applied in that case, per the rational analysis
provisions of Section A1.2(b) of the main Specification.
1.5
Local: Eq. 1.2.2-6
Local
1
Distortional
M test
My
0.5
0
0 1 2 3 4 5
λmax = M y M cr
The provisions of Appendix 1, applied to the beams of Section 1.1.1.2, are summarized in
Figure C-1.2.2-2. The controlling strength is determined either by Section 1.2.2.2, which
considers local buckling interaction with lateral-torsional buckling, or by Section 1.2.2.3, which
considers the distortional mode alone. The controlling strength (minimum predicted of the two
modes) is highlighted for the examined members by the choice of marker. Overall performance
of the method can be judged by examination of Figure C-1.2.2-2. The scatter shown in the data is
similar to that of the main Specification.
APPENDIX 1 REFERENCES
Acharya, V.V. and R.M. Schuster (1998), “Bending Tests of Hat Section with Multiple
Longitudinal Stiffeners,” Proceedings of the Fourteenth International Specialty Conference on
Cold-Formed Steel Structures, University of Missouri-Rolla, Rolla, MO, October, 1998.
American Institute of Steel Construction (2001), Manual of Steel Construction: Load and
Resistance Factor Design, 3rd Edition, American Institute of Steel Construction, Chicago,
IL.
American Iron and Steel Institute (2002), Cold-Formed Steel Design Manual, American Iron
and Steel Institute, Washington, DC.
Bambach, M.R., J.T. Merrick and G.J. Hancock (1998), “Distortional Buckling Formulae
for Thin Walled Channel and Z-Sections with Return Lips,” Proceedings of the 14th
International Specialty Conference on Cold-Formed Steel Structures, University of Missouri-
Rolla, Rolla, MO, October 1998, 21-38.
Bernard, E.S. (1993), “Flexural Behavior of Cold-Formed Profiled Steel Decking,” Ph.D.
Thesis, University of Sydney, Australia.
Cheung, Y.K. and L.G. Tham (1998), Finite Strip Method, CRC Press, 1998.
Cohen, J. M. (1987), “Local Buckling Behavior of Plate Elements,” Department of
Structural Engineering Report, Cornell University, Ithaca, NY, 1987.
Cook, R.D., D.S. Malkus and M.E. Plesha (1989), Concepts and Applications of Finite
Element Analysis, John Wiley & Sons, Third Edition.
Davies, J.M. and C. Jiang (1996), “Design of Thin-Walled Beams for Distortional
Buckling,” Proceedings of the Thirteenth International Specialty Conference on Cold-Formed
Steel Structures, University of Missouri-Rolla, Rolla, MO, October 1996, 141-154.
Davies, J.M., C. Jiang and V. Ungureanu (1998), “Buckling Mode Interaction in Cold-
Formed Steel Columns and Beams,” Proceedings of the Fourteenth International Specialty
Conference on Cold-Formed Steel Structures, University of Missouri-Rolla, Rolla, MO,
October 1998, 53-68.
Davies, J.M., P. Leach, D. Heinz (1994), “Second-Order Generalised Beam Theory.”
Journal of Constructional Steel Research, Elsevier, 31 (2-3) 221-242.
Desmond, T.P. (1977), “The Behavior and Design of Thin-Walled Compression Elements
with Longitudinal Stiffeners,” Ph.D. Thesis, Cornell University, Ithaca, NY, 1777.
Ellifritt, D., B. Glover and J. Hren (1997), “Distortional Buckling of Channels and Zees
Not Attached to Sheathing,” Report for the American Iron and Steel Institute,
Washington DC, 1997.
Elzein, A. (1991), Plate Stability by Boundary Element Method, Springer-Verlag, NY, 1991.
Galambos, T.V. (1998), Guide to Stability Design Criteria for Metal Structures, John Wiley &
Sons, the Fifth Edition.
Hancock, G.J. (1997), “Design for Distortional Buckling of Flexural Members,” Thin-
Walled Structures, 27(1), 3-12, Elsevier Science Ltd.
Hancock, G.J., Y.B. Kwon and E.S. Bernard (1994), “Strength Design Curves for Thin-
Walled Sections Undergoing Distortional Buckling,” Journal of Constructional Steel
Research, Elsevier, 31(2-3), 169-186.
Hancock, G.J., T.M. Murray and D.S. Ellifritt (2001), Cold-Formed Steel Structures to the
AISI Specification, Marcell-Dekker, New York, NY.
Hancock, G.J., C.A. Rogers and R.M. Schuster (1996) “Comparison of the Distortional
Buckling Method for Flexural Members with Tests.” Proceedings of the Thirteenth
International Specialty Conference on Cold-Formed Steel Structures, University of Missouri-
Rolla, Rolla, MO, October 1998, 125-140.
Harik, I.E., X. Liu and R. Ekambaram (1991), “Elastic Stability of Pates with Varying
Rigidities,” Computers and Structures, 38 (2) 161-168.
Höglund, T. (1980), “Design of Trapezoidal Sheeting Provided with Stiffeners in the
Flanges and Webs,” Swedish Council for Building Research, Stockholm, Sweden, D28:1980.
Schafer, B.W. (2002b), “Progress on the Direct Strength Method,” Proceedings of the 16th
International Specialty Conference on Cold-Formed Steel Structures, Orlando, FL. 647-662.
Schafer, B.W., Sarawit, A., Peköz, T. (2006). “Complex edge stiffeners for thin-walled
members.” Journal of Structural Engineering, ASCE, Vol. 132, No. 2, 212-226.
Schafer, B.W., T. Peköz, (1998), “Direct Strength Prediction of Cold-Formed Steel
Members using Numerical Elastic Buckling Solutions,” Proceedings of the Fourteenth
International Specialty Conference on Cold-Formed Steel Structures, University of Missouri-
Rolla, Rolla, MO, October 1998.
Schafer, B.W. and T. Peköz, (1999), “Laterally Braced Cold-Formed Steel Flexural
Members with Edge Stiffened Flanges.” Journal of Structural Engineering, ASCE, Vol. 125,
No. 2.
Schardt, R. (1989), Verallgemeinerte Technische Biegetheorie [Generalized Beam Theory],
Springer-Verlag, Berlin.
Schardt, R. and W. Schrade (1982), “Kaltprofil-Pfetten,” Institut Für Statik, Technische
Hochschule Darmstadt, Bericht Nr. 1, Darmstadt.
Schuster, R.M. (1992), “Testing of Perforated C-Stud Sections in Bending,” Report for the
Canadian Sheet Steel Building Institute, University of Waterloo, Waterloo Ontario,
Canada.
Shan, M., R.A. LaBoube and W. W. Yu (1994), “Behavior of Web Elements with
Openings Subjected to Bending, Shear and the Combination of Bending and Shear,”
Civil Engineering Study Structural Series, 94-2, Department of Civil Engineering,
University of Missouri-Rolla, Rolla, MO.
Silvestre, N. and D. Camotim (2002a), “First-Order Generalised Beam Theory for
Arbitrary Orthotropic Materials,” Thin-Walled Structures, Elsevier. Vol. 40, 755-789.
Silvestre, N. and D. Camotim (2002b), “Second-Order Generalised Beam Theory for
Arbitrary Orthotropic Materials,” Thin-Walled Structures, Elsevier. Vol. 40, 791-820.
Thomasson, P. (1978), “Thin-walled C-shaped Panels in Axial Compression,” Swedish
Council for Building Research, D1:1978, Stockholm, Sweden.
Willis, C.T. and B. Wallace (1990), “Behavior of Cold-Formed Steel Purlins under
Gravity Loading,” Journal of Structural Engineering, ASCE. Vol. 116, No. 8.
Yu, W.W. (2000), Cold-Formed Steel Design, John Wiley & Sons, Inc.
Zienkiewicz, O.C. and R.L. Taylor (1989), The Finite Element Method: Volume 1 Basic
Formulations and Linear Problems, McGraw Hill, the Fourth Edition.
Zienkiewicz, O.C. and R.L. Taylor (1991), The Finite Element Method: Volume 2 Solid and
Fluid Mechanics Dynamics and Non-linearity, McGraw Hill, the Fourth Edition.
Commentary on Appendix 2
Second-Order Analysis
2007 EDITION
Commentary on the 2007 North American Cold-Formed Steel Specification
APPENDIX 2 REFERENCES
American Institute of Steel Construction (2005), Specification for Structural Steel
Buildings, March 9, 2005.
Sarawit, A. (2003), Cold-Formed Steel Frame and Beam-Column Design, PhD Thesis,
and Research Report 03-03, Department of Civil and Environmental Engineering,
Cornell University, Ithaca, New York, March 2003.
Sarawit, A. And T. Pekoz (2006), “Notional Load Method for Industrial Steel Storage
Racks,” Thin-Walled Structures, Elserier, Vol. 44, No. 12, December 2006.
Commentary on Provisions
and Mexico
2007 EDITION
Commentary on the 2001 North American Cold-Formed Steel Specification
A1.1a Scope
In the 2007 edition of the Specification, both the Allowable Strength Design and the Load
and Resistance Factor Design are permitted to be used in a design.
A2.3a Ductility
The low ductility steel application is limited for curtain wall stud application in heavy
weight exterior walls in seismic areas with Design Categories D, E and F.
A3 Loads
A3.1 Nominal Loads
The Specification does not establish the dead, live, snow, wind, earthquake or other
loading requirements for which a structure should be designed. These loads are typically
covered by the applicable building code. Otherwise, the American Society of Civil Engineers
Standard, ASCE/SEI 7 (ASCE, 2005) should be used as the basis for design.
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 1999, AISC 2005).
When gravity and lateral 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.
C2 Tension Members
As described in Specification Section C2, the nominal tensile strength [resistance] of axially
loaded cold-formed steel tension members is determined either by yielding of the gross area of
the cross-section or by rupture of the net area of the cross section. At locations of connections,
the nominal tensile strength [resistance] is also limited by the capacities specified in Specification
Sections E2.7, E3, and E5 for tension in connected parts.
Yielding in the gross section indirectly provides a limit on the deformation that a tension
member can achieve. The definition of yielding in the gross section to determine the tensile
strength [resistance] is well established in hot-rolled steel construction.
For the LRFD Method, the resistance factor of φt = 0.75 used for rupture of the net section is
consistent with the φ factor used in the AISC Specification (AISC, 2005). The resistance factor
φt = 0.90 used for yielding in the gross section was also selected to be consistent with the AISC
Specification (AISC, 2005).
D6.1.2 Beams Having One Flange Fastened to a Standing Seam Roof System
For beams supporting a standing seam roof system, e.g. a roof purlin subjected to dead
plus live load, or uplift from wind load, the bending capacity is greater than the bending
strength of an unbraced member and may be equal to the bending strength of a fully
braced member. The bending capacity is governed by the nature of the loading, gravity or
uplift, and the nature of the particular standing seam roof system. Due to the availability
of many different types of standing seam roof systems, an analytical method for
determining positive and negative bending capacities has not been developed at the
present time. However, in order to resolve this issue relative to the gravity loading
condition, Section D6.1.2 was added in the 1996 edition of the AISI Specification for
determining the nominal flexural strength of beams having one flange fastened to a
standing seam roof system. In Specification Equation D6.1.2-1, the reduction factor, R, can
be determined by AISI S908 published by AISI (AISI, 2004). Application of the base test
method for uplift loading was subsequently validated after further analysis of the research
results.
pressure applied as uniform pressure in the E1592 test, without accounting for the reality
of the dynamic spatially-varying properties of the wind-induced pressures. The limits of
applicability of this factor (panel thickness and width) are conservatively listed based on
the scope of the research. The failure mode is restricted to those failures associated with
the load in the clip because this was how the research measured and compared the static
and dynamic capacities. The required strength factor of 0.67 is not permitted to be used
with other observed failures. In addition, the research does not support or confirm whether
interpolation would be appropriate between E1592 tests of the same roof system with
different spans, where one test meets the requirements, such as a clip failure, and another
test does not, such as a panel failure.
rather than direct bolt shear and bearing. An oversize or slotted hole is required for proper fit-
up due to offsets inherent in nested parts. Recent research (Bryant and Murray, 2001) has
shown that lapped and nested zee members with 1/2 in. (12.7 mm) diameter bolts without
washers and 9/16 in. x 7/8 in. (14.3 mm x 22.2 mm) slotted holes in the direction of stress can
develop the full moment in the lap.
4. The nominal tensile strength 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.
The presence of staggered or diagonal hole patterns in a bolted connection has long been
recognized as increasing the net section area for the limit state of rupture in the net section.
LaBoube and Yu (1995) summarized the findings of a limited study of the behavior of bolted
connections having staggered hole patterns. The research showed that when a staggered
hole pattern is present, the width of a rupture plane can be adjusted by use of s′2/4g.
Because of the lack of test data necessary for a more accurate design formulation, a
discontinuity between this Specification and the specifications or standards, stipulated in
Appendix A, may occur. The presence of a discontinuity should not be a significant design
issue because the use of the staggered hole patterns is not common in cold-formed steel
applications.
L
L
x
x
the test data with a corresponding value of β = 3.5 for LRFD. The Ω values are unchanged
from previous editions of the AISI ASD Specification.
B
g1
C
g2
D
E
s'
Figure C-E3.2-2 Flat Sheet Connections Having Staggered Holes
Note that when the required stress, f, in either shear or tension, is less than or equal to 20
per cent of the corresponding available stress, the effects of combined stress need not be
investigated.
For bolted connection design, the possibility of pullover of the connected sheet at the bolt
head, nut, or washer should also be considered when bolt tension is involved, especially for
thin sheathing material. For unsymmetrical sections, such as C- and Z-sections used as
purlins or girts, the problem is more severe because of the prying action resulting from
rotation of the member which occurs as a consequence of loading normal to the sheathing.
The designer should refer to applicable product code approvals, product specifications, other
literature, or tests.
For design tables and example problems on bolted connections, see Part IV of the Design
Manual (AISI, 2008).
E5 Rupture
Connection tests conducted by Birkemoe and Gilmor (1978) have shown that on coped
beams a tearing failure mode as shown in Figure C-E5-1(a) can occur along the perimeter of the
holes. Hardash and Bjorhovde (1985) have demonstrated these effects for tension members as
illustrated in Figure C-E5-1(b) and Figure C-E5-2. The provisions provided in Specification
Section E5 for shear rupture have been adopted from the AISC Specification (AISC, 1978). For
additional design information on tension rupture strength [resistance] and block shear rupture
strength [resistance] of connections (Figures C-E5-1 and C-E5-2), refer to the AISC Specifications
(AISC, 1989, 1999, and 2005).
Block shear is a limit state in which the resistance is determined by the sum of the shear
strength [resistance] on a failure path(s) parallel to the force and the tensile strength [resistance]
on the segment(s) perpendicular to the force, as shown in Figure C-E5-2. A comprehensive test
program does not exist regarding block shear for cold-formed steel members. However, a
limited study conducted at the University of Missouri-Rolla indicates that the AISC LRFD
equations may be applied to cold-formed steel members. The φ (LRFD) and Ω (ASD) values for
block shear were taken from the AISI 1996 edition of the Specification, and are based on the
performance of fillet welds. In calculating the net web area Awn, for coped beams, the web
depth is taken as the flat portion of the web as illustrated in Fig. C-E5-3.
The summary paper “AISC LRFD Rules for Block Shear in Bolted Connections – A Review”
(Kulak and Grondin, 2001) provides a summary of test data for block shear rupture strength. In
2004, Equations E5.3-1 and E5.3-2 were adopted for the limit state of block shear rupture for
bolted cold-formed steel connections because eccentricity in cold-formed steel sections is
usually small. In theory, provisions for block shear could also be applied to screw connections.
However, because the final placement location of self-drilling screws cannot be assured, a block
shear check is of little significance. Also, tests performed at the University of Missouri-Rolla
have indicated that the current design equations for shear and tilting provide a reasonably good
estimate of the connection performance for multiple screws in a pattern (LaBoube and Sokol,
2002).
Cope
Beam Failure by tearing
Failure by tearing out of shaded
out of shaded portion
Shear portion Shear
area Tensile
area
Tensile area
area
Po
(a) (b)
Figure C-E5-1 Failure Modes for Block Shear Rupture
Po
Po
Small tension
force
Large tension
force
Large shear
force
Small shear
force
Po
Po
(a) (b)
Figure C-E5-2 Block Shear Rupture in Tension
hwc
Commentary on Provisions
Applicable to Canada
2007 EDITION
Commentary on the 2007 North American Cold-Formed Steel Specification
A2.3.1a Ductility
The use of low ductility steel has been limited to curtain wall stud applications in
specific low seismic areas.
A3 Loads
The load provisions contained in Appendix B of CSA S136-01 were changed to be
compatible with the changes that are incorporated in Part 4 of the National Building Code of
Canada (NBC) 2005. This entails the following:
(1) The version of Limit States Design in NBC 2005 is based on the companion action format,
which is being adopted world-wide and is a more rational method of combining loads than
the previous version.
(2) NBC 1995 distinguished wind load for different categories of buildings using a return
period approach, an increase in design loads for earthquake based on building use by
means of an importance factor, and made no allowance for different snow loads based on
the occupancy of the structure. In NBC 2005, it was decided to harmonize the approach
used, and so the importance factor methodology was chosen for snow, wind and earthquake
loads.
C2 Tension Members
The general provisions for the design of tension members have not changed with respect to
the CSA Standard S136-01. The only change that was made involves staggered connections.
1
2
wg h 5
6
3
4 h=hole diameter
2
3
g
4
g
5 h h=hole diameter
g
6
g 7
8
s e
1
e1 2
h
g
3
h=hole diameter
g
4 5
6
s e2
Failure Path 1, 2, 3, 4, 5, 6
Lc = Lt + 0.6Lv
Lt = (s + e2 – 1.5h)
Lv = (e1 + 2g – 2.5h)
Lc = (s + e2 – 1.5h) + 0.6(e1 + 2g – 2.5h)
The provision regarding block tear-out of Section C2.2 was rewritten in accordance with
the latest research by Kulak and Grondin (2001). A new section on coped beams was also
added as per the recommendations by these authors.
spacing must be such that any stresses due to the rotation tendency are small enough so
that they will not significantly reduce the load-carrying capacity of the member. The
rotation must also be small enough (in the order of 2°) to be not objectionable as a service
requirement.
Based on tests and the study by Winter et al. (1949b), it was found that these
requirements are satisfied for any type of load if braces are provided at intervals of
one-quarter of the span, with the exception of concentrated loads requiring braces near the
point of application.
Fewer brace points may be used if it can be shown to be acceptable by rational analysis
or testing in accordance with Chapter F of the Specification, recognizing the variety of
conditions, including the case where loads are applied out of the plane of the web.
For sections used as purlins with a standing seam roof, the number of braces per bay is
often determined by rational analysis and/or testing. The requirement for a minimum
number of braces per bay is to recognize that predictability of the lateral support and
rotational restraint is limited on account of the many variables such as fasteners,
insulation, friction coefficients, and distortion of roof panels under load.
This provision extends the certification requirements to the welding of cold-formed members or
components to other construction, e.g., welding steel deck to structural steel framing.