You are on page 1of 18

EARTHQUAKE ENGINEERING AND STRUCTURAL DYNAMICS

Earthquake Engng Struct. Dyn. 2011; 40:789–806


Published online 30 September 2010 in Wiley Online Library (wileyonlinelibrary.com). DOI: 10.1002/eqe.1060

Three-dimensional modal pushover analysis of buildings subjected


to two components of ground motion, including its evaluation for
tall buildings

Juan C. Reyes1 and Anil K. Chopra2, ∗, †


1 Civil
and Environmental Engineering, Universidad de los Andes, Bogota, Colombia
2 Civil and Environmental Engineering, University of California, Berkeley, CA 94720-1710, U.S.A.

SUMMARY
The modal pushover analysis (MPA) procedure, presently restricted to one horizontal component of ground
motion, is extended to three-dimensional analysis of buildings—symmetric or unsymmetric in plan—
subjected to two horizontal components of ground motion, simultaneously. Also presented is a variant
of this method, called the practical modal pushover analysis (PMPA) procedure, which estimates seismic
demands directly from the earthquake response (or design) spectrum. Its accuracy in estimating seismic
demands for very tall buildings is evaluated, demonstrating that for nonlinear systems this procedure is
almost as accurate as the response spectrum analysis procedure is for linear systems. Thus, for practical
applications, the PMPA procedure offers an attractive alternative whereby seismic demands can be estimated
directly from the (elastic) design spectrum, thus avoiding the complications of selecting and scaling ground
motions for nonlinear response history analysis. Copyright 䉷 2010 John Wiley & Sons, Ltd.

Received 10 December 2009; Revised 22 July 2010; Accepted 23 July 2010

KEY WORDS: multi-component; pushover; tall buildings; two components

1. INTRODUCTION

The earthquake engineering profession has been moving away from traditional code procedures
to performance-based procedures for evaluating existing buildings and proposed designs of new
buildings. Both nonlinear response history analysis (RHA) and nonlinear static procedure (NSP)
are now used for estimating engineering demand parameters (EDPs)—floor displacements, story
drifts, internal forces, hinge rotations, etc.—in performance-based engineering of buildings.
Nonlinear RHA of the structure is usually implemented for a few (say, seven or less) ground
motion records, which are selected and scaled appropriately to obtain results close to the median
demand due to a large ensemble of ground motions. Selection and scaling of ground motions
is fraught with several unresolved issues. In contrast, NSP (or pushover) procedures estimate
seismic demands directly from the earthquake design spectrum for the building site, avoiding the
complications associated with selecting and scaling records.
To overcome the now well-known limitations of the NSP specified in most building evaluation
guidelines, researchers developed improved procedures [1–6]. Rooted in structural dynamics theory,
the modal pushover analysis (MPA) was developed to include the contributions of all ‘modes’
of vibration that contribute significantly to seismic demands [7]. The MPA procedure achieves
superior estimates of EDPs for steel and concrete moment-resisting-frame (MRF) buildings [8, 9],

∗ Correspondence to: Anil K. Chopra, Civil and Environmental Engineering, University of California, Berkeley, CA
94720-1710, U.S.A.
† E-mail: chopra@ce.berkeley.edu

Copyright 䉷 2010 John Wiley & Sons, Ltd.


790 J. C. REYES AND A. K. CHOPRA

while retaining the conceptual simplicity and computational attractiveness of standard nonlinear
static procedures. To date, the MPA procedure has been restricted to determine seismic demands
due to a single component of ground motion, although at least the two horizontal components of
ground motion should be considered in computing seismic demands by three-dimensional analysis
of multistory buildings. To address this issue, the first part of the paper extends the MPA procedure
[10] to three-dimensional analysis of buildings—symmetric or unsymmetric in plan—subjected to
two horizontal components of ground motion, simultaneously.
The second part of the paper is in response to the design and construction of several tall buildings
on the west coast of the U.S., a region known for its high seismicity. Because many of these
buildings use high-performance materials or exceed the height limits specified in building codes,
they are designed using an alternative performance-based procedure [11]. This approach entails
nonlinear RHA of a computer model of the building for an ensemble of multi-component ground
motions to guarantee serviceability and safety equivalent to that of the prescriptive code provisions
for ‘ordinary’ buildings. Based on a comprehensive investigation [12], this paper explores the
alternative of using the MPA procedure for estimating seismic demands for very tall buildings.

2. MPA FOR TWO COMPONENTS OF GROUND MOTION

2.1. Overall concept


The equations governing the displacements u of the N floor diaphragms (assumed to be rigid
in their own plane) of a building subjected to ground motion along two horizontal components
applied simultaneously are
Mü+cu̇+f S (u) = −Mix ü gx (t)−Mi y ü gy (t) (1)
in which M, a diagonal mass matrix of order 3N , includes three submatrices m, m, and Io ; m
is associated with the x-lateral and y-lateral degrees-of-freedom (DOFs), and Io with torsional
DOFs. c is the damping matrix and f S is the vector of resisting forces. The influence vectors ix
and i y associated with x and y ground motions are
⎡ ⎤ ⎡ ⎤
1 0
⎢ ⎥ ⎢ ⎥
ix = ⎣ 0⎦ , i y = ⎣ 1⎦ (2)
0 0
where 1 and 0 are vectors of dimension N with all elements equal to unity and zero, respectively.
Thus, the effective earthquake forces are
⎡ ⎤ ⎡ ⎤
m1 0
⎢ ⎥ ⎢ ⎥
peff (t) = −sx ü gx (t)−s y ü gy (t) = − ⎣ 0 ⎦ ü gx (t)− ⎣ m1⎦ ü gy (t) (3)
0 0
The spatial distribution of these forces over the building is defined by the vectors sx or s y ,
and the time variation by ü g (t) = ü gx (t) or ü gy (t). These force distributions can be expanded as a
summation of modal inertia force distributions sn [13]:

3N 
3N 
3N 
3N
sx = snx = nx Mn , sy = sny = ny Mn (4)
n=1 n=1 n=1 n=1

where n is the nth natural vibration mode of the system vibrating at small amplitudes (within its
linearly elastic range), consisting of three subvectors xn , yn , and n ,
T
Ln xn Mix for ü gx (t)
n = , Mn = Tn Mn , L n = (5)
Mn Tyn Mi y for ü gy (t)

Copyright 䉷 2010 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2011; 40:789–806
DOI: 10.1002/eqe
THREE-DIMENSIONAL MPA OF BUILDINGS 791

Thus,

peff,n (t) = −snx ü gx (t)−sny ü gy (t) (6)

is the nth-mode component of the effective earthquake forces.


In the MPA procedure, the peak response of the building to peff,n (t) = −snx ü gx (t)—or the peak
‘modal’ demand rnx —to one component of ground motion (say, the x-component) is determined
by a nonlinear static analysis of the structure subjected to lateral forces and torques defined by the
modal force distribution
⎡ ⎤
mxn
⎢ ⎥
s∗n = sign(n ) ⎢
⎣ m yn ⎦
⎥ (7)
I O n

at the peak roof displacement u r nx associated with the nth-‘mode’ inelastic SDF system; the
subscript x or y indicating the ground motion component is omitted from s∗n and n . Because the
signs of the peak modal demands rnx are crucial to the accuracy of the CQC rule, Equation (7) must
include the sign of n to obtain the correct algebraic sign for rnx . The peak modal demands rnx are
then combined by an appropriate modal combination rule, e.g. the CQC rule, to estimate the total
demand r x . The MPA procedure applied to linearly elastic systems subjected to one component of
ground motion is equivalent to the standard response spectrum analysis (RSA) procedure.
Such analyses are implemented for each component of the ground motion (x and y), inde-
pendently, and the responses due to individual components are combined using one of the
available multi-component combination rules [14–16]. This investigation selected the SRSS
multi-component combination rule assuming that the response quantities r x and r y are statistically
independent. This rule is accurate if ü gx (t) and ü gy (t) are along the principal axes of the ground
motion, or they are of equal intensity [17]. The first condition is usually not satisfied because
the recorded ground motion components were applied along the longitudinal and transverse axes
of the structure. Because the median response spectra for the two horizontal components of the
ensemble of ground motions considered in this investigation are of comparable intensity, the SRSS
rule is appropriate.

2.2. Step-by-step summary


The seismic demands—floors displacements and story drifts at the C.M.—for a symmetric- or
unsymmetric-plan building subjected to two horizontal components of ground motion can be
estimated by the MPA procedure, which is summarized in a step-by-step form (adapted from
Reference [12]):
1. Compute the natural frequencies, n , and modes, n , for linearly elastic vibration of the
building.
2. For the nth mode, develop the base shear–reference displacement, Vnx –u r nx , pushover curve
by nonlinear static analysis of the building and applying the force distribution s∗nx [Equation
(7)]. The reference point is located at the C.M. of the roof, but other floors may be chosen; the
component chosen is in the direction of the dominant motion in the mode being considered.
Gravity loads are applied before the lateral forces, causing roof displacement u rg .
3. Idealize the Vnx –u r nx pushover curve as a bilinear or trilinear curve, as appropriate, and
convert it into the force–deformation, Fsn /L n − Dn , relation for the nth-mode inelastic SDF
system by utilizing the following equations: Fsn /L n = Vn /Mn∗ and Dn = (u rn −u rg )/n rn ,
where Mn∗ is the effective modal mass of the nth-mode. Starting with this initial loading
curve, define the unloading and reloading branches appropriate for the structural system
and material being considered.
4. Calculate the reference displacement u r n,x = u rg +nx rn Dnx for the x-component of the
ground motion. The peak deformation Dnx of the nth mode inelastic SDF system, defined

Copyright 䉷 2010 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2011; 40:789–806
DOI: 10.1002/eqe
792 J. C. REYES AND A. K. CHOPRA

by the force–deformation relation developed in Step 3 and estimated damping ratio n , is


determined by nonlinear RHA.
5. From the pushover database (step 2), extract the response values rn+g,x due to the combined
effects of gravity and lateral forces at reference displacements u r n,x determined in Step 4.
6. Compute the dynamic response due to the nth ‘mode’: rnx =rn+g,x −r g , where rn+g,x are
the response quantities obtained in Step 5, and r g are the contributions of gravity loads
alone.
7. Repeat Steps 2–6 for as many modes as required for sufficient accuracy.
8. Determine the total dynamic responses for the x-component of the   ground motion by
combining the peak modal responses using the CQC rule, i.e. r x = ( i n inrix rnx )1/2 .
9. Repeat Steps 2–8 to determine the total dynamic response r y due to the y-component of
ground motion.
10. Combine

the responses r x and r y by the SRSS multi-component combination rule (i.e.
r = r x2 +r y2 ) to determine the dynamic response r , and, then, compute the total responses
r T =r g ±r ; where r g are the responses due to gravity loads.
Steps 1–10 lead to an estimate of floor displacements and story drifts at the C.M., but not for
member forces or end rotations.
The MPA procedure for estimating seismic demands for buildings subjected to one component
of ground motion has been extended to estimate member forces [18], but it does not lend itself to
automation in a large computer code. To overcome this disadvantage, this procedure is implemented
by first applying gravity loads to the structure, and then imposing a set of displacements that are
compatible with the MPA estimate of story drifts at the C.M. of the building model. This nonlinear
static analysis provides internal forces if both ends of an element develop plastic hinges. Otherwise,
internal forces are determined by implementing Steps 1–10 as described above.
The preceding procedure, extended to two components of ground motion, is implemented in the
next two steps:
11. Estimate other deformation quantities—such as story drifts—at locations other than the C.M.
and plastic hinge rotations from the story drifts. First apply gravity loads and then impose
at the C.M. a set of displacements u x and u y that are compatible with the drifts calculated
j j
in Step 10, i.e. u jx = i=1 Tjx and u jy = i=1 Tjy . Four combinations of these displacements
should be imposed: u x +u y , u x −u y , −u x +u y , −u x −u y . The largest of the resulting
four values of a response quantity is taken as the MPA estimate. Required for this purpose
is a computer program—such as SAP2000—that allows displacements to be imposed on the
structure instead of loads. In computer programs that do not allow imposing displacements—
such as PERFORM-3D—a model combining lateral forces with bidirectional ‘gap’ elements
at each floor could be implemented.
12. Using the plastic hinge rotations obtained in Step 11, internal forces are determined as follows.
If both ends of an element deform into the inelastic range, internal forces are those obtained
in Step 11. Otherwise, determine internal forces by implementing Steps 2–10. These results
are valid if the calculated bending moments, M, do not exceed the yield capacities, M y , of
the structural members. If M>M y , scale the internal forces by a factor equal to M y /M; this
situation may arise when the rotation calculated in Step 11 is just below the yield rotation.
The preceding computational steps were implemented for the two-component ground motions
and the median—defined as the geometric mean—of the resulting data set values for a response
quantity provides an estimate of the seismic demand.

2.3. Simplified MPA for practical applications


The MPA procedure summarized in Section 2.2 can be simplified into two ways. The first simpli-
fication comes in treating the building as linearly elastic in computing the response contributions
of ‘modes’ higher than the first three ‘modes,’ an approximation that is supported by Figure 9 that

Copyright 䉷 2010 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2011; 40:789–806
DOI: 10.1002/eqe
THREE-DIMENSIONAL MPA OF BUILDINGS 793

follows and by other work [12, 13, 19]; this leads to the modified MPA (MMPA) procedure [19].
Additionally, the contribution of torsional modes is negligible in the response of symmetric-plan
buildings to lateral ground motions; therefore, those modes can be ignored.
The second simplification comes in determining the median value of the peak deformation D̂n of
the nth-mode inelastic SDF system that is needed to obtain the reference displacement u rn . Instead
of using nonlinear RHA for each excitation, D̂n can be estimated by multiplying the median peak
deformation D̂no of the corresponding linear system, known from the design spectrum, by the
inelastic deformation ratio CRn . Several empirical equations for CRn are available [20–24]; the
equation selected here is taken from Reference [21]. Using this approach to estimate the median
Dn leads to the PMPA procedure, which permits estimation of seismic demands directly from the
design spectrum without having to perform any nonlinear RHA of the modal SDF systems for
each ground motion.

3. STRUCTURAL SYSTEMS AND MODELING

The structural systems considered are 48- and 62-story buildings taken from the Tall Building Initia-
tive project at the Pacific Earthquake Engineering Research Center (PEER) [http://peer.berkeley.
edu/∼yang/]. The buildings are identified by the letters CW (concrete wall) followed by the number
of stories. The lateral resisting system of the buildings is a ductile concrete core wall consisting of
two channel-shaped walls connected to rectangular walls by coupling beams, as shown in Figures 1
and 2. The channel-shaped walls of the CW48 building have openings that create L-shaped walls
at some stories (Figures 1(a) and 2(a)). The typical floors are 8-in-thick post-tensioned slabs span-
ning between the core and perimeter concrete columns. The total height of the CW48 and CW62
buildings is 471 and 630 ft, respectively. The height-to-width aspect ratio of the core of the 62-story
building is 12:1 in the long direction (x-direction) and 18:1 in the short direction (y-direction), as
shown in Figure 2. To increase the aspect ratio and stiffness of the system in the short direction,
concrete outrigger columns are included. The outrigger columns are connected to the core with
buckling restrained braces (BRBs) at the 28th and 51st floors. Additionally, this building has a
tuned liquid mass damper at the roof to help reduce sway to acceptable comfort levels during
strong wind [25].
These buildings were designed according to the 2001 San Francisco Building Code (SFBC) for
soil class Sd . Because these buildings exceed the height limit of 160 ft imposed in the SFBC for
concrete bearing wall systems, the buildings were designed by an alternative performance-based
procedure, allowed in Section 1629.10.1, to meet the equivalent criteria of Section 104.2.8 of
SFBC. The seismic forces were determined for site-specific design spectrum corresponding to a

xcm
xcm
CW48 coupling
beams CW62 coupling
beams
concrete
core wall
gravity outrigger
concrete columns columns
y
gravity core wall
y columns x
(b)
x
(a)

Figure 1. Schematic plans: (a) CW48 building and (b) CW62 building.

Copyright 䉷 2010 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2011; 40:789–806
DOI: 10.1002/eqe
794 J. C. REYES AND A. K. CHOPRA

concrete
shear wall

BRB
concrete outriggers
shear wall
coupling
beams

basement
z z z z

x y x y
(a) (b)

Figure 2. Elevations: (a) CW48 building and (b) CW62 building.

design basis earthquake (DBE) and for a maximum considered earthquake (MCE). The earthquake
forces for preliminary design were determined by linear RSA of the building with the design
spectrum reduced by a response modification factor of 4.5. Subsequently, the preliminary design
was refined based on the results of nonlinear RHA of the buildings for seven ground motions.
The core walls were detailed according to Chapter 21 of the ACI 318-99 code. To ensure a
ductile response of the system, the design included several features: (1) capacity design of the
core walls to avoid shear failure and guarantee a predominantly flexural behavior at plastic hinge
zones; (2) increased moment strength was assigned above the base of the CW48 building to ensure
that the plastic hinge forms at the base of the wall, and (3) capacity design to determine the joint
shear strength, in order to avoid brittle shear failure at the slab–column and slab–core joints.
The computer models used in this research were provided by a private company to the Tall
Building Initiative project of the PEER. Because these computer models reflect the current state
of professional practice in California, we decided to adopt them. Both buildings were modeled in
the PERFORM-3D computer program [26] using the following nonlinear models for the various
structural components. The shear wall element in PERFORM-3D is an area finite element with
inelastic fibers, including P-delta effects for both in-plane and out-plane deformations. With four
nodes and 24 DOF, this element has multiple layers that act independently, but are connected at
the element nodes, as shown in Figure 3. The behavior of the element is defined by the combined
behavior of four layers [26]: (1) Axial-bending layer for the vertical axis (Figures 3(a)) with its
cross-section composed of concrete and steel fibers with linear or nonlinear stress–strain relations;
(2) axial-bending layer for the horizontal axis (Figures 3(b)) with its cross-section assumed to be
linearly elastic, because this effect is secondary in a slender wall; (3) conventional shear layer
(Figures 3(c)), assuming constant shear stress and a uniform wall thickness; and (4) out-of-plane
bending (Figure 3(d)), modeled as elastic because this effect is secondary in concrete walls. These
layers have the following additional characteristics. First, in-plane deformations, such as axial strain,
shear strain, and curvature, are assumed constant along the element length. Second, the hysteretic
behavior of the fibers in the axial-bending layer and the material of the shear layer are represented
by a tri-linear model with in-cycle strength deterioration and cyclic stiffness degradation as shown
schematically in Figure 4; this figure does not show the specific model used for the concrete or
steel fibers.
The core walls of the buildings were modeled using this shear wall element considering inelastic
fibers with in-cycle strength deterioration, but without cyclic stiffness degradation or buckling of
the steel bars. Figure 5 shows the finite element mesh for the top stories of the CW48 building.

Copyright 䉷 2010 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2011; 40:789–806
DOI: 10.1002/eqe
THREE-DIMENSIONAL MPA OF BUILDINGS 795

concrete
wall

“+” “+” “+”

(a) Vertical (b) Horizontal (c) Conventional (d) Out-of-plane


axial-bending axial-bending shear bending

Figure 3. Parallel layers of the shear wall element in Perform-3D; adapted from Reference [26].

Force Force
excluding in-cycle excluding stiffness
strength deterioration degradation

Deformation Deformation

including
including in-cycle stiffness
strength deterioration degradation
(a) (b)

Figure 4. Tri-linear force–deformation relationships considering: (a) in-cycle strength


deterioration and (b) cyclic stiffness degradation.

core wall

coupling
beams

Figure 5. Finite element mesh for the top stories of the CW48 building.

Coupling beams, slabs, and girders were modeled by a linear one-dimensional element with
tri-linear plastic hinges at the ends, with ductility capacities specified according to FEMA 356 [20].
The plastic hinges of the coupling beams include in-cycle strength deterioration (Figures 4(a)) and
cyclic stiffness degradation associated with the unloading and reloading stiffnesses (Figures 4(b))
adjusted to reduce the area of the hysteresis loops by 40%.

Copyright 䉷 2010 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2011; 40:789–806
DOI: 10.1002/eqe
796 J. C. REYES AND A. K. CHOPRA

Figure 6. Natural periods and modes of vibration of CW48 (top row) and CW62 (bottom row) buildings;
shown in the three boxes are x-translational, y-translational, and torsional components of displacements.

The geometric nonlinear effects were considered by a standard P-Delta formulation for the
overall building using an equivalent leaning column to represent the gravity frames, and also locally
at the finite element level in the core wall.
The natural vibration periods and modes of the CW48 and CW62 buildings are shown in Figures 6
and 7, where xcm is the distance from the C.M. to a corner of the building (Figure 1). Figure 6
shows the height-wise variation of lateral displacements and torsional rotations, and Figure 7 shows
the motion of the roof in plan. Note: (1) lateral displacements dominate motion in the first mode of
lateral vibration in the x and y directions of both buildings, whereas torsional rotations dominate
motion in the third mode; (2) the fundamental vibration periods of the CW48 and CW62 buildings
are 4.08 and 4.50 s, respectively; the period of the dominantly torsional mode is much shorter
than that of the dominantly lateral modes; and (3) in modes 1 and 2, the CW48 building moves
simultaneously in the two lateral directions without torsion.
The damping of these buildings was modeled by the Rayleigh damping—a linear combination
of the mass and initial stiffness matrix—with its two constants selected to give 5% damping ratio
at the fundamental period of vibration T1 and a period of 0.1T1 . Damping ratios range from 2.9
to 7.2%, and from 2.9 to 7.6% for the first 12 vibration modes of the CW48 and CW62 buildings,
respectively.

4. GROUND MOTIONS

A total of 30 ground acceleration records from nine different earthquakes with magnitudes ranging
from 7.3 to 7.9 were selected according to the following criteria: (1) closest distance to the fault
<40 km; (2) average shear-wave velocity in upper 30 m of soil, Vs30 >200 m/s; and (3) longest
usable period >6.4 s. This is the period below which the response spectrum for the high-pass
filtered ground motion is unaffected by filtering of the data. This requirement ensures that selected
ground motions are appropriate for the analysis of these buildings with fundamental periods of 4.0
and 4.5 s. These selected records and their relevant data are listed in [12].
All the 30 records were scaled to represent the same seismic hazard defined by A(T1 ), the
pseudo-acceleration at the fundamental vibration period T1 of the structure. Both components of

Copyright 䉷 2010 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2011; 40:789–806
DOI: 10.1002/eqe
THREE-DIMENSIONAL MPA OF BUILDINGS 797

CW48

T1=4.08 sec T2=3.77 sec T3=2.65 sec

CW62

T1=4.50 sec T2=3.91 sec T3=2.85 sec

Figure 7. First triplet of periods and modes of vibration of CW48 and CW62 buildings
(only the motion at the roof is shown).

2
T1 = 4.08 sec

T1 = 4.50 sec
CW48 CW62
Pseudo-acceleration A, g

1.5

Hazard spectrum
1 Median spectrum
x-dir
y-dir
0.5

0
0 1 2 3 4 5 0 1 2 3 4 5
Natural vibration period Tn, sec Natural vibration period Tn, sec

Figure 8. Seismic hazard spectrum for building site corresponding to 2% probability of exceedance in
50 years (solid line), and the median response spectra of 30 scaled ground motions in the x and y
directions (dashed lines): CW48 and CW62 buildings.

a record were scaled by the same factor selected to match their geometric mean to the seismic
hazard. The geometric mean is defined as A(T1 ) = A x (T1 )× A y (T1 ), where A x (T1 ) and A y (T1 )
are the A(T1 ) values for the two horizontal components of the record. The selected seismic hazard
spectrum was determined as the average of three uniform hazard spectra for 2% probability of
exceedance in 50 years (return period of 2475 years), corresponding to three different attenuation
relationships [27–29] as a part of the ‘Next Generation of Ground-Motion Attenuation Models’
(NGA) project.
The values of A(T1 )2%/50 selected to define ground motion ensembles are 0.148g and 0.137g
for the CW48 and CW62 buildings, respectively. Figure 8 shows the median response spectra for
the ensemble of 30 ground motions along the x and y directions scaled to match A(T1 )2%/50 for
the two buildings, and the seismic hazard spectrum corresponding to 2% probability of exceedance
in 50 years. As imposed by the scaling criterion, the median pseudo-acceleration of the ground
motion ensemble at the fundamental vibration period of the building is matched to the seismic
hazard spectrum; because T1 differs for each structure, the scaling factors for ground motions and
hence their median spectra vary with the building.
After evaluating the MPA procedure for the aforementioned ground motions, which already
represent a low-exceedance-probability hazard for a highly seismic region, these excitations are

Copyright 䉷 2010 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2011; 40:789–806
DOI: 10.1002/eqe
798 J. C. REYES AND A. K. CHOPRA

scaled by a factor of 3.0 and 2.0 for CW48 and CW62 buildings, respectively, to test the procedure
for extreme seismic hazard; note that the resulting ground motions seem almost unrealistically
intense.

5. EVALUATION OF MPA PROCEDURE

5.1. Modal pushover curves and reference displacements


As a first step in evaluating the MPA procedure, this section presents the modal pushover curves
and the reference displacements for each building due to the scaled ground motions. Figure 9
shows pushover curves and their tri-linear idealization for each building associated with their
first and second ‘modes’ in each lateral direction, which are modes 1, 2, 4 and 5 of the three-
dimensional model of the building. Modes 3 and 6 are not shown because these are torsional
modes and their contribution to lateral displacements is negligible. Pushover analyses for the
first ‘mode’ of the CW62 building and the first two ‘modes’ of the CW48 building lead to the
following observations, with reference to the tri-linear idealization of the pushover curves, shown
schematically in Figure 10. At displacements up to u 1 , the building remains essentially elastic,
but cracking is initiated in some sections of the core wall. At the first yield displacement u 1 , the
reinforcement in the post-tensioned slabs starts to yield whereas the stresses at the extreme fiber
of the core wall have just reached the tensile strength of concrete. As displacement increases, the
tensile force in the concrete of the core wall and coupling beams is transferred to the steel until the
reinforcement yields at displacements around the second yield point u 2 ; yielding in the core wall
is concentrated in the first six stories above grade. The building reaches its maximum strength and
develops a collapse mechanism at a displacement between u 2 and u 3 ; nonlinear geometric effects
result in negative stiffness in the pushover curve at displacements smaller than u 3 . At displacements
larger than u 3 , coupling beams begin to loose strength, thus degrading the strength of the building
and leading to its eventual collapse. The second ‘mode’ of vibration of the CW62 building, which
is the first mode of lateral vibration in the y-direction, is less ductile than its first ‘mode’, the
first mode of lateral vibration in the x-direction (compare Figures 9(e) and (f)). At the first yield
displacement u 1 , yielding starts in the BRBs, slabs, coupling beams, and near the base of the wall.

"Mode" 1 "Mode" 2 "Mode" 4 "Mode" 5


.2 .4
Median=0.56

Median=0.55

Median=0.04

Median=0.04

CW48 CW48
Base shear/weight

.3

.1 .2

.1

0 1 2 0 1 2 0 0.2 0.4 0 0.2 0.4


(a) (b) (c) (d)
.2
Median=0.44

Median=0.45

Median=0.07

Median=0.08

CW62 CW62
Base shear/weight

.2
.1
.1

0 0.5 1 1.5 0 0.5 1 0 1.5 0.5 1 0 0.5 1


(e) (f) (g) (h)
Reference displacement/height,%

Figure 9. Modal pushover curves for the first four ‘modes’ of lateral vibration of CW48 and CW62
buildings; trilinear idealization is shown in dashed lines. The reference displacement due to 30 scaled
ground motions is identified and the median value is also noted.

Copyright 䉷 2010 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2011; 40:789–806
DOI: 10.1002/eqe
THREE-DIMENSIONAL MPA OF BUILDINGS 799

Idealized Actual

Slope 3

Vbn
Slope 2

Slope 1
u1 u2 u3 u rn

Figure 10. Pushover curve idealized as a tri-linear curve.

At displacements between u 1 and u 2 , sections of the wall around the 28th and 33rd stories yield.
From u 2 to u 3 , the building strength degrades progressively until it abruptly looses its strength
at u 3 due to damage in some coupling beams (Figure 9(f)). These coupling beams continue to
deteriorate at displacements larger than u 3 until they reach their maximum deformation capacity.
Figure 9 also identifies reference displacements due to each of the 30 scaled ground motions and
their median value; these reference displacements were determined in Step 4 of the MPA procedure
(Section 2.2). All ground motions drive both buildings beyond the first yield displacement u 1 in
the first and second ‘modes;’ the median displacement of the CW48 building exceeds its yield
displacement u 1 by factors of 3.8 and 3.5 in the x and y directions, respectively; for the CW62
building, these factors are 2.4 and 1.9. More than half of the excitations drive the CW62 building
beyond its first yield displacement u 1 in the second mode of lateral vibration in the x and y
directions, but the median displacement is only slightly larger than u 1 . Only a few ground motions
drive the CW48 building beyond its first yield displacement u 1 in its fourth and fifth ‘modes’, and
the median displacement is smaller than u 1 . These results show that the median displacement in
‘modes’ higher than the first pair of ‘modes’ is either close to or exceeds the first yield displacement
u 1 only by a modest amount. This is consistent with the results of past research on steel MRF
buildings [8].

5.2. Higher mode contributions in seismic demands


Figure 11 shows the median (over the ground motion ensemble) values of floor displacements
and story drifts in the x and y directions at the C.M., including a variable number of ‘modes’ in
MPA, superimposed with the ‘exact’ result from nonlinear RHA, for each building subjected to
both components of ground motion, simultaneously. Except for the rapid fluctuations near the top
of the CW62 building, which are due to reduction of core wall stiffness and presence of a tuned
liquid mass damper, electromechanical equipment and architectural setbacks, the variation of story
drifts over building height is typical of tall buildings. The observations presented in this section
and in Section 6 will exclude the upper three stories of the CW62 building and basements of both
buildings.
The first pair of ‘modes’ alone is adequate in estimating floor displacements; including higher
‘modes’ does not significantly improve this estimate (Figure 11(a)). Although the first pair of
‘modes’ alone is inadequate in estimating story drifts, with the second pair of ‘modes’ (fourth-
and fifth-‘mode’) included, story drifts estimated by the MPA procedure improve significantly, and
resemble nonlinear RHA results (Figure 11(b)). Despite this improved accuracy, notable discrep-
ancies between MPA and nonlinear RHA results remain for both buildings; these discrepancies
are discussed next.

5.3. Evaluation of MPA


The MPA procedure underestimates the x and y components of floor displacements for both
buildings, and the roof displacement is underestimated by about 16 and 12% for the CW48 and
CW62 buildings, respectively. The height-wise average underestimation of story drift is about 15%

Copyright 䉷 2010 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2011; 40:789–806
DOI: 10.1002/eqe
800 J. C. REYES AND A. K. CHOPRA

x-component y-component x-component y-component


48 48
CW48 CW48

36 36

Floor

Story
24 NL-RHA 24
MPA
12 1 pair 12
2 pairs
3 pairs
G G
0 0.5 1 0 0.5 1 0 1 2 0 1 2
60 CW62 60 CW62

40 40
Floor

Story
NL-RHA
MPA
20 1 pair 20
2 pairs
3 pairs
G G
0 0.5 1 0 0.5 1 0 1 2 0 1 2
(a ) Floor Displacement / Building Height, % (b ) Story Drift Ratio, %

Figure 11. Median (a) floor displacements and (b) story drifts at the C.M. of the CW48 (row 1) and
CW62 (row 2) buildings determined by nonlinear RHA and MPA, with a variable number of ‘modes’.

for the CW48 building and about 18% for the CW62 building. Notable discrepancies remain for
y-direction drifts in the upper part of the CW62 building where the underestimation is around
30%; recall that in this direction, the core wall of the building interacts with BRBs and outrigger
columns, developing plastic hinges at various locations over building height (Section 5.1).

5.4. Evaluation of MMPA


The results of Figure 9 and their interpretation suggested that the buildings could be treated as
linear in estimating contributions of modes higher than the first triplet of ‘modes’ (or first pair
of ‘modes’ for a symmetric-plan building) to seismic demands—MMPA—as used in Section 2.3
to estimate seismic demands for the CW48 and CW62 buildings; torsional modes were ignored
because their effective modal mass is negligible for the x- and y-direction excitations. Figure 12
shows the median values of x and y components of floor displacements and story drifts at the
C.M., determined by nonlinear RHA, MPA, and MMPA for both buildings. The seismic demands
estimated by MMPA and MPA are very close—MMPA values are slightly larger—implying that it
is valid to treat the buildings as linearly elastic in estimating higher-mode contributions to seismic
demands.

6. EVALUATION OF PMPA PROCEDURE

6.1. Floor displacements and story drifts at the C.M.


In implementing the PMPA procedure (Section 2.3), the median value of the peak deformation D̂n
of the nth-mode inelastic SDF system was estimated by multiplying the median peak deformation
D̂no of the corresponding linear system by the inelastic deformation ratio CRn ; D̂no was determined
from the median response spectrum for the ensemble of 30 ground motions corresponding to the
damping ratio appropriate for the particular mode. As expected, CRn was essentially equal to 1.0
for all modes of these long-period buildings.
Figure 12 shows the median values of the x and y components of floor displacements and story
drifts at the C.M., determined by nonlinear RHA, MPA, and PMPA procedures for both buildings.
In general, PMPA provides a larger estimate of seismic demands compared with MPA because

Copyright 䉷 2010 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2011; 40:789–806
DOI: 10.1002/eqe
THREE-DIMENSIONAL MPA OF BUILDINGS 801

x-component y-component x-component y-component


48 48
CW48 CW48

36 36

Floor

Story
24 24
NL-RHA
12 MPA 12
MMPA
PMPA
G G
0 0.5 1 0 0.5 1 0 1 2 0 1 2
60 CW62 60 CW62

40 40
Floor

Story
NL-RHA
20 MPA 20
MMPA
PMPA
G G
0 0.5 1 0 0.5 1 0 1 2 0 1 2
(a ) Floor Displacement / Building Height, % (b ) Story Drift Ratio, %

Figure 12. Median (a) floor displacements and (b) story drifts at the C.M. of the CW48 (row 1) and
CW62 (row 2) buildings determined by nonlinear RHA, MPA, MMPA, and PMPA.

Table I. Ratio of D̂n determined by empirical equation for CRn


and by nonlinear RHA of SDF systems.
Building First mode x-dir First mode y-dir
CW48 1.08 1.16
CW62 1.13 1.27

the empirical value of D̂n is larger than its value determined as the median of the Dn values
determined by nonlinear RHA of the ‘modal’ SDF system for 30 ground motions (Step 4 of MPA);
the ratio of the two for the first pair of ‘modes’ is shown in Table I. This is to be expected for
such long-period systems because the empirical equation does not permit values of CRn below 1.0,
whereas the exact data do fall below 1.0 [21]. Despite this overestimation in one step of PMPA,
the method underestimates the seismic demands (Figure 12); however, this underestimation is
primarily due to approximations inherent in modal combination and multi-component combination
rules, as demonstrated next.
The PMPA procedure for concrete buildings subjected to two horizontal components of ground
motion, simultaneously, is based on six approximations, the first three of which are also inherent in
the MPA procedure: (1) neglecting the coupling of ‘modal’ responses to peff,n (t) in Equation (6)—
associated with x or y excitation—which has been demonstrated to be weak for the structures
considered in this paper [12]; (2) using modal combination rules, originally developed for linear
systems, to determine the total response to one component of ground motion; (3) using multi-
component combination rules, originally derived for linear systems, to determine the response to
both components of ground motion, simultaneously; (4) estimating the peak deformation Dn of
the nth-mode inelastic SDF system from the elastic design spectrum using an empirical equation
for the inelastic deformation ratio; (5) treating the structure as linearly elastic in estimating higher-
mode contributions to seismic demand; and (6) ignoring cyclic stiffness degradation of structural
elements. Because approximations (2) and (3) are the only sources of error in the standard RSA
procedure for linear systems, the resulting error in the response of these systems serves as a
baseline for evaluating the additional errors in PMPA for nonlinear systems.

Copyright 䉷 2010 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2011; 40:789–806
DOI: 10.1002/eqe
802 J. C. REYES AND A. K. CHOPRA

CW48 building CW62 building


x-component y-component x-component y-component
48 60 CW62
CW48

36
40

Story
Story
24
20
12
RHA RHA
RSA RSA
G G
0 1 2 0 1 2 0 0.5 1 1.5 0 0.5 1 1.5
(a)
48 60 CW62
CW48

36
40

Story
Story

24
20
12
NL-RHA NL-RHA
PMPA PMPA
G G
0 1 2 0 1 2 0 0.5 1 1.5 0 0.5 1 1.5
(b) Story Drift Ratio, % Story Drift Ratio, %

Figure 13. Median story drifts at the C.M. for: (a) linearly elastic systems determined by
RSA and RHA procedures and (b) inelastic systems determined by PMPA and nonlinear RHA
procedures. Results are for CW48 and CW62 buildings.

Figure 13 compares the accuracy of PMPA in estimating the response of nonlinear systems
with that of RSA in estimating the response of linear systems. Observe that the RSA procedure
underestimates the median response for both buildings. This underestimation tends to be greater
in the upper stories of the buildings, consistent with the height-wise variation of contribution of
higher modes to response [13]. The height-wise average underestimation in story drifts is around
11% for both buildings. The errors introduced by the additional approximations in PMPA, which
are apparent by comparing parts (a) and (b) of Figure 13, are small. The height-wise average
underestimation in the story drifts determined by PMPA is 16 and 9% for the CW48 and the CW62
buildings, respectively.
The PMPA is an attractive method to estimate seismic demands for tall buildings due to two
horizontal components of ground motion, simultaneously, not only because it calculates Dn directly
from the design spectrum, but also because it leads to a slightly larger estimate of seismic demand,
thus reducing the unconservatism (relative to nonlinear RHA) of MPA results.

6.2. Other response quantities


The member forces and total end rotations corresponding to the median story drifts of Figure 12(b)
were estimated by implementing Steps 11 and 12 of the PMPA procedure (Section 2.2). Figure 14
presents such results for bending moments, shear forces, and plastic hinge rotations in the coupling
beams highlighted in Figure 1(a) of the CW48 building, together with the median values of these
responses determined by nonlinear RHA. The internal forces were estimated accurately, whereas
total rotations were underestimated just as the story drifts were underestimated.
The forces in the core wall can be estimated to a useful degree of accuracy by the PMPA
procedure. Figure 15 presents the height-wise variation of shear forces (Vx and Vy ) in the core
wall of the CW48 buildings, including a variable number of ‘modes’ in PMPA superimposed
with the results from nonlinear RHA. The first pair of ‘modes’ alone is grossly inadequate in
estimating internal forces; however, with the second and third pair of ‘modes’ included, internal
forces estimated by PMPA resemble nonlinear RHA results. By including the contributions of all
significant modes of vibration, PMPA is able to adequately capture the height-wise variation of
shear forces in the core shear wall. Thus, PMPA overcomes a well-known limitation of pushover

Copyright 䉷 2010 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2011; 40:789–806
DOI: 10.1002/eqe
THREE-DIMENSIONAL MPA OF BUILDINGS 803

48 48 48

36 36 36

Floor
Floor 24 24 24

12 12 12
NL-RHA NL-RHA NL-RHA
PMPA PMPA PMPA
G G G
0 1000 2000 3000 4000 0 500 1000 0 0.01 0.02
(a) Bending moment, kip-ft (b) Shear force, kips (c) Plastic hinge rotation, rad

Figure 14. (a) Bending moments; (b) shear forces; and (c) plastic rotations for the coupling beams of the
CW48 building highlighted in Figure 1(a), determined by nonlinear RHA and PMPA.

48 48

NL-RHA
36 36 PMPA
1 "pair"

Story
Story

24 24 2 "pairs"
3 "pairs"

12 12

G G
0 1 2 0 1 2
Vx , kips 4
x 10 Vy, kips x 10
4

Figure 15. Shear forces Vx and Vy for the core wall of the CW48 building, determined by nonlinear RHA
and PMPA with variable number of modes.

procedures with a single invariant force distribution that are unable to consider the important higher
mode effects after formation of local mechanisms [30].

7. EVALUATION OF THE PMPA PROCEDURE FOR EXTREME SEISMIC HAZARD

Figure 16 shows pushover curves and their tri-linear idealization for each building associated with
their first and second ‘modes’ in each lateral direction, wherein reference displacements due to
each of the 30 extreme ground motions (defined in Section 4) and their median value are identified.
These results lead to the following observations, with reference to the tri-linear idealization of
the pushover curves shown schematically in Figure 10. Almost all of the excitations drive the
CW48 building into the third slope of the pushover curve for the first- and second-‘mode’, with
median displacements equal to 11.9 to 12.7 times u 1 —the first yield displacement—or 1.8 to 2.0
times u 2 —the second yield displacement; however, the median displacement in the fourth and fifth
‘modes’ is only 2.1 to 2.3 times u 1 . The ground motions drive the CW62 building into the third
and second slopes of the pushover curve for the first and second ‘modes’, respectively; the median
displacement is 3.7 to 5.6 times u 1 or 0.7 to 1.6 times u 2 . The median displacement in the fourth
and fifth ‘modes’ is 2.7 to 3.1 times u 1 .
Figure 17 compares the accuracy of PMPA in estimating the story drifts of nonlinear systems
with that of RSA in estimating the response of linear systems. As mentioned in Section 6.1, the
height-wise average underestimation in RSA is around 11% for both buildings. The additional
approximations in PMPA do not increase the errors fortuitously in this case; the height-wise average
underestimation of story drifts determined by PMPA is reduced to 5 and 7% for the CW48 and
the CW62 buildings, respectively. Figure 18 presents the height-wise variation of shear forces
(Vx and Vy ) in the core wall of the CW48 building determined by PMPA and nonlinear RHA,

Copyright 䉷 2010 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2011; 40:789–806
DOI: 10.1002/eqe
804 J. C. REYES AND A. K. CHOPRA

"Mode" 1 "Mode" 2 "Mode" 4 "Mode" 5


0.2 0.4

Median=0.16

Median=0.15
CW48

Median=1.97

Median=1.97
CW48

Base shear/weight
0.15 0.3

0.1 0.2

0.05 0.1

0 0
0 1 2 30 1 2 3 0 0.2 0.4 0 0.2 0.4
(a) (b) (c) (d)
0.2
Median=1.09

Median=0.18
Median=0.19
CW62 CW62
Base shear/weight

0.15

Median=0.89
0.2
0.1
0.1
0.05

0 0
0 0.5 1 1.5 0 0.5 1 0 1.5
0.5 10 0.5 1
(e) (f ) (g) (h)
Reference displacement/height,%

Figure 16. Modal pushover curves for the first four ‘modes’ of lateral vibration of CW48 and CW62
buildings; trilinear idealization is shown in dashed lines. The reference displacement due to 30 extreme
ground motions is identified and the median value is also noted.

CW48 building CW62 building


x-component y-component x-component y-component
48 60 CW62
CW48

36
40
Story
Story

24
20
12
RHA RHA
RSA RSA
G G
0 2 4 6 0 2 4 6 0 2 4 0 2 4
(a)
48 60 CW62
CW48

36
40
Story

Story

24
20
12
NL-RHA NL-RHA
PMPA PMPA
G G
0 2 4 6 0 2 4 6 0 2 4 0 2 4
(b) Story Drift Ratio, % Story Drift Ratio, %

Figure 17. Median story drifts at the C.M. for: (a) linearly elastic systems determined by RSA and RHA
procedures and (b) inelastic systems determined by PMPA and nonlinear RHA procedures for an extreme
seismic intensity. Results are for CW48 and CW62 buildings.

where it is seen that the accuracy of PMPA is adequate. It is reassuring that, even for this extreme
seismic hazard, PMPA estimates the seismic demands to a useful degree of accuracy. As a side,
observe in Figure 17 the similarity between the story drifts for linearly elastic and inelastic systems.
This similarity seems to result for two reasons: (1) overall displacements of long-period systems
are known to be essentially independent of inelastic action and (2) yielding in the core wall is
essentially limited to the lower few stories above grade.

Copyright 䉷 2010 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2011; 40:789–806
DOI: 10.1002/eqe
THREE-DIMENSIONAL MPA OF BUILDINGS 805

48 48

36 36

Story

Story
24 24

12 12
NL-RHA
PMPA
G G
0 2 4 6 0 2 4 6
V , kips x 10
4 V , kips x 10
4
x y

Figure 18. Median shear forces Vx and Vy for the core wall of the CW48 building, determined by nonlinear
RHA and PMPA for an extreme seismic intensity.

8. CONCLUSIONS

The MPA procedure for estimating floor deformations and story drifts for buildings subjected to
one horizontal component of ground motion has been extended to estimate seismic demands—floor
displacements, story drifts, internal forces, and plastic hinge rotations—for buildings subjected to
two horizontal components of ground motion, simultaneously. The PMPA procedure, which is a
version of MPA that is especially suitable for practical application, has been developed to compute
seismic demands directly from the earthquake response (or design) spectrum. This procedure is
based on two principal simplifications: (1) the structure is treated as linearly elastic in estimating
higher-‘mode’ contributions to seismic demand; and (2) the median deformation of the nth-mode
inelastic SDF system is estimated directly from the design spectrum using an empirical equation
for the inelastic deformation ratio.
The median seismic demands for 48- and 62-story buildings with ductile concrete shear walls—
designed according to the alternative provisions of the 2001 SFBC—due to an ensemble of 30
two-component ground motions were computed by MPA, PMPA, and nonlinear RHA procedures
and compared. The presented results led to the following conclusions:
1. Although, as expected, the first ‘mode’ of lateral vibration in the x and y directions is
inadequate in estimating story drifts, with the second ‘mode’ of lateral vibration included,
the discrepancies (relative to nonlinear RHA results) reduce greatly and story drifts estimated
by MPA resemble nonlinear RHA results.
2. The MPA procedure may be simplified by treating the building as linearly elastic in estimating
higher-‘mode’ contributions to seismic demands. This approximation has been demonstrated
to be valid.
3. The modal combination and multi-component combination approximations inherent in RSA
of linearly elastic systems—a standard tool in structural engineering practice—may lead
to significant underestimation of story drift demands for tall buildings, especially in upper
stories.
4. For the selected tall buildings, the PMPA procedure for nonlinear systems is almost as accurate
as RSA for linear systems even for the extreme seismic hazard considered herein. Thus, PMPA
should be useful for practical application in estimating seismic demands—floor displacements,
story drifts, plastic hinge rotations, and internal forces—for tall buildings due to two horizontal
components of ground motion applied simultaneously. One attractive feature is that PMPA
estimates seismic demands directly from the (elastic) design spectrum, thus avoiding the
complications of selecting, scaling, and modifying ground motions for nonlinear RHA.

ACKNOWLEDGEMENTS
The first author acknowledge the fellowship from the Fulbright Commission, Colciencias, and the Univer-
sidad de los Andes, Bogota, Colombia to pursue a PhD degree in Structural Engineering at the University

Copyright 䉷 2010 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2011; 40:789–806
DOI: 10.1002/eqe
806 J. C. REYES AND A. K. CHOPRA

of California, Berkeley. The authors are most grateful to Dr Tony Yang from the PEER for providing the
structural models and ground motion data that served as the basis for this study.

REFERENCES
1. Bracci JM, Kunnath SK, Reinhorn AM. Seismic performance and retrofitt evaluation for reinforced concrete
structures. Journal of Structural Engineering (ASCE) 1997; 123:3–10.
2. Sasaki KK, Freeman SA, Paret TF. Multimode pushover procedure (MMP)—a method to identify the effects
of higher modes in a pushover analysis. Proceedings of the Sixth U.S. National Conference on Earthquake
Engineering, Seattle, Washington, 1998.
3. Kalkan E, Kunnath SK. Adaptative modal combination for nonlinear static analysis of building structures. Journal
of Structural Engineering (ASCE) 2006; 132:1721–1731.
4. Aydinoglu MN. An incremental response spectrum analysis procedure based on inelastic spectral displacements
for multi-mode seismic performance evaluation. Bulletin of Earthquake Engineering 2003; 1:3–36.
5. Kunnath SK. Identification of modal combinations for nonlinear static analysis of building structures. Computer-
Aided Civil and Infrastructure Engineering 2004; 19:282–295.
6. Poursha M, Khoshnoudian F, Moghadam AS. A consecutive modal pushover procedure for estimating the seismic
demands of tall buildings. Journal of Engineering Structures 2009; 31:591–599.
7. Chopra AK, Goel RK. A modal pushover analysis procedure for estimating seismic demands for buildings.
Earthquake Engineering and Structural Dynamics 2002; 31:561–582.
8. Goel RK, Chopra AK. Evaluation of modal and FEMA pushover analyses: SAC buildings. Earthquake Spectra
2004; 20:225–254.
9. Bobadilla H, Chopra, AK. Evaluation of the MPA procedure for estimating seismic demands: RC-SMRF buildings.
Earthquake Spectra 2008; 24(4):827–845.
10. Chopra AK, Goel RK. A modal pushover analysis procedure to estimate seismic demands for unsymmetric-plan
buildings. Earthquake Engineering and Structural Dynamics 2004; 33:903–927.
11. Moehle JP. The tall buildings initiative for alternative seismic design. The Structural Design of Tall and Special
Buildings 2007; 16:559–567.
12. Reyes JC. Estimating seismic demands for performance-based engineering of buildings. Ph.D. Dissertation,
Department of Civil and Environmental Engineering, University of California, Berkeley, CA, 2009.
13. Chopra AK. Dynamics of Structures: Theory and Applications to Earthquake Engineering (3rd edn). Prentice-Hall:
NJ, 2007.
14. Newmark NM. Vibration of structures induced by ground motion. Shock and Vibration Handbook (2nd edn).
McGraw Hill: New York, 1976.
15. Rosenblueth E, Contreras H. Approximate design for multicomponent earthquakes. Journal of the Engineering
Mechanics Division (ASCE) 1977; 103:881–893.
16. Menun C, Der Kiureghian AD. A replacement for the 30%, 40% and SRSS rules for multicomponent seismic
analysis. Earthquake Spectra 1998; 14(1):153–163.
17. Smeby W, Der Kiureghian AD. Modal combination rules for multicomponent earthquake excitation. Earthquake
Engineering and Structural Dynamics 1985; 13:1–12.
18. Goel AK, Chopra AK. Extension of modal pushover analysis to compute member forces. Earthquake Spectra
2005; 21(1):125–139.
19. Chopra AK, Goel RK, Chintanapakdee C. Evaluation of a modified MPA procedure assuming higher modes as
elastic to estimate seismic demands. Earthquake Spectra 2004; 20(3):757–778.
20. FEMA. Prestandard and commentary for the seismic rehabilitation of buildings. FEMA-356, Washington, DC, 2000.
21. Chopra AK, Chintanapakdee C. Inelastic deformation ratios for design and evaluation of structures: single-degree-
of-freedom bilinear systems. Journal of Structural Engineering (ASCE) 2004; 130:1309–1319.
22. Ruiz-Gracia J, Miranda E. Inelastic displacement ratios for evaluation of existing structures. Earthquake
Engineering and Structural Dynamics 2004; 32:1237–1258.
23. FEMA. Improvement of non-linear static seismic analysis procedures. FEMA-440, Washington, DC, 2005.
24. ASCE. Seismic rehabilitation of existing buildings. ASCE/SEI 41-06, Reston, VA, 2007.
25. Post NM. A sleek skyscraper in San Francisco raises the profile of performance-based design. Continuing
Education Center, McGraw-Hill Construction, 2008.
26. Computers and Structures (CSI), Inc. PERFORM 3D. User Guide v4, Non-linear Analysis and Performance
Assessment for 3D Structures. Computers and Structures, Inc.: Berkeley, CA, 2006.
27. Campbell KW, Bozorgnia Y. NGA ground motion model for the geometric mean horizontal component of PGA,
PGV, PGD and 5% damped linear elastic response spectra for periods ranging from 0.01 to 10 s. Earthquake
Spectra 2008; 24(1):139–171.
28. Boore DM, Atkinson GM. Ground-motion prediction equations for the average horizontal component of PGA,
PGV, and 5%-damped PSA at spectral periods between 0.01 s and 10.0 s. Earthquake Spectra 2008; 24(1):99–138.
29. Abrahamson N, Silva W. Summary of the Abrahamson & Silva NGA ground-motion relations. Earthquake
Spectra 2008; 24(1):67–97.
30. Krawinkler H, Seneviratna GDPK. Pros and cons of a pushover analysis of seismic performance evaluation.
Engineering Structures 1998; 20(4–6):452–464.

Copyright 䉷 2010 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2011; 40:789–806
DOI: 10.1002/eqe

You might also like