You are on page 1of 21

www.kssc.or.

kr
Steel Structures 6 (2006) 71-91

Stability Analysis and Design of Steel Building Frames


Using the 2005 AISC Specification
Donald W. White *, Andrea E. Surovek , Bulent N. Alemdar , Ching-Jen Chang ,
1, 2 3 4

Yoon Duk Kim and Garret H. Kuchenbecker


4 5

1
Structural Engineering, Mechanics and Materials, Georgia Institute of Technology, Atlanta, GA 30332-03551, USA
2
Civil and Environmental Engineering, South Dakota School of Mines and Technology, Rapid City, SD 57701, USA
3
RAM International, 2744 Loker Avenue West, Carlsbad, CA, 92010, USA
4
Structural Engineering, Mechanics and Materials, Georgia Institute of Technology, Atlanta, GA, USA
5
Hermanson Egge Engineering, Rapid City, SD, USA

Abstract
The 2005 AISC Specification reflects the latest advances in the stability analysis and design of structural steel buildings. The
new Specification defines the general requirements for stability analysis and design and gives engineers the freedom to select
or devise their own methods within these constraints. It also provides several specific procedures. This paper first gives an
overview of the elastic analysis and design procedures in AISC (2005) as well as specific second-order distributed plasticity
methods upon which, in part, these procedures are based. The relationship between the AISC elastic provisions and the refined
inelastic methods is explained. Secondly, the paper highlights one interpretation of the AISC inelastic analysis and design
provisions that greatly facilitates the application of elastic-plastic hinge methods of analysis. The paper closes by presenting
four basic examples selected to illustrate key characteristics of each of the methods.
Keywords: Advanced analysis, distributed plasticity analysis, direct analysis, effective length, plastic hinge analysis

1. Introduction fabrication and erection tolerances) relative to the


ideal configuration of the structure.
Chapter C of the AISC (2005a) Specification, Stability The above Chapter C statement gives engineers the
Analysis and Design, states that any analysis and design freedom to select or devise methods that are best suited
method that addresses the following effects on the overall for the various structure types encountered in practice. It
stability of the structure and its elements is permitted: allows for innovation as long as there is proper
1. Flexural, axial and shear deformations (in members, consideration of the physical effects that influence the
connections and other components), structural response.
2. Reduction in stiffness (and corresponding increases Since the 1961 AISC Specification, when the column
in deformations) due to residual stresses and material effective length concept was first introduced by AISC,
yielding,
3. P-∆ effects, which are the effects of axial loads P
acting through the relative transverse displacements
of the member ends ∆ (see Fig. 1),
4. P-δ effects, which for initially straight members
loaded by bending in a single plane, are due to the
member axial load acting through the transverse
bending displacements relative to the member chord
(see Fig. 1), and
5. P-∆o and P-δo effects, which are caused by the
member axial loads acting through unavoidable
initial ∆o and δo geometric imperfections (within

*Corresponding author
Tel: +404-894-5839, Fax: +404-894-2278 Figure 1. Second-order P-∆ and P-δ effects (from White
E-mail: dwhite@ce.gatech.edu and Kim (2006)).
72 Donald W. White et al.

American analysis and design methods have addressed all incipient elastic buckling (considering the interaction of
of the above effects in some fashion whenever they are the member with the rest of the structure). The parameter
deemed to have an important influence on the structural F e' is calculated by dividing this buckling stress by the
response. Also, there has always been implicit recognition factor of safety for column elastic buckling, 23/12 = 1.92.
that engineers can use their professional judgment to The axial stress e' is determined typically using the
F

disregard specific effects (e.g., member shear deformations, equation


connection deformations, etc.) whenever they are considered
negligible. Member yielding, residual stress effects, and Fe 12 E
π
2
(3)
geometric imperfection effects traditionally have been ' = ---------------------------
-
2
(KLb ⁄ rb )
addressed in the formulation of member design resistances,
23

and, with minor exceptions, have not been considered in where b b is the column effective slenderness ratio
KL /r

the analysis. In some cases, engineers have included a in the plane of bending, and is the effective length K

nominal out-of-plumbness effect in the analysis of gravity factor associated with the above buckling solution. Also,
load combinations, particularly if the geometry and loading this factor is used typically in calculating the column
K

are symmetric. Strictly speaking, this is not necessary for axial resistance a. If desired, e' = e/1.92 and a can be
F F F F

the in-plane strength assessment of beam-columns in the determined directly from the buckling analysis model.
prior AISC Specifications. However, this practice is The term m in Eqs. (1) and (2) is discussed subsequently.
C

necessary to determine -∆o effects on bracing forces,


P AISC (1986) LRFD was the first American Specification
beam moments, connection moments, and in-plane moments to refer explicitly to the calculation of second-order
for checking the out-of-plane resistance of beam-columns. moments from a structural analysis. This specification
Also, plastic design procedures incorporate material introduced the following two-equation format for the
yielding into the analysis, typically by the use of idealized beam-column resistance:
plastic hinge models in adequately braced compact-
section members. However, geometric imperfection and Pu Mu Pu
residual stress effects traditionally are not included in this ≤ for < (4a)
φ c P n φ b Mn φ c Pn
------------- + -----------
- 1.0 ---------- 0.2
2

type of analysis.
The 1961 AISC Specification, and other AISC Pu Mu Pu
Specifications up until 1986, relied strictly on the
8
≤ for ≥ (4b)
φ c Pn φ b Mn φ c Pn
---------- + -- -----------
- 1.0 ---------- 0.2
9

structural analysis for calculation of geometrically


only

linear (first-order) forces and moments in the idealized where u is defined as the maximum second-order
M

geometrically-perfect, nominally-elastic (or elastic-plastic) elastic moment along the member length. AISC (1986)
structure. For elastic analysis and design, second-order states that u may be determined from a second-order
M

( -∆ and -δ) effects were addressed very discreetly via


P P elastic analysis using factored loads. However, it also
the following beam-column strength interaction equation: provides a procedure that uses amplification factors to
calculate the second-order elastic moments from a first-
fa Cm fb order elastic analysis. This procedure is in fact an
≤ (1) approximate second-order analysis. The moments u in
Fa fa ⎞
----
- + -----------------------
- 1.0
M
⎛1 – ------ F Eqs. (4) are the second-order elastic moments in an
⎝ Fe ⎠ b
'

idealized model of the structure assuming initially perfect


The expression geometry and completely linear elastic material response.
In all the AISC Specifications from AISC (1961)
Cm through AISC (1989) ASD and AISC (1999) LRFD, the
AF (2) influence of geometric imperfections and residual stresses
fa ⎞
-----------------
- =

⎛ 1 – ------
⎝ Fe ⎠ '
was addressed solely within the calculation of the
member resistances ( a and b in ASD and n and n in
F F P M

in Eq. (1) amplifies the member flexural stresses b tof LRFD). The new provisions in AISC
direct analysis

account for the -∆ and -δ effects.P P (2005a) recognize that specific advantages can be
Generally, Eq. (2) gives only a coarse approximation of realized by moving an appropriate nominal consideration
the true second-order effects. LeMessurier (1977) and of these effects out of the resistance side and into the
others later addressed better ways to determine the structural analysis side of the design equations. By
amplified bending stresses in general rectangular frames, incorporating an appropriate nominal consideration of
all within the context of Allowable Stress Design (ASD). these effects in the analysis, the resistance side of the
However, even today (2005), engineers using the AISC design equations is greatly simplified, and the accuracy of
(1989) ASD provisions can apply Eq. (2) in ways that the design checks is generally improved. These attributes
significantly under-estimate the physical second-order are discussed further in the subsequent sections.
effects in certain cases. The accuracy hinges largely on All of the above analysis and design procedures are
the proper determination of e, the member axial stress at
F based inherently on the use of second-order analysis
Stability Analysis and Design of Steel Building Frames: The AISC (2005) Specification and Beyond 73

(first-order elastic analysis with amplifiers being one type as 0.85 in Eqs. (1) and (2) for frames subject to joint
of second-order elastic analysis). Elastic analysis does not translation. This m value typically underestimates the
C

include the consideration of the member resistances in sidesway amplification effects (Salmon and Johnson, 1996).
itself. Therefore, all of the above elastic methods must Nevertheless, the ASD moment amplifier summarized in
include member resistance equations. However, the Eq. (2) is still conservative in many practical cases. This
method of analysis and the equations for checking the is because the predominant second-order effects are often
member resistances are inextricably linked. Changes in associated solely with the structure sidesway. Equation
the analysis calculation of the required strengths (e.g., a f (1) applies a single amplifier indiscriminately to the total
and b m/(1 − a/ e') in Eq. (1) or u and u in Eqs. (4))
f C f F P M flexural stresses from both sidesway and non-sway
can lead to simplifications in the calculation of the displacements. The amplification factor procedure in
member resistances (typically a in Eq. (1) or n in Eqs.
F P AISC (1999) LRFD and AISC (2005a) is more accurate,
(4)). Specifically, if the structural analysis is configured but involves a cumbersome subdivision of the analysis
to provide an appropriate representation of the internal into separate no-translation (nt) and lateral-translation (lt)
member forces, the in-plane resistance of the structure parts. Kuchenbecker . (2004) and White
et al . (2005a
et al

can be checked entirely on a cross-section by cross- & b) outline an amplified first-order elastic analysis
section basis. approach that provides good accuracy for rectangular
AISC (2005a) adopts the equations from AISC (1999) framing. This approach avoids the above cumbersome
LRFD as a base representation of the beam-column attributes of the AISC (1999 & 2005a) amplification factor
resistances for all of its analysis and design procedures. procedure.
However, the notation is generalized such that the In many cases, Eqs. (5) provide a more liberal
equations apply to both ASD and LRFD: characterization of the beam-column resistances than the
multiple beam-column strength curves in AISC ASD
P M
-----r + 8---------r ≤ 1.0
P
for -----r ≥ 0.2 (5b, AISC H1-1a) (1989). Also, AISC (2005a) provides the following
P c 9 Mc P
1

c
alternative interaction equation to characterize the out-of-
plane flexural-torsional buckling resistance of doubly-
P M
-------r- + 8---------r ≤ 1.0 for
P
----- < 0.2
r
(5a, AISC H1-1b) symmetric I-section members subjected to axial compression
2 P c 9 Mc P c
and major-axis bending moment:
The terms in these equations are defined as follows: P M 2
-------r + ⎛--------r ⎞ ≤ 1.0 (6, AISC H1-2)
r = the required axial compressive strength, determined
P Pco ⎝Mcx⎠
in ASD by analyzing the structure under 1.6 times the
ASD load combinations and then dividing the results by where co is the out-of-plane column strength and cx
P M

1.6, or determined in LRFD by analyzing the structure is the flexural strength with respect to lateral-torsional
under the LRFD load combinations. buckling. White and Kim (2006) provide a detailed
Mr = the required flexural strength, determined in ASD discussion of the background to Eq. (6) and explain that
by analyzing the structure under 1.6 times the ASD load this equation should be applied only for doubly-symmetric
combinations and then dividing the results by 1.6, or compact I-section members.
determined in LRFD by analyzing the structure under the In the subsequent developments, it is useful to consider
LRFD load combinations. the characterization of separate in-plane and out-of-plane
c = the allowable or design axial compressive strength,
P beam-column resistances. This can be accomplished by
given by n Ωc in ASD or by φc n in LRFD, where n is
P / P P using Eqs. (5) with different definitions of c and c, or
P M

the nominal compressive resistance determined in accordance by using Eqs. (5) to characterize the in-plane strength and
with Chapter E. Eq. (6), where it is applicable, to characterize the out-of-
Mc = the allowable or design flexural strength, given by plane strength. The in-plane resistance is estimated with
n Ωb in ASD or by φb n in LRFD, where
M / M n is the M Eqs. (5) by neglecting: (1) out-of-plane flexural, torsional
corresponding nominal resistance determined in accordance or flexural-torsional buckling in the calculation of c and
P

with Chapter F. (2) lateral-torsional buckling in the calculation of c. The


M

φc and φb = resistance factors for axial compression and out-of-plane resistance is estimated with Eqs. (5) by using
bending, both equal to 0.9. P c = co and using the governing resistance from Chapter
P

Ωc and Ωb = factors of safety for axial compression and F for c (White and Kim, 2006).
M

bending, both equal to 1.67. For checking the in-plane and out-of-plane strength of
The above 1.6 factor for ASD is smaller than the general I-section members, Eqs. (5) must be used. However,
column safety factor of 1.92 in the AISC ASD (1989) for doubly-symmetric compact I-section members subjected
amplification of the flexural stresses (see Eqs. (1) through to in-plane major-axis bending and large axial loads, Eq.
(3)). However, ASD-H1 also states that m shall be taken C (6) provides a more liberal assessment of the out-of-plane
The AISC (2005a) equation numbers are denoted by “AISC” followed by the equation number.
1
74 Donald W. White et al.

flexural-torsional buckling strength. AISC (2005a) allows


the engineer to neglect out-of-plane moments whenever
Mry/Mcy is smaller than 0.05 in the out-of-plane direction,

where Mry and Mcy are the required moment and the
corresponding resistance in this direction. Otherwise, an
extended form of Eqs. (5) must be used that includes the
out-of-plane bending.

2. Overview of Stability Analysis and Design


Procedures
2.1. Design by distributed plasticity analysis

This section discusses the requirements that must be


satisfied for strength design using a refined second-order
inelastic frame analysis in which the spread of yielding is
tracked explicitly through the cross-section and along the
member length. This type of analysis is referred to in this Figure 2.Lehigh (Galambos and Ketter 1959) residual
paper as a distributed plasticity analysis. stress pattern.
For hot-rolled compact I-section members, ASCE (1997),
Deierlein (2003), and Surovek-Maleck and White (2003
& 2004) have shown that distributed plasticity analysis
closely replicates the in-plane AISC LRFD beam-column
strengths based on an exact inelastic effective length, for
a comprehensive range of end conditions, when the
following nominal geometric imperfections, residual stresses
and material idealizations are included in the analysis:
• A sinusoidal or parabolic out-of-straightness with a
maximum amplitude of δo = L/1000, where L is the
unsupported length in the plane of bending.
• An out-of-plumbness of ∆o = L/500, the maximum
tolerance specified in the AISC (2005b) Code of Standard
Practice.
• The Lehigh (Galambos and Ketter, 1959) residual
stress pattern shown in Fig. 2.
• An elastic-perfectly plastic material stress-strain Figure 3. Example cantilever beam-columns.
response.
These results are not surprising, since Eqs. (5) were the plane of bending. Martinez-Garcia (2002), Surovek-
originally developed in part based on calibration to results Maleck and White (2003), and Deierlein (2003) summarize
from this type of analysis (ASCE, 1997; Surovek-Maleck the results from other more comprehensive studies.
and White, 2004). Since the AISC (2005a) analysis and If the above distributed plasticity analysis is to be used
design procedures do not make any distinction between in an AISC (2005a) LRFD context, the resistance factors
hot-rolled and general built-up members, the above φc = φb = 0.9 must be included. One way of doing this is
nominal geometric imperfections and residual stresses are to determine the nominal beam-column strength curves as
sufficient to capture the Specification requirements for all shown in Fig. 4, and then to multiply both the abscissa
compact I-section member types. Figure 3 defines two and the ordinate by φc = φb = 0.9 to obtain the final
simple cases that illustrate the above findings. These are member design resistance. However, identical results are
W10 × 60 cantilever beam-columns with L/r = 40 subjected obtained if both the yield strength Fy and the elastic
to axial and transverse loads at their free ends. In the first modulus E are factored by 0.9. If only the yield strength
case, the member is subjected to major-axis bending, Fy is factored by 0.9, the design strengths are overestimated

while in the second case it is subjected to minor-axis for highly slender columns that fail by elastic buckling.
bending. Figure 4 compares the nominal strength interaction The factoring of both E and Fy by 0.9 up front is
curves obtained from the above type of distributed preferred, since this approach facilitates the general
plasticity analysis to the corresponding AISC (2005a) inelastic analysis and design of structural systems. There
beam-column strength curve. The AISC curve is obtained is no straightforward way of applying distributed plasticity
using the AISC (2005a) effective length method with K = analysis, or any other form of inelastic analysis, in the
2, and is the same for both the minor and major-axis context of ASD. In as such, AISC (2005a) disallows the
bending examples, since their L/r values are the same in use of inelastic analysis for this approach. Most of the
Stability Analysis and Design of Steel Building Frames: The AISC (2005) Specification and Beyond 75

values producing the smallest total resistance should be


selected. For structures in which the postbuckling response
is asymmetric, with one direction corresponding to an
unstable postbuckling response and the other corresponding
to a stable one, the direction associated with the unstable
postbuckling response gives the smallest total resistance.
Although they can be programmed, the above out-of-
straightness considerations involve a level of complexity
that many engineers would find unacceptable. The α r < P

0.15 eL rule allows the engineer to disregard these


P

considerations in many practical situations.


The appropriate direction for the out-of-plumbness is
typically much easier to specify than the appropriate
direction for out-of-straightness. For the overall assessment
of building frames, it is sufficient in the vast majority of
cases to specify the out-of-plumbness in the total net
direction that the structure sways under the applied
Figure 4. Nominal strength curves by distributed plasticity loadings.
analysis versus the AISC (2005a) effective length method At the present time (2005), distributed plasticity analysis
for the example W10 × 60 cantilever beam-columns. has been applied most commonly in research studies
investigating the resistance of frames composed of
subsequent discussions in this paper are phrased in the adequately braced compact-section members. However,
context of LRFD. in design practice, the member strengths can be governed
The distributed plasticity analysis solution captures the by out-of-plane buckling, flange or web local buckling, or
in-plane design resistances of compact I-section members combinations of these strength limits. These limit states
completely when it includes the above attributes. Therefore, cannot be captured by a planar distributed plasticity
the in-plane beam, column and beam-column resistance analysis. Although some progress has been made on 3D
checks are automatically satisfied for these member types distributed plasticity methods (e.g., see Pi and Trahair
if the distributed plasticity analysis shows that the (1994), Izzudin and Smith (1996), Teh and Clarke (1998),
structure supports the design loadings. The engineer does Battini and Pacoste (2002) and Nukala and White (2004)
not need to perform any separate evaluation of the in- among others), the complexity and cost of the analysis is
plane member resistances. The Australian Standard significantly greater. Furthermore, the appropriate handling
AS4100 (SAA, 1990) was the first to explicitly permit of residual stresses, geometric imperfections, warping
this type of analysis and design. This Standard coined the continuity at beam-to-column joints, local-overall member
term advanced analysis , in a very specific context, to buckling interactions, restraint from and interaction with
denote an analysis that supersedes the Specification member floor slabs, and other important attributes that can influence
strength checks. Subsequently, various SSRC publications the 3D response have not been studied thoroughly at this
(e.g., White and Chen (1993)) as well as other research time. In fact, many of these effects are considered in
papers and reports have adopted this terminology. rather simplistic ways in ordinary design practices, e.g.,
In many practical steel building structures, member out-of-plane and spatial beam-column resistances are
out-of-straightness has little to no effect on the frame based typically on the assumption of unrestrained warping
resistance. White and Nukala (1997) explain that when at the member ends. Nevertheless, in frames subjected
α r is less than eL/7 ≅ 0.15 eL, where
P P P predominantly to in-plane loading, the authors assert that
α = 1.6 for ASD or 1.0 for LFRD and distributed plasticity analysis can be utilized to determine
eL = π using the moment of inertia I in the plane an accurate estimate of the internal forces in the structure.
2 2
P EI/L

of bending, These forces then can be checked against Specification


out-of-straightness effects can be neglected generally in member resistance equations corresponding to any of the
the distributed plasticity analysis (i.e., δo can be taken equal above 3D limit states not included in the analysis.
to zero). Otherwise, one must determine the appropriate Various analysis refinements are possible relative to the
direction for the member out-of-straightness. Usually, the above procedures. For instance, a number of approaches
strength is reduced the most due to member out-of- have been suggested for reducing the nominal out-of-
straightness if δo is specified in the direction of the plumbness relative to base values, e.g., see the Commentary
member deformations (relative to its chord) due to the of AISC (2005), White et al. (2003) and CEN (2003).
applied loads. However, in some cases, it is advisable to Also, other nominal residual stress distributions as well as
specify the various member δo values in the pattern of the more complete stress-strain models (e.g., models including
fundamental buckling mode obtained from an eigenvalue strain-hardening) can be incorporated within the distributed
buckling analysis. In these cases, the direction of the δo plasticity analysis. However, these considerations introduce
76 Donald W. White et al.

additional complexities that must be addressed. The applied at this level. Table 1 emphasizes the use of
above procedures fully satisfy the base requirements of notional lateral loads. However, in general cases where
the AISC (2005a) Specification. questions may arise about the appropriate calculation of
these loads, one can always use the more fundamental
2.2. Elastic analysis and design methods in AISC out-of-plumb geometry. For example, the total base shear
(2005a) due to any out-of-plumbness is always zero, and thus the
The AISC (2005a) Specification defines three specific total base shear due to the notional loads also must be
elastic analysis and design methods. These are: zero. As shown in Fig. 5, the notional loads arise from the
1. The direct analysis method, detailed in Appendix 7, sum of the P∆ shear forces above and below each level.
o

2. The effective length method, detailed in Section The P∆ shear at the base of the structure offsets the sum
o

C2.2a and of all the horizontal notional loads in the analysis model.
3. The first-order analysis method, detailed in Section Explicit modeling of the out-of-plumbness also removes
C2.2b. the need to calculate different notional loads for different
Table 1, summarizes the key attributes of each of these load combinations.
methods. Within the restrictions specified on their usage, The direct analysis method provides an improved
and provided that effects such as connection rotations or representation of the structure’s distributed plasticity
member axial and shear deformations are properly forces and moments at the strength limit of the most
considered in the analysis when these attributes are critical member or members. Due to this improvement in
important, each of the above methods is intended to the calculation of the internal forces and moments, AISC
comprehensively address all of the effects listed at the (2005a) bases its calculation of P , the column nominal
ni

beginning of Section 1. The following subsections provide strength for checking the in-plane resistance in Eqs. (5),
an overview of these AISC (2005a) procedures. The on the actual unsupported length in the plane of bending.
reader is referred to Deierlein (2004), Nair (2005a), Nair In short, the need to calculate in-plane effective length
(2005b) and White and Kim (2006) for additional (K) factors is eliminated.
discussions. Interestingly, the use of P = P for members with
ni y

compact cross-section elements was considered in the


2.2.1. Direct analysis method development of the direct analysis approach (Maleck,
The direct analysis method is the only one of the above 2001). Although this is a viable option, it requires the
three procedures that is generally applicable to all types modeling of out-of-straightness in the analysis for members
of frames. This method involves two simple modifications subjected to large axial compression (to properly capture
to the second-order elastic analysis: (1) the use of a in-plane limit states dominated by non-sway column
nominal reduced elastic stiffness and (2) the inclusion of flexural buckling). The modeling of member out-of-
a nominal initial out-of-plumbness. These two devices are straightness adds an additional level of complexity to the
adjustments to the analysis that approximate the internal analysis, and in many steel structures, P based on the
ni

forces and moments from the type of distributed plasticity actual unsupported length is only slightly smaller than P . y

analysis explained in the previous section. The reduced Therefore, AISC (2005a) uses P based on the actual
ni

elastic stiffness is taken as 0.8 of the nominal elastic unsupported length (K = 1) to capture the influence of
stiffness of the structure, except in members subjected to potential in-plane non-sway column flexural buckling.
large axial loads of αP > 0.5P , where the member
r y For certain member types, such as tapered members,
flexural rigidity is taken as 0.8 times the column inelastic there are significant advantages to using the cross-section
stiffness reduction factor τ (see Table 1). The base
b axial strength rather than the nominal buckling strength
nominal out-of-plumbness is taken as ∆ = L/500, the o P as the axial strength term in the beam-column
ni

same value as discussed previously for distributed plasticity interaction check (White and Kim, 2006). In cases where
analysis. However, the direct analysis provisions permit the member axial loads are small and the cross-section is
the use of a smaller nominal out-of-plumbness where compact, the column and beam-column resistances are
justified. For instance, when the sidesway amplification represented accurately using P = P , without the inclusion
ni y

∆2nd/∆1st is smaller than 1.5, AISC (2005a) permits the of any member out-of-straightness in the analysis. This
out-of-plumbness effect to be neglected when the simplification is appropriate whenever αP < 0.1P . For
r eL

associated notional load (see the discussion below) is members with cross-section elements that are slender
smaller than the corresponding applied lateral load. under axial compression, White and Kim (2006) show
Many engineers may prefer to model the above out-of- that P may be taken as QP when the above limit is
ni y

plumbness effects by using notional lateral loads. As satisfied, where Q is the AISC form factor accounting for
shown in Fig. 5, if the framing above and below a given local buckling effects with the web edge stress f taken as
vertical load elevation has the same ∆ /L, the out-of-
o F.
y

plumbness effect can be represented accurately by AISC (2005a) introduces a plethora of rules intended to
applying a notional lateral load of N = Y ∆ /L at the level
i i o allow (and provide limits on) the use of various idealizations
under consideration, where Y is the total vertical load
i and approximations (see Table 1). This characteristic is
Stability Analysis and Design of Steel Building Frames: The AISC (2005) Specification and Beyond 77

Table 1. Summary of specific AISC (2005a) elastic analysis-design procedures adapted from White and Kim (2006)
Direct analysis (Appendix 7) Effective length (Section C2.2a) First-order analysis (Section C2.2b)
Limitations on the None ∆2nd/∆1st ≤ 1.5 ∆2nd/∆1st ≤ 1.5, αPr ≤ 0.5Py
use of the method (See Note 5)
Type of analysis Second-order (See Note 1) Second-order (See Note 1) First-order, B1 is applied to the
member total moment
Structure geometry Nominal (See Note 2) Nominal Nominal
used in the analysis
Notional load to be ≤0.002 i Minimum if ∆2nd/∆1st
Y

applied in the analysis (See Note 2) if ∆2nd/∆1st > 1.5


1.5 Additive 0.002 i minimum 2.1(∆/ )Yi ≥ 0.0042 i additive
Y L Y

0.8 * Nominal, except eff = EI

0.8τb when α r > 0.5 y


EI P P Nominal Nominal
(See Note 3)
Effective stiffness τb = 4[α r y(1 − α r y)]
P /P P /P

used in the analysis Use of τb = 1 is permitted in all


cases if additional notional loads
of 0.001 i are applied, additive
Y

to other lateral loads


Pni in moment-frame columns is
Pni is based on the unsupported based on a buckling analysis or
length in the plane of bending, the corresponding effective
L i (i.e., = 1)
K length ; ni in all other cases
KL P

is based on i = i (i.e., = 1).


KL L K
In-plane flexural ni is based on the unsupported
P

buckling strength ni P
If α r < 0.15 eL, or if a member
P P
length in the plane of bending, i L
out-of-straightness of 0.001L or
the equivalent notional loading If ∆2nd/∆1st < 1.1, K may be taken
is included in the analysis, ni P equal to one in all cases.
may be taken equal to y P

(See Note 4)
P no is based on the unsupported length in the out-of-plane direction, o L
Out-of-plane flexural
buckling strength no Alternatively, no may be based on an out-of-plane buckling analysis or the corresponding effective
P
P

length o (see Note 4)


KL

General Note. ∆2nd/∆1stt is the ratio of the 2nd-order drift to the 1st-order drift (for rectangular frames, ∆2nd/∆1st may be taken as B2 calculated
by Section C2.1b). ∆/L is the largest 1st-order drift from all the stories in the structure. In structures that have flexible diaphragms, the ∆/L in
each story is taken as the average drift weighted in proportion to the vertical load, or alternatively, the maximum drift. All ∆2nd/∆1st and ∆/L
ratios shall be calculated using the LRFD load combinations or using a factor of α = 1.6 applied to the gravity loads in ASD. The factor a is
1.0 for LRFD and 1.6 for ASD. The term Y is the total gravity load applied at a given level of the structure. P is the member elastic
buckling resistance based on the actual unsupported length in the plane of bending, π2EI/L2 for prismatic members.
i eL

Note 1. Any legitimate method of second-order analysis that includes both P∆ and Pδ effects is permitted, including 1st-order analysis with
amplifiers. For αP < 0.15P , a P-large delta (P-∆) analysis using one element per member generally provides an accurate solution for the
sidesway displacements and the corresponding internal second-order forces and moments. However, for members with αP > 0.05P ,
r eL

either multiple elements must be used per member to obtain accurate second-order internal moments (unconservative error less than or
r eL

equal to 5%) in general from a P-large delta analysis, or a P-small delta amplifier must be applied to the element internal moments. Accurate
general P-∆ analysis solutions may be obtained by maintaining αP < 0.05P , where P = π2EI/l2 is the Euler buckling load in the plane of
bending based on the element length l. Second-order analysis methods that directly include both P-∆ and P-δ effects at the element level
r el el

generally provide better accuracy than P-large delta analysis procedures. The target of 5% maximum unconservative error is based on the
original development of the AISC LRFD beam-column strength equations (ASCE 1997; Surovek-Maleck and White 2004a).
Note 2. A nominal initial out-of-plumbness of ∆o/L = 0.002 may be used directly in lieu of applying 0.002Y minimum or additive notional
loads.
i

Note 3. The nominal stiffness and geometry should be employed for checking serviceability limit states. The reduced effective stiffness and
the notional loads or nominal initial out-of-plumbness are required only in considering strength limit states.
Note 4. AISC (2005) does not explicitly state this provision in the context of the direct analysis method. This provision is encompassed
within the Chapter C requirements for general stability analysis and design, which allow any method of analysis and design that addresses
the effects listed at the beginning of Section 1.
Note 5. The largest unconservative error associated with the limit αP < 0.1P is approximately 5% and occurs for a simply-supported,
concentrically loaded column with zero moment and αP = 0.1P = φ P . The target of 5% maximum unconservative error is based on the
r eL

original development of the AISC LRFD beam-column strength equations (ASCE 1997; Surovek-Maleck and White 2004a).
r eL c y

Note 6. The 1st-order analysis method does not account for the influence of significant axial compression in the rafters or beams of portal
frames. Therefore, this method strictly should not be applied for the analysis and design of the primary moment frames in these types of
structures.
78 Donald W. White et al.

length method when the second-order amplification of the


sidesway displacements is larger than 1.5, i.e., ∆2nd/∆1st >
1.5 (based on the nominal elastic stiffness of the
structure). This is because the effective length method
significantly underestimates the internal forces and
moments in certain cases when this limit is exceeded
(Deierlein, 2003 & 2004; Kuchenbecker et al., 2004;
White and Kim, 2006). For structures with ∆2nd/∆1st > 1.5,
AISC (2005a) requires the use of the direct analysis
method. Correspondingly, when using the direct analysis
approach with structures having ∆2nd/∆1st < 1.5, AISC
(2005a) allows the engineer to apply the notional lateral
loads (or the corresponding nominal out-of-plumbness) as
minimum values solely in the gravity-only load combinations.
For column and beam-column in-plane strength assessment
in moment frames, the effective length approach focuses
on the calculation of the member axial stresses Fei at
incipient buckling of an appropriately selected model (the
subscript “i” is used to denote in-plane flexural buckling).
This buckling model is usually some type of subassembly
that is isolated from the rest of the structural system
(ASCE, 1997). Engineers often handle the elastic buckling
stresses (Fei) implicitly, via the corresponding column
effective lengths KLi. The effective length is related to the
underlying elastic buckling stress via the relationship
Figure 5. Relationship between notional lateral loads and 2
nominal out-of-plumbness. Fei = ------π------E-------2- (7a)
(KLi ⁄ ri)
typical of design Specifications and is not new to these or
AISC provisions. One can avoid the need to consider all 2 2
of these caveats by using the direct analysis method, ( π E ) ⁄ ( Li ⁄ r i )
calculating Pn as the applicable column strength based
Ki = ------------------------------- (7b)
Fei
on the actual unsupported length, in-plane or out-of-
plane, including an elastic stiffness reduction factor of 0.8 In the effective length method, the effects of residual
or 0.8τb in the analysis as applicable, and including a stresses, P-∆o effects and P-δo effects are addressed
uniform initial out-of-plumbness of 0.002L in the analysis implicitly by the calculation of Pni from the column
relative to the perfect geometry of the structure. strength equations. These equations can be written in
terms of either KLi or Fei (AISC 2005a). Unfortunately,
2.2.2. Effective length method the selection of an appropriate subassembly buckling
The AISC (2005a) effective length method is the same model generally requires considerable skill and judgment.
as the traditional AISC method of analysis and design, As a result, there is a wide range of different buckling
but with the addition of a notional minimum lateral load models and K factor equations. In certain cases, subtle
for gravity-only load combinations. This minimum lateral differences in the models can produce radically different
load accounts for the influence of nominal geometric results (ASCE, 1997).
imperfections on the brace forces, beam moments, In particular, one should note that a rigorous buckling
connection moments and in-plane moments used for out- analysis of the entire structure does not necessarily
of-plane strength design of beam-columns. In actuality, provide an appropriate Fei or Ki (ASCE, 1997). Members
the effects of any physical out-of-plumbness are present that have small axial stress Fei at the buckling limit
for all load combinations. However, these effects are (relative to π2E/(Li/ri)2) tend to have large values for Ki
overwhelmed by the effects of the primary lateral loads in from Eq. (7b). In some cases, these large Ki values are
all the ASCE 7 (ASCE, 2005) lateral load combinations, justified while in other cases they are not. If the member
as long as the structure’s sidesway amplification is not is indeed participating in the governing buckling mode, a
excessive. Therefore, in the AISC (2005a) effective length large Ki is justified. If the member is largely undergoing
method, the notional lateral loads are specified solely as rigid-body motion in the governing buckling mode, or if
minimum lateral loads in gravity-only load combinations. it has a relatively light axial load and is predominantly
AISC (2005a) does not allow the use of the effective serving to restrain the buckling of other members, a large
Stability Analysis and Design of Steel Building Frames: The AISC (2005) Specification and Beyond 79

Ki value often is not justified. The distinction between often significantly more accurate. Therefore, the first-
these two situations requires engineering judgment. Some order analysis method is not considered further in this
of the situations requiring the greatest exercise of paper.
judgment to avoid excessively large K values include: (1)
columns in the upper stories of tall buildings, (2) columns 2.2.4. General comments
with highly flexible and/or weak connections and (3) Both the direct analysis and effective length methods
beams or rafters in portal frames, which may have require a second-order elastic analysis. However, any
significant axial compression due to the horizontal thrust second-order elastic analysis procedure is sufficient,
from the base of the frame. including first-order analysis with amplifiers, assuming
There is no simple way of quantifying the relative that the procedure is sufficiently accurate or conservative.
participation of a given member in the overall buckling The above methods differ in the way that they handle
of the structure or subassembly under consideration. geometric imperfection and distributed yielding effects in
Quantifying the participation requires an analysis of the the second-order analysis model and in the member
sensitivity of the buckling load to variations in the resistance equations.
member sizes. Even if one conducted such an analysis, The beam-column out-of-plane resistance check is the
there is no established metric for judging when Eq. (7b) same in both of the above methods, albeit with different
should or should not be used. Engineers typically base values of Pr and Mr. In AISC (2005a), the simplest out-
their effective length calculations on story-by-story models of-plane beam-column resistance check is given by Eqs.
to avoid the first of the above situations. They idealize (5) but with Pn = Pno, where Pno is the out-of-plane
columns with weak and/or flexible connections as pin- flexural, torsional or flexural-torsional buckling strength
ended leaner columns with K = 1 to avoid the second of of the member as a concentrically-loaded column. Eq. (6)
the above situations. Lastly, many engineers utilize K = 1 generally provides a more liberal estimate of the out-of-
for design of the beams or rafters in portal frames for the plane flexural-torsional resistance of doubly-symmetric
axial compression effects, although Eq. (7b) may suggest compact I-section members.
K > 1 based on the Fei from an eigenvalue buckling
analysis. 2.3. Design by direct elastic-plastic hinge analysis

The direct analysis method provides a more straightforward Since 1961, the AISC Specifications have permitted the
and accurate way of addressing frame in-plane stability use of plastic analysis and design in cases where members
considerations. By including an appropriately reduced subjected to plastic hinging satisfy requirements that
nominal elastic stiffness, an appropriate nominal out-of- ensure their ductility. However, the AISC Specifications
plumbness of the structure, and an appropriate out-of- from 1969 through 1999 have also generally required the
straightness (for members subjected to high axial loads) engineer to supplement the plastic analysis by beam-
in the analysis, the member in-plane length effects can be column strength interaction checks in which the axial
removed entirely from the resistance side of the design resistance term is based on a member effective length.
equations. The member in-plane column strength Pni may This practice adds significant complexity to the AISC
be taken simply as Py for members that satisfy the plastic analysis and design procedures. Furthermore, at
previously discussed caveats. In-plane stability is addressed best, the resulting beam-column interaction equations
by estimating the actual required internal cross-section provide only an approximate assessment of the frame
strengths Pr and Mr directly from the analysis, and by stability behavior under progressive plastic hinge formation.
comparing these required strengths against appropriate In many cases, they overly restrict the forces and moments
cross-section based resistances. Alternatively, to avoid the in sway frame columns (Ziemian et al., 1992; McGuire,
need to include out-of-straightness effects in members 1995).
with large axial loads, Pni may be calculated using the The AISC (2005a) direct analysis method can be
actual in-plane unsupported length Li (K = 1). extended to provide an attractive alternative to the above
procedures. The extension is very simple - for members
2.2.3. First-order analysis method that satisfy separate requirements to ensure sufficiently
The first-order analysis method, summarized in Table ductile response (i.e., sufficient rotation capacity), moment
1, is implicitly a simplified conservative application of redistribution is allowed based on the assumption of
the direct analysis approach, targeted at rectangular or elastic-perfectly plastic hinge behavior at the limit of the
tiered building frames. White and Kim (2006) detail the member resistance from Eqs. (5). This extension satisfies
conservative assumptions invoked in the development of the AISC (2005a) Appendix I provisions for inelastic
this procedure. Although the first-order analysis method analysis and design. The separate ductility requirements
can be useful for simplified analysis and design of some in AISC (2005a) Appendix 1 include restrictions on:
types of frames, this method is really just a direct analysis • The material yield strength Fy,
with a number of simplifying assumptions. There are • The flange and web slenderness values bf/2tf and hp/tw,
numerous other ways to apply direct analysis using an • The member out-of-plane slenderness Lb/ry,
approximate second-order analysis, many of which are • The magnitude of the axial force αPr (α = 1.0 for
80 Donald W. White et al.

LRFD loadings) and The third example is a hypothetical industrial building


• The connection details. frame that meets representative service drift requirements.
These restrictions tend to preclude the formation of However, the behavior of the lateral load resisting system
plastic hinges in beams and beam-columns with noncompact is sensitive to stability considerations to the extent that
flanges or webs, in beams where the resistance is governed the structure’s limit load is reached when the first column
by out-of-plane lateral-torsional buckling, and in beam- plastic hinge forms in the direct elastic-plastic hinge
columns with out-of-plane unbraced lengths large enough analysis. Lastly, a fourth example is provided in which the
such that the resistance is governed by Eq. (6). Therefore, lateral load resisting columns are subjected to substantial
for members subjected to in-plane loading and where non-sway gravity load moments. This frame exhibits
inelastic redistribution is allowed, the strength either is some reserve strength beyond the first plastic hinge for its
given or is accurately approximated by Eqs. (5) with n M critical load combination. All of these examples are
= p and with n = ni calculated using the actual member
M P P individual members or single-story rectangular structures
unsupported length in the plane of bending. Furthermore, with either fully-restrained (FR) or simple connections.
when α r is less than 0.1 eL, ni = y is an acceptable
P P P P The reader is referred to Maleck (2001), Martinez-Garcia
approximation. The approximation ni = y also can be
P P (2002), Deierlein (2003), Surovek (2005), White
et al. et

used for general I-section members as long as an appropriate al. (2005a & b) and White and Kim (2006) for consideration
out-of-straightness is included in the second-order of a broader range of structure types including multi-story
analysis. Members that do not satisfy all the requirements frames, braced and combined framing systems, trussed
necessary to ensure ductile response must be designed framing, PR frames, gabled frames, and frames with slender
elastically in the manner discussed in Section 2.2.1. element section members, singly-symmetric members,
The above type of analysis and design is referred to in nonprismatic members, and/or members with non-
this paper as the direct elastic-plastic hinge method. In constant axial load along their lengths.
this method, the elastic stiffness of the structure is reduced In all the following examples, the appropriate stiffness
to account for distributed yielding effects neglected in the reduction factor in the direct analysis is 0.8. That is, yP/P

elastic-plastic hinge idealization. Also, a nominal initial is always less than 0.5, and therefore τb = 1.0. The
out-of-plumbness is included to account for geometric distributed plasticity analysis is conducted using a factor
imperfection effects. These devices eliminate the need to of 0.9 on the elastic stiffness and the strength y in all
E F

calculate and apply column effective lengths in the cases. The axial force is close to or smaller than 0.1 eL
P P

context of inelastic design, as long as the second-order in all the above problems. Therefore, the column out-of-
effects are captured in the elastic-plastic hinge analysis. straightness is neglected and, to illustrate the validity of
Furthermore, these devices allow the engineer to take this option, ni is taken equal to y in the direct analysis
P P

advantage of second-order elastic-plastic hinge analysis solutions. A nominal column out-of-straightness is included
software. This kind of software is becoming more and in the direction of the member curvature due to the applied
more available in engineering practice. loadings in all of the distributed plasticity solutions, to
Other limits are possible within which the engineer may demonstrate that the direct analysis solutions give accurate
be permitted to perform a classical rigid-plastic analysis predictions for these cases with out-of-straightness
and design, e.g., see King (2001) and Davies and Brown neglected. All of the direct and distributed plasticity
(1996). These limits are not addressed in this study. analyses are conducted using an out-of-plumbness of
∆o = 0.002 , specified in the direction of the drift under
L

3. Illustrative Examples the applied loads. For the effective length method, a
second-order elastic analysis is conducted using the nominal
This section provides a number of basic examples elastic stiffness and the initially perfect geometry.
aimed at illustrating the relative merits of the various The second-order elastic and distributed plasticity
analysis and design procedures outlined in Section 2. The solutions are conducted using the GT-Sabre software system
first example addresses the strength predictions for one of (Chang 2005). A flexibility-based element (Alemdar and
the major-axis bending cases of the Fig. 3 cantilever White 2005), which is based on an equilibrium-based
beam-columns. Since this is a statically determinate distribution of moments and a correspondingly accurate
structure, inelastic analysis and design does not provide representation of inelastic curvatures (including the influence
any advantage. This example is representative of numerous of transverse distributed loads), is used for the beams.
nonredundant stability critical benchmark problems The mixed element developed by Alemdar and White
considered in the development of the direct analysis (2005), which is capable of accurate modeling of the
method (Deierlein 2003; Surovek-Maleck and White second-order moments and inelastic curvatures, is used
2003 & 2004). The second example is a nonredundant for the beam-columns. A small modulus of 0.001 is E

portal frame previously posed by LeMessurier (1977). assumed for the yielded material. The direct elastic-
This frame exhibits significant sway under gravity loadings, plastic hinge solutions are conducted using the Mastan2
due to lack of symmetry of its geometry. Also, the beam software system (McGuire et al ., 2000). For major-axis
governs its maximum resistance rather than the columns. bending, the Mastan2 yield surface, with axial force and
Stability Analysis and Design of Steel Building Frames: The AISC (2005) Specification and Beyond 81

Figure 7. Ratio of the effective moment of inertia Ie at


the distributed plasticity limit load to the elastic cross-
section moment of inertia I, example cantilever beam-
column subjected to major-axis bending.
Figure 6. Comparison of effective length and direct analysis
method beam-column strength interaction calculations to
distributed plasticity analysis, example cantilever beam- (0.626Py) based on K = 2, or KL/rx = 80. The anchor point
column subjected to major-axis bending, adapted from on the horizontal axis is φbMn = 0.9Mp for both of the
White and Kim (2006). above strength curves.
The direct analysis provides a reasonably accurate
bending moment anchor points of 0.9Py and 0.9Mp, is estimate of the distributed plasticity internal force and
slightly more liberal than the yield surface corresponding moment up to the predicted maximum resistance. The
to Eqs. (5) with Pc = 0.9Py and Mc = 0.9Mp. force point trace from the distributed plasticity analysis
At the strength design loads, ∆ /∆ , which is
2nd 1st indicates a larger moment than determined by the second-
equivalent to the sidesway amplifier B in AISC (2005a),
2 order elastic analysis of the nominally-elastic initially-
is larger than 1.5 in all of these examples. Therefore, perfect structure by the effective length method. This is
AISC (2005a) disallows the use of the effective length due to the stiffness reduction factor of φc = 0.9 in the
method for all of these structures. The effective length distributed plasticity analysis as well as the effects of
method is applied in the examples to illustrate the distributed yielding as the maximum strength limit is
implications of its use for these types of frames. approached. Figure 7 shows the ratio of the effective
moment of inertia Ie for the beam-column (i.e., the
3.1. Cantilever column moment of inertia of its elastic core) to the elastic cross-
Figure 6 compares the results predicted by the direct section moment of inertia at the distributed plasticity
analysis and the effective length methods to the results analysis limit load. The member experiences significant
from distributed plasticity analysis for the Fig. 3 beam- yielding at the strength limit, but a full plastic hinge has
column subjected to major-axis bending and the specific not yet formed at its base. The maximum value of P on
loading of H = 0.01P. The direct (elastic) and direct the force-point trace from the distributed plasticity
elastic-plastic hinge analysis methods are one and the analysis corresponds to the limit load in the refined
same up to the maximum strength limit for this example, inelastic solution. Correspondingly, the design strength
since the structure is statically determinate. Figure 6 for the effective length and the direct analysis methods is
shows the force-point traces at the column base (i.e., the defined by the intersection of their force-point traces and
variation of the member base moment and axial force for the corresponding beam-column strength curves. Both the
increasing levels of the applied loads) determined by the effective length and direct analysis provisions are calibrated
direct analysis, effective length and distributed plasticity such that these intersection points give an accurate to
analysis solutions. conservative estimate of the “actual” maximum strength
Figure 6 also shows the factored beam-column strength represented by the distributed plasticity solution.
curves for the effective length and the direct analysis The direct analysis method accounts for all the key
methods (with φ = φ = 0.9). In this problem, the member
c b attributes that influence the in-plane system stability effects
in-plane strength governs in both the direct analysis and directly within the analysis. Hence, its force point trace
the effective length solutions if the member is braced at may be compared against the cross-section strength limit
its top and bottom in the out-of-plane direction, K = 0.7 is given by Eqs. (5) with Pni = Py and Mn = Mp. Conversely,
used for the calculation of Pno, and Eq. (6) is used for the the effective length method accounts for the in-plane
out-of-plane strength assessment. Therefore, Fig. 6 focuses system stability effects by reducing the member axial
only on the in-plane resistance. The direct analysis in- strength Pni via the effective length KL, or the elastic
plane strength curve is based on φcPni = 0.9Py (although P buckling stress Fei, obtained implicitly or explicitly from
is slightly larger than 0.1PeL). The corresponding axial an appropriately configured buckling analysis.
strength for the effective length method is φcPni = 0.9 Table 2 summarizes the results at the calculated strength
82 Donald W. White et al.

Table 2. Summary of calculated design strengths, cantilever beam-column example


Pmax (kN) Mmax (kN-m) Mmax/HL Pmax/Pmax (Distributed Plasticity)

Traditional effective length 1590 119 1.68 0.96


Direct analysis 1730 217 2.79 1.05
Distributed plasticity analysis 1650 198 2.68

limits by each of the above methods. The ratios of the


base moments Mmax = HL + P(∆ + ∆o) to the primary
moment HL indicate the magnitude of the second-order
effects. The axial load at the direct analysis strength limit,
which is representative of the strength in terms of the
total applied load, is 5% higher than that obtained from
the distributed plasticity analysis. Conversely, the axial
load at the effective length method beam-column strength
limit is 4% smaller than that obtained from the distributed
plasticity solution. Both of these estimates are within the
targeted upper bound of 5% unconservative error relative
to distributed plasticity analysis established in the original
development of the AISC LRFD beam-column strength Figure 8. LeMessurier’s (1977) Example 3.
equations (ASCE 1997; Surovek-Maleck and White 2004).
The difference in the calculated internal moments is
much larger. This difference is expected since the effective
length approach compensates for the underestimation of
the actual moments by reducing the value of the axial
resistance term Pni, whereas the direct analysis method
imposes additional requirements on the analysis to obtain
an improved estimate of the actual internal moments.
This more accurate calculation of the internal moments
also influences the design of the restraining members and
their connections. For instance, in this example, the direct
analysis base moments are more representative of the
actual moments required at this position to support the
applied loads at the limit of the member resistance. In
other words, the direct analysis method provides a direct Figure 9. Design load fraction versus the story drift for
calculation of the required strengths for all of the structural LeMessurier’s Example 3.
components. Conversely, the effective length approach
generally needs supplementary rules for calculation of the
required component strengths. In AISC (2005a), these resistance is governed by in-plane limit states. A live-to-
supplementary rules are: (1) A minimum notional lateral dead load ratio of 3.0 is assumed for the LRFD
load is applied in gravity-only load combinations, and (2) calculations. The loading 1.2∆ + 1.6Lr + 0.8W, with the
The effective length method is limited to frames having wind applied in the direction of the sway under the
∆2nd/∆1st < 1.5 (see Table 1). gravity load, is the most critical of the ASCE 7-05 load
combinations for this structure. The following discussions
3.2. LeMessurier’s (1977) Example 3 focus on the behavior of the frame under this load
The structure shown in Fig. 8 is a market shed deliberately combination.
designed to have slender columns. LeMessurier (1977) Figure 9 shows the fraction of the design load versus
presents this frame as his Example 3. He explains that the drift ∆/L from the different analysis methods, Fig. 10
column CD is on an open side and is slender for architectural shows the moment diagrams and the diagrams of the
reasons. Column AB is selected to limit the first-order effective moment of inertia Ie at the distributed plasticity
drift under nominal wind to a maximum of 25.4 mm (1 analysis limit load, Fig. 11 shows the deflected shape, the
in). The discussions below focus on checking the maximum location of the single plastic hinge and the design load
resistance of this frame by LRFD using the various fraction at the formation of the plastic hinge from the
analysis and design procedures. The girder is assumed to direct elastic-plastic hinge analysis, Fig. 12 shows the
be adequately braced in the out-of-plane direction such applied fraction of the design load versus the girder
that Mn = Mp. Column AB is braced at its ends in the out- maximum bending moment from the different analysis
of-plane direction. This bracing is sufficient such that its methods, and Fig. 13 shows the force-point trace at the
Stability Analysis and Design of Steel Building Frames: The AISC (2005) Specification and Beyond 83

Figure 10. Diagrams of member moments and effective moment of inertia Ie at the distributed plasticity analysis limit
load, LeMessurier’s Example 3.

Figure 11. Deflected shape, location of plastic hinge, and


design load fraction at the formation of the plastic hinge
from direct elastic-plastic hinge analysis, LeMessurier’s
Example 3.

Figure 13.Force-point trace at the top of column AB


obtained from the different analysis-design methods,
LeMessurier’s Example 3.

discontinuity in the direct analysis load-deflection curve


(Fig. 9) occurs when a plastic hinge forms at the top of
column AB. The resulting overall shape of the direct
analysis load-deflection curve matches quite well with the
distributed plasticity analysis results.
Figure 12. Applied fraction of the design load versus the Due to the smaller drift predicted in the second-order
girder maximum bending moment obtained from the different
analysis-design methods, LeMessurier’s Example 3. elastic analysis by the effective length method, the
maximum girder moments are slightly smaller for a given
design load fraction. As a result, the effective length
top of column AB from the different analysis methods. method predicts a maximum strength of the structure (at
One can observe that the direct elastic-plastic hinge the formation of a plastic hinge with a resistance of 0.9Mp
analysis provides a reasonable estimate of the distributed in the girder) at 0.987 of the loading 1.2∆ + 1.6Lr + 0.8W.
plasticity solution. Due to the use of the smaller stiffness The plastic hinge and distributed plasticity analysis methods
reduction factor in the plastic hinge analysis (0.8 versus predict corresponding strength limits of 0.975 and 0.959
0.9), the elastic deflections are slightly larger than those of the design loading. Although there is substantial
predicted by the distributed plasticity analysis. However, distributed yielding along the length of the girder at the
the sway deflection at the limit load is slightly smaller in distributed plasticity strength limit (see Fig. 10), this
the plastic hinge analysis. The maximum resistances in all yielding has a small impact on the girder maximum
of the analysis solutions correspond to the formation of a second-order moment.
plastic hinge in the girder. The second major slope
84 Donald W. White et al.

Figure 14. Maleck’s (2001) 11-bay industrial building frame.

3.3. Maleck’s (2001) Industrial Building Frame


Figure 14 shows an example 11-bay industrial building
frame developed by Maleck (2001) and studied further in
Martinez-Garcia (2002), Deierlein (2003) and Surovek-
Maleck and White (2004). Five gravity columns are
located on each side of the interior moment frame. In
practice, the exterior bays of this type of frame might be
designed with simply-supported joist girders, hence the
large number of gravity columns. Large gravity loads are
specified to simulate conditions in some single-story industrial
buildings, such as automobile plants (Springfield, 1991),
in which large equipment loads are located on the roof.
Significant -∆ effects are transmitted to the moment
P

frame from the large number of gravity (leaner) columns.


However, the design lateral loadings from wind, etc. are Figure 15. Design load fraction versus the story drift for
Maleck’s frame.
relatively small.
The same girder size (W27x84) is used throughout the
structure in Fig. 14. This size is selected to resist the calculations.
moments in the simply-supported exterior bays of the Figures 15 through 19 present the same results as Figs.
frame. Although a 24 × 84 has sufficient strength to
W 9 through 13 for LeMessurier’s (1977) Example 3
withstand the design gravity loads, the deeper 27 × 84W structure, but illustrate the responses for Maleck’s (2001)
profile is selected to reduce the lateral drift. The tops of frame. Figure 19 shows the force-point trace at the top of
the two 10 × 49 interior columns are rigidly connected
W the left-hand lateral load resisting column. This location
to the middle 27 × 84 girder. The middle girder is
W has the most critical combination of column axial force
continuous over the tops of these columns, and the and bending moment. Figure 18 shows the girder negative
columns are attached to the bottom of the girder by end- bending moment versus the design load fraction just to
plate connections. Base restraint is provided for the the left of the left-hand column. This is the location of the
lateral-load resisting columns to limit the drift under a largest girder bending moment.
service loading of 1.0∆ + 0.5 + 0.7 to /400. The service
Lr W L Again, the direct elastic-plastic hinge analysis provides
drift is calculated on the nominally-elastic geometrically- a reasonable estimate of the distributed plasticity analysis
perfect frame, that is, no stiffness reduction or geometric solution. Due to the smaller stiffness reduction factor in
imperfections are applied for the service load analysis. the plastic hinge analysis (0.8 versus 0.9), the elastic
The strength design of the girders and the columns in deflections are slightly larger than those predicted by the
the above structure is governed by the LRFD gravity load distributed plasticity analysis. Also, the sway deflection
combination 1.2∆ + 1.6 . The ASCE7-05 required inclusion
Lr at the limit load is slightly larger in the plastic hinge
of 0.8W with this combination is neglected to accentuate analysis. The maximum resistance in the plastic hinge
the frame stability effects. The following discussions analysis (1.155 of the design load level) corresponds to
focus on the frame behavior under this load combination. the formation of a plastic hinge at the top of the left-hand
The columns are again governed by the frame in-plane column. Subsequently, plastic hinges form at the other
stability limit states when Eq. (6) is used for the out-of- ends of the lateral load resisting columns within the post-
plane strength checks. The reader is referred to Surovek- peak range of the response at 1.115, 1.075 and 0.930 of
Maleck and White (2004) and to Kuchenbecker .
et al the design load level (see Fig. 17). If the example frame
(2004) for further details of the strength and service load were not so sensitive to stability effects, there would be
Stability Analysis and Design of Steel Building Frames: The AISC (2005) Specification and Beyond 85

Figure 16. Diagrams of member moments M and effective moment of inertia Ie at the distributed plasticity analysis limit
load, Maleck’s frame, 1.2∆ + 1.6L . r

Figure 17.Deflected shape, plastic hinge locations, and


design load fraction at the formation of the plastic hinges
predicted by direct elastic-plastic hinge analysis, Maleck’s
frame, 1.2∆ + 1.6L .
r

some reserve strength associated with the redistribution


of the left-hand column moments from its top to its base,
as well as the redistribution of moments from the left-
hand column to the right-hand column. Of course, elastic
analysis and design discounts this reserve system strength.
The distributed plasticity analysis predicts a similar Figure 18. Applied fraction of the design load versus the
response; however, at the maximum load level (at 1.158 maximum girder bending moment obtained from the
of the design loading), there are no fully-formed plastic different analysis-design methods, Maleck’s frame, 1.2∆ +
hinges in the structure (see Fig. 16). The limit load is 1.6L .
r

reached due to a combination of P-∆ effects along with


the progressive softening of the columns and girders due 2.12 using Eq. (C-C2-5) of the AISC (2005a) Commentary.
to the spread of plasticity through their cross-sections and The conservatism of this strength check is due to two
along their lengths. At the distributed plasticity analysis causes: (1) a substantial fraction of the column internal
limit load, the girders are significantly yielded on the left- moments are due to non-sway gravity loading and (2) the
hand side of each of the columns (see Fig. 16). However, AISC (2005a) effective length method requires the inclusion
this yielding is highly localized due to the moment of a minimum notional lateral load with gravity-only load
gradient in the girders at these positions. combinations. The original calibration of the effective
Again, the second-order elastic analysis of the perfect length method in AISC LRFD (1986) did not include any
nominally-elastic structure by the effective length method sway frames that were subjected to gravity loads causing
results in significantly smaller overall drift at the predicted non-sway member moments. The loadings in the original
strength limit (see Fig. 15). The corresponding force- benchmark solutions were all applied directly at the
point trace for the columns intersects the column in-plane beam-column joints such that the non-sway moments
strength curve at a design load fraction of 0.946. The were zero (ASCE 1997; Surovek-Maleck and White
effective length method beam-column strength for this 2003; Deierlein 2003). The above conservatism of the
frame is based on a column effective length factor of K = AISC (2005a) effective length method strength checks,
86 Donald W. White et al.

Figure 20.AISC Specification Committee Design Problem


13 (DP-13).
Figure 19. Force-point trace at the top of left-hand lateral-
load resisting column obtained from the different analysis-
design methods, Maleck’s frame, 1.2∆ + 1.6L . r
this method does a good job of estimating the maximum
overall capacity of the frame. This is evidenced by the
due to the significant non-sway column moments, is similarity of the peak ordinate values for the direct and
much larger in the next example (see Section 3.4). Also, distributed plasticity analysis methods in Figs. 15 and 19.
the original AISC (2005a) beam-column strengths were
calibrated to give accurate to conservative predictions of 3.4. AISC Specification Committee (2004) Design
the results from distributed plasticity analysis without the Problem 13 (DP-13)
inclusion of any notional lateral loads. However, even if Figure 20 shows a single story rectangular frame posed
the beam-column strength checks in the example frame as one of many design problems during the final Specification
are conducted with zero notional lateral load, a conservative Committee validation and checking of the AISC (2005a)
estimate of 1.119 of the design loading still is obtained provisions. The interior columns in this structure are all
for the maximum frame resistance. leaner columns. All the beam-to-column connections are
Although the above effective length method checks are simple except for the connections to the exterior columns,
conservative for the in-plane strength of the columns of which are fully-restrained. The exterior columns are
Maleck’s frame, the moments for checking the out-of- subjected to relatively light axial loads, whereas they
plane strength of these members as well as the design of experience substantial gravity load moments in addition
the end-plate connections and column bases are dramatically to the wind load moments. Also, the frame has significant
underestimated. Figure 19 shows that the maximum second-order effects, i.e., amplification of the member
moment at the top of the left-hand column is only 54.6 sidesway internal bending moments. The columns are
kN-m when the beam-column interaction equation is braced in the out-of-plane direction at their base and at
intersected by the force-point trace of the effective length the roof height. Simple base conditions are assumed in
method using zero notional load. However, this moment both the in-plane and out-of-plane directions, and simple
is 186.9 kN-m at the distributed plasticity analysis limit connections are assumed in the out-of-plane direction at
load (242% larger). The minimum lateral load requirement the column tops. This problem is considered using the
in the AISC (2005a) effective length method increases the following LRFD load combinations:
predicted value to 100 kN-m when the beam-column Load Case 1 (LC1): 1.2D + 1.6S
strength condition is reached. This moment is also Load Case 2 (LC2): 1.2D + 0.5S + 1.6W
significantly smaller than the moment predicted by the Figures 21 and 22 show the applied fraction of the
distributed plasticity analysis. However, in this case, the design loads versus the story drift by the distributed
conservatism of the in-plane strength check prevents the plasticity analysis, direct elastic-plastic hinge analysis and
beam-columns from reaching a state in which their in- effective length methods for these load combinations. The
plane bending moments increase substantially with only a distributed plasticity analysis gives a maximum in-plane
minor increase in the applied load. The force-point trace capacity of 1.13 times the factored design load level for
for the AISC (2005a) effective length method is a reasonable LC1 and 1.06 of the factored design load level for LC2.
approximation of the result predicted by the distributed Each of the figures shows two curves determined from a
plasticity analysis up to the point at which its in-plane direct elastic-plastic hinge analysis, one in which a
beam-column strength condition is reached. The direct stiffness reduction factor (SRF) of only 0.9 (the same as
elastic-plastic hinge analysis gives a conservative estimate the value used in the distributed plasticity analysis) is
of the connection moment at the top of the left-hand employed, and the other in which a SRF of 0.8 (derived
column of 231.7 kN-m at its limit load (24% larger than from the AISC (2005a) direct analysis provisions) is
the value from the distributed plasticity analysis). However, employed. For LC2, the direct analysis matches the
Stability Analysis and Design of Steel Building Frames: The AISC (2005) Specification and Beyond 87

Surovek-Maleck and White 2004)).


The load-drift curves for the effective length method
analysis indicate a stiffer load deflection response. These
curves end at the load level at which the right-hand
exterior beam-column force-point trace intersects the
effective length method strength envelope. As indicated
in the previous example, the original AISC effective length
method (no notional minimum lateral load) tends to be
conservative for beam-columns in which a large fraction
of the bending moments are associated with non-sway
gravity loading. To isolate the conservatism associated with
the original effective length calculations, the effective
length solution for LC1 is conducted without including
Design load fraction versus the story drift for
the minimum notional load required by AISC (2005a).
Figure 21.

DP-13 load case 1 (LC1). Since LC2 includes a nonzero lateral load, there is no
additional notional lateral load requirement for this load
case in AISC (2005a). The original effective length
procedure, which does not include any notional lateral
load, suggests that only 0.78 and 0.82 of the design loads
can be applied to the DP-13 frame for LC1 and LC2
respectively. That is, the effective length method gives a
capacity of only 0.78/1.13 = 0.69 and 0.82/1.06 = 0.77 of
the in-plane capacity from the distributed plasticity
analyses for LC1 and LC2. If the AISC (2005a) minimum
lateral load is included for LC1, the predicted capacity is
reduced only a slight additional amount to 0.76 of the
design load level.
Table 3 summarizes the fractions of the design loads
giving a value of 1.0 for the in-plane strength check from
Eqs. (5) for the AISC (2005a) direct (elastic) and effective
Figure 22.Design load fraction versus the story drift for length methods, as well as the maximum capacities
DP-13 load case 2 (LC2). predicted by the direct elastic-plastic hinge analysis using
Mastan2, and the distributed plasticity analysis using GT-
Sabre. The direct (elastic) analysis method gives a slightly
distributed plasticity solution closely when the larger SRF smaller maximum strength than the direct elastic-plastic
of 0.9 is employed. However, for LC1, the direct analysis hinge analysis using Mastan2 for LC1 (1.17 versus 1.19).
over-predicts the maximum load capacity by 12% when This is due to the slightly more liberal cross-section plastic
the SRF is taken as 0.9. Use of the recommended SRF of hinge strength assumed in Mastan2 compared to Eqs. (5).
0.8 results in a predicted maximum load capacity of 1.19 For LC2, the direct (elastic) analysis force-point trace
of the design load level for LC1 (5% larger than the reaches Eqs. (5) in the right-hand exterior column at a
distributed plasticity analysis solution, which is equal to design load fraction of 0.95. This value is comparable to
the acceptable tolerance on the unconservative error the first plastic hinge formation at the top of the right-
established in the original development of the LRFD hand column at 0.96 of the design loading in Mastan2.
beam-column strength interaction equations (ASCE 1997; However, as indicated in Table 3 and illustrated by the
Table 3. Applied fraction of the design loads giving a value of 1.0 for the in-plane strength check from Eqs. (5) for the
AISC (2005a) direct (elastic) and effective length methods with and without minimum lateral loads, maximum capacities
predicted by direct elastic-plastic hinge analysis using Mastan2, and maximum capacities predicted by distributed plasticity
analysis using GT-Sabre, DP-13 LC1
LC1 LC2
Direct (elastic) analysis 1.17 0.95
AISC (2005a) effective length 0.76 0.82
Effective length with zero N
i 0.78 0.82
Direct elastic-plastic hinge analysis first-hinge strength, Mastan2 1.19 0.96
Direct elastic-plastic hinge analysis limit load, Mastan2 1.19 1.02
Distributed plasticity analysis 1.13 1.06
88 Donald W. White et al.

Figure 23. Plastic hinges and deflected shape from direct elastic-plastic hinge analysis and effective moment of inertia Ie
at the distributed plasticity analysis limit load, DP-13 LC2.

Figure 24. Plastic hinges and deflected shape from direct elastic-plastic hinge analysis and effective moment of inertia Ie
at the distributed plasticity analysis limit load, DP-13 LC1.

direct elastic-plastic hinge load-drift curve in Fig. 22, the


maximum capacity of the frame is not reached under the
LC2 loading until a design load fraction of 1.02 by the
direct elastic-plastic hinge analysis (versus 1.06 by the
distributed plasticity solution).
Figure 23 illustrates the behavior at the strength limit
predicted by the direct elastic-plastic hinge and distributed
plasticity solutions for this load combination. The effective
moment of inertia Ie is reduced to zero at the top of the
right-hand column at the limit load in the distributed
plasticity analysis. Also, the bending rigidities are
significantly reduced in the positive bending region of the
left-most girder and at the top of the left-hand column at
this load level. However, significant portions of these
cross-sections are still elastic. The direct-elastic plastic Figure Force-point traces for the right-hand beam-
25.

hinge analysis predicts the formation of a sway mechanism, column by the original effective length method, the direct
by plastic hinging at the tops of each of the exterior analysis method and the distributed plasticity analysis
method versus the effective length and direct analysis
columns, at its predicted capacity of 1.02 of LC2. strength curves, DP-13 LC1.
Figure 24 shows the corresponding predictions for
LC1. In this case, no plastic hinges have fully formed at
any location in the frame (i.e, Ie/I > 0 everywhere) at the Fig. 25. This figure compares the force-point trace at the
predicted capacity of 1.13 of the design loading in the top of the right-hand column predicted by the original
distributed plasticity analysis. The direct elastic-plastic effective length method (no notional minimum lateral
hinge solution reaches its limit load (at 1.19 of the design load) and the AISC (2005a) direct (elastic) analysis
loading) when a hinge forms at the top of the right-hand method to the corresponding force-point trace from the
column. distributed plasticity analysis for LC1. The AISC (2005a)
The key reason for the underestimation of the load strength envelopes (Eqs. (5)) are shown as dashed curves
capacities by the effective length method is illustrated in in the figure. The significant in-plane stability effects in
Stability Analysis and Design of Steel Building Frames: The AISC (2005) Specification and Beyond 89

this example result in a P of only 0.098P for LC1.


n(KL) y obtain the same result as determined by analyzing a
The corresponding effective length factor for the right- member or structure with nominal E and F and y

hand column is 5.02 using a rigorous sidesway buckling subsequently factoring the abscissa and ordinate of the
analysis (comparable values are obtained using the resulting strength curves by φ and φ . The appropriate
b c

equations in the AISC (2005a) Commentary, as long as nominal residual stress, geometric imperfection, and
the leaner column effects and the effect of the simple material stress-strain idealizations for use of distributed
connections at the inside ends of the exterior girders are plasticity analysis in a design context are discussed. Also,
correctly incorporated). a simple limit on the magnitude of the axial force is
The direct analysis method gives a substantially larger provided within which member out-of-straightness may
estimate of the in-plane resistance because it focuses on be neglected in a distributed plasticity analysis.
a more realistic estimate of the internal moments and the Secondly, three specific elastic methods of analysis and
corresponding member resistances. In determining the design detailed in AISC (2005a) are discussed in the
anchor point P for its beam-column strength envelope,
n(KL) context of the above refined inelastic analysis and design
the effective length method overemphasizes the response procedure. These are the direct analysis, the effective
of the structure to idealized loads causing uniform axial length and the first-order analysis methods. The new
compression in all of the columns. The physical strength AISC (2005a) direct analysis method provides a better
of the DP-13 frame is dominated by its amplified internal approximation of the results from refined distributed
moments, a large fraction of which is not related to the plasticity analysis. By improving the accuracy of the
sidesway of the structure. The limit of the DP-13 frame analysis, the direct analysis method allows for greater
resistance is tied essentially to the flexural resistance of simplicity in the design checks by eliminating the need
the right-hand beam-column and the amplified internal for effective length factors. Also, a more accurate
moments in this member, not by a column failure under characterization of the overall strength is attained in
concentrically-applied axial loads. general.
When the effective length method is applied to the DP- Thirdly, the paper presents a basic extension of the
13 frame, the design resistance is governed by the in- direct (elastic) analysis method that satisfies the intent of
plane strength of the leeward beam-column (Eqs. (5)) for the Specification and captures the beneficial effects of
both LC1 and LC2 (rather than the out-of-plane strength inelastic redistribution from compact adequately-braced
as represented by Eq. (6)). However, the out-of-plane members after the individual member resistances are
strength based on Eq. (6) governs slightly relative to the reached. This extension, termed the direct elastic-plastic
in-plane strength in both the distributed plasticity analysis hinge analysis method, is obtained simply by approximating
as well as the direct analysis methods. Eq. (6) is intersected the response of these types of members by an elastic-
by the direct (elastic) analysis force-point trace for the perfectly plastic hinge model at the Specification limit of
right-hand column at 1.16 and 0.94 of LC1 and LC2 the member resistance.
respectively. The right-hand column does not satisfy the Lastly, four examples are provided that illustrate key
AISC (2005a) unbraced length requirements for inelastic characteristics of each of the above analysis and design
design using braces only at its ends. Additional out-of-plane methods. The direct elastic and elastic-plastic hinge analysis
bracing would be required to satisfy these requirements methods provide accurate elastic and inelastic approximations
such that the inelastic reserve strength illustrated by Fig. of the refined distributed plasticity solutions. The effective
22 can be utilized based on the AISC requirements. length analysis and design method is shown to give
accurate to conservative solutions for the in-plane beam-
4. Conclusions column resistances in all of the example cases. However,
the effective length method significantly underestimates
This paper explains the AISC (2005a) stability analysis the internal moments that the beam, connection, and out-
and design provisions with an emphasis on their relationship of-plane beam-column resistances must accommodate in
to refined second-order inelastic analysis methods and on certain cases. Furthermore, the effective length method is
how they can facilitate the use of second-order inelastic shown to give conservative solutions for beam-columns
analysis in practical design. in sway frames where a large percentage of the moments
First, an appropriate application of second-order distributed are due to non-sway gravity loading. AISC (2005a)
plasticity analysis is discussed for refined assessment of requires the use of a notional minimum lateral load in
adequately braced compact I-section members and structural gravity-only load combinations, and limits the application
systems. The base AISC (2005a) beam-column strength of the effective length method to frames having ∆2nd/∆1st
interaction equations have been developed in part by < 1.5 to control these errors. The AISC (2005a) direct
calibration against the results from refined inelastic elastic analysis method is generally applicable to all types
analyses of this type. The paper recommends that for of frame structures. The direct elastic-plastic hinge method
distributed plasticity analysis, both the material elastic is a useful extension of this method that facilitates
stiffnesses (E) and strengths (F ) should be reduced up
y simplified inelastic analysis and design.
front by the factors φb = φc = 0.9. This is necessary to
90 Donald W. White et al.

Acknowledgments 5811-5831.
CEN (2003). “Eurocode 3: Design of Steel Structures - Part
Many of the concepts discussed in this paper have 1-1: General Rules and Rules for Buildings,” Final Draft
benefited greatly from discussions with the members of prEN 1993-1-1:2003 E, European Committee for
the AISC Technical Committee 10 (TC10), the Structural Standardization.
Stability Research Council (SSRC) Task Group 4 on Chang, C.-J. (2005). GT-Sabre User Manual, School of Civil
Frames, and the former SSRC Task Group 29 on Inelastic and Environmental Engineering, Georgia Institute of
Analysis for Frame Design. Professor J. Yura of the Technology, Atlanta, GA.
Davies, J.M. and Brown, B.A. (1996). “Plastic Design to BS
University of Texas at Austin, Dr. S. Nair of Teng and 5950,” The Steel Construction Institute, Blackwell
Associates, Inc. and Prof. G.G. Deierlien of Stanford Science, U.K., pp. 326.
University are thanked for their guidance of the TC10 Deierlein, G. (2003). “Background and Illustrative Examples
efforts toward the implementation of the AISC (2005a) on Proposed Direct Analysis Method for Stablity Design
provisions. Various portions of this work were funded by of Moment Frames,” Report on behalf of AISC TC10,
the National Science Foundation, the Georgia Institute of July 13 2003, pp. 17.
Technology, the American Society of Civil Engineers, Deierlein, G. (2004). “Stable Improvements: Direct Analysis
and the Metal Building Manufacturers Association. The Method for Stability Design of Steel-Framed Buildings,”
funding by these organizations is gratefully acknowledged. Structural Engineer, November, pp. 24-28.
The opinions, findings and conclusions expressed in this Galambos, T.V. and Ketter, R.L. (1959). “Columns Under
paper are the authors’ and do not necessarily reflect the Combined Bending and Thrust,” Journal of the
views of the above individuals, groups and organizations. Engineering Mechanics Division, ASCE, 85(EM2), pp.
135-152.
Izzudin, B.A., and Smith, D.L, (1996) “Large-Displacement
References Analysis of Elasto-plastic Thin-Walled Frames.
I:Formulation and Implementation”, Journal of structural
AISC (2005a). Specification for Structural Steel Buildings, Engineering, Vol. 122, No. 8, pp. 905-914
American Institute of Steel Construction, Chicago, IL. King, C. (2001). “In-Plane Stability of Portal Frames to BS
AISC (2005b). Code of Standard Practice for Steel Buildings 5950-1:2000,” SCI Publication P292, The Steel Construction
and Bridges, American Institute of Steel Construction, Institute, Berkshire, U.K., pp. 213.
Inc., Chicago, IL. Kuchenbecker, G.H., White, D.W. and Surovek-Maleck,
AISC (1999). Load and Resistance Factor Design A.E. (2004). “Simplified Design of Building Frames
Specification for Structural Steel Buildings, American using First-Order Analysis and K = 1.0,” Proceedings,
Institute of Steel Construction, Chicago, IL. SSRC Annual Technical Sessions, April, pp. 20.
AISC (1989). Specification for Structural Steel Buildings: LeMessurier, W. J. (1977). “A Practical Method of Second
Allowable Stress Design and Plastic Design, 9 Ed., th
Order Analysis. Part 2: Rigid Frames,” Engineering
American Institute of Steel Construction, Chicago, IL. Journal, AISC, 13(4), pp. 89-96.
AISC (1986). Load and Resistance Factor Design Maleck, A.E. and White, D.W. (2003). “Direct Analysis
Specification for Structural Steel Buildings, American Approach for the Assessment of Frame Stability:
Institute of Steel Construction, Chicago, IL. Verification Studies,” Proceedings, SSRC Annual
AISC (1969). Specification for the Design, Fabrication and Technical Sessions, pp. 18.
Erection of Structural Steel for Buildings, American Maleck, A.E. (2001). Second-Order Inelastic and Modified
Institute of Steel Construction, Chicago, IL. Elastic Analysis and Design Evaluation of Planar Steel
AISC (1961). Specification for the Design, Fabrication and Frames, Ph.D. Dissertation, Georgia Institute of
Erection of Structural Steel for Buildings, American Technology, pp. 579.
Institute of Steel Construction, New York, NY. Martinez-Garcia, J.M. (2002). “Benchmark Studies to
Alemdar, B.N. and White, D.W (2005), “Displacement, Evaluate New Provisions for Frame Stability Using
Flexibility and Mixed Beam-Column Finite-Element Second-Order Analysis,” M.S. Thesis, School of Civil
Formulations for Distributed Plasticity Analysis,” Journal Engineering, Bucknell Univ., pp. 241.
of Structural Engineering, ASCE, 131(12), pp. 1811- McGuire, W. (1995). “Inelastic Analysis and Design in
1819. Steel, A Critique,” Restructuring America and Beyond,
ASCE (2005). Minimum Design Loads for Buildings and Proceedings of Structures Congress XIII, M. Sanayei
Other Structures, SEI/ASCE 7-05, ASCE, Reston, VA. (ed.), ASCE, pp. 1829-1832.
ASCE (1997). Effective Length and Notional Load Approaches McGuire, W., Gallagher, R.H. and Ziemian, R.D. (2000).
for Assessing Frame Stability: Implications for American Matrix Structural Analysis, with Mastan2, Wiley, New
Steel Design, Task Committee on Effective Length, York.
Technical Committee on Load and Resistance Factor Nair, R.S. (2005a). “Stability and Analysis Provisions of the
Design, Structural Engineering Institute, American 2005 AISC Specification for Steel Buildings,”
Society of Civil Engineers, pp. 442. Proceedings, Structures Congress 2005, ASCE, pp. 3.
Battini, J.-M. and Pacoste, C. (2002). “Plastic Instability of Nair, R.S. (2005b), “Stability and Analysis,” Modern Steel
Beam Structures Using Co-rotational Elements,” Computer Construction, September 2005.
Methods in Applied Mechanics and Engineering, 191, pp. Nukala, P.K.V.V. and White, D.W. (2004). “A Mixed Finite
Stability Analysis and Design of Steel Building Frames: The AISC (2005) Specification and Beyond 91

Element Formulation for Three-Dimensional Nonlinear Systems,” Engineering Journal, AISC, to appear.
Analysis of Frames,” Computer Methods in Applied White, D.W., Surovek-Maleck, A.E. and Chang, C.-J.
Mechanics and Engineering, 193, pp. 2507-2545. (2005b). “Direct Analysis and Design Using Amplified
Pi, Y.L. and Trahair N.S. (1994), “Nonlinear Inelastic First-Order Analysis, Part 2 - Moment Frames and
Analysis of Steel Beam-columns Theory.”, J. Struct. General Rectangular Framing Systems,” Engineering
Engrg., ASCE, 120(7), pp. 2041-2061. Journal, AISC, to appear.
SAA (1990), Steel Structures, AS4100-1990, Standards White, D.W. and Kim, Y.D. (2006). “A Prototype
Association of Australia, Australian Institute of Steel Application of the AISC (2005) Stability Analysis and
Construction, Sydney, Australia. Design Provisions to Metal Building Structural Systems,”
Salmon, C.G. and Johnson, J.E. (1996). Steel Structures, Report to Metal Building Manufacturers Association,
Design and Behavior, 4th Ed., Prentice Hall, NJ, pp. 1024. January 2006, pp. 156.
Springfield, J. (1991). “Limits on Second-Order Elastic White, D.W., Surovek-Maleck, A.E. and Kim, S.-C. (2003).
Analysis,” Proceedings, Annual Technical Session, “Direct Analysis and Design Using Amplified First-Order
Structural Stability Research Council, Chicago, IL, 89-99. Analysis, Part 1 - Combined Braced and Gravity Framing
Surovek-Maleck, A.E., White, D.W. and Leon, R.T. (2005). Systems,” Structural Engineering, Mechanics and Materials
“Direct Analysis for Design Evaluation of Partially- Report No. 42, School of Civil and Environmental
Restrained Steel Framing Systems,” Journal of Structural Engineering, Georgia Institute of Technology, Atlanta,
Engineering, ASCE, to appear. GA., pp. 34.
Surovek-Maleck A.E. and White, D.W. (2004). “Alternative White, D.W. and Nukala, P.K.V.V. (1997). “Recent
Approaches for Elastic Analysis and Design of Steel Advances in Methods for Inelastic Frame Analysis:
Frames. I: Overview,” Journal of Structural Engineering, Implicatinos for Design and a Look Toward the Future,”
ASCE, 130(8), pp. 1186-1196. Proceedings, National Steel Construction Conference,
Surovek-Maleck, A. and White, D.W. (2003). “Direct American Institute of Steel Construction, pp. 43-1 to 45-
Analysis Approach for the Assessment of Frame 24.
Stability: Verification Studies,” Proceedings, SSRC White, D.W. and Chen, W.F (1993). Plastic Hinge Based
Annual Technical Sessions, pp. 18. Methods for Advanced Analysis and Design of Steel
Teh, L., and Clarke, M.J., (1998), “Plastic-Zone Analysis of Frames - An Assessment of the State-of-the-Art, Structural
3D Steel Frames Using Beam Elements”, Journal of Stability Research Council, University of Missouri-Rolla,
Structural Engineering, Vol. 125, No. 11, pp. 1328-1337 Rolla, MO, pp. 299.
White, D.W., Surovek-Maleck, A.E. and Kim, S.-C. (2005a). Ziemian, R.D., McGuire, W. and Deierlein, G.G. (1992).
“Direct Analysis and Design Using Amplified First-Order “Inelastic Limit States Design. Part I: Planar Frame
Analysis, Part 1 - Combined Braced and Gravity Framing Studies,” 118(9), pp. 2532-2549.

You might also like