Вы находитесь на странице: 1из 61

872 STABILITY UNDER SEISMIC LOADING

has occurred, similar to the situation for beams in chevron bracing (Fig. 19.16d ).
This represents an undesirable stability condition and such bracing configuration is
not permitted when high-ductility demand is expected.

19.2.4 Buckling Restrained Braced Frames (BRBFs)


Buckling restrained braced (BRB) frames are concentrically braced frames that are
built with braces that are specially designed to yield in both tension and com-
pression without buckling. A typical BRB member is illustrated in Fig. 19.17.
A steel core plate with a reduced cross section is inserted in a steel tube. The
volume between the tube and the core is filled with mortar or concrete. At both
ends, the core plate extends outside of the tube and is connected to the framework.
The core is covered with an unbonding material prior to filling the tube such that the
core can axially deform freely and, hence, axial loading on the brace is essentially
resisted by the core. The unsupported core projections at both ends of the brace
are typically made larger in size and/or stiffened to avoid buckling in compression.
Under seismic loading, yielding is expected to take place in both compression and
tension in the reduced segment of the core, Lc in Fig. 19.17. A small gap must be
left between the concrete fill and the core to allow for transverse expansion of the
core upon compression yielding.
Main stability issues for BRB frames include local buckling (ripping) of the
brace core, overall brace buckling, buckling at the end connections, and buckling
of the beams and columns. Design for stability of gusset plates, beams, and columns
is very much the same as that for CBFs and only the main differences are discussed
herein. In most countries, BRB members are proprietary products that are devel-
oped and supplied by specialized manufacturers to comply with project-specific
performance requirements determined by the design engineer (e.g., minimum and
maximum axial strength, cyclic deformation capacities, minimum axial stiffness,
etc.). Local buckling and overall brace stability are therefore aspects that are gener-
ally addressed by the suppliers using techniques and methods specially developed
for their products. Because a large variety of BRB systems have been proposed and
used in the last two decades, only the main concepts and recent research findings
that are common to most systems will be introduced using the simple configuration
shown in Fig. 19.17. Additional information on the system and on seismic design

Lc A B C
Splice Unbonding
Steel core plate Stiffener
Material
(typ.) (typ.) Splice
Steel plate
Core (typ.)
Outer Steel Tube Mortar
Fill
Outer
Steel Tube
Section A Section B Section C

FIGURE 19.17 Buckling restrained bracing member.


DESIGN FOR LOCAL AND MEMBER STABILITY 873

procedures can be found in Uang and Nakashima (2004), Sabelli (2004), López and
Sabelli (2004), and Xie (2005). Among the codes examined, buckling restrained
braced frames have been introduced only in the AISC seismic provisions and only
these provisions are referenced in the discussion.

Brace Local and Global Buckling The axial load deformation P−δ hysteretic
response of a buckling restrained brace is illustrated in Fig. 19.18a. Figure 19.18b
shows a core of rectangular cross section of a BRB member after testing. Evidence
of local buckling and friction response between the core and the buckling restraining
mechanism can be observed. Local buckling of the core as constrained by the buck-
ling restraining mechanism is schematically illustrated in Fig. 19.18c. The inelastic
cyclic response is stable and nearly symmetrical in tension and compression and
shows marked strain-hardening behavior. The peak tension and compression forces
are measured in cyclic tests at maximum anticipated deformations. The force Tmax
is defined as ωPyc , where Pyc is the yield load of the core (= Ac Fyc ) and ω is
a factor that essentially accounts for strain hardening (ω ≈ 1.2 to 1.4, depending
on the steel material and the imposed deformation history). The load Cmax equals
βωPyc , where β mainly accounts for the longitudinal friction that develops between
the core and the restraining system when pushing the locally buckled core into the
brace (β ≈ 1.05 to 1.2).
The minimum stiffness per unit length γmin of the encasing material can be
obtained by comparing the maximum anticipated compression√ force to the buck-
ling load of a beam on an elastic foundation: Pcr = 2 γmin Et Ic , where Et is the
P/Py

d/dy
(a) (b)

Gap Concrete Fill


P P EIr, dr

P P
P P a0 Core
L

(c) (d )

FIGURE 19.18 Buckling restrained bracing member: (a) hysteretic response; (b) evidence
of core local buckling after cyclic testing; (c) core local buckling; and (d ) overall brace
buckling.
874 STABILITY UNDER SEISMIC LOADING

tangent modulus of steel (Et ≈ 0.02E ) and Ic is the moment of inertia of the core
in the direction of interest (Black et al., 2004). The stiffness of the encasing mate-
rial must be reduced to account for the local transverse flexibility of the member
containing the material (e.g., walls of tubes) and of the unbonding material used
between the core and the restraining system, if any. The number of contact points
between the core and the restraining system increases (wavelength reduces) as the
axial compression load or deformation is increased. The magnitude of the contact
forces or pressures between the core and the restraining system, normal to the
longitudinal brace axis, depends on a number of factors including the brace axial
load (or axial deformation), the spacing of the local buckles (local mode shape), the
initial gap left between the core and the restraining mechanism, and the actual local
stiffness of the restraining system. The amplitude of the friction forces depends on
the contact forces and conditions at the interface. Low-friction, unbonding material
can be used at the interface to ease sliding of the core. Only limited information
has been published on this complex interaction between the brace core and the
restraining mechanism. Furthermore, the available data often pertain to particular
BRB systems or details and the findings may or may not apply directly to other
designs. Iwata et al. (2000) conducted tests on BRB members with the same elastic
buckling-to-yield-strength ratio, Per /Pyc (see definition below) but different buck-
ling restraining systems. The absence of an unbonding material led to less uniform
strain demand and tensile fracture of the core. Failure of the restraining mech-
anism due to the outward forces induced by core local buckling was observed.
Usami et al. (2003) developed a model to predict the buckling response of a steel
core incorporating the flexibility of the unbonding material under monotonic and
cyclic loading conditions. Iwata and Murai (2006) examined the influence of the
core thickness and core width-to-thickness ratio together with the Per /Pyc ratio on
the local buckling response of the core and the cumulative ductility and the nor-
malized plastic strain energy capacity of the brace. No clear trend could be found
for the influence of the core geometry except that greater deformation capacity
was achieved by specimens having the thickest core or the core with the small-
est width-to-thickness ratio when compared to specimens with otherwise identical
parameters. The positive impact of increasing the thickness of square restrainer
tubes on the buckling behavior of the core was studied analytically and experimen-
tally by Matsui et al. (2008) and through testing by Ju et al. (2009). Tests by Ding
et al. (2009) showed that the normal forces are reduced when keeping the gap thick-
ness to a minimum. Wei and Tsai (2008) carried out tests and developed equations
to characterize the interaction between strength, stiffness, and thickness of the con-
crete fill material and the width-to-thickness ratio of the core plate in concrete-filled
BRB members. It was concluded that the concrete thickness must be increased if
the strength of the concrete is reduced or the width-to-thickness ratio of the core is
increased.
The contribution of the concrete fill is typically ignored when examining the
overall buckling response of BRB members. Assuming that the axial load in the
restraining system is low enough to avoid inelastic response, the overall buckling
DESIGN FOR LOCAL AND MEMBER STABILITY 875

strength of the system, Per , is given by

π 2 EIr
Per = (19.19)
L2

in which Ir is the moment of inertia of the restraining system and L is the length
of the brace. In past studies, the latter has generally been taken equal to the length
of the restraining mechanism (Fig. 19.18d ) or the overall brace dimensions includ-
ing end connections. As a minimum, the load Per must be greater than the load
Cmax ; in addition, imperfections and friction response will induce axial stresses
in the restraining system that must be limited to prevent yielding. Neglecting the
contribution of the concrete fill, the maximum bending stress σbr due to an initial
out-of-straightness ao under the load Cmax in a restraining system with symmetrical
cross-section depth dr and moment of inertia Ir can be obtained from (Inoue et al.,
2004; Wada and Nakashima, 2004)

dr Cmax ao
σbr = Mr with Mr = (19.20)
2Ir 1 − Cmax /Per

With Cmax = βωPyc , Eq. 19.20 may be rewritten as


  
Per π 2 E  ao  dr
≥ βω 1 + (19.21)
Pyc 2σbr L L

If the axial stress in the restraining system needs to be limited to φFyr , the stress
σbr in Eq. 19.21 could be taken equal to Fyr [φ − ω(β − 1)Pyc /Pyr ], where Pyr is
the yield axial load of the restraining system. The second term in the brackets
accounts for the additional axial stress induced by frictional response. For Fyr =
300 MPa, φ = 0.9, ω = 1.2, β = 1.1, ao /L = 0.001, dr /L = 0.03, Pyc /Pyr = 0.75,
and σbr = 0.81Fyr , a minimum ratio Per /Pyc = 1.48 is required to achieve the
desired behavior. Usami et al. (2008) suggested that the influence of the thickness
of the gap and axial load eccentricity could also be considered by replacing ao in
Eq. 19.21 by the sum of these geometrical dimensions. Early tests by Watanabe et
al. (1988) showed the influence of the Per /Pyc ratio on the performance of BRB
members and they suggested that a minimum ratio of 1.5 should be used in design.
Iwata and Murai (2006) found that the cumulative ductility and the normalized
plastic strain energy capacity of BRB members up to the point where their compres-
sion resistance started to drop both increased linearly with the Per /Pyc ratio. They
suggested that these relationships could be used in design by selecting a restraining
mechanism capable of achieving the expected demand from earthquakes. In their
test program, one of the specimens had Per /Pyc < 1.0 (0.9) and that specimen
failed by global buckling early in the test. Four other specimens had the same core
plate thickness but with a higher Per /Pyc ratio (1.6, 2.3, 3.2, and 4.5) obtained
by increasing the depth of the restraining mechanism. These specimens failed by
876 STABILITY UNDER SEISMIC LOADING

local buckling of the core plate after several inelastic cycles at large axial defor-
mations. The increase in local stiffness of the restraining system resulting from
the increase in depth may also have contributed to the observed improvement in
inelastic performance with Per /Pyc .

Buckling at End Connections and of Beams and Columns In typical BRB


design, the portion of the core protruding at the end of the restraining member has
a larger cross section compared to that of the core. Axial yielding is then confined
to the reduced segment of the core. When flat core plates are used, stiffeners can
be added to prevent buckling of the protruding core portion. The portion of a core
with a cruciform cross section that extends outside of the restraining mechanism
may buckle in torsion when yielding in compression. Based on incremental theory
of plasticity, Black et al. (2004) developed an equation for the critical stress,
 
Et π 2 bc2 3Fyc tc2
σcr = 2
+1+ (19.22)
3 3 l Et bc2

where tc and bc are the thickness and width, respectively, of each of the four flanges
of the core and l is the maximum anticipated protruding length of the steel core.
Unless specifically detailed to accommodate rotations, the connections of BRB
members to gusset plates are generally fixed and flexural demand is expected on
the bracing members when the frame deforms laterally (López et al., 2004; Tsai and
Hsiao, 2008). This may cause inelastic rotation demand to concentrate in the steel
core extension, just outside of the restraining system, and BRB members must be
detailed to prevent this type of instability. In AISC seismic provisions, qualifying
cyclic testing is required for BRB members and subassemblage tests on specimens
with connections to the framing members are specified to investigate the response
of BRB members including this rotation demand.
Out-of-plane instability has been observed at the gusset-to-brace connections
in past BRB tests by Tsai and Hsiao (2008) and Ma et al. (2008). In the former
study, gusset plates buckled in the early phase of the test program. The test was
halted to allow the installation of stiffeners at the free edges of the gusset plates
so that the tests could then be completed without further connection instability.
The gusset plates had been initially designed for compression using the method
proposed by Thornton that was described earlier for CBFs, which assumed an
effective length equal to 0.65 times the maximum of the lengths l1 , l2 , l3 (see
Fig. 19.13g). Experimental observations (buckled shapes and buckling loads) indi-
cated that an effective-length factor of 2.0 would have been more appropriate in the
absence of free edge stiffeners. Koetaka et al. (2008) developed design criteria to
prevent out-of-plane buckling of BRB members at the central connection of V and
inverted-V bracing. The method accounts for the lateral stiffness of the connection
as well as for the beam torsional stiffness. Cyclic tests were performed to validate
the proposed procedure. For the same connection, Chou and Chen (2009) recom-
mended that the BRB core terminations be extended beyond the gusset bending
DESIGN FOR LOCAL AND MEMBER STABILITY 877

line and that stiffeners be used along gusset free edges. Chou and Chen further
provided guidelines for the design of the stiffeners.
Similar to CBFs, beams and columns must be designed to carry the tributary
gravity loads together with lateral load effects corresponding to the development
of Cmax and Tmax in the braces. In V or inverted-V bracing, the beams must be
designed such that the full brace capacities can be reached at small deflections.
As was the case for CBFs, inelastic flexural demand is expected in multistory
BRB frame columns (Sabelli et al., 2003) and it is required in the AISC seismic
provisions that the stringent seismic compactness requirements be satisfied for these
components. In the study by Richards (2009), the axial load demand from the
amplified seismic loads, as prescribed in U.S. codes, underestimates the expected
seismic demand in the uppermost floors of BRB frames. In this case, Richards
indicates that this may have a limited impact on response as top-story columns
typically possess excess strength capacity.

19.2.5 Eccentrically Braced Steel Frames (EBFs)


Inelastic response in EBFs is constrained to ductile link beam segments created at
the beam ends, next to the columns when using single-diagonal members, or at the
beam midspans when a chevron bracing configuration is employed (Fig. 19.19a).
Yielding in the beam link may develop in the form of inelastic shear deformations
in the web of the beam, plastic rotation at the ends of the links, or a combination
thereof, depending on the length of the link e relative to the ratio of the link plastic
moment capacity Mp to the link plastic shear capacity Vp . The first two inelastic
deformation modes are illustrated in Fig. 19.4e. The plastic deformation imposed
on link beam segments is defined by the plastic rotation γ shown in Fig. 19.19a.
This plastic rotation depends on the link length relative to the bay width. Therefore,
by selecting the link length e and a beam section with capacities Mp and Vp , the
designer can control both the type of yielding mechanism (shear or flexure) and the
amount of plastic deformation imposed on the links for a given inelastic story drift
θp . Buckling of the web is more critical in shorter, shear-dominated links whereas
flange buckling is more an issue in longer, flexure-critical links. Both phenomena
are accentuated when γ is increased. Local buckling of the link beam segments
is discussed in the following section, together with lateral stability requirements
for the beam segments. Upon yielding, the link segments impose forces to the
beam segments outside of the links, the bracing members, and the columns. These
structural elements must be designed to withstand these forces combined with
gravity loading without instability failure modes, as will also be discussed in the
subsequent section. For simplicity, the description herein is limited to the chevron
bracing configuration with the link at the beam midspan. Additional information
on the seismic design of EBFs can be found in AISC (2006).

Buckling of Link Beam Segments In EBFs, the eccentricity introduced at the


brace-to-beam connections causes the axial loads from the braces to induce high
shear forces and reversed curvature bending moments in the link beam segments.
878 STABILITY UNDER SEISMIC LOADING

e e
Link (typ.) Link (typ.) gp = qp L /e
qp gp = q L /ep qp

gp (rad.)
L L e/(Mp /Vp)

(a) (c)

> 1.6 Mp / Vp
≤ 1.6 Mp / Vp > 2.6 Mp /Vp
≤ 2.6 Mp / Vp
s s s
M s s ≤ 5 Mp /Vp
V V

e
M
V
1.5 bf 1.5 bf 1.5 bf 1.5 bf

Intermediate Stiffeners > 5 Mp /Vp


M s 52 tw − d/5 End Stiffener (typ.)
30 tw − d/5

V = 2 M/e
gp (rad.) Lateral
0.02 0.08
Bracing (typ.)

(b) (d )

FIGURE 19.19 (a) Inelastic response of single-diagonal and chevron-type EBFs; (b)
seismic-induced shear forces and bending moments in links of symmetrical chevron EBFs;
(c) link plastic rotation capacity (AISC, 2005a); and (d ) link stiffeners and lateral bracing
requirements (AISC, 2005a).

An illustration of this loading condition imposed on a link specimen is shown


in Fig. 19.20a (Okazaki and Engelhardt, 2007), where a short link is yielding
essentially in shear. Buckling in shear is visible as this test specimen violated
the minimum stiffener spacing prescribed in current code provisions, which will
described later. Nevertheless, the performance of this link in Fig. 19.20b is very
stable, with slight strength degradation after reaching a total rotation of approxi-
mately 0.10 rad. Links deforming in shear can show significant strain hardening
with shear forces reaching 1.4Vp to 1.5Vp at large deformations. Longer links with
beams yielding in flexure behave in a manner similar to beams in moment-resisting
frame beam-to-column tests, with bending moments reaching 1.1Mp to 1.2Mp at
plastic hinges.
Extensive test programs and analytical studies were performed in the 1980s and
early 1990s to characterize the inelastic behavior of link beams and develop seismic
design methodologies for EBF systems (e.g., Hjelmstad and Popov, 1983; Malley
and Popov, 1984; Kasai and Popov, 1986a,b; Engelhardt and Popov, 1992). The
links were defined as short or long, depending on the ratio Mp /Vp . For links in
symmetrical chevron bracing, a point of zero moment exists at the link midlength
under seismic loads and M = V (e/2), as illustrated in Fig. 19.19b. The limiting
case where shear and bending moment reach simultaneously their respective plastic
DESIGN FOR LOCAL AND MEMBER STABILITY 879

(a) (b)

FIGURE 19.20 EBF link response. (a) Shear yielding and buckling in EBF link. (b)
Hysteretic response. Courtesy of T. Okazaki (University of Minnesota).

resistances Vp and Mp corresponds to e = 2Mp /Vp . Using the hardened capacities,


short links are defined as those with e ≤ 2 (1.2 Mp )/(1.5 Vp ) = 1.6Mp /Vp , whereas
long links have been defined as links by a length exceeding 2.6Mp /Vp (AISC,
CSA-S16) or 3.0 Mp /Vp (EC8, NZS3404). Short links subjected to inelastic cyclic
loading with increasing amplitude eventually develop web buckling, which in turn
results in strength degradation. Test programs, however, demonstrated that this
buckling response could be delayed by adding regularly spaced web stiffeners and
this technique has been adopted in link design: in most codes, the plastic rotation
capacity of short links is set equal to 0.08 rad, provided that web stiffeners are
used. The spacing of these intermediate stiffeners, s in Fig. 19.19d , depends on the
plastic rotation demand γp obtained from the anticipated plastic interstory drifts
θp (γp and θp are shown Fig. 19.19a). For instance, in the AISC, CSA-S16, and
EC8 seismic provisions, the stiffener spacing varies linearly from to 52tw − 0.2d to
30tw − 0.2d when γp increases from 0.02 to 0.08 rad, as illustrated in Fig. 19.19d
(tw and d are the link beam web thickness and depth, respectively). Slightly larger
spacing is permitted in NZS3404. Detailing requirements for the stiffeners are also
given in codes. Longer links yielding in flexure have lower rotation capacity and
the limit has been set to 0.02 rad in AISC, CSA-S16, and EC8 seismic codes and to
0.03 rad in NZS3404. The rotation capacity varies linearly for intermediate links, as
shown in Fig. 19.19c. As the shear inelastic deformation demand decreases when
increasing eVp /Mp , the need for web stiffeners becomes less critical and these
stiffeners are no longer required for long links. As shown in Fig. 19.19d , web
stiffeners, however, are still required at a distance of 1.5 times the beam flange
width from the link ends for intermediate and long links, with the requirement
1.6Mp /Vp < e ≤ 5.0Mp /Vp . These stiffeners control flange buckling in the region
where high rotational demand is anticipated. Flangebuckling is also controlled by
limiting the beam flange slenderness bf /2tf to 0.30 E /Fy .
As would be expected, the presence of an axial compressive load P in link beams
reduces their shear and moment resistances. It also represents a more critical sta-
bility condition and the length of the links must be limited to prevent instability
880 STABILITY UNDER SEISMIC LOADING

if P exceeds the link beam axial yield capacity AFy . The limit depends on the
ratio of the axial stress (P/A) to the web shear stress V /Aw in the link (with A =
link beam cross-sectional area and Aw = tw (d − 2tf ). When this ratio is less than
0.3, the length is limited to the short link limit e ≤ 1.6Mp /Vp . If the ratio exceeds
0.3, this limit is reduced further by multiplying it by [(1.15 − 0.5(P/A)/(V /Aw )].
These limits have been adopted in the 2005 AISC seismic provisions. As in
beam-columns, axial loading must be taken into account when verifying the link
web slenderness against compactness requirements.
Most of the experimental research supporting the code design requirements for
proper inelastic link response was carried out on I-shaped links made of ASTM
A36 steel with Fy = 248 MPa. Additional series of tests recently completed by
Okazaki et al. (2005) and Okazaki and Engelhardt (2007) showed that these code
rules still apply for links made of the higher strength steel ASTM A992 (Fy =
345 MPa) now commonly used in the United States for seismic applications. It
is recommended to terminate the stiffener to beam web welds at some distance
below the k -area of the beams to delay link web fracture in this region where steel
exhibits higher strength and lower toughness due to shape-straightening processes
typically employed during manufacturing. Chao et al. (2006) proposed to improve
the resistance against web fracture by replacing the intermediate web stiffeners by a
single horizontal stiffener placed at the link midheight between the end stiffeners.
This scheme has been verified through finite element analysis simulations that
include ductile fracture initiation modeling, but experimental validation has not yet
been carried out to check the ability of this stiffener to control web buckling. For
short links, the test programs by Okazaki et al. (2005) and Okazaki and Engelhardt
(2007), together with the analytical work by Richards  and Uang (2005), suggested
that the limit on bf /2tf could be extended to 0.38 E /Fy for short links and this
change was included in the 2005 AISC seismic provisions.
Pairs of web stiffeners are required at both ends of the link beams. Lateral
beam bracing is also required at both top and bottom flanges at the link ends.
Experimental work by Hjelmstad and Lee (1989) emphasized the necessity for
robust lateral bracing at the link ends. In the AISC seismic provisions, the bracing
must be designed for a force of 0.06 times the link beam expected plastic moment
divided by the distance between flange centroids. Minimum stiffness must also be
provided in order to limit the bracing deformations to 0.45% of e under the bracing
design loads. A similar minimum bracing strength level is prescribed in CSA-S16
and EC8 (0.06 times the expected flange yield strength). In NZS3404, top and
bottom lateral bracing must be provided atthe ends of the link, as well as along
the link, at a spacing not exceeding 1.06ry E /Fy . Such bracing must be designed
for 2.5% of the axial force in the link flange at the point of restraint and minimum
bracing stiffness must also be provided to limit the lateral beam flange displacement
to 4 mm under the design bracing force. Tests by Ricles and Popov (1989) showed
that lateral bracing of the link top flange can be provided by a composite floor, but
the slab itself is not sufficient to prevent lateral–torsional buckling of the beam and
independent bracing must be provided at the beam bottom flange at the link ends.
Transverse beams framing from one side at both ends of the link have been found
DESIGN FOR LOCAL AND MEMBER STABILITY 881

to provide effective lateral bracing. Berman and Bruneau (2008b) recently proposed
and experimentally validated link beams made of built-up rectangular hollow cross
sections that do not require lateral bracing. This could represent a very attractive
alternative for cases when lateral bracing is difficult to provide, as would be the case
along an exterior column line, or adjacent to an elevator or stairway. Built-up links
also have the advantage that links can be sized to exactly match the design force
demand and, hence, minimize the force demand when carrying out the capacity
design verifications.

Buckling of Beams and Columns In order to achieve the intended yield-


ing mechanism, the beams outside of the links, the braces, and the columns in
EBF systems must be designed for the gravity loads plus the axial forces and
bending moments induced by the links reaching their expected strain-hardened
yield strength. Applications of such a capacity design procedure are given in AISC
(2006) and Chao and Goel (2006). In this design phase, codes generally assume
that the links will develop forces corresponding to approximately 1.25 to 1.30 times
their nominal resistance determined with the expected yield strength. Although this
overstrength factor is lower than the strain-hardening response observed for short
links, the difference is assumed to be compensated by factors such as the fact
that factored resistances are used in design. For the beam segment outside of the
link, the overstrength factor in AISC seismic provisions is reduced to 1.1, which
assumes that the beam strength will be increased by the presence of a floor slab
and that limited yielding of that beam segment would not impair the global per-
formance of the system. Beam segments outside of the links can be subjected to
high combined axial load and bending moment demand and the verification against
instability failure modes may pose difficulties in practice. The flexural demand can
be reduced if short links are used instead of long links. Brace members can also be
rigidly connected to the beams such that part of the bending moment induced by
the links is resisted by the braces (Koboevic and Redwood, 1997). Another pos-
sible solution consists of using different shapes for the beam links and the beam
segments outside of the links, as proposed by Stratan and Dubina et al. (2004) and
Mansour et al. (2006).
For column design, the reduced expected link strength assuming an overstrength
factor of 1.1 is also specified in AISC seismic provisions to account for the fact that
several links may contribute to the column axial load at the level under considera-
tion. As described earlier for CBFs, this approach may underestimate the demand
on columns in low-rise structures and at the uppermost floors in tall multistory
frames (Richards, 2009). Koboevic and Redwood (1997) proposed to use higher
link overstrength for the design of the upper tier columns and this approach has
been adopted in CSA-S16. Past dynamic time history analyses of multistory EBFs
showed that the columns are also subjected to flexural demand, as is also observed
in CBFs and BRB frames. Kasai and Han (1997) proposed that this demand be
accommodated in design-bending by reducing the axial column strength by 15%.
Koboevic and Redwood (1997) suggested increasing this allowance to 40% in
the top two stories of multistory buildings due to higher localized story drifts
882 STABILITY UNDER SEISMIC LOADING

resulting from higher mode response. These recommendations have been included
in CSA-S16 by limiting the interaction equation for columns to 0.85 at every level
except in the top two levels where the limit is set equal to 0.65.

19.3 GLOBAL SYSTEM STABILITY (P− EFFECTS)

Structures must be capable of safely carrying the gravity loads they support while
undergoing large inelastic lateral deformations resulting from strong earthquake
ground motions. The design objective essentially consists in controlling the struc-
ture lateral displacements such that the inelastic demand on individual yielding
components remains within acceptable limits and global instability of the entire
structure or any part thereof is prevented. Achieving this performance objective
in day-to-day practice still represents a formidable challenge in view of the com-
plex and random nature of the seismic excitation and the difficulties in adequately
predicting the dynamic nonlinear response of structures with proper consideration
of all sources of geometric and material nonlinearities, damage accumulation, and
consequent strength degradation and so on. In this regard, current code provisions
for P− effects in seismic design are generally based on simplistic models that may
not properly address the actual inelastic seismic stability issues observed in building
structures. Significant research efforts have been devoted in the last two decades
to better understand dynamic seismic instability and the prediction of structural
collapse. A vast portion of this work has been carried out within the wider context
of the development of methodologies for probabilistic assessment of the seismic
performance of structures and performance-based design strategies for buildings.
While remarkable advances have been accomplished on various aspects related to
global structural seismic response no simple and reliable design methods have been
developed yet to ensure stable seismic inelastic response.
This section focuses on dynamic seismic stability. Basic concepts are first
reviewed by examining P− effects on the response of simple single-degree-
of-freedom (SDOF) models subjected to static and dynamic seismic horizontal
loading. Elastic and inelastic responses are examined. SDOF systems are useful to
understand the effects of gravity on seismic response and assess the influence of
key parameters, such as the hysteretic behavior of the structural system. Findings
from studies on SDOF systems are obviously applicable to single-story structures,
but they can also be extended to multistory buildings. For instance, most steel seis-
mic force-resisting systems exhibit shear-dominated global inelastic response; that
is, the lateral load response at a given level essentially depends on the properties
of the structural components at that level. This situation favors the application of
SDOF model results on a per-story basis in multistory structures, as is currently
done in seismic code provisions for P− effects. Results from SDOF systems can
also be used, with modifications, when adopting design strategies where the global
structure is reduced to an equivalent SDOF system. This is the case in the study
of individual collapse mechanisms or when adopting displacement-based design
approaches.
GLOBAL SYSTEM STABILITY (P− EFFECTS) 883

After reviewing the SDOF system, past studies on seismic stability effects on
multistory structures are presented with emphasis on analysis methods. Building
code provisions for seismic P− effects are reviewed and discussed, and strate-
gies that have been proposed to enhance the seismic stability of steel seismic
force-resisting systems are presented. Given that this section cannot provide detailed
coverage of this topic, additional information on the design for global seismic
stability can be found in Gioncu (2000), Krawinkler (2006), Villaverde (2007),
Zareian and Krawinkler (2007), and ATC (2008). Only the effects of horizontal
ground motions are discussed in this section. Spears and Charney (2006) reviewed
past studies on and examined the influence of vertical ground motions on seismic
structural stability.

19.3.1 P− Effects under Static Loading


Figure 19.21a shows a SDOF system subjected to a static horizontal load V while
supporting a gravity load P. The seismic force-resisting system is represented by
the horizontal spring with an elastic stiffness k and the force resisted by the system
is Fs = k, where  is the horizontal displacement at the top of the structure.
If the gravity load is omitted, Fso = V and the deflection is o = Fso /k = V /k .
When P is included, equilibrium in the deformed configuration leads to
   
P P P Po
V = k− =k 1−  = k (1 − θ )  with θ = =
h kh kh Vh
(19.23)
In this expression, θ is referred to as the stability coefficient, a parameter that
relates the geometric stiffness (P/h) to the system elastic stiffness k . The presence
of gravity loads results in a reduced effective lateral stiffness of k (1 − θ ) and both
the deflection o and the force Fs can be determined by amplifying the results from
a first-order analysis without P− effects by 1/(1 − θ ) (Fig. 19.21b) to obtain

V o
= = (19.24)
k (1 − θ) 1−θ
V
Fs = k  = (19.25)
1−θ

The elastic buckling load of the structure, Pcr , is the load that reduces the
effective lateral stiffness to zero, that is, when Pcr = kh, θ = 1.0. As will be
discussed later, other forms of the stability coefficient are reported in the litera-
ture or used in building codes. The same concepts also apply to more complex
multi-degree-of-freedom systems, except that the above are expressed in a matrix
format as described in Chapter 16 and in more detail by McGuire et al. (2000).
In seismic design, however, the SDOF representation is commonly applied on
a per-story basis, with V being the story shear,  the interstory drift, P the
total gravity loads carried by the columns at the story, and h (or hs ) the story
height.
884 STABILITY UNDER SEISMIC LOADING

Δ Δ
Fs
P P V Fs P P
V V
V F°s
k k 1
h k h
k–P/h k
Δ°
1 Δ 1 V
P/h Δ Δ
1 Δ° Δ Vh + PΔ
(a) (b) (c)

Δ Δ
P/2 P/2 P P
V/2 V/2
V

V
(
L

+
d

PΔ h
h h

Ld
(

a
V/2 V/2 V + PΔ
h h
a
P – Vh + PΔ P + Vh + PΔ P – Vh + PΔ Vh + PΔ
2 L 2 L a a

Mb (e)
Ma
Mb = Vh + PΔ 1 + Δ ( ( Fs V
2 L
Ma = Vh + PΔ 1 – Δ ( ( 1 ak
1
2 L ≈ Vh + PΔ Fsy
ak Vy
1
Vh + PΔ 2 (a – q)k
≈ V′y 1
2 k
k (1 – q)k
(d )
1 1
Δy Δres Δ Δmax Δ
Δy qk Δres
Δmax 1

(f)

FIGURE 19.21 P− effects under static loading: (a) SDOF model with lateral
load-resisting system; (b) elastic response under lateral loading; (c) SDOF pole-type model;
(d ) moment-resisting frame; (e) Braced frame; and (f ) inelastic response under lateral
loading.

Gravity loads acting on a laterally deformed structure produce an additional


overturning moment P− that must be resisted by the lateral force-resisting sys-
tem. This is illustrated in Figs. 19.21c to 19.21e for three different systems: a
cantilevered column system, a single-bay moment-resisting frame at a given level
in a structure, and a braced frame system. In the first case, resistance to lateral load
is provided by bending of the column, and the base moment is amplified by the
P− moment. The beam moments in the moment-resisting frame are increased by
the presence of the gravity loads. In the braced frame, the additional overturning
moment induces additional axial load in the diagonal bracing member resisting
lateral loads. In all three cases shown, the story shear and the total base horizontal
reaction remain equal to the external lateral load V . Hence, the effects of the lat-
eral loads, that is, member forces in the seismic force-resisting system and lateral
GLOBAL SYSTEM STABILITY (P− EFFECTS) 885

deflections, are amplified by the P− moments, not the lateral loads themselves or
the story shears.
In Fig. 19.21f , the system has a defined yield strength Fsy and yield deformation
y . Upon increasing the lateral load V in the elastic range, P− effects are as
above with a reduced elastic lateral stiffness k (1 − θ ) and a force Fs larger than
the applied lateral load V . Once y is reached, yielding is initiated at a lateral
load Vy , that is, at a reduced lateral load compared to the yield resistance Fsy or
the load Vy the system would resist in the absence of gravity loads. Therefore,
P− causes a reduction of the effective resistance offered by the system to lateral
loading. If the cantilever column in Fig. 19.21c has a flexural resistance of V y h,
yielding will occur under a lower lateral load in the presence of gravity loads.
Similarly, if the beam in Fig. 19.21d or the brace in Fig. 19.21e is sized to yield
under a lateral load Vy without gravity loads, the effective lateral yield strength of
these systems will be reduced by Py /h if the structure supports a load P. On this
basis, a logical design solution would consist of increasing the yield strength Fsy
such that the system has sufficient capacity to resist the anticipated lateral load V ,
which can be done using Eq. 19.25 with Fsy = V /(1 − θ ).
While this solution is appropriate when yielding is prevented under defined static
loading, it does not guarantee adequate performance when the structure is expected
to deform in the inelastic range under seismic-induced lateral loading. If the system
has an elastic–perfectly plastic Fs − behavior (α = 0 in Fig. 19.21f ), the effec-
tive lateral stiffness becomes negative upon yielding, equal to −θ k (or −P/h),
when including gravity load effects. This represents an unstable condition under
static loading because the lateral load must be reduced to maintain equilibrium as
the effective lateral resistance decreases when the lateral displacement is increased,
from V = Fsy − P/h. Lateral loads due to seismic ground motions are dynamic
inertia forces, not static loads, and their amplitude and direction change continually
as the structure oscillates. In addition, damping forces are also induced in dynamic
response and these forces also contribute to equilibrium. Therefore, stable inelas-
tic response can still be achieved under an earthquake for this system, although
a negative-yielding stiffness generally causes the structure to progressively drift
toward one direction—a behavior that can eventually lead to large displacements
and structural collapse, as will be discussed in the next section. In Fig. 19.21f , the
postyield negative stiffness beyond y can be reduced and even cancelled out if the
seismic-resisting system exhibits a positive-yielding stiffness αk . This is another
possible design option to counteract P− effects and it is examined in the next
section. The figure also shows that residual permanent drifts (res ) are likely after
the structure has been laterally deformed in the inelastic range. This may represent
a stability issue in the case of aftershocks and, as such, requires examination.

19.3.2 P− Effects on Dynamic Seismic Response

Single-Degree-of-Freedom Systems The reduction in lateral elastic stiff-


ness due to gravity loads lengthens the period of vibration of a structure and,
886 STABILITY UNDER SEISMIC LOADING

thereby, its response to dynamic loading. For the system shown in Fig. 19.22a
with a seismic weight W and lateral stiffness k , the period T  including P− is

 W T W
T = 2π =√ where T = 2π (19.26)
g k (1 − θ ) 1−θ gk

where g is the acceleration due to gravity (g = 9.81 m/s2 ) and T is the period of
vibration of the system without consideration of P− effects. This change in period,

P V
Fsy
W k, c
+
h T=1.0 s V
5% damping
Δ
h = 4.0 m − +
Fsy = 0.15W V >V
q = 0.10 −Fsy

(a) (f) (g) (h)

(b)

(i) (j)

(c)

(k) (l)

(d)

(m) (n)

(e)

FIGURE 19.22 (a) SDOF model for dynamic analysis; (b−e) effects of gravity and hys-
teretic behavior on time history drift response; (f, g) design and ground motion response
spectra; (h) influence of lateral strength on maximum displacement demand on the EP
model; (I −n) Lateral force–deformation responses of SDOF systems with different hys-
teretic behavior.
GLOBAL SYSTEM STABILITY (P− EFFECTS) 887

however, is generally small and its effect on seismic response is not significant
because the stability coefficient θ is generally less than 0.10 and very rarely exceeds
0.20 in actual structures.
Figures 19.22b to 19.22e show the normalized computed drift (/h) time his-
tory response of SDOF systems with a period T of 1.0 sec, viscous damping of 5%
of the critical damping, and θ = 0.10 under the M6.7-1994 Northridge earthquake
record. The response assuming elastic behavior (EL model) is given in Fig. 19.22b
for the cases where P− effects are omitted and included in the analysis. The
difference is very small and is mainly characterized by a time shift in the decay-
ing (free-vibration) response that follows the strong ground motion part of the
earthquake record. Analytical studies on similar systems (e.g., Bernal, 1987) and
on more complex structures (e.g., Cheng and Tseng, 1973) showed that period
changes due to the P− effects and cause the lateral displacements and forces to
increase or decrease, but generally not by substantial amounts.
Most studies on the seismic stability of yielding structures have been performed
using elastic–perfectly plastic (EP) and bilinear (BL) hysteretic models. The two
models are identical except that the latter exhibits a positive load–deformation slope
upon yielding (α > 0 in Fig. 19.21f ). Strain hardening of steel can contribute to
this increase in strength with increasing lateral deformations. For complex structural
systems, α can also represent progressive yielding taking place in the structure as
the lateral displacement is increased (the components that are still elastic attract
additional loads before they eventually yield). In the example of Fig. 19.22, the
lateral strength Fsy of the EP and BL systems was determined using the elastic and
inelastic design spectra (EDS and IDS) in Fig. 19.22f . The structure is assumed to
have a global ductility of 4.0 and the EDS and IDS ordinates at T = 1.0 sec are
0.6W and 0.15W , respectively. The 5% damped acceleration response spectrum
of the ground motion is shown in the figure and the demand from the seismic
record matches the EDS well at the model period of 1.0 sec. Inelastic time history
dynamic analyses of the EP model are first carried out for values of Fsy varying
stepwise from 0.6W (elastic) to 0.15W (design). In Fig. 19.22g, the peak lateral
displacements /h computed for the different resistance levels are plotted for the
cases with and without P−. Nearly identical results were obtained for the two
cases until the lateral strength Fsy approaches 0.25W to 0.30W . Above that point,
the response is in accordance with the equal-displacement principle, that is, the
peak horizontal displacements are not influenced much by the lateral strength of
the system. When further reducing Fsy , the response with consideration of gravity
loads starts to deviate significantly from the case without P−, with excessive
drifts and total collapse when approaching the design level. Montgomery (1981)
observed this behavior for several structures and concluded that P− has negligible
effects on displacements until the lateral resistance is decreased to a critical level
below which displacements increase very rapidly.
The time history response of the EP model with Fsy = 0.15W is presented in
Fig. 19.22c for the two cases with and without P−. Figure 19.22i shows the
hysteretic lateral load–lateral displacement (V-vs.-/h) and lateral force–lateral
displacement (Fs -vs.-/h) response of the model with gravity loads. The response
888 STABILITY UNDER SEISMIC LOADING

of the EP model supporting gravity loads is characterized by gradual drifting


toward the positive-displacement side after a first large-amplitude inelastic excur-
sion occurred in that direction. The reduced resistance at yield and the nega-
tive effective postyield stiffness of the system can be observed in the hysteretic
(V-vs.-/h) curve of Fig. 19.22i , together with the one-sided progressive shift-
ing of the structure. Once the structure has yielded in one direction, it becomes
unsymmetrical as the lateral load required to further yield the structure in the
same direction is gradually decreased by the term P/h (noting V = Fsy − P/h)
whereas the horizontal load that must be applied to initiate yielding in the opposite
direction, toward the initial position, is augmented by P/h. This bias in lateral
strength is schematically illustrated in Fig. 19.22h and it is clear that such a dif-
ference in yield strength encourages subsequent yielding excursions to develop
in the same direction, further increasing the displacements reached in previous
yielding events. In Fig. 19.22i , yielding only occurred in the positive direction
after the first large yielding event in that direction. This behavior can also be
examined from the point of view of loading: Once the structure has displaced
by yielding to a deformation of , the overturning P moment adds to subse-
quent seismic loads acting in the same direction but reduces seismic load effects
acting in the opposite direction. The phenomenon becomes more pronounced as
the displacement increases and it accelerates to cause structural collapse as the
displacement approaches c = Fsy h/P (or y /θ ), the point beyond which the
system cannot support the static gravity loads acting alone. The P− effects are
more important when the strength of the structure relative to the force demand from
the ground motion is decreased or when θ is increased. The detrimental effect of
increasing the stability coefficient on the lateral displacements and spectral accel-
erations that can be resisted before collapse was confirmed by shake table tests on
reduced-scale SDOF steel frames by Vian and Bruneau (2003). Structural collapse
was also recently reproduced using hybrid simulation techniques (Schellenberg et
al., 2008). It is however difficult to generalize on the magnitude of the P− dis-
placement amplification as it heavily depends on the characteristics of the ground
motion (duration, frequency content, presence of pulses, etc.) and how the struc-
ture responds to that motion (Araki and Hjelmstad, 2000; Williamson, 2003; Bernal
et al., 2006).
To illustrate the benefits of selecting a system with strain-hardening response
upon yielding, plots of the time history responses of BL systems with α values of
0.02 and 0.10 (BL02 and BL10) are provided in Fig. 19.22c. In this case, P−
effects gradually reduce as α is increased and are eventually completely mitigated
when α = θ . Husid (1967) and Jennings and Husid (1968) carried out a pioneering
study in this area by developing expressions to predict the time to collapse for EP
and BL systems. For EP systems, the time to failure was found to increase linearly
with h and to be proportional to the square of the strength of the structure relative
to the ground motion intensity. It also lengthens when increasing the ratio α/θ and
no collapse was observed when that ratio was equal to 1.0. Several subsequent
studies confirmed the stability of BL models with a positive effective yield slope
GLOBAL SYSTEM STABILITY (P− EFFECTS) 889

(α − θ > 0). MacRae (1994) extended this concept and proposed a criterion to
verify the stability of any hysteretic shape.
Providing sufficient strain-hardening response to achieve a stable response may
not be practical in all cases, and an alternative approach consists of increasing the
system lateral strength to compensate for P− effects. In Fig. 19.22c, the peak
displacement reached by the EP model without P− effects is max = 3.3% of h.
The strength of the model is therefore increased to Fsy = 1.35Fsy (= 0.203W ) such
that the new strengthened system (EPS) possesses an effective lateral resistance of
0.15W when max is reached (Fsy − Pmax /h = 0.15W ). It is assumed that the
drift of the EPS model will remain less than or equal to the drift of the EP model
without P− effects. In Fig. 19.22e, the EPS model underwent a peak drift of 4.9%
of h, much larger (48%) than the target value of 3.3% of h. This result is typical
and shows that compensating for dynamic P− effects using methods based on
static equilibrium, even when considering the inelastically deformed configuration,
may not be sufficient to control the structure deformation or ductility (Humar et
al., 2006). Bernal (1987) performed a parametric study on EP SDOF models to
determine the strength amplification factor, ϕ = Fsy /Fsy , that is needed to reach
the same ductility μ with a resistance Fsy when supporting the gravity load as
that obtained with a resistance Fsy when P− effects are neglected. Both Fsy and
Fsy are determined by successive iterations until the same target μ is attained.
The parameters μ, θ , and T were varied. Limited correlation was found between
the period and ϕ for the range studied (T = 0.2 sec to T = 2.0 sec) and, from
regression analysis, an empirical expression was proposed for ϕ as a function of
μ and θ
1 + ψ (μ − 1) θ
ϕ= (19.27)
1−θ
where ψ is a constant (1.87 and 2.69 for mean and mean plus one standard deviation
results, respectively). The equation satisfies the limiting conditions ϕ = 1.0 when
θ = 0, ϕ = 1/(1 − θ ) when μ = 1.0 (elastic response, Eq. 19.25), and ϕ = ∞
when θ = 1.0 (elastic critical load is reached). This study showed that large ampli-
fication can be obtained for ductile structures. For instance, for θ = 0.10, the mean
value of ϕ varies from 1.53 to 2.15 for μ comprised between 3.0 and 6.0. Maz-
zolani and Piluso (1993) reported on similar investigations performed in Italy that
led to the same expression for ϕ except that an exponent ψ2 is applied to the
term (μ − 1). For T > 0.5 sec, the mean amplification using this expression can be
obtained with ψ = 0.62 and ψ2 = 1.45, showing a greater influence of the duc-
tility. Mazzolani and Piluso (1996) developed more refined expressions for ϕ that
can be used to control ductility parameters that account for damage accumulation.
Fenwick et al. (1992) proposed expressions for a P− amplification factor β that
must be applied to the lateral strength increase required at a ductility of μ, as
determined from static analysis, to obtain the dynamic strength amplification ϕ:
 
Pμy
ϕ =1+β = 1 + β (μθ ) (19.28)
Fsy h
890 STABILITY UNDER SEISMIC LOADING

The factor β depends on the building period, the soil type (as it affects the
ground motion signature), and the ductility. This approach has been adopted in
the New Zealand code (Davidson and Fenwick, 2004) and is discussed later. For
μ ≥ 3.5, a value of β = 2.0 is proposed for periods up to 2.0 sec (2.5 sec for
structures on soft soils), confirming again the amplification of P− effects under
dynamic seismic loading. Yamazaki and Endo (2004) developed strength ampli-
fication factors based on the ratio of the seismic energy input required to reach
a target ductility with and without consideration of P− effects. This approach
was found to give amplification levels nearly identical to those obtained from Eq.
19.27. Using the above expressions, values of θ can be determined to maintain the
amplification small enough to be neglected. For instance, as suggested by Bernal
(1987), if ϕ < 1.1 is deemed appropriate for no amplification, θ must be limited
to 0.015 for a ductility of 4.0 when using Eq. 19.28 with ψ = 1.87, which is very
restrictive.
The yielding components in a seismic structural system can experience strength
degradation beyond a given deformation level. Such degradation may result in a
negative effective lateral stiffness or a reduction in strength with cycles. Both of
these situations can have detrimental impacts when combined with P− effects.
This is illustrated in Figs. 19.22d and 19.22l where strength degradation was imple-
mented in the EPS model to form the EPSD model. In this case, the strength
degradation was set to initiate at a drift of 0.027h and produce a slow linearly
decreasing resistance such that the strength reduces by 10% at 0.04h. This simple
modification to the hysteretic response provoked collapse of the structure, empha-
sizing the importance of including this behavior in collapse analysis.
The interaction between strength degradation and gravity load effects has been
studied analytically in past research (e.g., Della Corte et al., 2002; Williamson,
2003; Ibarra et al., 2005) and its impact on collapse response was demonstrated in
shake table tests on steel MRFs (Rodgers and Mahin, 2006, 2008; Lignos et al.,
2008; Suita et al., 2008). The deformation at which strength degradation begins, the
slope of the descending (softening) branch, and the rate of cyclic degradation are
key parameters influencing the response of structures. These aspects are discussed
further in the next section.
Initial out-of-plumb of structures may also affect their seismic stability, as illus-
trated in Fig. 19.22e with the EPSD–0 case. This model is identical to the EPSD
model except that an initial deflection of h/500 corresponding to typical erection
tolerances for steel structures was specified. This initial displacement creates a
biased situation similar to the condition that exists after a first yielding excursion
has occurred in one direction. Hence, seismic-induced drifts will tend to develop
in the direction of 0 . An even more critical situation exists for structures that
have been permanently deformed after an earthquake, as past studies showed that
residual story drifts can be much larger than erection tolerances (e.g., Kawashima
et al., 1998; Ruiz-Garcı́a and Miranda, 2006). Limited studies have been carried out
to predict the response of such initially deformed structures. In one study, Bernal
(1987) derived from static equilibrium an upper limit on the product of μ and θ for
EP systems that must be met if the structure needs to carry a gravity load Pu after
GLOBAL SYSTEM STABILITY (P− EFFECTS) 891

the earthquake. The limit is given by μθ = Ps /Pu . It assumes that the structure
reached a displacement equal to μy while supporting a gravity load Ps during the
earthquake. According to this criterion, the stability coefficient would need to be
limited to 0.10 if, for example, Pu /Ps was equal to 2.5 and the anticipated ductility
was 4.0.
Figures 19.22e and 19.22m show the response of a SDOF model representing
a dual seismic force-resisting system composed of two framing systems exhibiting
elastic–perfectly plastic behavior. In this example, the two frames have the same
lateral strength Fsy = 0.075W , but the stiffness of one frame is four times the
stiffness of the other frame. These stiffnesses have been adjusted such that T =
1.0 sec when considering the total initial lateral stiffness, which is identical to the
original EP system. Yielding of the stiff frames occurs at a displacement of 0.58%
of h whereas the more flexible frame remains elastic up to 2.33% of h. Hence, the
flexible frame effectively provides α = 0.20, larger than θ in the 0.58% to 2.33%
of h displacement range. As shown, the seismic response of that dual-EP system is
stable, even if its total lateral strength is the same (0.15W ) as the EP model, which
collapsed in Figs. 19.22d and 19.22i . This is attributed to the contribution of the
flexible frame that provides the system with positive stiffness up to a drift of 2.33%
of h, thereby reducing the duration and the number of yielding excursions during
which an effective negative postyield stiffness prevails. Compared to the EPS model
(EP model with resistance increased to 0.203W ), this design would represent an
attractive solution because it would likely be lighter and impose lower forces on the
other structural components located along the seismic load path while experiencing
comparable lateral deformations. This example shows the benefits of designing for
progressive yielding in the seismic load-resisting system or adding a back-up elastic
system designed to yield at large deformations. The former application has been
investigated by Tagawa et al. (2004) for steel MRFs, as will be presented below.
The response of a system with self-centering or recentering hysteretic properties
is shown in Figs. 19.22e and 19.22n. These systems are forced to return to the unde-
formed position after each yielding excursion, which translates into a flag-shaped
hysteretic curve. This behavior can be achieved by different means, such as using
specialized materials displaying this particular hysteretic response, allowing the
structure to rock, or introducing components connected by or built with preten-
sioned tendons. Compared to EP models, these systems lack energy dissipation
capacity, but it has been demonstrated that this shortcoming can be compensated
for by their ability to recenter (Christopoulos et al., 2002a). Of more significance
for seismic stability, the increased displacement due to P− effects during each
individual yielding excursion is eradicated when the deformation is reduced to zero
upon load reversal. Thus, the drifting or ratcheting behavior responsible for col-
lapse is prevented, resulting in a more symmetrical response without residual drifts,
as can be observed in Figs. 19.22e and 19.22n. Further discussion is presented at
the end of this chapter on possible applications of this system to achieve stable
seismic behavior.
Realizing that the influence of gravity loads remains small until the lateral
strength of a structure is significantly reduced and dynamic instability is likely
892 STABILITY UNDER SEISMIC LOADING

(Fig. 19.22d), Bernal (1992) proposed an alternative treatment of P− effects


oriented toward collapse prevention instead of ductility control. In this case, the
minimum resistance level that is needed for stable inelastic seismic response is
determined from analysis, rather than the strength amplification factor required to
meet a given target ductility. From parametric studies on EP and degrading models,
expressions were proposed for collapse spectral ordinates (or a minimum Fsy /W
ratio required to prevent collapse) as a function of the peak ground displacement
and velocity, θ and T . For the EP model, the duration of strong ground shaking was
also found to influence the strength at onset of instability. These ordinates increase
with ground motion amplitude (and duration), as well as with θ , but decrease with
the period. Asari et al. (2000) developed similar collapse spectra that also pro-
vide minimum strength values to achieve target ductility levels. In design, these
spectra can be used to set the minimum lateral resistance necessary to achieve a
desired safety margin against collapse. Alternatively, maximum values of the seis-
mic response modification coefficient R can be determined from the ratios of the
elastic seismic force demand from earthquakes to the lateral strength at which col-
lapse occurs (Rutenberg and De Stefano, 2000; Miranda and Akkar, 2003). These
critical values of R decrease when θ is increased and generally increase with T .
For bilinear systems, the strength causing collapse generally decreases (R increases)
when increasing the postyield stiffness ratio α. It must be noted that θ typically
increases with T (Bernal, 1992), which reduces the apparent positive influence of
the period on minimum strength requirements that is observed when θ and T are
considered separately.

P− Effects on Multistory Structures Experimental investigations and


numerical simulations of the inelastic seismic response of multistory steel seismic
force-resisting systems showed that the peak interstory drift demand generally
varies along the height of buildings and the difference between peak roof drifts
and peak interstory drifts accentuates when increasing the intensity of the ground
motion relative to the lateral capacity of the structure (e.g., Roeder et al., 1987;
Whittaker et al., 1987; Gupta and Krawinkler, 2000a; Marino and Nakashima,
2006; Chen et al., 2008). Figure 19.23a illustrates this phenomenon for an
eight-story BRB frame subjected to two different ground motion levels. In the
figure, the time histories of the horizontal displacement at each level, the vertical
profile of the horizontal displacements at three points in time when the roof
displacement reaches a maximum value, and the first-story (base) shear versus
interstory drift hysteretic response are given for two ground motion intensities.
As shown, the fairly uniform demand pattern observed under the lower shaking
level changes to a one-sided response with more pronounced interstory drifts
and residual deformations in the bottom levels when increasing the amplitude of
earthquake motion.
A number of reasons contribute to nonuniform interstory drift demand along the
height of structures. One reason is that inelastic deformations of the yielding com-
ponents in steel frame systems (plastic hinging in beams in MRFs, plate yielding in
SPSWs, brace yielding or buckling in CBFs or BRBFs, or yielding in EBF links)
GLOBAL SYSTEM STABILITY (P− EFFECTS) 893

(a) (b)

Leaning
Δ Columns

{
V
A
Δ

1.0 Without P-Δ


With P-Δ

{
Δ Leaning
CDF

Δ { Columns

0.0
0.0 1.0 2.0 3.0
Ground Motion Intensity

(c) (d)

FIGURE 19.23 (a) Response of an eight-story BRB frame under increasing ground motion
levels (time histories of story horizontal displacements, horizontal displacement profiles, and
story shear vs. story drift response at level 1). (b) three-story collapse mechanism in an
eight-story BRB frame; (c) incremental dynamic analysis results; and (d ) non-linear static
incremental analysis.

essentially translate into story shear deformations that can develop nearly indepen-
dently of the story shear deformations taking place in adjacent stories. Columns
of both the seismic and gravity load resisting systems are the only structural com-
ponents that can contribute, through bending, to minimizing differences between
stories, but they are generally not specifically designed for this particular purpose
(Medina and Krawinkler, 2005). Another reason is that the vertical distribution of
the horizontal demand (force or deformation) at any time during a particular earth-
quake can differ significantly from the lateral force pattern assumed in design when
assigning story shear resistance to the structure. The design pattern represents an
envelope, not the specific loading patterns that will be applied at various times dur-
ing an earthquake. These actual time varying patterns can create large deformations
at specific locations along the building height and there is no guarantee that the
sequence and characteristics of such loading events will lead to the same inelastic
894 STABILITY UNDER SEISMIC LOADING

deformation demand at every story at the end of a particular ground motion (e.g.,
Tjondro et al., 1992).
Amplification of lateral deformations and progressive drifting due to P− may
take place at stories where large interstory drifts have developed and a negative
effective story shear stiffness exists, accentuating further the variations in interstory
drifts between stories. This can lead to dynamic instability response characterized
by the formation of a collapse mechanism with a vertical profile defined by the
incremental inelastic story shear deformations due to P− effects, thereby involving
only one or a few adjacent stories in the region where large inelastic deformations
had already developed (Bernal, 1992, 1998; Gupta and Krawinkler, 2000b; Trem-
blay and Poncet, 2005). The collapse induces plastic hinging of the continuous
columns at the top and bottom ends of the mechanism, as illustrated in Fig. 19.23b
for a three-story mechanism in a BRB frame assuming pinned brace-to-beam and
beam-to-column connections. This behavior also occurs in steel MRFs, even if
strong column–weak beam design was applied, as was recently demonstrated in
the shake table test program by Lignos et al. (2008).
Occurrence of dynamic seismic instability depends on the numerous factors
that dictate the overall inelastic response of the structure to the ground motion,
including dynamic properties of the structure (periods and mode shapes, damp-
ing), structure lateral strength and vertical distribution of the story shear strength,
hysteretic behavior of the yielding components, amplitude and vertical distribution
of the gravity loads, continuity conditions and flexural properties of the columns,
characteristics of the ground motion, and the local inelastic response to the large
deformations developing at the levels involved in the collapse mechanism. Because
inelastic structural deformations before and during collapse are essentially in the
form of story shear deformations and the hysteretic behavior of the yielding com-
ponents can be transposed into the story shear–story shear deformation response,
the key parameters influencing frame stability can be related to each other on a
story basis (e.g., stability coefficient, gravity load vs. story shear strength, etc.) and
interstory drifts can be used to characterize both the structural response and the
demand on the yielding components.
In past studies, inelastic deformation demand tends to concentrate at the build-
ing base, where the ground motion is input into the structure and the gravity
loads are maximum, or in the upper levels as a result of higher vibration mode
response inducing large story shear demand (e.g., Marino and Nakashima, 2006).
Furthermore, comparative studies on identical buildings having different heights
consistently showed that the deviation between peak interstory drifts and peak
roof drifts increases when the height of the building is increased (e.g., Gupta and
Krawinkler, 2000a; Sabelli et al., 2003; MacRae et al., 2004; Karavasilis et al.,
2007; Tremblay et al., 2008b). There are many reasons that may explain this trend.
As the structure height is increased, the relative contribution of the higher vibra-
tion modes augments, increasing uneven demand along the frame height. Gravity
loads in the lower levels also increase with the building height whereas the design
seismic load to seismic weight ratio diminishes due to the shape of current code
design spectra. This creates a more critical situation for P− effects with higher
GLOBAL SYSTEM STABILITY (P− EFFECTS) 895

values of the term P/h relative to Fsy for a given deformation  or smaller
values of the interstory drift c where the effective lateral stiffness reduces to zero
and instability is likely (= Fsy h/P for EP models). A recent probabilistic study
of the collapse response of 4- to 16-story steel MRFs by Krawinkler and Zareian
(2007) seemed to confirm this tendency. It was found that the risk of collapse was
much higher for the taller (longer period) frames and the authors suggested that a
minimum lateral resistance be prescribed in codes such that sufficient margin exists
against seismic instability. In this regard, a minimum design base shear ratio V/W
expressed as a function of the seismic ground motion intensity at the site, has been
reintroduced in ASCE 7 code to address this issue (ATC, 2008; ASCE, 2008).
As was observed for SDOF systems, comparative simulations performed on mul-
tistory building models with and without P− effects showed that gravity loads
have negligible effects when the structure experiences limited inelastic deforma-
tions (e.g., Tjondro et al., 1992; Jin and El-Tawil, 2005; Bhowmick et al., 2009).
Montgomery (1981) concluded from his study on MRFs of up to 10 stories and
with elastic–plastic hysteretic behavior that P− effects were significant only when
V /W was less than 0.10 or the interstory drifts were greater than 2.0 times the yield
interstory drift. Fenwick et al. (1992) found that the ductility demand on ductile
reinforced concrete frames could be well controlled by applying the P− ampli-
fication factor developed for SDOF systems (Eq. 19.28). For structures near the
collapse state, Tremblay et al. (1999) noted that increasing the lateral capacity using
strength amplification factors such as those developed for SDOF systems or sug-
gested in codes (see below) could reduce the interstory drifts to the values obtained
when neglecting P− effects in the analysis. These observations suggest that ampli-
fying the seismic load effects in design is necessary when approaching instability
or limiting the ductility demand when large inelastic deformations occur but no
augmentation would be needed when sufficient lateral strength is already present.
When properly detailed, yielding components of most steel seismic force-
resisting systems can develop substantial strain-hardening cyclic response up to
large interstory drift deformations. In particular, shear yielding in short EBF
links, axial yielding in BRB members, and plastic rotation in beams can be
very pronounced. As described earlier for SDOF systems, such strain-hardening
response contributes to counter dynamic P− effects (e.g., Uetani and Tagawa,
1998) and should be modeled in stability analyses (e.g., Ricles and Popov, 1994;
Byfield et al., 2005). Strain hardening may also improve the uniformity in the
yielding demand over adjacent floors because a hardening response at a floor
undergoing large inelastic deformations requires gradually increasing lateral loads
which may eventually trigger yielding in adjacent floors and, thereby, lead to
more uniform inelastic demand. In contrast, deteriorating hysteretic response is
likely to aggravate the concentration of story shear deformations and exacerbate
collapse response when approaching instability. This may have a major impact in
collapse prediction analysis, that is, when pushing the structure to its limit, either
statically or dynamically.
In steel MRFs, monotonic and cyclic strength degradation occurs as a result
of local buckling or ductile fracture (see Fig. 19.7b) and several models have
896 STABILITY UNDER SEISMIC LOADING

been developed and used to simulate this effect in performance assessment studies
including stability effects (e.g., Mehanny and Deierlein, 2001; Della Corte et al.,
2002; Ibarra et al., 2005; Kazantzi et al., 2008; Krawinkler and Zareian, 2007;
Liao et al., 2007; Rodgers and Mahin, 2008). Degradation in well-detailed joints
typically begins at 0.02 to 0.03 rad of plastic rotation and progressively takes
place as the number of cycles and amplitude of deformation are increased. Ibarra
and Krawinkler (2005) noted that the deformation where deterioration begins and
the rate of degradation beyond that point had the largest effect on the collapse
capacity of structures. Strength degradation is generally less pronounced for short
links of eccentrically braced steel frames or buckling restrained braces. Conversely,
buckling and fracture of steel bracing members can result in a marked reduction
in story shear resistance and have detrimental impact on the seismic stability of
braced frame structures. Brace buckling initiates at smaller interstory drifts (less
than 0.005hs ) compared to beam-to-column joints in MRFs and, thereby, has an
earlier effect on frame response during an earthquake. The drop in brace compres-
sive strength after buckling varies with the brace overall slenderness, with greater
reductions for intermediate slenderness (see Fig. 19.14b). Because brace slender-
ness varies along the structure height, the variation in strength degradation may
lead to greater concentration of inelastic damage.
Negative effects of brace buckling are minimized when slender braces sized not
to buckle under the design seismic story shear are used in tension–compression
pairs at every level (Tremblay, 2000, Karavasilis et al., 2007); in each direction,
the slender tension-acting diagonal possesses sufficient reserve strength in tension
to compensate for the strength reduction experienced by the buckled compression
brace. This is not the case in chevron bracing when the beams are not specifically
designed to resist the unbalanced brace loads illustrated in Fig. 19.16d . In this
case, the postbuckling deterioration in brace compressive strength translates nearly
entirely into story shear response degradation because the force developing in the
tension-acting brace is limited by the capacity of the beams. The impact of this
behavior on chevron-framing stability has been demonstrated experimentally (e.g.,
Fukuta et al., 1989) as well as in several numerical studies (e.g., Khatib et al.,
1988; Remennikov and Walpole, 1998b; Tremblay and Robert, 2001). If braces
are prone to low-cycle fatigue fracture (Eqs. 19.16 to 19.18), fracture is likely
to occur at levels where large story deformations have already developed, creating
more favorable conditions for the formation of a collapse mechanism. As described
earlier, numerical brace models that could reproduce global buckling and fracture
have been developed and can be used in braced frame seismic stability analyses
(e.g., Uriz and Mahin, 2004; Daravan and Far, 2009).
Steel moment-resisting frames generally possess high lateral overstrength due to
design requirements other than those related to the resistance of seismic loads; such
requirements may include drift limits, meeting a weak beam–strong column design
philosophy, gravity loads, and so on. Such extra-resistance reduces the extent of
yielding, the likelihood of large interstory drifts, and initiation of a degradation sit-
uation and thus contributes to preventing dynamic instability. Similar overstrength
exists in steel plate shear walls when the infill plate thickness is often governed
GLOBAL SYSTEM STABILITY (P− EFFECTS) 897

by handling or welding processes rather than lateral strength requirements. In both


systems, the columns are inherently strong and stiff in bending, which also helps
in mitigating drift concentrations.
Braced steel frames, CBFs, BRBFs, or EBFs generally possess lower over-
strength and their columns are of smaller sizes, resulting in greater likelihood
for story response. For CBFs, this weakness contributes to a relatively limited
strain-hardening response and strength degradation due to brace buckling, and this
is compensated for by most building codes specifying lower seismic force modifi-
cation factors. Column continuity can also have marked positive effects for braced
steel frames, as was observed in several past studies (e.g., Tremblay and Stiemer,
1994; Martinelli et al., 2000; MacRae et al., 2004; Kimura and MacRae, 2006). Sim-
ilarly, a combination with moment-resisting steel frames has been found to better
the distribution of inelastic demand in multistory bracing systems (e.g., Whittaker
et al., 1987; Shibata, 1988; Bertero et al., 1989; Hassan and Goel, 1991; Martenelli
et al., 1996; Kiggins and Uang, 2006). Moment-resisting frames can be created by
using rigid beam-to-column connections within the bracing bents and/or the gravity
load supporting structure.
In view of the significant influence of the ground motion signature on stability
response and the high variability in ground motion characteristics, randomness in
ground motion must be accounted for in stability analysis. Similarly, the inherent
variation in structural properties and the sensitivity of the response to modeling
assumptions should also be included in the analysis. For instance, correlative time
history analyses of shake table tests showed that asymmetric drift accumulations in
CBFs were very sensitive to slight variations in modeling parameters (Broderick et
al., 2008) or that collapse results were affected by damping assumptions (Vian and
Bruneau, 2003). Gupta and Krawinkler (2000b) observed that refining an analyt-
ical steel MRF model by adding other dependable components (gravity columns,
orthogonal frame columns, shear connections, and floor slabs) to a basic centerline
representation can favourably impact the seismic stability assessment.
Methodologies have been recently developed to include these effects and obtain
probabilistic estimates of the risk of collapse. In these methods, the ground motion
intensity and the deformation at collapse are first obtained through incremental
dynamic analysis (IDA) of the structure when subjected to a suite of ground motion
records (Vamvatsikos and Cornell, 2002). For each record, nonlinear dynamic time
history analysis is performed at stepwise increasing ground motion amplitudes
until structure collapse occurs, with the maximum peak interstory drift value along
the building height plotted as a function of the ground motion intensity mea-
sure. Figure 19.23c shows an example of IDA curves obtained for 10 ground
motion records. Each dot in the graph corresponds to the point where collapse
occurred under each record, that is, when interstory drifts become excessive. The
structure model should include the frame studied together with the tributary lean-
ing gravity columns supporting their gravity loads, as illustrated in Fig. 19.23d .
Three-dimensional representation of the entire structure can also be used to better
capture the contribution of all components as well as possible in-plane torsional
response. Of course, geometric nonlinearities and gravity loads expected to be
898 STABILITY UNDER SEISMIC LOADING

present during earthquakes should be included. Hysteretic behavior of the key


yielding components should be specified, based on expected material properties and
with proper consideration of strain hardening and deteriorating responses, together
with a realistic representation of damping.
In the procedure proposed in FEMA 350 (FEMA, 2000a; Lee and Foutch,
2002), the level of confidence against collapse can be determined by compar-
ing the interstory drifts at collapse (capacity) to those at the ultimate limit-state
(collapse prevention) ground motion level (demand) from the IDA curves. The
methodology accounts for the slope of the seismic hazard at the site, variations
in capacity and demand, as well as uncertainties in the prediction of the demand
(damping, material properties, etc.). From the IDA results, it is also possible to
develop collapse fragility curves that correspond to the cumulative distribution
function (CDF) or probability of collapse as a function of the intensity of ground
motion, as illustrated in the bottom part of Fig. 19.23c. These curves are used in
the ATC-63 methodology for the assessment of the seismic performance of struc-
tural systems (ATC, 2008). A collapse margin ratio is determined by comparing
the ground motion intensity corresponding to a 50% probability of collapse to the
ultimate limit-state (maximum considered earthquake) ground motion level. The
probability of collapse at the ultimate limit-state ground motion level is also ver-
ified. In the procedure, the collapse margin ratio is modified to account for the
spectral shape of rare extreme earthquakes and the accepted collapse margin ratio
is defined with consideration of the variability in ground motions and uncertainties
and the quality of the data and models used in the analysis. Selection and scaling of
ground motions and accounting for modeling uncertainties are of key importance
in these methodologies. Additional information on these two aspects can be found
respectively in Haselton and Baker (2006) and Liel et al. (2009).
The above procedure can be used in design to determine a minimum lateral
strength level required to achieve sufficient margin against collapse. These
techniques, however, require intensive analysis and are mainly intended for the
development and validation of design provisions for structural systems and may
not be currently (2009) suitable for day-to-day design. Alternatively, incremental
static (pushover) analysis can be used to identify potential collapse mechanisms
and the lateral capacity of multistory structures. In such an analysis, the structure
is subjected to a pattern of monotonically increasing horizontal loads, and the
total lateral load–roof displacement response is plotted (Fig. 19.23d ). Guidance
on such analysis can be found in FEMA 356 (FEMA, 2000b). It is recommended
in FEMA 356 and EC8 that at least two different load patterns be considered
as an attempt to bound the solution; these include an inverted triangular pattern
(according to the design seismic load pattern) and a uniform pattern (loads
proportional to floor seismic weight). The analysis in each case should be run
until a complete mechanism has formed.
A pushover analysis is particularly useful to determine the displacement and lat-
eral capacity at the point where the effective lateral stiffness becomes negative and
drifting of the response is likely (point A in Fig. 19.23d ). The roof displacement at
point A can be compared to the anticipated roof displacement demand established
GLOBAL SYSTEM STABILITY (P− EFFECTS) 899

from the design spectrum and an equivalent SDOF system, as described in FEMA
356. Global instability is unlikely if the stiffness does not become negative under
the anticipated deformation demand (Gupta and Krawinkler, 2000b; Humar et al.,
2006). Greater lateral strength and deformation capacity before collapse can gener-
ally be achieved when the collapse mechanism mobilizes a larger number of stories
and yielding components (Bernal, 1998); in general, the structure can be corrected
as necessary to meet the anticipated demand. Compared to localized collapse mech-
anisms, a global mechanism is also characterized by a less pronounced negative
P− stiffness beyond point A, which should lessen seismic P− effects in case of
large deformations (Mazzolani and Piluso, 1996).
Other methods also based on the reduction of the structure to an equivalent
SDOF system have been proposed for design purposes. Bernal (1992, 1998) sug-
gested a method to identify the critical collapse mechanism for a steel MRF and
evaluate the lateral strength and stability coefficient associated with that mecha-
nism. The mechanism profile is assumed to correspond to the incremental lateral
deformations between elastic and near-collapse conditions from a pushover anal-
ysis of the structure under a uniform load pattern. Safety against collapse is then
verified by comparing the lateral strength capacity of the mechanism to SDOF col-
lapse spectral values, as described earlier. Mazzolani and Piluso (1996) proposed to
use a kinematic plastic analysis method to evaluate the lateral resistance of poten-
tial collapse mechanisms in steel MRFs. The method can then be used to control
the governing mechanism and assess local ductility demand. Adam et al. (2004)
proposed using a nonlinear dynamic time history analysis to examine the stability
of an equivalent SDOF system having properties derived from a pushover analy-
sis of the entire structure. The conditions for stability found from the equivalent
SDOF system can then be transformed back into the structure domain to assess the
stability of the structure.
It must be recognized that methods based on static analysis procedures have
limitations (Krawinkler and Seneviratna, 1998; Krawinkler, 2006) and may not pre-
dict the collapse mechanism that would actually form under dynamically applied
ground motions. In particular, static analysis cannot reproduce the dynamic drift-
ing or ratcheting response due to gravity loads under cyclic loading, the strength
deterioration or fracture of yielding components due to cyclic loading, and, more
importantly, the continuously varying horizontal load patterns, including higher
mode effects, that contribute to drift concentrations. Although enhanced static anal-
ysis methods, such as the modal incremental analysis that accounts for higher mode
response, may improve the prediction compared to a single load pattern analysis,
the technique may still fail in predicting accurately the demand on structures that
deform significantly in the inelastic range with marked negative stiffness, as is the
case when investigating dynamic P− instability (Goel and Chopra, 2004).

19.3.3 Design for Seismic Stability


Building Code Requirements for Global Stability Recent research on
global seismic stability of steel structures has not yet resulted in simple and
900 STABILITY UNDER SEISMIC LOADING

generally applicable methods that could be implemented in codes to account for


P− effects in seismic design. Current methods proposed in codes essentially
require that seismic-induced member forces be amplified based on the value of
stability coefficients. The calculations are performed on a per-story basis and the
amplification of seismic effects is generally based on static equilibrium. In all codes
reviewed, P− effects can be ignored if the stability coefficient is below a certain
value, and the design must be corrected if the stability coefficient exceeds a max-
imum value. Although similarities exist, there are also marked differences in the
treatment of the topic within many building codes. This is probably a consequence
of the rapid evolution of knowledge in this field that has taken place in the last two
decades, and it is expected that more consistent modifications will be implemented
in the near future to better reflect these recent developments.
In the United States, the ASCE 7-05 standard (ASCE, 2005) is generally the ref-
erence document for loads and general design requirements for building structures.
In ASCE 7, a stability coefficient is evaluated at every story,

P (Cd )
θASCE = (19.29)
Vhs Cd
where P is the total gravity load above the level under consideration, Cd is an
amplification factor to account for inelastic deformations,  is the elastic deflection
under the design seismic loads, V is the story shear due to the design seismic
loads, and hs is the story height. The Cd -factor varies depending upon the seismic
force-resisting system. For most systems, it is approximately equal to 0.7 times the
force modification factor R. It is important to note that the Cd -factor is contained
in both the numerator and denominator in Eq. 19.29, which means that the stability
coefficient is in fact based on elastic response. According to ASCE 7, P− effects
can be ignored in design if θ is less than 0.10. If the stability coefficient exceeds
0.10, amplification of displacements and member forces due to P− effects must
be determined using a rational analysis. No such method is given in ASCE 7,
but it is permitted to determine P− amplification by multiplying the seismic
effects by 1/(1 − θ ), thus assuming elastic response amplification. This approach is
believed to give a reasonable estimate provided that the negative stiffness region is
avoided at the target displacement (Krawinkler, 2000). It is noted here that strength
amplification only applies to the design of the yielding components of seismic
force-resisting systems. The design loads for the capacity-protected components
are driven and limited by the capacity of the yielding components and need not be
amplified for P− effects.
An upper limit for θ is specified in ASCE 7,

0.5
θmax = ≤ 0.25 (19.30)
β Cd

In this equation, β is the story shear demand-to-capacity ratio. This ratio reflects
the overstrength present in the structure and is typically less than 1.0. If θ exceeds
θmax , ASCE indicates that the structure may become unstable during strong ground
GLOBAL SYSTEM STABILITY (P− EFFECTS) 901

motions and must be redesigned. By setting V = βVy in Eqs. 19.29 and 19.30, this
verification becomes

P (Cd )
≤ min (0.5; 0.25βCd ) (19.31)
Vy hs

In this equation, the left-hand side corresponds to the ratio of the story moment due
to gravity loads acting on the inelastically deformed structure to the overturning
moment capacity. This ratio also represents the loss in lateral strength caused by
P− (which equals PCd /hs ) relative to the structure actual story shear resis-
tance Vy . For most ductile systems, the limit will be equal to or slightly less than
0.5 because the factor Cd in ASCE 7 varies between 4.0 and 5.5 and system over-
strength typically varies between 1.5 and 2.5. If the limit is exceeded and because
P and hs are parameters that cannot be modified, the designer has the option of
increasing either the stiffness (reducing ) or the lateral strength Vy .
A similar upper limit exists in FEMA-350 (FEMA, 2000a), in which the param-
eter  is used to characterize P− effects at every level,

P (R)
= (19.32)
Vy hs

where the product R represents the expected story drift in the yielded condition.
In FEMA-350, it is suggested to ignore P− effects if  ≤ 0.1 and to redesign
the structure if the factor exceeds 0.30. No amplification is required for inter-
mediate values of . Taking Cd = 0.70R, the limit of  = 0.3 corresponds to
PCd Vy hs = 0.21, which is more restrictive than the ASCE 7 limit. The FEMA
approach is recommended in the commentary to the 2005 AISC seismic provisions.
The format in EC8 is similar to ASCE 7, except that inelastic deformations
are included in the interstory drift values used in the calculation of the stability
coefficient,

P (q)
θEC8 = (19.33)
Vhs

where q is the behavior (ductility) factor. For the high-ductility-class systems, it


takes a value of 6.5 for redundant moment-resisting frames, 6.0 for eccentrically
braced frames, 4.0 for concentrically braced frames, and 2.5 for frames with chevron
bracing. The P− effects need not be taken into account if θ is less than 0.10.
If θ is between 0.1 and 0.2, second-order effects must be included and this can
be done by multiplying lateral load effects by 1/(1 − θ ), that is, based on elastic
response, but using a reduced stiffness accounting for inelastic deformations. If θ
exceeds 0.3, the structure must be stiffened to meet this upper limit. Compared to
ASCE 7, the EC8 provisions for P− effects are more severe due to the q-factor
included in θ , and the difference is more pronounced for the more ductile systems.
Moreover, EC8 also requires that the formation of a soft story plastic mechanism
902 STABILITY UNDER SEISMIC LOADING

be prevented to avoid excessive demand on the columns. It is suggested that this


be achieved by performing a static incremental analysis as discussed earlier.
In the National Building Code of Canada (NRCC, 2005), the design seismic
loads are determined by dividing the elastic earthquake forces by the product Ro Rd ,
where Ro and Rd are the dependable overtrength and ductility force modification
factors, respectively. The Rd -factor varies from 1.0 for the most brittle systems to
5.0 for the most ductile ones such as type D (ductile) moment-resisting frames.
For this system, the product Ro Rd = 7.5 (Mitchell et al., 2003). The anticipated
interstory drifts including inelastic deformations are obtained by multiplying the
story drifts under the design seismic loads by Ro Rd . As in FEMA 350, the stability
coefficient corresponds to the story shear due to the gravity loads acting on the
deformed structure divided by the lateral resistance Vy taken equal to R o V, which
results in
P (Ro Rd ) P (Rd )
θNBCC = = (19.34)
Ro V hs V hs

In NBCC, P− effects need not be considered if θ is less than 0.10. The
structures must be stiffened or strengthened if the stability coefficient exceeds 0.40,
which is similar to the requirements in EC8. For intermediate θ values, however, the
seismic-induced forces at every level must be multiplied by 1 + θ , which means
that additional lateral resistance must be supplied to fully compensate for P−
effects (the effective lateral strength is equal to Ro V at the anticipated deformed
interstory drift).
In the New Zealand Standard NZ 1170 (NZS, 2004), the stability coefficient at
each level is obtained from
P (μ)
θNZS = (19.35)
Vy hs

where μ is the structural ductility factor and, similar to FEMA 350 and NBCC,
the anticipated deflections including inelastic response μ are used together with
the story shear strength Vy . The ductility factor varies from 1.0 to 6.0. The latter
value applies for the more ductile categories of MRFs, CBFs, or EBFs. In the
determination of θ , the potential for a concentration of inelastic demand over the
building height must be considered as the inelastic interstory drift μ must not
be less than the value obtained from any possible mechanism with an amplitude
at a roof level equal to μ times the elastic roof displacement. For concentrically
braced steel frames, the so-computed interstory drifts must be increased further to
account for the greater tendency of the system to form a sway mechanism, includ-
ing 1.5 times for tension–compression X-bracing and 2.0 times for tension-only
and chevron-bracing systems. In regular structures, however, when a design pro-
cedure is used to prevent story mechanisms, as is the case when adopting a strong
column–weak beam design approach for ductile moment-resisting frames, the sta-
bility coefficient need not be calculated in the upper half of the structure.
GLOBAL SYSTEM STABILITY (P− EFFECTS) 903

In NZ 1170, the structure must be redesigned if the stability coefficient at any


level, where applicable, exceeds 0.3. For lower values of θ , P− effects must be
considered in the calculation of the seismic-induced forces except if the fundamen-
tal period is less than 0.4 sec (0.6 sec if the structure height is less than 15 m) or the
computed values of θ do not exceed 0.10 over the building height. Two methods
are provided to account for P− effects.
In the first method, the seismic effects for the entire structure are amplified by
the single factor 1 + kp W /V , where W is the total seismic weight, V is the design
base shear, and kp is given by

kp = 0.015 + 0.0075 (μ − 1) for 0.015 < kp < 0.03 (19.36)

In this method, amplification is independent of the anticipated deflections. For


ductile structures, it only varies according to the amplitude of the seismic load ratio
V /W (kp = 0.03 for μ ≥ 3.0). For instance, amplification would range between
1.15 and 1.6 for V /W between 0.20 and 0.05, respectively. This method thus
yields greater amplification when higher inelastic response is expected (μ large)
and for structures designed for lower seismic loads.
In the alternative method, a story shear force equal to βP(μ)/hs is added to
compensate for P− effects, where β is the P− amplification factor that accounts
for P− effects on dynamic response (as described earlier with Eq. 19.28). This
factor is plotted in Fig. 19.24a and its value depends on the period of the structure,
the type of soil (site class), and the structural ductility factor. For ductile systems,
the amplification factor equals 2.0 for systems in the low- and intermediate-period
range. This method is identical to the approach used in Canada, except that the
dynamic nature of P− effects on inelastic structures responding to seismic motions
is included through the β-factor and the inelastic interstory drifts μ must account
for the possibility of concentration of inelastic deformations along the building
height. For high-ductility systems, the NZ provisions are significantly more severe
Vu /V

(a) (b)

FIGURE 19.24 (a) P− amplification factor in NZS1170 Standard (NZS, 2004); (b) story
shear overstrength from tension–compression (T/C) and tension-only (T/O) brace designs.
904 STABILITY UNDER SEISMIC LOADING

than those included in ASCE 7. Additional background information on the method


can be found in Davidson and Fenwick (2004).

Design Strategies for Enhanced Seismic Stability As discussed in previ-


ous sections, P− effects on the seismic response are generally small and can be
neglected when the structure response is nearly elastic and/or far from dynamic
instability conditions. Verifying the safety margin against collapse after a prelimi-
nary design has been performed could be one of the first steps in design for seismic
stability. If P− is a problem, increasing the lateral strength is a solution, but other
options are reviewed herein that may be considered.
The potential for collapse by instability can be verified by performing a static
incremental (pushover) analysis (thereby verifying that the anticipated target dis-
placement is less than the displacement where negative stiffness starts) or by
adopting approaches such as those described earlier and proposed by Bernal (1992),
Mazzolani and Piluso (1996), or Adam et al. (2004). When assessing the capac-
ity of steel MRFs, consideration should be given to the strength of the joint panel
zones as these components could govern the frame lateral capacity (Krawinkler and
Mohasseb, 1987; Tsai et al., 1995). Using the proposed methods, the lateral strength
of the structure can be increased as necessary to achieve the required margin against
possible instability. This can be done by trial and error or by using strength ampli-
fication factors from SDOF systems using equivalent SDOF properties reflecting
specific structural behavior. For this purpose, one may also consider the use of the
amplification approach proposed for the displacement-based design method (e.g.,
Asimakopoulos et al., 2007; Pettinga and Priestley, 2007). As described earlier,
all these methods may not capture potential inelastic demand concentrations due
to dynamic effects and caution should be exercised for cases of tall structures and
structures exhibiting strength and stiffness irregularities or strength degradation or
when the structure is subjected to large inelastic deformations and/or is close to
the point of dynamic instability.
For multistory structures, many of the solutions proposed to mitigate P− effects
share the same objective of distributing more uniformly the inelastic demand over
the structure height and, thereby, prevent the formation of localized mechanisms.
Achieving this goal offers several advantages. In particular, a greater number of
yielding components will be mobilized in the structure, leading to a greater energy
dissipation capacity of the structure being exploited without imposing excessive
demand on individual components and, thereby, better control of lateral displace-
ments. Uniform drift demand also results in a more uniform ratio of the P−
overturning moment to the story overturning moment capacity (= P/Vhs ) over
the building height as P and V increase toward the base. Other options can be
considered to limit P− effects are (i) providing the structure with larger lat-
eral displacement capacity before the development of a negative effective lateral
stiffness, (ii) using back-up systems, or (iii) taking advantage of newly developed
self-centering technologies. Those are also briefly introduced below. It must be
noted that most of these methods or structural systems may have not been fully
validated and may have not been approved by regulating authorities, and, of course,
GLOBAL SYSTEM STABILITY (P− EFFECTS) 905

it is the responsibility of the design engineer to ensure the method or system selected
will provide for at least the level of protection implicitly assumed or expected in
the applicable building code.
Inelastic interstory drifts are closely related to the supplied story shear resistance,
and drift concentrations could be minimized by closely matching the story shear
demand without abrupt changes in capacity along the structure height. For instance,
Rossi and Lombardo (2007) observed for EBF structures that determining the story
shear demand from a spectrum analysis, rather than from an equivalent static force
procedure, and providing uniform link overstrength along the building height led
to better collapse capacities. In EC8, a global yielding mechanism is encouraged
by requiring that the actual strength of the links including strain hardening at any
floor not exceed the minimum specified value by more than 25%. Consideration
may be given to vertical story shear distributions that have been recently proposed
to achieve more uniform yielding along the height of structures (Chao and Goel,
2006; Park and Medina, 2007).
Taking advantage of the flexural stiffness and strength of continuous columns
is another means of reducing variations in interstory drifts in multistory structures.
The columns must be designed, however, to fulfill this function. Even for steel
MRFs designed according to the weak beam–strong column code requirements for
preventing story sway mechanisms, additional column capacity must be provided
to prevent plastic hinging in columns resulting from differences in interstory drifts
developing between adjacent floors during an earthquake (Krawinkler, 2006). Eval-
uating this flexural demand on columns is not straightforward because the bending
moments have to be determined from a nonlinear dynamic time history analysis
and they depend on the deformation profile of the structure, which in turn depends
on the flexural stiffness of the columns. A simpler approach has been proposed that
consists of sizing the columns to achieve a target collapse mechanism. For steel
MRFs, Mazzolani and Piluso (1996) recommend that a global collapse mechanism
could be achieved by proportioning the beams for gravity loads and, subsequently,
designing the columns such that the global mechanism exhibits the lowest lateral
capacity among all other possible collapse mechanisms. The calculations in their
procedure are based on the plastic analysis method described earlier. The approach
resulted in an enhanced inelastic demand distribution compared to a current code
design procedure for a five-story building (Mazzolani and Piluso, 1997). To enhance
this method, Neri (2000) developed a solution for the base columns such that itera-
tions are avoided when including P− in the calculations. Mastrandrea et al. (2003)
applied the method to eccentrically braced steel frames. Ghersi et al. (2006) com-
bined this global collapse strategy to the capacity spectrum (N2) method (Fajfar,
1999) to ensure that the structure has a minimum lateral strength consistent with the
expected seismic demand. The approach has been verified by comparing the roof
displacements from the method to those obtained from nonlinear dynamic analyses
for a number of frames having different properties. Although not reported in the
literature, a similar approach (adjusting column sizes to achieve a desired collapse
mechanism) could probably be adopted using the stability check procedure pro-
posed by Bernal (1992, 1998). Application of the method to tall structures, when
906 STABILITY UNDER SEISMIC LOADING

the response is dominated by higher vibration modes, has not been reported in the
literature. More recently, Ohsaki et al. (2007) used a multiobjective optimization
technique for steel MRFs to simultaneously minimize steel tonnage and maximize
the dissipated energy such that a uniform interstory drift is obtained at collapse
of the structure. In this procedure, collapse is determined through an incremental
dynamic analysis, thus accounting for structural dynamic response. Validation of
the method was carried out for a five-story building.
Continuity of the columns that are not part of the moment-resisting frames
(gravity columns or columns part of perpendicular frames) can also improve the
drift response of steel MRFs (Gupta and Krawinkler, 2000b; Tagawa et al., 2004),
and the above methods should be extended to include these columns because they
will experience the same displacement demand during an earthquake.
The stability of braced steel frames can also be enhanced by exploiting column
continuity. For these structures, MacRae et al. (2004) and Kimura and MacRae
(2006) developed equations to predict the drift concentration factor (ratio of max-
imum interstory drift to roof story drift) and the bending moment demand on
continuous columns as a function of the flexural stiffness of the columns, the duc-
tility of the structure, the number of stories, and the vertical distribution of the
story shears. These expressions have been validated through nonlinear dynamic
analysis for frames up to 20 stories. In view of limitations imposed by assumptions
used in their development (e.g., uniform column size over the building height),
further refinement may be needed, however, before they can be used efficiently in
practice. In CSA-S16, minimum column continuity is required for concentrically
braced steel frames, both for the gravity columns and the bracing bent columns, and
minimum flexural strength is required for the bracing bent columns of braced frame
systems to delay plastic hinging and maintain the capacity to distribute vertically
the inelastic demand.
For steel MRFs, Gupta and Krawinkler (2000b) propose providing a flexible
back-up system with sufficient story stiffness to overcome the negative P− effects
on lateral stiffness (i.e., provide α ≥ θ in Fig. 19.21f ) and suggest that the gravity
load system could be used for this function. This solution should be applicable
in practice and could be extended to other systems as well. For instance, the
benefits of adding an MRF back-up system to braced steel frames have long been
recognized in codes by means of relaxations (extended height limits, higher force
modification factors, etc.). In order to be effective, such a parallel system should
be designed to remain essentially elastic over the anticipated deflection range.
Gupta and Krawinkler also proposed providing MRFs with more redundancy and
adjusting the strength of members such that plastic hinging occurs at widely spaced
interstory drifts, and the displacement when a full mechanism forms increases. The
benefit of this concept was illustrated in Fig. 19.22e for a simple SDOF dual system
exhibiting trilinear response.
In order to develop this behavior, Tagawa et al. (2004) inserted pinned beam-
to-column connections at one end of all beams located in the exterior bays of a
nine-story MRF. When column continuity was ignored in the analysis, the peak
interstory drifts were generally reduced when compared to the response of the
GLOBAL SYSTEM STABILITY (P− EFFECTS) 907

reference frame with rigid connections that had higher lateral stiffness and strength.
For the frame studied, the benefits of the multilinear response were surpassed by
those resulting from providing column continuity. Similarly, Charney and Atlayan
(2008) proposed an MRF design using members and connections with a variety
of detailing rules, including those associated with ordinary (OMF), intermediate
(IMF), and special moment frames (SMF). The more ductile components were
designed to yield at force levels well below the design basis earthquake, whereas
the less ductile ones could remain elastic under this seismic demand. The cost
of this design should be less than the conventional SMF solution because only a
limited number of elements and connections have special detailing. Preliminary
studies of a nine-story MRF structure showed that the margin against collapse was
improved when compared to employing a conventional SMF system. The authors
did indicate that this strategy resulted in a larger ductility demand on the ductile
components and that this aspect would require further examination.
As discussed earlier, concentrically braced steel frames may be more vulnera-
ble to a sway mechanism due to story strength degradation associated with brace
buckling, and several provisions have been implemented in codes to compensate
for this weakness. For instance, minimum symmetry requirements are provided
to avoid one-sided response; in this regard, story shear is shared between tension-
and compression-acting braces or tension-acting braces with comparable resistances
supplied in each direction. More stringent height limits apply to the braced frames
designed assuming higher ductility (or using higher Rd - or μ-factors) in CSA-S16
and NZS3404, which is consistent with the fact that P− effects are more pro-
nounced in taller braced frames or when greater inelastic demand is expected.
Lower height limits are also specified in both codes when T/O bracing design is
used, that is, the braces are designed assuming that the total story shear is entirely
resisted by the tension-acting braces. This design allows for slender braces that
have lower energy dissipation capacity compared to tension–compression braces,
which causes larger interstory drifts. In NZS3404, the prescribed height limits also
reduce if the brace slenderness is increased, and the design seismic loads must
be amplified if the building height or the brace slenderness is increased, reflecting
the influence of brace energy dissipation on braced frame response (Remennikov
and Walpole, 1997). Furthermore, 50% amplification (braces effective in tension
and compression) and 100% amplification (tension-only and chevron bracing) must
be applied to interstory drift predictions from code seismic loads in view of the
tendency for soft-story response in braced steel frames.
Tension–compression bracing in which braces are designed not to buckle under
the design seismic loads may possess story shear overstrength due to the reserve
tension capacity available in the tension-acting braces. This overstrength can be
expressed as the ratio of the ultimate story shear capacity Vu contributed by the
brace postbuckling compression strength Cu and the tension brace tensile strength
Tu to the design story shear V . This ratio is plotted against brace slenderness in
Fig. 19.24b assuming that the brace design strength exactly matches the demand
V . Abrupt variations in Vu /V along the building height are likely to result in
drift concentrations, and selecting braces to minimize such variations should be
908 STABILITY UNDER SEISMIC LOADING

considered in design. In Japan, the braces are designed such that Vu equals or
exceeds the story shear demand V (Tada et al., 2003; Marino et al., 2006). With
this approach, it is therefore easier to achieve uniformity in the Vu /V ratio. This
method, however, will generally lead to lower lateral overstrength compared to
cases where brace buckling is prevented under the design story shear, and as a
result, increased seismic design loads (lower force modification factors) may be
needed to achieve similar seismic performance.
In Fig. 19.24b, the variation of Vu /V with brace slenderness for T/C bracing
shows that higher story shear capacity after brace buckling, that is, strain-hardening
response, can be obtained if more slender braces are selected in design. This strategy
can have positive impacts on braced frame stability by limiting interstory drifts
and triggering brace buckling and yielding in several adjacent stories (Lacerte and
Tremblay, 2006). In addition, longer brace fracture life would also be obtained by
increasing λ. Brace slenderness, however, should be kept within a reasonable range
because slender braces will lead to the use of additional bracing steel tonnage as
well as higher design loads for brace connections and adjacent beam and column
members. In EC8, the contribution of the compression-acting braces is ignored
in brace design (T/O design). In this case, beneficial story shear overstrength is
obtained by selecting less slender braces (Fig. 19.24b). The resulting extra-capacity
provided by the compression braces, however, must be considered in the capacity
design check.
Several innovative bracing configurations have been proposed to achieve more
uniform interstory drift demand in braced steel frames. Some of them are illustrated
in Fig. 19.25. For chevron bracing, Khatib et al. (1988) suggested adding a zipper
column to tie together the beams and prevent a story mechanism (Fig. 19.25a). The
further addition of a hat truss as shown in Fig. 19.25b allows for the full develop-
ment of the tension brace capacity along the building height and further enhances
the robustness of the system (Yang et al., 2008a,b). Rai and Goel (2003) found that
the seismic stability of chevron bracing could be improved significantly by using
a two-story X (split-X) configuration (Fig. 19.25c). This configuration has become
very popular for CBF applications in North America. Mazzolani et al. (1994) pro-
posed to connect two CBF bracing bents by weak beams rigidly connected to the
inner columns (Fig. 19.25d ). The CBFs would be designed to remain elastic such
that a global collapse mechanism can be achieved with the energy dissipation con-
centrated in plastic hinges in the beam links. In Fig. 19.25e, vertical ties are added
between the link ends of an EBF to form two braced substructures coupled by
the links (Ghersi et al., 2000, 2003; Rossi, 2007). With this configuration, more
uniform and effective exploitation of the overall energy dissipation capacity of the
links can be achieved while mitigating the potential for story mechanisms.
Figure 19.25f shows a similar concept for buckling-restrained braced frames.
The frame is divided into two parts, including an energy-dissipating frame
and an elastic braced frame to prevent damage concentration (Tremblay, 2003;
Merzouq and Tremblay, 2006). Mega braced frames built with conventional or
buckling-restrained braces spanning over several levels along the exterior walls of
GLOBAL SYSTEM STABILITY (P− EFFECTS) 909

Zippr Hat Truss Plastic


Column (typ.) Hinge (typ.)

(a) (b) (c) (d)

Link Elastic Energy


Tie BRB Gravity Loads
(typ.) Truss (typ.) Dissipation
(typ.) (typ.) Mechanism (typ.)

Tendons

Energy
Dissipation
Tendons Mechanism
(typ.)

(e) (f) (g) (h)

FIGURE 19.25 (a) Zipper column in chevron bracing; (b) suspended zipper system in
chevron bracing; (c) two-story X-bracing; (d ) coupled CBFs; (e) tied EBF; (f ) dual BRB
frame; (g) posttensioned MRF; and (h) controlled rocking CBF.

buildings have also been shown effective in developing uniform drift demand (Di
Sarno and Elnashai, 2009).
By virtue of their capacity to mitigate residual deformations and permanent
damage, systems exhibiting self-centering hysteretic behavior have attracted much
attention in the last decade by the design profession. As illustrated in Fig. 19.22e,
recentering properties may also be used to counter P− effects. In the systems
studied to date, the required hysteretic response has been essentially obtained by
connecting framing components with either (i) tendons or bars made of superelastic
shape memory (SMA) alloys that exhibit flag shape hysteretic behavior or (ii)
tensioned steel tendons or bars acting in combination with a dedicated energy
dissipation mechanism (yielding, friction etc.). Bracing members built with slid-
ing components and tendons have also been designed to produce the desired
self-centering behavior in compression and tension. SMA tendons have been used
in beam-to-column moment connections (e.g., Graesser and Cozzarelli, 1991; Ocel
et al., 2004; Sepúlveda et al., 2008) or in the fabrication of self-centering braces
(e.g., Dolce et al., 2000; Auricchio et al., 2006; McCormick et al., 2007; Zhang
and Zhu, 2008; Zhu and Zhang, 2008). Christopoulos et al. (2008) also proposed a
bracing member with pretensioned tendons and friction energy dissipation capacity.
Moment-resisting frames built with horizontal posttensioned steel tendons that run
parallel to the beams have been examined in several past research programs (e.g.,
910 STABILITY UNDER SEISMIC LOADING

Ricles et al., 2001, 2002; Christopoulos et al., 2002b; Garlock et al., 2005; Tsai
et al., 2007). As shown in Fig. 19.25g, the beams rotate relative to the face of
the columns, which creates a gap and imposes elongation to the tendons. Energy
is dissipated through yielding or friction in components connecting the beam to
the columns. Rocking of the structure with respect to the foundations represents
another means of developing self-centering capability (Fig. 19.25h). Recentering is
provided by gravity loads acting alone or in combination with supplemental down-
ward forces using posttensioned steel tendons. Energy dissipation can be introduced
to control the motion and reduce impact forces when the columns land back on the
foundations. Analytical studies and test programs have been conducted on several
rocking systems for steel structures (Midorikawa et al., 2006, 2008; Azuhata et al.,
2006; Ikenaga et al., 2006; Ishihara et al., 2006; Sause et al., 2006; Pollino and
Bruneau, 2007, 2008; Eatherton et al., 2008; Tremblay et al., 2008c).

REFERENCES

Adam, C., Ibarra, L., and Krawinkler, H. (2004), “Evaluation of P-Delta Effects in
Non-Deteriorating MDOF Structures from Equivalent SDOF Systems,” Paper No. 3407,
Proc. 13th World Conf. on Earthquake Eng., Vancouver, Canada.
AIJ (1975), Guide to the Plastic Design of Steel Structures, Architectural Institute of Japan,
Tokyo, Japan.
AISC (2005a), Seismic Provisions for Structural Steel Buildings, Including Supplement No.
1 , ANSI/AISC 341-05, American Institute of Steel Construction, Chicago, IL.
AISC (2005b), Prequalified Connections for Special and Intermediate Steel Moment Frames
for Seismic Applications, ANSI/AISC 358-05, American Institute of Steel Construction,
Chicago, IL.
AISC (2005c), Specification for Structural Steel Buildings, American Institute of Steel Con-
struction, ANSI/AISC 360-05, Chicago, IL.
AISC (2006), Seismic Design Manual , American Institute of Steel Construction and Struc-
tural Educational Council, Chicago, IL.
Alinia, M.M., and Dastfan, M. (2007), “Cyclic Behaviour, Deformability and Rigidity of
Stiffened Steel Shear Panels,” J. Constr. Steel Res., Vol. 63, No. 4, pp. 554–563.
Araki, Y., and Hjelmstad, K. D. (2000), “Criteria for Assessing Dynamic Collapse of Elasto-
plastic Structural Systems,” Earthquake Eng. Struct. Dyn., Vol. 29, No. 8, pp. 1177–1198.
Arlekar, J.N., and Murty, C. V. R. (2002), “P-V-M Interaction Curves for Seismic Design
of Steel Column Base Connections,” AISC Eng. J., Vol. 39, No. 3, pp. 154–165.
ASCE (2005), Minimum Design Loads for Buildings and Other Structures, Includes Supple-
ment No. 1 , ASCE/SEI 7-05, ASCE, Reston, VA.
ASCE (2008), Supplement No. 2 to ASCE 7-05 , ASCE, Reston, VA.
Asari, T., Inoue, K., and Ishiyama, Y. (2000), “Evaluation of Destructiveness of Earthquake
Motions by Collapse Base Shear Coefficient Spectra,” Paper No. 1537, Proc. 12th World
Conf. Earthquake Eng., Auckland, NZ.
Asimakopoulos, A. V., Karabalis, D. L., and Beskos, D. E. (2007), “Inclusion of P− Effect
in Displacement-Based Seismic Design of Steel Moment Resisting Frames,” Earthquake
Eng. Struct. Dyn., Vol. 36, No. 14, pp. 2171–2188.
REFERENCES 911

Aslani, F., and Goel, S. C. (1991), “Stitch Spacing and Local Buckling in Seismic-Resistant
Double-Angle Braces.” ASCE J. Struct. Eng., Vol. 117, No. 8, pp. 2442–2463.
Aslani, F., and Goel, S. C. (1992a), “Stitch Spacing and End Fixity in Seismic Resistant
Boxed Angle Braces,” ASCE J. Struct. Eng., Vol. 118, No. 10, pp. 2872–2889.
Aslani, F., and Goel, S. C. (1992b), “Analytical Criteria for Stitch Strength of Built-up
Compression Members,” AISC Eng. J., Vol. 29, No. 3, pp. 102–110.
Astaneh-Asl, A. (1998), Seismic Behavior and Design of Gusset Plates, Steel Tips, Structural
Steel Educational Council, Moraga, CA.
Astaneh-Asl, A. (2001), Seismic Behavior and Design of Steel Shear Walls, Steel Tips,
Structural Steel Educational Council, Moraga, CA.
Astaneh-Asl, A., Cochran, M. L., and Sabelli, R. (2006), Seismic Detailing of Gusset Plates
for Special Concentrically Braced Frames, Steel Tips, Structural Steel Education Council,
Moraga, CA.
Astaneh-Asl, A., and Goel, S. C. (1984), “Cyclic In-Plane Buckling of Double Angle Brac-
ing,” ASCE J. Struct. Eng., Vol. 110, No. 9, pp. 2036–2055.
Astaneh-Asl, A., Goel, S. C., and Hanson, R. D. (1985), “Cyclic Out-of-Plane Buckling of
Double-Angle Bracing,” ASCE J. Struct. Eng., Vol. 111, No. 5, pp. 1135–1153.
Astaneh-Asl, A., Goel, S. C., and Hanson, R. D. (1986), “Earthquake-Resistant Design of
Double-Angle Bracing,” AISC Eng. J., Vol. 23, No. 4, pp. 133–147.
ATC (2008), ATC-63, Recommended Methodology for Quantification of Building System
Performance and Response Parameters, 90 % Interim Draft Report , Applied Technology
Council, Redwood City, CA.
Auricchio, F., Fugazza, D., and Desroches, R. (2006), “Earthquake Performance of Steel
Frames with Nitinol Braces,” J. Earthquake Eng., Vol. 10, No. 1, pp. 45–66.
Azuhata, T., Ishihara, T., Midorikawa, M., and Wada, M. (2006), “Seismic Response Reduc-
tion of Steel Frames with Multi-Spans by Applying Rocking Structural System,” paper
presented at the Proc. STESSA 2006 Conf., Yokohama, Japan, pp. 659–664.
Ballio, G., and Perotti, F. (1987), “Cyclic Behaviour of Axially Loaded Members: Numerical
Simulation and Experimental Verification,” J. Constr. Steel Res., Vol. 7, No. 1, pp. 3–41.
Berman, J. W., and Bruneau, M. (2003), “Plastic Analysis and Design of Steel Plate Shear
Walls,” ASCE J. Struct. Eng., Vol. 129, No. 11, pp. 1448–1456.
Berman, J. W., and Bruneau, M. (2004), “Steel Plate Shear Walls Are Not Plate Girders,”
AISC Eng. J., Vol. 41, No. 3, pp. 95–106.
Berman, J. W., and Bruneau, M. (2005), “Experimental Investigation of Light-Gauge Steel
Plate Shear Walls,” ASCE J. Struct. Eng., Vol. 131, No. 2, pp. 259–267.
Berman, J. W., and Bruneau, M. (2008a), “Capacity Design of Vertical Boundary Elements
in Steel Plate Shear Walls,” AISC Eng. J., Vol. 45, No. 3, pp. 57–71.
Berman, J. W., and Bruneau, M. (2008b), “Tubular Links for Eccentrically Braced Frames
I: Finite Element Parametric Study; II: Experimental Verification,” ASCE J. Struct. Eng.,
Vol. 134, No. 5, pp. 692–712.
Bernal, D. (1987), “Amplification Factors for Inelastic Dynamics P− Effects in Earthquake
Analysis,” Earthquake Eng. Struct. Dynamics, Vol. 15, No. 5, pp. 635–651.
Bernal, D. (1992), “Instability of Buildings Subjected to Earthquakes,” ASCE J. Struct. Eng.,
Vol. 118, No. 8, pp. 2239–2260.
Bernal, D. (1998), “Instability of Building During Seismic Response,” Eng. Struct., Vol. 20,
Nos. 4–6, pp. 496–502.
912 STABILITY UNDER SEISMIC LOADING

Bernal, D., Nasseri, A., and Bulut, Y. (2006), “Instability Inducing Potential of Near Fault
Ground Motions,” in Proc. of the SMIP06 Seminar on Utilization of Strong-Motion Data
(Ed. M. Huang ), California Strong Motion Instrumentation Program, Oakland, CA, pp.
41–62.
Bertero, V. V., Anderson, J., and Krawinkler, H. (1994), “Performance of Steel Build-
ing Structures During the January 17, 1994 Northridge Earthquake,” Report No.
UBC/EERC-94/09, Earthquake Engineering Research Center, Univ. of California, Berke-
ley, CA.
Bertero, V. V., Uang, C.-M., Llopiz, C. R., and Igarashi, K. (1989), “Earthquake Simulator
Testing of Concentric Braced Dual System,” ASCE J. Struct. Eng., Vol. 115, No. 8, pp.
1877–1894.
Bhowmick, A. K., Driver, R. G., and Grondin, G. Y. (2009), “Seismic Analysis of Steel
Plate Shear Walls Considering Strain Rate and P–Delta Effects,” J. Constr. Steel Res.,
Vol. 65, No. 5, pp. 1149–1159.
Black, C. J., Makris, N., and Aiken, I. D. (2004), “Component Testing, Seismic Evaluation
and Characterization of Buckling-Restrained Braces,” ASCE J. Struct. Eng., Vol. 130,
No. 6, pp. 880–894.
Black, R. G., Wenger, W. A. B., and Popov, E. P. (1980), “Inelastic Buckling of Steel Struts
under Cyclic Load Reversals”, Report UCB/EERC-80/40, Earthquake Eng. Research
Center, Univ. of California, Berkeley, CA.
Bozorgnia, Y., and Campbell, K. W. (2004), “Engineering Characterization of Ground
Motions,” in Earthquake Engineering—From Engineering Seismology to Performance-
Based Engineering (Eds. Y. Bozorgnia and V. V. Bertero), CRC Press, Boca Raton, FL,
Chap. 5.
Broderick, B. M., Elghazouli, A. Y., and Goggins, J. (2008), “Earthquake Testing and
Response Analysis of Concentrically-Braced Sub-Frames,” J. Constr. Steel Res., Vol. 64,
No. 9, pp. 997–1007.
Broderick, B. M., Goggins, J. M., and Elghazouli, A. Y. (2005), “Cyclic Performance of Steel
and Composite Bracing Members,” J. Constr. Steel Res., Vol. 61, No. 4, pp. 493–514.
Bruneau, M., Uang, C.-M., and Whittaker, A. S. (1998), Ductile Design of Steel Structures,
McGraw-Hill, New York.
Byfield, M. P., Davies, J. M., and Dhanalakshmi, M. (2005), “Calculation of the Strain
Hardening Behaviour of Steel Structures Based on Mill Tests,” J. Constr. Steel Res., Vol.
61, No. 2, pp. 133–150.
Caccese, V., Elgaaly, M., and Chen, R. (1993), “Experimental Study of Thin Steel-Plate
Shear Walls Under Cyclic Load,” ASCE J. Struct. Eng.., Vol. 119, No. 2, pp. 573–587.
Calderoni, B., Giubileo, C., and De Martino, A. (2006), “Assessment of Hysteretic Cyclic
Behaviour of Plastic Hinge in Cold-Formed Steel Beams,” Proc. STESSA 2006 , Yoko-
hama, Japan, pp. 185–190.
Calderoni, B., De Martino, A. Formisano, A., and Fiorino, L. (2009), “Cold-Formed Steel
Beams Under Monotonic and Cyclic Loading: Experimental Investigation,” J. Constr.
Steel Res., Vol. 65 No. 1, pp. 219–227.
Carter, C. J. (2003), Stiffening of Wide-Flange Columns at Moment Connections: Wind and
Seismic Applications, Steel Design Guide 13, 2nd printing, American Institute of Steel
Construction, Chicago, IL.
REFERENCES 913

Castiglioni, C. A., (2005), “Effects of the Loading History on the Local Buckling Behavior
and Failure Mode of Welded Beam-to-Column Joints in Moment-Resisting Steel Frames,”
ASCE J. Eng. Mech., Vol. 6, No. 1, pp. 568–585.
Castiglioni, C. A., Mouzakis, H. P., and Carydis, P. G. (2007), “Constant and Variable
Amplitude Cyclic Behavior of Welded Steel Beam-to-Column Connections,” J. Earth-
quake Eng., Vol. 11, No. 6, pp. 876–902.
CEN. (2002), prEN 1993-1-8:E, Eurocode 3: Design of Steel Structures—Part 1.8 : Design
of Joints, Comité Européen de Normalisation, (CEN), European Committee for Stan-
dardization, Brussels, Belgium.
CEN. (2004), EN 1998-1:2004:E, Eurocode 8: Design of Structures for Earthquake
Resistance—Part 1: General Rules, Seismic Actions and Rules for Buildings, Comité
Européen de Normalisation, (CEN), European Committee for Standardization, Brussels,
Belgium.
Chambers, J. J., and Ernst, C. J. (2004), Brace Frame Gusset Plate Research—Phase I
Literature Review , Report, Dept. of Civil and Environmental Eng., University of Utah,
UT.
Chao, S. H., and Goel, S. C. (2006), “Performance-Based Seismic Design of Eccentrically
Braced Frames Using Target Drift and Yield Mechanism as Performance Criteria,” AISC
Eng. J., Vol. 43, No. 3, pp. 173–200.
Chao, S.-H., Khandelwal, K., and El-Tawil, S. (2006), “Ductile Web Fracture Initiation in
Steel Shear Links,” ASCE J. Struct. Eng., Vol. 132, No. 8, pp. 1192–1200.
Charney, F. A., and Atlayan, O. (2008), “Hybrid Moment Resisting Steel Frames,” Paper
No. 05-06-0164, Proc. 14th World Conf. on Earthquake Eng., Beijing, China.
Chen, C.-H., Lai, J. W., and Mahin, S. A., (2008), “Seismic Performance Assessment of
Concentrically Braced Steel Frame Buildings,” Paper No. 05-05-0158, Proc. 14th World
Conf. on Earthquake Eng., Beijing, China.
Chen, S.-J., and Tu, C.-T. (2004), “Experimental Study of Jumbo Size Reduced Beam
Section Connections Using High-Strength Steel,” ASCE J. Struct. Eng., Vol. 130, No. 4,
pp. 582–587.
Chen, Z., Ge, H., and Usami, T. (2006), “Hysteretic Model of Stiffened Shear Panel
Dampers,” ASCE J. Struct. Eng., Vol. 132, No. 3, pp. 478–483.
Cheng, F. Y., and Tseng, W. H. (1973), “Dynamic Instability and Ultimate Capacity of
Inelastic Systems Parametrically Excited by Earthquakes: Part 1,” Tech. Report, National
Science Foundation, August (available from U.S. Department of Commerce, National
Technical Information Service, Springfield, VA, NTIS Access No. PB261096jAS).
Chi, B., and Uang, C.-M. (2002), “Cyclic Response and Design Recommendations of
Reduced Beam Section Moment Connections with Deep Columns,” ASCE J. Struct.
Eng., Vol. 128, No. 4, pp. 464–473.
Chi, W.-M., Kanvinde, A. M., and Deierlein, G. G. (2006), “Prediction of Ductile Fracture
in Steel Connections Using SMCS Criterion,” ASCE J. Struct. Eng., Vol. 132, No. 2, pp.
171–181.
Choi, J., Stojadinovic, B., and Goel, S. C. (2003), “Design of Free Flange Connections,”
AISC Eng. J., Vol. 40, No. 1, pp. 24–41.
Chopra, A. K. (2001), Dynamics of Structures—Theory and Applications to Earthquake
Engineering, 2nd ed., Prentice Hall, Upper Saddle River, NJ.
914 STABILITY UNDER SEISMIC LOADING

Chou, C.-C., and Chen, P.-J. (2009), “Compressive Behavior of Central Gusset Plate Con-
nections for a Buckling-Restrained Braced Frame,” J. Constr. Steel Res., Vol. 65, No. 5,
pp. 1138–1148.
Christopoulos, C., Filiatrault, A., and Folz, B. (2002a), “Seismic Response of Self-Centering
Hysteretic SDOF Systems,” Earthquake Eng. Struct. Dyn., Vol. 31, No. 5, pp. 1131–1150.
Christopoulos, C., Filiatrault, A., Uang, C. M., and Folz, B. (2002b), “Post-Tensioned Energy
Dissipating Connections for Moment-Resisting Steel Frames,” ASCE J. Struct. Eng., Vol.
128, No. 9, pp. 1111–1120.
Christopoulos, C., Tremblay, R., Kim, H.-J., and Lacerte, M. (2008), “Development and
Validation of a Self-Centering Energy Dissipative (SCED) Bracing System for the Seismic
Resistance of Structures,” ASCE J. Struct. Eng., Vol. 134, No. 1, pp. 96–107.
Chusilp, P., and Usami, T. (2002), “New Elastic Stability Formulas for Multiple-Stiffened
Shear Panels,” ASCE J. Struct. Eng., Vol. 128, No. 6, pp. 833–836.
Ciutina, A. L., and Dubina, D. (2008), “Column Web Stiffening of Steel Beam-to-Column
Joints Subjected to Seismic Actions,” ASCE J. Struct. Eng., Vol. 134, No. 3, pp. 505–510.
Cochran, M. L., and Honeck, W. C. (2004), Design of Special Concentric Braced Frames
(with Comments on Ordinary Concentric Braced Frames), Steel Tips, Structural Steel
Education Council, Moraga, CA.
CSA. (2001), “CAN/CSA-S16.1 Limits States Design of Steel Construction, Including
CSA-S16S1-05 Supplement No. 1,” Canadian Standards Association, Willowdale, ON.
Davaran, A. (2001), “Effective Length Factor for Discontinuous X-Bracing Systems,” ASCE
J. Struct. Eng., Vol. 127, No. 2, pp. 106–112.
Davaran, A., and Far, N. E. (2009), “An Inelastic Model for Low Cycle Fatigue Prediction
in Steel Braces,” J. Constr. Steel Res., Vol. 65, No. 3, pp. 523–530.
Davaran, A., and Hoveidae, N. (2009), “Effect of Mid-Connection Detail on the Behavior
of X-Bracing Systems,” J. Constr. Steel Res., Vol. 65, No. 2, pp. 290–298.
Davidson, B. J., and Fenwick, R. C. (2004), “Proposed P-Delta Design for the Australian and
New Zealand Loadings Standard,” Proc. 2004 Annual Stability Conf., Structural Stability
Research Council, Long Beach, CA, pp. 581–602.
Della Corte, G., De Matteis, G., Landolfo, R., and Mazzolani, F. M. (2002), “Seismic
Analysis of MR Steel Frames Based on Refined Hysteretic Models of Connections,” J.
Constr. Steel Res., Vol. 58, No. 10, pp. 1331–1345.
Ding, Y., Zhang, Y., and Zhao, J. (2009), “Tests of Hysteretic Behavior for Unbonded Steel
Plate Brace Encased in Reinforced Concrete Panel,” J. Constr. Steel Res., Vol. 65, No.
5, pp. 1160–1170.
Di Sarno, L., and Elnashai, A. S. (2009), “Bracing Systems for Seismic Retrofitting of Steel
Frames,” J. Constr. Steel Res., Vol. 65, No. 2, pp. 452–465.
Di Sarno, L., Elnashai, A. S., Nethercot, D. A. (2008), “Seismic Response of Stainless Steel
Braced Frames,” J. Constr. Steel Res., Vol. 64, Nos. 7–8, pp. 914–925.
Dolce, M., Cardone, D., and Marnetto, R. (2000), “Implementation and Testing of Passive
Control Devices Based on Shape Memory Alloys,” Earthquake Eng. Struct. Dyn., Vol.
29, No. 7, pp. 945–968.
Dowswell, B. (2006), “Effective Length Factors for Gusset Plate Buckling,” AISC Eng. J.,
Vol. 43, No. 2, pp. 91–101.
Driver, R. G., Kulak, G. L., Kennedy, D. J. L., and Elwi, A. E. (1998), “Cyclic Test of
Four-Story Steel Plate Shear Wall,” ASCE J. Struct. Eng., Vol. 124, No. 2, pp. 112–120.
REFERENCES 915

Dubina, D. (2004), “Ductility and Seismic Performance of Thin-Walled Cold-Formed Steel


Structures,” Int. J. Steel Struct., Vol. 4, No. 4, pp. 209–222.
Eatherton, M. R., Hajjar, J. F., Deierlein, G. G., Krawinkler, H., Billington, S., and
Ma, X. (2008), “Controlled Rocking of Steel-Framed Buildings with Replaceable
Energy-Dissipating Fuses,” Paper No. 05-06-0026, Proc. 14th World Conf. on Earthquake
Eng., Beijing, China.
Elchalakani, M., Zhao, X.-L., and Grzebieta, R. (2003), “Tests of Cold-Formed Circular
Tubular Braces under Cyclic Axial Loading,” ASCE J. Struct. Eng., Vol. 129, No. 4, pp.
507–514.
El-Tawil, S., and Deierlein, G. G. (1999), “Strength and Ductility of Concrete Encased
Composite Columns,” ASCE J. Struct. Eng., Vol. 125, No. 9, pp. 1009–1019.
El-Tawil, S., Vidarsson, E., Mikesell, T., and Kunnath, S. K. (1999), “Inelastic Behavior
and Design of Steel Panel Zones,” ASCE J. Struct. Eng., Vol. 125, No. 2, pp. 183–193.
El-Tayem, A. A., and Goel, S. C. (1986), “Effective Length Factor for the Design of
X-Bracing Systems,” AISC Eng. J., Vol. 23, No. 1, pp. 41–45.
Engelhardt, M. D., and Popov, E. P. (1992), “Experimental Performance of Long Links
in Eccentrically Braced Steel Frames,” ASCE J. Struct. Eng., Vol. 118, No. 11, pp.
3067–3088.
Engelhardt, M. D., Winneberger, T., Zekany, J., and Potyraj, T. J. (1998), “Eexperimental
Investigation of Dog Bone Moment Connections,” AISC En. J., Vol. 35, No. 4, pp.
128–139.
Fahmy, M., Goel, S. C., and Stojadinovic, B. (2000), “Hysteretic Behavior of Steel Moment
Resisting Column Bases,” Proc. STESSA 2000 Conf., Montreal, Canada, pp. 191–197.
Fajfar, P. (1999), “Capacity Spectrum Method Based on Inelastic Demand Spectra,” Earth-
quake Eng. Struct. Dyn., Vol. 28, No. 9, pp. 979–993.
FEMA. (2000a), FEMA-350, Recommended Seismic Design Criteria for New Steel
Moment-Frame Buildings, SAC Joint Venture for the Federal Emergency Management
Agency, Washington, DC.
FEMA. (2000b), FEMA 356 Prestandard and Commentary for the Seismic Rehabilitation of
Buildings, ASCE, Federal Emergency Management Agency, Washington, DC.
Fell, B. V., Kanwinde, A. M., Deierlein, G. G., and Myers, A. T. (2009), “Experimental
Investigation of Inelastic Cyclic Buckling and Fracture of Steel Braces,” ASCE J. Struct.
Eng., Vol. 135, No. 1, pp. 1–19.
Fell, B. V., Kanwinde, A. M., Deierlein, G. G., Myers, A. T., and Fu, X. (2006), Buckling
and Fracture of Concentric Braces under Inelastic Cyclic Loading, Steel Tips, Structural
Steel Education Council, Moraga, CA.
Fenwick, R. C., Davidson, B. J., and Chung, B. T. (1992), “P-Delta Actions in Seismic
Resistant Structures,” Bull. New Zeal. Nat. Soc. Earthquake Eng., Vol. 25, No. 1, pp.
56–69.
Formisano, A., Faggiano, B., Landolfo, R., and Mazzolani, F. M. (2006), “Ductile
Behavioural Classes of Steel Members for Seismic Design,” Proc. STESSA 2006 Conf.,
Yokohama, Japan, pp. 225–232.
Foutch, D. A., Goel, S. C., and Roeder, C. W. (1987), “Seismic Testing of Full-Scale Steel
Building—Part I,” ASCE J. Struct. Eng., Vol. 113, No. 11, pp. 2111–2129.
916 STABILITY UNDER SEISMIC LOADING

Fukumoto, Y., and Itoh, Y. (1992), “Width-to-Thickness Ratios for Plate Elements in
Earthquake Engineering Design of Steel Structures,” in Stability and Ductility of Steel
Structures Under Cyclic Loading (Ed., Y. Fukumoto and G. Lee), CRC Press, Boca
Raton, FL, pp. 285–296.
Fukumoto, Y., and Lee, G. (Eds.) (1992), Stability and Ductility of Steel Structures Under
Cyclic Loading, CRC Press, Boca Raton, FL.
Fukuta, T., Nishiyama, I., Yamanouchi, H., and Kato, B. (1989), “Seismic Performance
of Steel Frames with Inverted V Braces,” ASCE J. Struct. Eng., Vol. 115, No. 8, pp.
2016–2028.
Garlock, M. M., Ricles, J. M., Sause, R., and Zhao, C. (2005), “Experimental Studies of
Full-Scale Posttensioned Steel Connections,” ASCE J. Struct. Eng., Vol. 131, No. 3, pp.
438–448.
Garlock, M. M., Ricles, J. M., and Sause, R. (2003), “Cyclic Load Tests and Analysis of
Bolted Top-and-Seat Angle Connections,” ASCE J. Struct. Eng., Vol. 129, No. 12, pp.
1615–1625.
Ger, J.-F., Cheng., F. Y., and Lu, L.-W. (1993), “Collapse Behavior of Pino Suarez Build-
ing During 1985 Mexico City Earthquake,” ASCE J. Struct. Eng., Vol. 119, No. 3, pp.
852–870.
Ghersi, A., Marino, E. M., and Neri, F. (2006), “A Seismic Design Method for High Ductility
Steel Frames,” Proc. STESSA 2006 Conf., Yokohama, Japan, pp. 33–38.
Ghersi, A., Neri, F., Rossi, P. P., and Perretti, A. (2000), “Seismic Response of Tied and
Trussed Eccentrically Braced Frames,” Proc. STESSA 2000 Conf., Montreal, Canada, pp.
495–502.
Ghersi, A., Pantano, S., and Rossi, P. P. (2003), “On the Design of Tied Braced Frames,”
Proc. STESSA 2003 Conf., Naples, Italy, pp. 413–419.
Ghobarah, A., Korol, R. M., and Osman, A. (1992), “Cyclic Behavior of Extended End-Plate
Joints,” ASCE J. Struct. Eng., Vol. 118, No. 5, pp. 1333–1353.
Gioncu, V. (2000), “Framed Structures. Ductility and Seismic Response. General Report,”
J. Constr. Steel Res., Vol. 55, Nos. 1–3, pp. 125–154.
Gioncu, V., and Petcu, D. (1997), “Available Rotation Capacity of Wide-Flange Beams and
Beam-Columns: Part 1. Theoretical Approaches and Part 2. Experimental and Numerical
Tests,” J. Constr. Steel Res., Vol. 43, Nos. 1–3, pp. 161–244.
Goel, S. C., and Chao, S.-H. (2008), Performance-Based Plastic Design: Earthquake-
Resistant Steel Structures, ICC Publications, Country Club Hills, IL.
Goel, R. K., and Chopra, A. K. (2004), “Evaluation of Modal and FEMA Pushover Analyses:
SAC Buildings,” Earthquake Spectra, Vol. 20, No. 1, pp. 225–254.
Goel, S. C., and El-Tayem, A. A. (1986), “Cyclic Load Behavior of Angle X-Bracing,”
ASCE J. Struct. Eng. J., Vol. 112, No. 11, pp. 2528–2539.
Goel, S. C., and Xu, P. (1989), “Brace Connections for Ductility,” Proc. ASCE Structures
Congress, San Francisco, CA, pp. 559–568.
Goggins, J. M., Broderick, B. M., Elghazouli, A. Y., and Lucas, A. S. (2006), “Behaviour
of Tubular Steel Members Under Cyclic Axial Loading,” J. Constr. Steel Res., Vol. 62,
Nos. 1–2, pp. 121–131.
Graesser, E. J., and Cozzarelli, F. A. (1991), “Shape-Memory Alloys as New Materials for
Aseismic Isolation,” ASCE J. Eng. Mech., Vol. 117, No. 11, pp. 2590–2608.
REFERENCES 917

Green, P. S., Sause, R., and Ricles, J. M. (2002), “Strength and Ductility of HPS Flexural
Members,” J. Constr. Steel Res., Vol. 58, Nos. 5–8, pp. 907–941.
Gugerli, H. (1982), “Inelastic Cyclic Behavior of Steel Bracing Members”, Ph.D. thesis,
University of Michigan, Ann Arbor, MI.
Guo, B., Gu, Q., and Liu, F. (2006), “Experimental Behavior of Stiffened and Unstiffened
End-Plate Connections under Cyclic Loading,” ASCE J. Struct. Eng., Vol. 132, No. 9,
pp. 1352–1357.
Gupta, A., and Krawinkler, H. (2000a). “Estimation of Seismic Drift Demands for Frame
Structures,” Earthquake Eng. Struct. Dyn., Vol. 29, No. 9, pp. 1287–1305.
Gupta, A., and Krawinkler, H. (2000b). “Dynamic P− Effects for Flexible Inelastic Steel
Structures,” ASCE J. Struct. Eng., Vol. 126, No. 1, pp. 145–154.
Haddad, M. A. (2004), “Design of Concentrically Braced Steel Frames for Earthquakes,”
Ph.D. thesis, Dept. of Civil Engineering, University of Calgary, Calgary, Alberta, Canada.
Han, S.-W., Kim, W. T., and Foutch, D. A. (2007), “Seismic Behavior of HSS Bracing
Members according to Width–Thickness Ratio under Symmetric Cyclic Loading,” ASCE
J. Struct. Eng., Vol. 133, No. 2, pp. 264–273.
Haselton, C. B., and Baker, J. W. (2006), “Ground Motion Intensity Measures for Collapse
Capacity Predictions: Choice of Optimal Spectral Period and Effect of Spectral Shape,”
Paper No. 461, Proc. 8th U.S. Nat. Conf. Earthquake Eng., San Francisco, CA.
Hassan, O. F., and Goel, S. C. (1991), “Modeling of Bracing Members and Seismic Behavior
of Concentrically Braced Steel Structures,” Research Report UMCE 91-1, Dept. of Civil
Engineering, University of Michigan, Ann Arbor, MI.
Hjelmstad, K. D., and Popov, E. P. (1983), “Cyclic Behavior and Design of Link beams,”
ASCE J. Struct. Eng., Vol. 109, No. 10, pp. 2387–2403.
Hjelmstad, K. S., and Lee, S.-G. (1989), “Lateral Buckling of Beams in Eccentrically-Braced
Frames,” J. Constr. Steel Res., Vol. 14, No. 4, pp. 251–272.
Huang, Y., and Mahin, S. A. (2008), “Evaluation of Steel Structure Deterioration with Cyclic
Damaged Plasticity,” Paper No. 14-0240, Proc. 14th World Conf. on Earthquake Eng.,
Beijing, China.
Huang, Z., Varmat, A. H., Arichadran, R. S., and Cordero, A. (2004), “Effects of Inelas-
tic Local Buckling on the Seismic Behavior of Members and Frames,” Proc. 2004
Annual Stability Conference, Structural Stability Research Council, Long Beach, CA,
pp. 329–351.
Humar, J., Mahgoub, M., and Ghorbanie-Asl, M. (2006), “Effect of Second-Order Forces
on Seismic Response,” Can. J. Civ. Eng., Vol. 33, No. 6, pp. 692–706.
Husid, R. (1967), “Gravity Effects on the Earthquake Response of Yielding Structures,”
Ph.D. thesis, California Institute of Technology, Pasadena, CA.
Ibarra, L. F., and Krawinkler, H. (2005), “Global Collapse on Frame Structures Under
Seismic Excitations,” Report No. 152, John A. Blume Earthquake Engineering Research
Center, Stanford University, San Francisco, CA.
Ibarra, L. F., Medina, R. A., and Krawinkler, H. (2005), “Hysteretic Models That Incorporate
Strength and Stiffness Deterioration,” Earthquake Eng. Struct. Dyn., Vol. 34, No. 12, pp.
1489–1511.
Ikenaga, M., Nagae, T., Nakashima, M., and Suita, K. (2006), “Development of Column
Bases Having Self-Centering and Damping Capability,” Proc. STESSA 2006 Conf., Yoko-
hama, Japan, pp. 703–709.
918 STABILITY UNDER SEISMIC LOADING

Inai, E., Mukai, A., Kai, M., Tokinoya, H., Fukumoto, T., and Mori, K. (2004), “Behavior
of Concrete-Filled Steel Tube Beam Columns,” ASCE J. Struct. Eng., Vol. 130, No. 2,
pp. 189–202.
Inoue, K., Sawaizumi, S., and Higashibata, Y. (2004), “Stiffening Requirements for
Unbonded Braces Encased in Concrete Panels,” ASCE J. Struct. Eng., Vol. 127, No.
6, pp. 712–719.
Inoue, K., Suita, K., Takeuchi, I., Chusilp, P., Nakashima, M., and Zhou, F. (2006),
“Seismic-Resistant Weld-Free Steel Frame Buildings with Mechanical Joints and Hys-
teretic Dampers,” ASCE J. Struct. Eng., Vol. 132, No. 6, pp. 864–872.
Ishihara, T., Azuhata, T., and M. Midorikawa, M. (2006), “Effect of Dampers on Dynamic
Behavior of Structures Allowed to Uplift,” Proc. STESSA 2006 Conf., Yokohama, Japan,
pp. 709–715.
Ishizawa, T., and Iura, M. (2005), “Analysis of Tubular Steel Bridge Piers,” Earthquake
Eng. Struct. Dyn., Vol. 34, No. 8, pp. 985–1004.
Iwata, M., Kato, T., and Wada, A. (2000), “Buckling Restrained Braces as Hysteretic
Dampers,” Proc. STESSA 2000, Montreal, Canada, pp. 33–38.
Iwata, M., and Murai, M. (2006), “Buckling-Restrained Brace Using Steel Mortar Planks;
Performance Evaluation as a Hysteretic Damper,” Earthquake Eng. Struct. Dyn., Vol. 35,
No. 14, pp. 1807–1826.
Jain, A. K., Goel, S. C., and Hanson, R. D. (1978), “Inelastic Response of Restrained Steel
Tubes,” ASCE J. Struct. Div., Vol. 104, No. ST6, pp. 897–910.
Jain, A. K., Goel, S. C., and Hanson, R. D. (1980), “Hysteretic Cycles of Axially Loaded
Steel Members,” ASCE J. Struct. Div., Vol. 106, No. ST8, pp. 1777–1795.
Jennings, P. C., and Husid, R. (1968), “Collapse of Yielding Structures During Earthquakes,”
ASCE J. Eng. Mech. Div., Vol. 94, No. 5, pp. 1045–1065.
Jin, J., and El-Tawil, S. (2005), “Seismic Performance of Steel Frames with Reduced Beam
Section Connections,” J. Constr. Steel Res., Vol. 61, No. 4, pp. 453–471.
Jones, S. L., Fry, G. T., and Engelhardt, M. D. (2002), “Experimental Evaluation of Cycli-
cally Loaded Reduced Beam Section Moment Connections,” ASCE J. Struct. Eng., Vol.
128, No. 4, pp. 441–451.
Ju, Y. K., Kim, M.-H., Kim, J., and Kim, S.-D. (2009), “Component Tests of Buckling-
Restrained Braces with Unconstrained Length,” Eng. Struct., Vol. 31, No. 2, pp. 507–516.
Kanvinde, A. M., and Deierlein, G. G. (2007), “A Cyclic Void Growth Model to Assess
Ductile Fracture in Structural Steels Due to Ultra Low Cycle Fatigue,” ASCE J. Eng.
Mech., Vol. 133, No. 6, pp. 701–712.
Karavasilis, T. L., Bazeos, N., and Beskos, D. E. (2007), “Estimation of Seismic Drift and
Ductility Demands in Planar Regular X-Braced Steel Frames,” Earthquake Eng. Struct.
Dyn., Vol. 36, No. 15, pp. 2273–2289.
Kasai, K., and Han, X. (1997), “New EBF Design Method and Application: Redesign and
Analysis of US-Japan EBF,” Proc. STESSA 1997 Conf., Kyoto, Japan, pp. 242–249.
Kasai, K., and Popov, E. P. (1986a), “General Behavior of WF Steel Shear Link Beams,”
ASCE J. Struct. Eng., Vol. 112, No. 2, pp. 362–382.
Kasai, K., and Popov, E. P. (1986b), “Cyclic Web Buckling Control for Shear Link Beams,”
ASCE J. Struct. Eng., Vol. 112, No. 3, pp. 505–523.
Kato, B. (1989), “Rotation Capacity of H-Section Members as Determined by Local Buck-
ling,” J. Constr. Steel Res., Vol. 13, Nos. 2-3, pp. 95–109.
REFERENCES 919

Kawashima, K., Hasegawa, K., MacRae, G. A. and Ikeuchi, T. (1992), “Ductility of Steel
Bridge Piers from Dynamic Loading Tests,” in Stability of Steel Structures Under Cyclic
Loading (Eds. T. Fukumoto and G. C. Lee), CRC Press, Boca Raton, FL.
Kawashima, K., MacRae, G. A., Hoshikuma, J. I., and Nagaya, K. (1998), “Residual Dis-
placement Response Spectrum,” ASCE J. Struct. Eng., Vol. 124, No. 6, pp. 523–530.
Kazantzi, A. K., Righiniotis, T. D., and Chryssanthopoulos, M. K. (2008), “The Effect of
Joint Ductility on the Seismic Fragility of a Regular Moment Resisting Steel Frame
Designed to EC8 Provisions,” J. Constr. Steel Res., Vol. 64, No. 9, pp. 987–996.
Kemp, A. R. (1996), “Inelastic Local and Lateral Buckling in Design Codes,” ASCE J.
Struct. Eng., Vol. 122, No. 4, pp. 374–382.
Khatib, I. F., Mahin, S. A., and Pister, K. S. (1988), “Seismic Behavior of Concentrically
Braced Steel Frames.” Report UCB/EERC-88/01, Earthquake Engineering Research Cen-
ter, University of California, Berkeley, CA.
Kiggins, S., and Uang, C.-M. (2006), “Reducing Residual Drift of Buckling-Restrained
Braced Frames as a Dual System,” Eng. Struct., Vol. 28, No. 11, pp. 1525–1532.
Kim, H. I., and Goel, S. C. (1996). “Upgrading of Braced Frames for Potential Local
Failures,” ASCE J. Struct. Eng., Vol. 122, No. 5, pp. 470–475.
Kimura, Y., and MacRae, G. A. (2006), “Effect of Column Flexural Characteristic on Seis-
mic Behavior of Braced Frame with Fixed Column Base,” Proc. STESSA 2006 Conf.,
Yokohama, Japan, pp. 437–443.
Kitipornchai, S., and Finch, D. L. (1986), “Stiffness Requirements for Cross Bracing,” ASCE
J. Struct. Eng., Vol. 112, No. 12, pp. 2702–2707.
Koboevic, S., and Redwood, R. (1997), “Design and Seismic Response of Shear Critical
Eccentrically Braced Frames,” Can. J. Civ. Eng., Vol. 24, No. 5, pp. 761–777.
Koetaka, Y., Kinoshita, T., Inoue, K., and Iitani, K. (2008), “Criteria of Buckling-Restrained
Braces to Prevent Out-of-Plane Buckling,” Paper no. 05-05-0121, Proc. 14th World Conf.
Earthquake Eng., Beijing, China.
Krawinkler, H. (2000), “State of the Art Report on Systems Performance of Steel Moment
Frames Subject to Earthquake Ground Shaking,” Report No. FEMA-355C, SAC Joint
Venture for the Federal Emergency Management Agency, Washington, DC.
Krawinkler, H. (2006), “Importance of Good Nonlinear Analysis,” Struct. Design Tall Spec.
Build., Vol. 15, pp. 515–531.
Krawinkler, H., and Mohasseb, S. (1987), “Effects of Panel Zone Deformations on Seismic
Response,” J. Constr. Steel Res., Vol. 8, 1987, pp. 233–250
Krawinkler, H., and Seneviratna, G. D. P. K. (1998), “Pros and Cons of a Pushover Analysis
for Seismic Performance Evaluation,” J. Eng. Struct., Vol. 20, Nos. 4–6, pp. 452–464.
Krawinkler, H., and Zareian, F. (2007), “Prediction of Collapse—How Realistic and Prac-
tical Is It, and What Can We Learn From It?” Struct. Design Tall Spec. Build., Vol. 16,
pp. 633–653.
Kuwamura, H. (1992), “High Performance Steels for Earthquake Resistant Building Struc-
tures in Japan,” In: Stability and Ductility of Steel Structures Under Cyclic Loading (Eds.
Y. Fukumoto and G. Lee), CRC Press, Boca Raton, FL, pp. 49–60.
Lacerte, M., and Tremblay, R. (2006), “Making Use of Brace Overstrength to Improve the
Seismic Response of Multi-Storey Split-X Concentrically Braced Steel Frames,” Can. J.
Civ. Eng., Vol. 33, No. 8, pp. 1005–1021.
920 STABILITY UNDER SEISMIC LOADING

Lee, D., Cotton, S. C., Hajjar, J. F., Dexter, R. J., and Ye, Y. (2005), “Cyclic Behav-
ior of Steel Moment-Resisting Connections Reinforced by Alternative Column Stiffener
Details. I. Connection Performance and Continuity Plate Detailing,” AISC Eng. J., Vol. 42,
No. 4, pp. 189–214.
Lee, G. C., and Lee, E. T. (1994), “Local Buckling of Steel Sections Under Cyclic Loading,”
J. Constr. Steel Res., Vol. 29, Nos. 1–3, pp. 55–70.
Lee, K., and Bruneau, M. (2005), “Energy Dissipation of Compression Members in Con-
centrically Braced Frames,” ASCE, J. Struct. Eng., Vol. 131, No. 4, pp. 552–559.
Lee, K., and Bruneau, M. (2008), “Seismic Vulnerability Evaluation of Axially Loaded Steel
Built-Up Laced Members I: Experimental Results & II: Evaluation,” Earthquake Eng.
Eng. Vib., Vol. 7, No. 2, pp.113–136.
Lee, K., and Foutch, D. A. (2002), “Performance Evaluation of New Steel Frame Buildings
for Seismic Loads,” Earthquake Eng. Struct. Dyn., Vol. 31, No. 3, pp. 653–670.
Lee, K., and Stojadinović, B. (2003), “Seismic Bracing Requirements for US Steel Moment
Connections,” Proc. 2003 Annual Technical Session & Meeting, Structural Stability
Research Council, Baltimore, MD, pp. 85–104.
Lee, K., and Stojadinović, B. (2007), “Rotation Capacity of Pre-Qualified Moment Connec-
tions: A Yield Line Approach,” Paper No. 34, Proc. ASCE Structures Congress 2007,
Long Beach, CA.
Lee, S., and Goel, S. C. (1987), “Seismic Behavior of Hollow and Concrete-Filled Square
Tubular Bracing Members,” Research Report UMCE 87-11, Dept. of Civil Engineering,
University of Michigan, Ann Arbor, MI.
Leelataviwat, S., Goel, S. C., and Stojadinovic, B. (1999), “Toward Performance-Based
Seismic Design of Structures,” Earthquake Spectra, Vol. 15, No. 3, pp. 435–461.
Lehman, D. E., Roeder, C. W., Herman, D., Johnson, S., and Kotulka, B. (2008), “Improved
Seismic Performance of Gusset Plate Connections,” ASCE J. Struct. Eng., Vol. 134, No.
6, pp. 890–901.
Liao, K. W., Wen, Y.-K., and Foutch, D. A. (2007), “Evaluation of 3D Steel Moment Frames
under Earthquake Excitations. I: Modeling,” ASCE J. Struct. Eng., Vol. 133, No. 3, pp.
462–470.
Liel, A. B., Haselton, C. B., Deierlein, G. G., and Baker, J. W. (2009), “Incorporating
Modeling Uncertainties in the Assessment of Seismic Collapse Risk of Buildings,” Struct.
Safety, Vol. 31, No. 2, pp. 197–211.
Lignos, D. G., Krawinkler, H., and Whittaker, A. S. (2008), “Shaking Table Collapse Tests of
Two Scale Models of a 4-Story Moment Resisting Steel Frame,” Paper No. 12-01-0186,
Proc. 14th World Conf. Earthquake Eng., Beijing, China.
Lin, C.-H., Tsai, K.-C., Wei, C.-Y., Wang, K.-J., Tsai, C.-Y., Wu, A.-C., Powell, J., Clark,
K., and Roeder, C. (2009), “Cyclic Tests of a Full-Scale 2-Story Steel Concentrically
Braced Frame,” Proc. 5th Int. Symposium on Steel Structures, Seoul, Korea (in press).
Liu, Z., and Goel, S. C. (1987), “Investigation of Concrete-Filled Steel Tubes Under Cyclic
Bending and Buckling,” Research Report No. 87-3, Department of Civil Engineering,
University of Michigan, Ann Arbor, MI.
Liu, Z., and Goel, S. C. (1988), “Cyclic Load Behavior of Concrete-Filled Tubular Braces,”
ASCE J. Struct. Eng., Vol. 114, No. 7, pp. 1488–1506.
López, W. A., Gwie, D. S, Lauck, T. W., and Saunders, M. (2004), “Structural Design and
Experimental Verification of a Buckling-Restrained Braced Frame System,” AISC Eng.
J., Vol. 41, No. 4, pp. 177–186.
REFERENCES 921

López, W. A., and Sabelli, R. (2004), Seismic Design of Buckling-Restrained Braced Frames,
Steel Tips, Structural Steel Education Council, Moraga, CA.
Lubell, A. S., Prion, H. G., Ventura, C., and Rezai, M. (2000), “Unstiffened Steel Plate
Shear Wall Performance under Cyclic Loading,” ASCE, J. Struct. Eng., Vol. 126, No. 4,
pp. 453–460.
Ma, N., Wu, B., Zhao, J., Li, H., Ou, J., and Yang, W. (2008), “Full Scale Test of All-Steel
Buckling Restrained Braces,” Paper No. 11-0208, Proc. 14th World Conf. Earthquake
Eng., Beijing, China.
MacRae, G. A. (1990), “The Seismic Response of Steel Frames,” Research Report No. 90-6,
University of Canterbury, Christchurch, New Zealand
MacRae, G. A. (1994), “P-Delta Effects on Single-Degree-of-Freedom Structures in Earth-
quakes,” Earthquake Spectra, Vol. 10, No. 3, pp. 539–568.
MacRae, G. A., Kimura, Y., and Roeder, C. (2004), “Effect of Column Stiffness on Braced
Frame Seismic Behavior,” ASCE J. Struct. Eng., Vol. 130, No. 3, pp. 381–391.
MacRae, G. A., and Tagawa, H. (2001), “Seismic Behavior of 3D Steel Moment Frame with
Biaxial Columns,” ASCE J. Struct. Eng., Vol. 127, No. 5, pp. 490–497.
Malley, J., and Popov, E. P. (1984), “Shear Links in Eccentrically Braced Steel Frames,”
ASCE J. Struct. Eng., Vol. 110, No. 9, pp. 2275–2295.
Mamaghani, I. H. P. (2008), “Seismic Design and Ductility Evaluation of Thin-Walled Steel
Bridge Piers of Box Sections,” Transportation Res. Record: J. Transportation Res. Board ,
No. 2050, pp. 137–142.
Mansour, N., Christopoulos, C., and Tremblay, R. (2006), “Seismic Design of EBF Steel
Frames Using Replaceable Nonlinear Links,” Proc. STESSA 2006 Conf., Yokohama,
Japan, pp. 745–750.
Marino, E. M., Ghersi, A., and Nakashima, M. (2006), “Design of Chevron Braced Frames
Based on Ultimate Lateral Strength I: Theory; II Evaluation of the Behavior Factor,”
Proc. STESSA 2006 Conf., Yokohama, Japan, pp. 51–65.
Marino, E. M., and Nakashima, M. (2006), “Seismic Performance and New Design Procedure
for Chevron-Braced Frames,” Earthquake Eng. Struct. Dyn., Vol. 35, No. 4, pp. 433–452.
Marson, J., and Bruneau, M. (2004), “Cyclic Testing of Concrete-Filled Circular Steel Bridge
Piers Having Encased Fixed-Based Detail,” ASCE J. Bridge Eng., Vol. 9, No. 1, pp.
14–23.
Martinelli, L., Mulas, M. G., and Perotti, F. (1996), “The Seismic Response of Concentrically
Braced Moment-Resisting Steel Frames,” Earthquake Eng. Struct. Dyn., Vol. 25, No. 11,
pp. 1275–1299.
Martinelli, L., Perotti, F., and Bozzi, A. (2000), “Seismic Demand and Response of a
14-Story Concentrically Braced Steel Buildings,” Proc. STESSA 2000 Conf., Montreal,
Canada, pp. 327–334.
Mastrandrea, L., Montuori, R., and Piluso, V. (2003), “Failure Mode Control of Seismic
Resistant EB-Frames,” Proc. STESSA 2003 Conf., Naples, Italy, pp. 435–441.
Matsui, R., Takeuchi, T., Hajjar, J. F., Nishimoto, K., and Aiken, I. (2008), “Local Buck-
ling Restraint Condition for Core Plates in Buckling Restrained Braces,” Paper No.
05-05-0055, Proc. 14th World Conf. On Earthquake Eng., Beijing, China.
Mazzolani, F. M., Georgescu, D., and Astaneh-Asl, A. (1994), “Remarks on Behaviour
of Concentrically and Eccentrically Braced Steel Frames,” Proc. STESSA 1994 Conf.,
Timisoara, Romania, pp. 310–322.
922 STABILITY UNDER SEISMIC LOADING

Mazzolani, F. M., and Piluso, V. (1993), “P− Effect on Seismic Resistant Steel Structures,”
Proc. 1993 Annual Technical Session, Structural Stability Research Council, Milwaukee,
WI, pp. 271–282.
Mazzolani, F. M., and Piluso, V. (1996), Theory and Design of Seismic Resistant Steel
Frames, E & FN Spon, London.
Mazzolani, F. M., and Piluso, V. (1997), “Platic Design of Seismic Resistant Steel Frames,”
Earthquake Eng. Struct. Dyn., Vol. 26, No. 2, pp. 167–191.
McCormick, J., DesRoches, R., Fugazza, D., and Auricchio, F. (2007), “Making Seismic
Assessment of Concentrically Braced Steel Frames with Shape Memory Alloy Braces,”
ASCE J. Struct. Eng., Vol. 133, No. 6, pp. 862–870.
McGuire, W., Gallagher, R. H., and Ziemian, R. D. (2000), Matrix Structural Analysis, 2nd
ed., John Wiley and Sons, Inc., Hoboken, NJ.
Medina, R. A., and Krawinkler, H. (2005) “Strength Demand Issues Relevant for the Seismic
Design of Moment-Resisting Frames,” Earthquake Spectra, Vol. 21, No. 2, pp. 415–439.
Mehanny, S. S. F., and Deierlein, G. G. (2001), “Seismic Damage and Collapse Assessment
of Composite Moment Frames, ASCE J Struct. Eng., Vol. 127, No. 9, pp. 1045–1053.
Merzouq, S., and Tremblay, R. (2006), “Seismic Design of Dual Concentrically Braced
Steel Frames for Stable Seismic Performance for Multi-Storey Buildings,” Paper No.
1909, Proc. 8th U.S. Nat. Conf. on Earthquake Eng., San Francisco, CA.
Midorikawa, M., Azuhata, T., Ishira, T., and Wada, A. (2006), “Shaking Table Tests on
Seismic Response of Steel Braced Frames with Column Uplift,” Earthquake Eng. Struct.
Dyn., Vol. 35, No. 14, pp. 1767–1785.
Midorikawa, M., Toyomaki, S., Hori, H., Asari, T., Azuhata, T., and Ishihara, T. (2008),
“Seismic Response of Six-Story Eccentrically Braced Steel Frames with Column Partially
Allowed to Uplift,” Paper No. ST17-03-006, Proc. 14th World Conf. on Earthquake Eng.
Miranda, E., and Akkar, S. D. (2003), “Dynamic Instability of Simple Structural Systems,”
ASCE J. Struct. Eng., Vol. 129, 1, pp. 1722–1726.
Mitchell, D., Tremblay, R., Karacabeyli, E., Paultre, P., Saatcioglu, M., and Anderson, D.
(2003), “Seismic Force Modification Factors for the Proposed 2005 NBCC,” Can. J. Civ.
Eng., Vol. 30, No. 2, pp. 308–327.
Moldovan, A., Petcu, D., and Gioncu, V. (2000), “Calibration of Thin-Walled Members
Ductility,” Proc. STESSA 2000 Conf., Montreal, Canada, pp. 63–71.
Montgomery, C. J. (1981), “Influence of P-Delta Effects on Seismic Design,” Can. J. Civ.
Eng., Vol. 8, No. 1, pp. 31–43.
Montgomery, J., and Medhekar, M. (2001), “Unstiffened Steel Plate Shear Wall Performance
under Cyclic Loading : Discussion,” ASCE J. Struct. Eng., Vol. 127, No. 8, p. 973.
Moon, J., Yoon, K.-Y., Han, T.-S., and Lee, H.-E., (2008), “Out-of-Plane Buckling and
Design of X-Bracing Systems with Discontinuous Diagonals,” J. Constr. Steel Res., Vol.
64, No. 3, pp. 285–294.
Moslehi Tabar, A., and Deylam, A. (2006), “Investigation of Major Parameters Affect-
ing Instability of Steel Beams with RBS Moment Connections,” Steel and Composite
Structures, Vol. 6, No. 3, pp. 203–219.
Murray, T. M., and Sumner, E. A. (2003), Extended End-Plate Moment
Connections—Seismic and Wind Applications, 2nd ed., AISC Design Guide No.
4, American Institute of Steel Construction, Chicago, IL.
REFERENCES 923

Nakashima, M., Kanao, I., and Liu, D. (2002), “Lateral Instability and Lateral Bracing of
Steel Beams Subjected to Cyclic Loading,” ASCE J. Struct. Eng., Vol. 128, No. 10, pp.
1308–1316.
Nakashima, M., and Liu, D. (2005), “Instability and Complete Failure of Steel Columns
Subjected to Cyclic Loading,” ASCE J. Eng. Mech., Vol. 131, No. 6, pp. 559–567.
Nakashima, M., Liu, D., and Kanao, I. (2003), “Lateral–Torsional and Local Instability of
Steel Beams Subjected to Large Cyclic Loading,” Int. J. Steel Struct., Vol. 3, No. 3, pp.
179–189.
Nakashima, M., Matsumiya, T., Suita, K., and Liu, D. (2006), “Test on Full-Scale
Three-Storey Steel Moment Frame and Assessment of Ability of Numerical Simula-
tion to Trace Cyclic Inelastic Behavior,” Earthquake Eng. Struct. Dyn., Vol. 35, No. 1,
pp. 3–19.
Nakashima, M., and Wakabayashi, M. (1992), “Analysis and Design of Steel Braces and
Braced Frames in Building Structures,” in Stability and Ductility of Steel Structures under
Cyclic Loading (Eds.) Y. Fukumoto and G. C. Lee), CRC Press, Boca Raton, FL, pp.
309–321.
Nast, T. E., Grondin, G. Y., and Cheng, R. J. J. (1999), Cyclic Behaviour of Stiffened
Gusset Plate-Brace Member Assemblies, Struct. Eng. Report No. 229, Dept. of Civil
Engineering, University of Alberta, Edmonton, Alberta, Canada.
Neri, F. (2000), “Influence of P− Effect on Proposed Procedure for Seismic Design of
Steel Frames,” Proc. STESSA 2000 Conf., Montreal, Canada, pp. 511–518.
Neuss, C. F., and Maison, B. F. (1984), “Analysis for P− Effects in Seismic Response of
Buildings,” Comput. Struct., Vol. 19, pp. 369–380.
Newell, J. D., and Uang, C.-M. (2008), “Cyclic Behavior of Steel Wide-Flange Columns
Subjected to Large Drift,” ASCE J. Struct. Eng., Vol. 134, No. 8, pp. 1334–1342.
Nip, K. H., Gardner, L., and Elghazouli, A. Y. (2008), “Comparative Study of Stainless
Steel and Carbon Steel Tubular Members Subjected to Cyclic Loading,” Proc. XII Int.
Symposium on Tubular Structures, Shanghai, pp. 219–226.
NRCC (2005), National Building Code of Canada, 12th ed., National Research Council of
Canada Ottawa, Canada.
NZS (2007), NZ 3404:Parts 1 and 2:1997, Steel Structures Standard, Incorporating Amend-
ment No. 1 and Amendment No. 2(2007), New Zealand Standards, Wellington, New
Zealand.
NZS (2004), NZ 1170.5:2004, Structural Design Actions, Part 5: Earthquake Actions—New
Zealand , New Zealand Standards, Wellington, New Zealand.
Ocel, J. DesRoches, R., Leon, R. T., Hess, W. G., Krumme, R., Hayes, J. R., and Sweeney,
S. (2004), “Steel Beam-Column Connections Using Shape Memory Alloys,” ASCE J.
Struct. Eng., Vol. 130, No. 5, pp. 732–740.
Ohsaki, M., Kinoshita, T., and Pan, P. (2007), “Multiobjective Heuristic Approaches to
Seismic Design of Steel Frames with Standard Sections,” Earthquake Eng. Struct. Dyn.,
Vol. 36, No., pp. 1481–1495.
Okazaki, T., Arce, G., Ryu, C., and Engelhardt, M. D. (2005), “Experimental Study of Local
Buckling, Overstrength, and Fracture of Links in Eccentrically Braced Frames,” ASCE
J. Struct. Eng., Vol. 131, No. 10, pp. 1526–535.
924 STABILITY UNDER SEISMIC LOADING

Okazaki, T., and Engelhardt, M. D. (2007), “Cyclic Loading Behavior of EBF Links Con-
structed of ASTM A992 Steel,” J. Constr. Steel Res., Vol. 63, No. 6, pp. 751–765.
Okazaki, T., Liu, D., Nakashima, M., and Engelhardt, M. (2006), “Stability Requirements
for Beams in Seismic Steel Moment Frames,” ASCE J. Struct. Eng., Vol. 132, No. 9, pp.
1334–1342.
Osteraas, J., and Krawinkler, H. (1989), “The Mexico Earthquake of September 19,
1985—Behavior of Steel Buildings,” Earthquake Spectra, Vol. 5, No. 1, pp. 51–88.
Osteraas, J., and Krawinkler, H. (1990), “Strength and Ductility Considerations in Seismic
Design”, Rep. No. 90, The John A. Blume Earthquake Engineering Research Center,
Dept. of Civil and Environmental Engineering, Stanford University, Stanford, CA.
Park, H.-G., Kwack, J.-H., Jeon, S.-W., Kim, W.-K., and Choi, I.-R. (2007), “Framed Steel
Plate Wall Behavior under Cyclic Lateral Loading,” ASCE J. Struct. Eng., Vol. 133, No.
3, pp. 378–388.
Park, K., and Medina, R. A. (2007), “Conceptual Seismic Design of Regular Frames Based
on the Concept of Uniform Damage,” ASCE J. Struct. Eng., Vol. 133, No. 7, pp. 945–955.
Park, Y. S., Iwai, S., Kameda, H., and Nonaka, T. (1996), “Very Low Cycle Failure Process
of Steel Angle Members,” ASCE J. Struct. Eng., Vol. 122, No. 2, pp. 133–141.
Petcu, D., and Gioncu, V. (2003), “Computer Program for Available Ductility Analysis of
Steel Structures,” Comput. Struct., Vol. 81, Nos. 22–23, pp. 2149–2164.
Pettinga, J. D., and Priestley, M. J. N. (2007), Accounting for P-Delta Effects in Structures
When Using Direct Displacement-Based Design, Research Report No. ROSE -2007/02,
IUSS Press, Pavia, Italy.
Picard, A., and Beaulieu, D. (1987), “Design of Diagonal Cross Bracings Part 1: Theoretical
Study,” AISC Eng. J., Vol. 24, No. 3, pp. 122–126.
Picard, A., and Beaulieu, D. (1988), “Design of Diagonal Cross-Bracings Part 2: Experi-
mental Study,” AISC Eng. J., Vol. 25, No. 4, pp. 156–160.
Plumier, A. (1997), “The Dogbone: Back to the Future,” AISC Eng. J., Vol. 34, No. 2, pp.
61–67.
Plumier, A. (2000), “General Report on Local Ductility,” J. Constr. Steel Res., Vol. 55, Nos.
1–3, pp. 91–107.
Pollino, M., and Bruneau, M. (2007), “Experimental Study of the Controlled Rocking
Response of Steel Braced Frames,” Paper No. 46, Proc. ASCE Struct. Congress, Long
Beach, CA.
Pollino, M., and Bruneau, M. (2008), “Dynamic Seismic Response of Controlled Rocking
Bridge Steel-Truss Piers,” Eng. Struct., Vol. 30, No. 6, pp. 1667–1676.
Pons, H. F. (1997), “Dynamic Behavior of Tubular Bracing Members with Single Plate
Concentric Connections,” M.Sc. Thesis, University of Toronto, Toronto, Ontario Canada.
Popov, E. P., Bertero, V. V., and Chandramouli, S. (1975), “Hysteretic Behavior of Steel
Columns,” Report No. UCB/EERC-75/11, Earthquake Engineering Research Center, Uni-
versity of California, Berkeley, CA.
Popov, E. P., and Black, R. G. (1981), “Steel Struts Under Severe Cyclic Loadings,” ASCE
J. Struct. Eng. Div., Vol. 107, No. ST9, pp. 1857–1881.
Popov, E. P., and Pinkney, R. B. (1967), “Behavior of Steel Building Connec-
tions Subjected to Repeated Inelastic Strain Reversal, Experimental Data,” Reports
REFERENCES 925

UCB/SESM-1967/30&31, Dept. of Civil Engineering, University of California, Berkeley,


CA.
Popov, E. P., and Stephen, R. M., (1970), “Cyclic Loading of Full Size Steel Connections,”
UCB/EERC Report 70-3, University of California, Berkeley, CA.
Porter, F. L., and Powell, G. H. (1971), “Static and Dynamic Analysis of Inelastic Struc-
tures,” Report No. EERC 71-3, Earthquake Engineering Research Center, University of
California, Berkeley, CA.
Qu, B., Bruneau, M., Lin, C.-H., and Tsai, K.-C. (2008), “Testing of Full-Scale Two-Story
Steel Plate Shear Wall with Reduced Beam Section Connections and Composite Floors,”
ASCE J. Struct. Eng., Vol. 134, No. 3, pp. 364–373.
Rai, D. C., and Goel, S. C. (2003), “Seismic Evaluation and Upgrading of Chevron Braced
Frames,” J. Constr. Steel Res., Vol. 59, No. 8, pp. 971–994.
Redwood, R. G., and Channagiri, V. S. (1991), “Earthquake Resistant Design of Concentri-
cally Braced Steel Frames,” Can. J. Civ. Eng., Vol. 18, No. 5, pp. 839–850.
Remennikov, A. M., and Walpole, W. R. (1997), “Analytical Prediction of Seismic Behaviour
for Concentrically-Braced Steel Systems,” Earthquake Eng. Struct. Dyn., Vol. 26, No. 8,
pp. 859–874.
Remennikov, A. M., and Walpole, W. R. (1998a), “A Note on Compression Strength Reduc-
tion Factor for a Buckled Strut in Seismic-Resisting Braced System,” Eng. Struct., Vol.
20, No. 8, pp. 779–782.
Remennikov, A. M., and Walpole, W. R. (1998b), “Seismic Behavior and Deterministic
Design Procedures for Steel V-Braced Frames,” Earthquake Spectra, Vol. 14, No. 2, pp.
335–355.
Richards, P., Okazaki, T., Engelhardt, M., and Uang, C.-M. (2007), “Impact of Recent
Research Findings on Eccentrically Braced Frame Design,” AISC Eng. J., Vol. 44, No.
1, pp. 41–53.
Richards, P., and Uang, C.-M. (2005), “Effect of Flange Width-Thickness Ratio on Eccen-
trically Braced Frames Link Cyclic Rotation Capacity,” ASCE J. Struct. Eng., Vol. 131,
No. 10, pp. 1546–1552.
Richards, P. W. (2009), “Seismic Column Demands in Ductile Braced Frames,” ASCE J.
Struct. Eng., Vol. 135, No. 1, pp. 33–41.
Ricles, J. M., Sause, R., Peng, S. W., and Lu, L. W. (2002), “Experimental Evaluation of
Earthquake Resistant Posttensioned Steel Connections,” ASCE J. Struct. Eng., Vol. 128,
No. 7, pp. 850–859.
Ricles, J. M., and Paboojian, S. D. (1994), “Seismic Performance of Steel-Encased Com-
posite Columns,” ASCE J. Struct. Eng., Vol. 120, No. 8, pp. 2474–2494.
Ricles, J. M., and Popov, E. P. (1989), “Composite Action in Eccentrically Braced Frames,”
ASCE J. Struct. Eng., Vol. 115, No. 8, pp. 2046–2066.
Ricles, J. M., and Popov, E. P. (1994), “Inelastic Link Element for EBF Seismic Analysis,”
ASCE J. Struct. Eng., Vol. 120, No. 2, pp. 441–463.
Ricles, J. M., Sause, R., Garlock, M. M., and Zhao, C. (2001), “Post-Tensioned Seismic-
Resistant Connections for Steel Frames,” ASCE J. Struct. Eng., Vol. 127, No. 2,
pp. 113–121.
Ricles, J. M., Sause, R., and Green, P. S. (1998), “High Strength Steel: Implications of
Material and Geometric Characteristics on Inelastic Flexural Behavior,” Eng. Struct.,
Vol. 95, Nos. 4–6, pp. 323–335.
926 STABILITY UNDER SEISMIC LOADING

Roberts, T. M., and Sabouri-Ghomi, S. (1992), “Hysteretic Characteristics of Unstiffened


Perforated Steel Plate Shear Panels,” Thin Walled Struct., Vol. 14, No. 2, pp. 139–151.
Rodgers, J. E., and Mahin, S. A. (2006), “Effects of Connection Fractures on Global Behavior
of Steel Moment Frames Subjected to Earthquakes,” ASCE J Struct. Eng., Vol. 132, No.
1, pp. 78–88.
Rodgers, J. E., and Mahin, S. A. (2008), “Effects of Deformation Softening on Seismic
Behavior of Steel Moment Frames,” Paper No. 12-01-0093, Proc. 14th World Conf.
Earthquake Eng., Beijing, China.
Roeder, C. W. (1989), “Seismic Behavior of Concentrically Braced Frame,” ASCE J. Struct.
Eng., Vol. 115, No. 8, pp. 1837–1857.
Roeder, C. W. (2000), FEMA-355D State of the Art Report on Connection Performance,
prepared for the SAC Joint Venture, SAC Joint Venture, Federal Emergency Management
Agency, Washington, DC.
Roeder, C. W. (2002), “General Issues Influencing Connection Performance,” ASCE J. Struct.
Eng., Vol. 128, No. 4, pp. 420–428.
Roeder, C. W., Foutch, D. A., and Goel, S. C. (1987), “Seismic Testing of Full-Scale Steel
Building—Part II,” ASCE J. Struct. Eng., Vol. 113, No. 11, pp. 2130–2145.
Roeder, C. W., Schneider, S. P., and Carpenter, J. E. (1993), “Seismic Behavior of
Moment-Resisting Steel Frames: Analytical Study,” ASCE J. Struct. Eng., Vol. 119, No.
6, pp. 1866–1884.
Rossi, P. P. (2007), “A Design Procedure for Tied Braced Frames,” Earthquake Eng. Struct.
Dyn., Vol. 36, No. 14, pp. 2227–2248.
Rossi, P. P., and Lombardo, A. (2007), “Influence of the Link Overstrength Factor on the
Seismic Behavior of Eccentrically Braced Frames,” J. Constr. Steel Res., Vol. 63, No.
11, pp. 1529–1545.
Ruiz-Garcı́a, J., and Miranda, E. (2006), “Evaluation of Residual Drift Demands in Regu-
lar Multi-Storey Frames for Performance-Based Seismic Assessment,” Earthquake Eng.
Struct. Dyn., Vol. 35, No. 13, pp. 1609–1629.
Rutenberg, A., and De Stefano, M. (2000), “Approximate Stability Bounds on the Seismic
Force Reduction Factor,” Paper No. 1635, Proc. 12th World Conf. Earthquake Eng.,
Auckland, New Zealand.
Sabelli, R. (2001), “Research on Improving the Design and Analysis of Earthquake Resistant
Braced Frames,” 2000 NEHRP Fellowship in Earthquake Hazard Reduction, Final Report
to FEMA and EERI Earthquake Eng. Res. Institute, Oahland, CA.
Sabelli, R. (2004), “Recommended Provisions for Buckling-Restrained Braced Frames,”
AISC Eng. J., Vol. 41, No. 4, pp. 155–175.
Sabelli, R., and Bruneau, M. (2007), Steel Plate Shear Walls, Steel Design Guide 20, Amer-
ican Institute of Steel Construction, Chicago, IL.
Sabelli, R., and Hohbach, D. (1999), “Design of Cross-Braced Frames for Predictable Buck-
ling Behavior,” ASCE J. Struct. Eng., Vol. 125, No. 2, pp. 163–168.
Sabelli, R., Mahin, S., and Chang, C. (2003), “Seismic Demands on Steel Braced
Frame Buildings with Buckling Restrained Braces,” Eng. Struct., Vol. 25, No. 5,
pp. 656–666.
Sabouri-Ghomi, A., and Gholhaki, M. (2008), “Ductility of Thin Steel Plate Shear Walls,”
Asian J. Civ. Eng. (Building and Housing), Vol. 9, No. 2, pp. 153–166.
REFERENCES 927

Sabouri-Ghomi, S., Kharrazi M. H. K., Mam-Azizi, S.-E-D., and Sajadi, R. A. (2008),


“Buckling Behavior Improvement of Steel Plate Shear Wall Systems,” Struct. Design
Tall Special Buildings, Vol. 17, No. 4, pp. 823–837.
Sabouri-Ghomi, S., and Roberts, T. M. (1991), “Nonlinear Dynamic Analysis of Thin Steel
Plate Shear Walls,” Comput. Struct., Vol. 39, Nos. l–2, pp. 121–127.
Sakimoto, T., and Watanabe, H. (1997), “Simplified Methods of Analysis for Cyclic Behav-
ior of Steel Box Beam-Columns,” Proc. 1997 Annual Technical Meeting and Session,
Structural Stability Research Council, Toronto, Ontario, Canada, pp. 265–276.
Salmanpour, A. H., and Arbabi, F. (2008), “Seismic Reliability of Concentrically Braced
Steel Frame,” Paper No. 05-05-0062, Proc. 14th World Conf. on Earthquake Eng., Beijing,
China.
Sato, A., and Uang, C.-M. (2009), “Seismic Design Procedure Development for Cold-Formed
Steel—Special Bolted Moment Frames,” J. Constr. Steel Res., Vol. 65, No. 4, pp.
860–868.
Sause, R., Ricles, J. M., Roke, D., Seo, C.-Y., and Lee, K.-S. (2006), “Self-Centering Seismic
Resistant Steel Concentrically-Braced Frames,” Proc. STESSA 2006 Conf., Yokohama,
Japan, pp. 85–90.
Schellenberg, A., Yang, T. Y., Mahin, S. A., and Stojadinovic, B. (2008), “Hybrid Simulation
of Structural Collapse,” Paper No. S16-02-004, Proc. 14 World Conf. Earthauke Eng.,
Beijing, China.
Schneider, S. P., Roeder, C. W., and Carpenter, J. E. (1993), “Seismic Behavior of
Moment-Resisting Steel Frames: Experimental Investigation,” ASCE J. Struct. Eng., Vol.
119, No. 6, pp. 1885–1902.
Segal, F., Levy, R., and Rutenberg, A. (1994), “Design of Imperfect Cross-Bracings,” ASCE
J. Eng. Mech., Vol. 120, No. 5, pp. 1057–1075.
Sepúlveda, J., Boroschek, R., Herrera, R., Moroni, O., and Sarrazin, M. (2008), “Steel
Beam– Column Connection Using Copper-Based Shape Memory Alloy Dampers,” J.
Constr. Steel Res., Vol. 64, No. 4, pp. 429–435.
Shaback, J. B., and Brown, T. (2003), “Behaviour of Square Hollow Structural Steel Braces
with End Connections Under Reversed Cyclic Axial Loading,” Can. J. Civ. Eng., Vol.
30, No. 4, pp. 745–753.
Shen, J.-H. J., Astaneh-Asl, A., and McCallen, D. B. (2002), “Use of Deep Columns in
Special Steel Moment Frames,” Steel Tip Report No. 24, Structural Steel Educational
Council, Moraga, CA.
Shibata, M. (1988), “Hysteretic Behavior of Multi-Story Braced Frames,” Proc. 9th World
Conf. Earthquake Eng., Vol. IV, Tokyo, Japan, pp. 249–254.
Shishkin, J. J., Driver, R. G., and Grondin, G. Y. (2005), “Analysis of Steel Plate Shear
Walls Using the Modified Strip Model,” Struct. Eng. Report No. 261, Dept. of Civil
Engineering, University of Alberta, Edmonton, Alberta, Canada.
Spears, P. W., and Charney, F. A. (2006), “The Effect of Vertical Ground Accelerations on
the Dynamic Stability of Simple Structures,” Proc. 2006 Annual Stability Conference,
Structural Stability Research Council, San Antonio, TX, pp. 443–461.
928 STABILITY UNDER SEISMIC LOADING

Stoman, S. H. (1989), “Effective Length Spectra for Cross Bracings,” ASCE J. Struct. Eng.,
Vol. 115, No. 12, pp. 3112–3122.
Stratan, A., and Dubina, D. (2004), “Bolted Links for Eccentrically Braced Steel Frames,”
Proc. of Fifth Int. Workshop on Connections in Steel Structures V , Amsterdam, The
Netherlands, pp. 223–232.
Suita, K., Yamada, S., Tada, M., Kasai, K., Matsuoka, Y., and Sato, E. (2008), “Results
of Recent E-Defense Tests on Full-Scale Steel Buildings: Part 1—Collapse Experiments
on 4-Story Moment Frames,” Paper No. 266, Proc. ASCE Structures Congress 2008 ,
Vancouver, BC.
Sun, F. F., Liu, G. R., and Li, G. Q. (2008), “An Analytical Model for Composite Steel
Plate Wall,” Paper No. 05-05-0097. Proc. 14th World Conf. on Earthquake Conf., Beijing,
China.
Susantha, K. A. S., Ge, H., and Usami, T. (2002), “Cyclic Analysis and Capacity Prediction
of Concrete-Filled Steel Box Columns,” Earthquake Eng. Struct. Dyn., Vol. 31, No. 2,
pp. 195–216.
Swanson, J. A., and Leon, R. T. (2000), “Bolted Steel Connections: Tests on T-Stub Com-
ponents,” ASCE J. Struct. Eng., Vol. 126, No. 1, pp. 50–56.
Tada, M., Fukui, T., Nakashima, M., and Roeder, C. W. (2003), “Comparison of Strength
Capacity for Steel Building Structures in the United States and Japan,” Earthquake Eng.
Eng. Seismol., Vol. 4, No. 1, pp. 1–13.
Tada, M., Ohsaki, M., Yamada, S., Motoyui, S., and Kasai, K., (2007), “E-Defense Tests on
Full-Scale Steel Buildings Part 3—Analytical Simulation of Collapse,” Paper No. 19,
Proc. ASCE Structures Congress 2007, Long Beach, CA.
Tagawa, H., MacRae, G. A., and Lowes, L. (2004), “Evaluation and Mitigation of P−
Effects on 2D MDOF Frame Behavior,” Proc. 2004 Annual Stability Conference, Struc-
tural Stability Research Council, Long Beach, CA, pp. 499–518.
Takanashi, K. (1973), “Inelastic Lateral Buckling of Steel Beams Subjected to Repeated and
Reversed Loadings,” Proc. 5th World Conf. on Earthquake Eng., Vol. 1, Rome, Italy, pp.
795–798.
Tang, X., and Goel, S. C. (1989), “Brace Fractures and Analysis of Phase I Structure,” ASCE
J. Struct. Eng., Vol. 115, No. 8, pp. 1960–1976.
Thevendran, V., and Wang, C. M. (1993), “Stability of Nonsymmetric Cross-Bracing Sys-
tems,” ASCE J. Struct. Eng., Vol. 119, No. 1, pp. 169–180.
Thorburn, L. J., Kulak, G. L., and Montgomery, C. J. (1983), “Analysis of Steel Plate Shear
Walls,” Struct. Eng. Report No. 107, Dept. of Civil Engineering, University of Alberta,
Edmonton, Alberta, Canada.
Thornton, W. A. (1984), “Bracing Connections for Heavy Construction,” AISC Eng. J., Vol.
21, No. 3, pp. 139–148.
Timler, P. A., and Kulak, G. L. (1983), “Experimental Study of Steel Plate Shear Walls,”
Struct. Eng. Report No. 114, Dept. of Civil Engineering, University of Alberta, Edmonton,
Alberta, Canada.
Tjondro, J. A., Moss, P. J., and Carr, A. J. (1992), “Seismic P−δ Effects in Medium Height
Moment Resisting Steel Frames,” Eng. Struct., Vol. 14, No. 2, pp. 75–90.
REFERENCES 929

Tremblay, R. (2000), “Influence of Brace Slenderness on the Seismic Response of Concen-


trically Braced Steel Frames,” Proc. STESSA 2000 Conf., Montréal, Quebec, Canada, pp.
527–534.
Tremblay, R. (2002), “Inelastic Seismic Response of Steel Bracing Members,” J. Constr.
Steel Res., Vol. 58, Nos. 5–8, pp. 665–701.
Tremblay, R. (2003), “Achieving a Stable Inelastic Seismic Response for Concentrically
Braced Steel Frames,” AISC Eng. J., Vol. 40, No. 2, pp. 111–129.
Tremblay, R., Archambault, M. H., and Filiatrault, A. (2003), “Seismic Performance of
Concentrically Braced Steel Frames Made with Rectangular Hollow Bracing Members,”
ASCE J. Struct. Eng., Vol. 129, No. 12, pp. 1626–1636.
Tremblay, R., Bruneau, M., Nakashima, M., Prion, H. G. L., Filiatrault, A., and DeVall, R.
(1996), “Seismic Design of Steel Buildings: Lessons from the 1995 Hyogoken-Nanbu
Earthquake,” Can. J. Civ. Eng., Vol. 23, No. 3, pp. 727–756.
Tremblay, R., Côté, B., and Léger, P. (1999), “An Evaluation of P− Amplification Factors
in Multistorey Steel Moment Resisting Frames,” Can. J. Civ. Eng., Vol. 26, No. 5, pp.
535–548.
Tremblay, R., Haddad, M., Martinez, G., Richard, J., and Moffatt, K. (2008a), “Inelastic
Cyclic Testing of Large Size Steel Bracing Members,”, Paper No. 05-05-0071, Proc.
14th World Conf. Earthquake Eng., Beijing, China.
Tremblay, R., Lacerte, M., and Christopoulos, C. (2008b), “Seismic Response of
Multi-Storey Buildings with Self-Centering Energy Dissipative Steel Braces,” ASCE J.
Struct. Eng., Vol. 134, No. 1, pp. 108–120.
Tremblay, R., Poirier, L.-P., Bouaanani, N., Leclerc, M., René, V., Fronteddu, L., and Rivest,
S. (2008c), “Innovative Viscously Damped Rocking Braced Steel Frames,” Paper No.
05-01-0527, Proc. 14th World Conf. on Earthquake Eng., Beijing, China.
Tremblay, R., and Poncet, L. (2005), “Seismic Performance of Concentrically Braced Steel
Frames in Multi-Storey Buildings with Mass Irregularity,” ASCE J. Struct. Eng., Vol.
131, No. 9, pp. 1363–1375.
Tremblay, R., and Robert, N. (2000), “Design of Low- and Medium-Rise Chevron Braced
Steel Frames,” Can. J. Civ. Eng., Vol. 27, No. 6, pp. 1192–1206.
Tremblay, R. and Robert, N. (2001), “Seismic Performance of Low- and Medium-Rise
Chevron Braced Steel Frames,” Can. J. Civ. Eng., Vol. 28, No. 4, pp. 699–714.
Tremblay, R., and Stiemer, S. F. (1994), “Back-Up Stiffness for Improving the Stabil-
ity of Multi-Storey Braced Frames Under Seismic Loading,” Proc. 1994 SSRC Annual
Task Group Technical Session, Structural Stability Research Council, Bethlehem, PA, pp.
311–325.
Tremblay, R., Tchebotarev, N., and Filiatrault, A. (1997), “Seismic Performance of RBS
Connections for Steel Moment Resisting Frames: Influence of Loading Rate and Floor
Slab,” Proc. STESSA ’97 Conf., Kyoto, Japan, pp. 664–671.
Tremblay, R., Timler, P., Bruneau, M., and Filiatrault, A. (1995), “Performance of Steel
Structures During the January 17, 1994 Northridge Earthquake,” Can. J. Civ. Eng., Vol.
22, No. 2, pp. 338–360.
930 STABILITY UNDER SEISMIC LOADING

Tsai, K.-C., Chou, C.-C., Lin, C.-L., Chen, P.-C., and Jhang, S.-J. (2007), “Seismic
Self-Centering Steel Beam-to-Column Moment Connections Using Bolted Friction
Devices,” Earthquake Eng. Struct. Dyn., Vol. 37, No. 4, pp. 624–645.
Tsai, K.-C., and Hsiao, P.-C. (2008), “Pseudo-Dynamic Test of a Full-Scale CFT/BRB
Frame—Part II: Seismic Performance of Buckling-Restrained Braces and Connections,”
Earthquake Eng. Struct. Dyn., Vol. 37, No. 7, pp. 1099–1115.
Tsai, K. C., Wu, S., and Popov, E. P. (1995), “Experimental Performance of Seismic Steel
Beam-Column Moment Joints,” ASCE J. Struct. Eng., Vol. 121, No. 6, pp. 925–931.
Uang, C.-M., and Fan, C.-C. (2001), “Cyclic Stability Criteria for Steel Moment Connections
with Reduced Beam Section,” ASCE J. Struct. Eng., Vol. 127, No. 9, pp. 1021–1027.
Uang, C.-M., and Nakashima, M. (2004), “Steel Buckling-Restrained Braced Frames,” In
Earthquake Engineering—From Engineering Seismology to Performance-Based Engi-
neering (Eds. Y. Bozorgnia and V. V. Bertero), CRC Press, Boca Raton, FL,
Chap. 16.
Uetani, K., and Tagawa, H. (1998), “Criteria for Suppression of Deformation Concentration
of Building Frames Under Severe Earthquakes,” Eng. Struct., Vol. 20, Nos. 4–6, pp.
372–383.
Uriz, P. (2005), “Towards Earthquake Resistant Design of Concentrically Braced Steel Struc-
tures,” Ph.D. thesis. Dept. of Civil Engineering, University of California, Berkeley, CA.
Uriz, P., Filippou, F. C., and Mahin, S. A. (2008), “Model for Cyclic Inelastic Buckling of
Steel Braces,” ASCE Struct. Eng., Vol. 134, No. 4, pp. 619–628.
Uriz, P., and Mahin, S. A. (2004), “Seismic Vulnerability Assessment of Concentrically
Braced Steel Frames,” J. Int. Steel Struct., Vol. 4, No. 4, pp. 240–248.
Usami, T., Gao, S., and Ge, H. (2000), “Stiffened Steel Box Columns. Part 2: Ductility
Evaluation,” Earthquake Eng. Struct. Dyn., Vol. 29, No. 11, pp. 1707–1722.
Usami, T., Ge, H. B., and Kasai, A. (2008), “Overall Buckling Prevention Condition of
Buckling Restrained Braces as a Structural Control Damper,” Paper No. 05-05-0128,
Proc. 14th World Conf. on Earthquake Eng., Beijing, China.
Usami, T., Kasai, A., and Kato, M. (2003), “Behavior of Buckling-Restrained Brace Mem-
bers,” Proc. STESSA 2003 , Naples, Italy, pp. 211–216.
Vamvatsikos, D., and Cornell, C. A. (2002), “Incremental Dynamic Analysis,” Earthquake
Eng. Struct. Dyn., Vol. 31, No. 3, pp. 491–514.
Varma, A. H., Ricles, J. M., Sause, R., and Lu, L-W. (2002), “Seismic Behavior and Mod-
eling of High-Strength Composite Concrete-Filled Steel Tube (CFT) Beam-Columns,” J.
Constr. Steel Res., Vol. 58, Nos. 5–8, pp. 725–758.
Varma, A. H., Ricles, J. M. Sause, R., and Lu, L.-W. (2004), “Seismic Behavior and Design
of High-Strength Square Concrete-Filled Steel Tube Beam-Columns,” ASCE J. Struct.
Eng., Vol. 130, No. 2, pp. 169–179.
Vian, D., and Bruneau, M. (2003), “Tests to Structural Collapse of Single Degree of Freedom
Frames Subjected to Earthquake Excitations,” ASCE J. Struct. Eng., Vol. 129, No. 12,
pp. 1676–1685.
REFERENCES 931

Vian, D., and Bruneau, M. (2004), “Testing of Special LYS Steel Plate Shear Walls,” Paper
No. 978, Proc. 13th World Conf. on Earthquake Eng., Vancouver, Canada.
Villaverde, R. (2007), “Methods to Assess the Seismic Collapse Capacity of Building Struc-
tures: State of the Art,” ASCE J. Struct. Eng., Vol. 133, 1, pp. 57–66.
Wada, A., and Nakashima, M. (2004), “From Infancy to Maturity of Buckling Restrained
Braces Research,” Paper No. 1732, Proc. 13th World Conf. on Earthquake Eng., Vancou-
ver, Canada.
Walpole, W. R. (1996), “Behaviour of Cold-Formed Steel RHS Members Under Cyclic
Loading,” Research Rep. No. 96-4, University of Canterbury, Christchurch, New Zealand.
Wang, D. Q., and Boresi, A. P. (1992), “Theoretical Study of Stability Criteria for X-Bracing
Systems,” ASCE J. Eng. Mech., Vol. 118, No. 7, pp. 1357–1364.
Watanabe, A., Hitomi, Y., Saeki, E., Wada, A., and Fujimoto, M. (1988), “Properties of
Brace Encased in Buckling-Restraining Concrete and Steel Tube,” Proc. 9th World Conf.
on Earthquake Eng., Vol. IV, Tokyo, Japan, pp. 719–724.
Wei, C.-Y., and Tsai, K.-C. (2008), “Local Buckling of Buckling Restrained Braces,” Paper
No. 05-05-0139, Proc. 14th World Conf. on Earthquake Eng., Beijing, China.
Whittaker, A. S., Uang, C.-M., and Bertero, V. V. (1987), “Earthquake Simulation Tests
and Associated Studies of a 0.3-Scale Model of a Six-Story Eccentrically Braced Steel
Structure,” Report No. UCB/EERC-87/02, Earthquake Eng. Research Center, University
of California, Berkeley, CA.
Williamson, E. B. (2003), “Evaluation of Damage and P− Effects for Systems Under
Earthquake Excitation,” ASCE J. Struct. Eng., Vol. 129, No. 8, pp. 1036–1046.
Xie, Q. (2005), “State of the Art of Buckling-Restrained Braces in Asia,” J. Constr. Steel
Res., Vol. 61, No. 6, pp. 727–748.
Yamazaki, S., and Endo, K. (2004), “Stability Ratio and Dynamic P-D Effects in Inelas-
tic Earthquake Responses,” Paper No. 466, Proc. 13th World Conf. Earthquake Eng.,
Vancouver, British Columbia.
Yang, C. S., Leon, R. T., and Desroches, R. (2008a), “Pushover Response of a Braced Frame
with Suspended Zipper Struts,” ASCE J. Struct. Eng., Vol. 134, No. 10, pp. 1619–1626.
Yang, C. S., Leon, R. T., and Desroches, R. (2008b), “Design and Behavior of Zipper-Braced
Frames,” Eng. Struct., Vol. 30, No. 4, pp. 1092–1100.
Yoo, J.-H., Roeder, C. W., and Lehman, D. E. (2008), “Analytical Performance Simulation
of Special Concentrically Braced Frames,” ASCE J. Struct. Eng., Vol. 134, No. 6, pp.
881–889.
Zareian, F., and Krawinkler, H. (2007), “Assessment of Probability of Collapse and Design
for Collapse Safety,” Earthquake Eng. Struct. Dyn., Vol. 36, No. 13, pp. 1901–1914.
Zhang, X., and Ricles, J. M. (2006a), “Experimental Evaluation of Reduced Beam Section
Connections to Deep Columns,” ASCE J. Struct. Eng., Vol. 132, No. 3, pp. 346–357.
Zhang, X., and Ricles, J. M. (2006b), “Seismic Behavior of Reduced Beam Section Moment
Connections to Deep Columns,” ASCE J. Struct. Eng., Vol. 132, No. 3, pp. 358–367.
932 STABILITY UNDER SEISMIC LOADING

Zhang, Y., and Zhu, S. (2008), “Seismic Response Control of Building Structures with
Superelastic Shape Memory Alloy Wire Dampers,” ASCE J. Struct. Eng., Vol. 134, No.
3, pp. 240–251.
Zhao, Q., and Astaneh-Asl, A. (2004), “Cyclic Behavior of Traditional and Innovative Com-
posite Shear Walls,” ASCE J. Struct. Eng., Vol. 130, No. 2, pp. 271–284.
Zhao, X.-L., Grzebieta, R., and Lee, C. (2002), “Void-Filled Cold-Formed Rectangular Hol-
low Section Braces Subjected to Large Deformation Cyclic Axial Loading,” ASCE J.
Struct. Eng., Vol. 128, No. 6, pp. 746–753.
Zheng, Y., Usami, T., and Ge. H. (2000), “Ductility of Thin-Walled Steel Box Stub-
Columns,” ASCE J. Struct. Eng., Vol. 126, No. 11, pp. 1304–1311.
Zhu, S., and Zhang, Y. (2008), “Seismic Analysis of Concentrically Braced Frame Systems
with Self-Centering Friction Damping Braces,” ASCE J. Struct. Eng., Vol. 134, No. 1,
pp. 121–131.

Вам также может понравиться