Vous êtes sur la page 1sur 18

Engineering Structures 23 (2001) 407424

www.elsevier.com/locate/engstruct

Static pushover versus dynamic collapse analysis of RC buildings


A.M. Mwafy, A.S. Elnashai

Department of Civil and Environmental Engineering, Imperial College, Imperial College Road, London SW7 2BU, UK
Received 2 February 2000; received in revised form 23 May 2000; accepted 26 May 2000

Abstract
Owing to the simplicity of inelastic static pushover analysis compared to inelastic dynamic analysis, the study of this technique
has been the subject of many investigations in recent years. In this paper, the validity and the applicability of this technique are
assessed by comparison with dynamic pushover idealised envelopes obtained from incremental dynamic collapse analysis. This
is undertaken using natural and artificial earthquake records imposed on 12 RC buildings of different characteristics. This involves
successive scaling and application of each accelerogram followed by assessment of the maximum response, up to the achievement
of the structural collapse. The results of over one hundred inelastic dynamic analyses using a detailed 2D modelling approach for
each of the twelve RC buildings have been utilised to develop the dynamic pushover envelopes and compare these with the static
pushover results with different load patterns. Good correlation is obtained between the calculated idealised envelopes of the dynamic
analyses and static pushover results for a defined class of structure. Where discrepancies were observed, extensive investigations
based on Fourier amplitude analysis of the response were undertaken and conservative assumptions were recommended. 2001
Elsevier Science Ltd. All rights reserved.
Keywords: Pushover analysis; Timehistory collapse analysis; RC buildings; Fourier amplitude analysis

1. Introduction
Inelastic timehistory analysis is a powerful tool for
the study of structural seismic response. A set of carefully selected ground motion records can give an accurate evaluation of the anticipated seismic performance of
structures. Despite the fact that the accuracy and
efficiency of the computational tools have increased substantially, there are still some reservations about the
dynamic inelastic analysis, which are mainly related to
its complexity and suitability for practical design applications. Moreover, the calculated inelastic dynamic
response is quite sensitive to the characteristics of the
input motions, thus the selection of a suite of representative acceleration timehistories is mandatory. This
increases the computational effort significantly. The
inelastic static pushover analysis is a simple option for
estimating the strength capacity in the post-elastic range.
The technique may be also used to highlight potential
weak areas in the structure. This procedure involves

* Corresponding author. Fax: +44 207 594 6053.


E-mail address: a.elnashai@ic.ac.uk (A.S. Elnashai).

applying a predefined lateral load pattern which is distributed along the building height. The lateral forces are
then monotonically increased in constant proportion with
a displacement control at the top of the building until a
certain level of deformation is reached. The target top
displacement may be the deformation expected in the
design earthquake in case of designing a new structure,
or the drift corresponding to structural collapse for
assessment purposes. The method allows tracing the
sequence of yielding and failure on the member and the
structure levels as well as the progress of the overall
capacity curve of the structure.
The static pushover procedure has been presented and
developed over the past twenty years by Saiidi and
Sozen [1], Fajfar and Gaspersic [2] and Bracci et al. [3],
among others. The method is also described and recommended as a tool for design and assessment purposes
by the National Earthquake Hazard Reduction Program
NEHRP (FEMA 273) [4] guidelines for the seismic
rehabilitation of existing buildings. Moreover, the technique is accepted by the Structural Engineers Association of California SEAOC (Vision 2000) [5] among
other analysis procedures with various level of complexity. This analysis procedure is selected for its

0141-0296/01/$ - see front matter 2001 Elsevier Science Ltd. All rights reserved.
PII: S 0 1 4 1 - 0 2 9 6 ( 0 0 ) 0 0 0 6 8 - 7

408

A.M. Mwafy, A.S. Elnashai / Engineering Structures 23 (2001) 407424

applicablity to performance-based seismic design


approaches and can be used at different design levels to
verify the performance targets. Finally, it is clear from
recent discussions in code-drafting committees in Europe that this approach is likely to be recommended in
future codes.
The technique has been evaluated in several previous
studies [610], to some extent, with different emphasis.
In most of the previous work, only comparative studies
between dynamic and static pushover analysis have been
assessed at certain loading levels, i.e. design level, or at
equal top displacement (roof displacement from pushover equal to the maximum dynamic roof displacement).
The results have been presented mainly in terms of global quantities, i.e. deformations, calculated hysteretic
energy and structural damage indices. The main aim of
this paper is to develop complete pushover-like load
displacement curves from incremental dynamic analysis
up to collapse for a range of structural configurations
representing the most common types of RC building,
including different structural systems, building heights,
design acceleration and level of ductility. The dynamic
pushover envelopes are then compared with the force
deformation curves obtained from inelastic static pushover analysis considering different lateral loading patterns. The procedure offers an opportunity for full comparisons between the two methods of analysis up to
ultimate collapse.

representative of a seismic design code applicable to


more than one country. While the second and third
groups are regular in plan and in elevation, the first
group exhibits two sources of irregularity in elevation.
The first storey has a greater height than the remaining
ones and severance at the first storey of some intermediate columns, which are supported by long span beams.
The geometric characteristics of the structures are illustrated in Fig. 1.
The overall plan dimensions of the configurations considered are 15 m20 m. The total heights are 25.5, 36
and 24 m for groups 1, 2 and 3, respectively, with equal
storey heights of 3 m except the first storey of group 1,
which is 4.5 m high. While the lateral force resisting
system for groups 1 and 2 is moment frames, group 3
possesses both a central core extending over the full
height and moment frames on the perimeter. The floor
system is solid slab in groups 1 and 2, and a waffle slab
in group 3. Live loads and loading from floor finishes
and partitions are both assumed to be 2.0 kN/m2. All
buildings are assumed to be founded on medium soil
type B of EC8 (firm). The cross section capacities have
been computed by considering a characteristic cylinder
strength of 25 N/mm2 for concrete and a characteristic
yield strength of 500 N/mm2 for both longitudinal and
transverse steel. More details regarding member cross
section sizes and reinforcements are given in Fardis [12].

3. Modelling approach and assumptions


2. Description of the buildings
In order to achieve the aforementioned objectives,
twelve RC buildings are considered, split into three
groups: sets of four 8-storey irregular frame, four 12storey regular frame and four 8-storey dual frame-wall
structures. Within each group, combination of two
design ground accelerations (0.15 and 0.30 g) and three
design ductility classes (High, Medium and Low) lead
to the four cases mentioned above. The selection of four
cases for each configuration is motivated by the desire
to compare the performance of structures design according to a ductility class set of rules but for different
ground acceleration and for the same ground acceleration but different ductility class rules. The value of the
force reduction factor (behaviour factor q in EC8 and
response modification factor R in UBC) increases and
rigorous standards on member detailing requirements are
imposed for higher ductility classes. Table 1 shows the
definition of the set of structures under consideration
where the elastic force reduction factors used in the
design as well as the observed elastic fundamental period, obtained from elastic free vibration analyses, are
also given.
Each building has been designed and detailed in
accordance with Eurocode 8 [11], Parts 1-1 to 1-3, as a

The inelastic analyses have been performed using the


adaptive static and dynamic structural analysis program
ADAPTIC, a program developed at Imperial College
[13] for the nonlinear analysis of steel, reinforced concrete and composite structures under static and dynamic
loading. The program utilises the layered fibre
approach for inelastic RC frame analysis and has the
capability of predicting the large displacement response
of elastic and inelastic plane and space frames. It has
also the feature of representing the spread of inelasticity
within the member cross section and along the member
length. It is widely accepted that this technique is more
accurate than the pointhinge models mainly used in
many other programs, especially when large axial force
variations exist. The program has been verified elsewhere [1416].
To accurately predict the inelastic seismic response of
the structure with sufficient accuracy, due care has been
given to create detailed and efficient models of the structures, taking into account all necessary geometric and
strength characteristics of columns, beams and beam
column connections. Towards minimising the computational requirements and the volume of input and output
data to be handled, an effort was made to select powerful
two-dimensional models that can provide, with appropri-

A.M. Mwafy, A.S. Elnashai / Engineering Structures 23 (2001) 407424

409

Table 1
Definition of the structural systems under analysis
Group
1

Reference
name

No. of storeys and structural


system

Ductility
class

IF-H030
IF-M030
IF-M015
IF-L015
RF-H030
RF-M030
RF-M015
RF-L015
FW-H030
FW-M030
FW-M015
FW-L015

8-storey irregular frame

High
Medium
Medium
Low
High
Medium
Medium
Low
High
Medium
Medium
Low

12-storey regular frame

8-storey regular frame-wall

Design acc. (g)

Force red. factor

Elas. fund. period (s)

0.30

4.00
3.00
3.00
2.00
5.00
3.75
3.75
2.50
3.50
2.625
2.625
1.75

0.674
0.654
0.719
0.723
0.857
0.893
0.920
0.913
0.538
0.533
0.592
0.588

0.15
0.30
0.15
0.30
0.15

Fig. 1. Plane and cross sectional elevation of the buildings: (a) 8-storey irregular frame buildings; (b) 12-storey regular frame buildings; (c) 8storey regular frame-wall buildings.

ate selection of parameter values, acceptable representation of the cyclic inelastic behaviour on member and
structure levels, while guaranteeing numerical stability.
The choice of two-dimensional modelling may be also
justified in the light of satisfying basic code requirements

for such type of modelling. Two-dimensional analyses


are undertaken in one direction only (global X-direction
of frame structures and global Z-direction of frame-wall
ones). This is supported by the fact that conservative
response parameters will be obtained as a result of the

410

A.M. Mwafy, A.S. Elnashai / Engineering Structures 23 (2001) 407424

domination of gravity loads in long beam spans in the


frame structures. On the other hand, the critical strain
and shear criteria are expected to occur in the coupling
beams in the dual structural systems. Taking advantage
of symmetry, only the interaction of two distinct frames
(one internal and another external) is considered. Both
lateral load-resisting frames are assembled using an
overlay approach, which is illustrated in Fig. 2 for the
irregular frame structure. The two frames are coupled
appropriately with regard to translational and rotational
degrees of freedom by 2D joint elements to meet the
assumption of infinite in-plane stiffness of the slab in
the normal direction.
The structural mesh utilises three elements for beam
column members, the lengths of which are determined
on the basis of the critical member lengths. These
lengths are determined according to EC8 provisions for
different ductility classes. The ends of horizontal
elements within the beamcolumn joints are considered
rigid. Consequently, two elements are added to each
beam at its extremities. Furthermore, shear spring connection elements are introduced to represent the shear
stiffness of the beamcolumn connection. To simplify
calculations of the shear stiffness of the joint, the force
deformation relationship for both concrete and steel
reinforcement within the joint is assumed to be linear
elastic. Despite the simplicity of the joint modelling, global structural response obtained have been extensively
compared and checked with analyses performed by Salvitti and Elnashai [17] and Panagiotakos and Fardis [18].
These show a good conceptual agreement with the current modelling results since the drift values are on the
whole higher than the values by Salvitti and Elnashai
[17] where no provision for beamcolumn connection
behaviour was made. For the sake of brevity, only some
results of the comparison are shown in Fig. 3. The results
of the current study, for two of the 12-storey frame

Fig. 2.

buildings, are between results of the rigid beamcolumn


joint modelling of the former and the flexible, one member lumped plasticity modelling, of the latter where bar
slip effects within the joints and member shear deformations are considered.
Modelling of the core is achieved by making use of
two flexural elements, for each wall at each storey, in
order to account for splicing of bars at mid-storey height.
The elements are located at the centroid of the core Ushaped cross section and connected with beams at each
storey level using two rigid links. In addition, five
elements are used to represent each coupling beam, with
bidiagonal reinforcement represented by vectorial resolution of the inclined reinforcement area along the longitudinal and transverse directions. The same method is
utilised to represent the bidiagonal shear reinforcement
in some other beams and in the lower two storeys of the
core of the FW-H030 and FW-M030 building.
Reinforced concrete column-section and T-section are
utilised for modelling of columns and beams, respectively. Both sections, taken from ADAPTIC library,
allow the geometrical definition of the section as well
as that of the confined concrete region within it. Taking
into account the available cross sections in ADAPTIC
library, a reasonable approximation is made to replace
the original U-shaped section of the core of the framewall structures by a T-section, with the same stiffness
properties. The approximation may be justified in the
light of the two-dimensional modelling which neglects
torsion and the regularity of the structure both in terms
of stiffness and strength. Reinforcement patterns are
varied for each section as a function of stirrup spacing
in accordance with those specified in the design. Confinement factors are evaluated as described in Eurocode
8, and varied along the member length according to the
arrangement of transverse reinforcements. The effective
slab width participating in beam deformation is taken as

The overlay technique considered and description of the beamcolumn joint modelling.

A.M. Mwafy, A.S. Elnashai / Engineering Structures 23 (2001) 407424

Fig. 3.

411

Verification of modelling assumptions (a) RF-H030 building and (b) RF-L015 building (average for four artificial accelerograms).

the beam width plus 7% of the clear span of the structural member on either side of the web. Horizontal and
vertical structural members including core walls are
modelled using two-dimensional cubic elasto-plastic
beamcolumn elements, where a cubic shape function is
used for the transverse displacement [13]. This formulation is intended to represent short lengths of RC
elements, consequently, axial strain is assumed to be
constant along the element length. The numerical integration of the governing equations for this element is
performed over two Gauss sections, which have a fixed
position within the element length. The inelastic
response of the cross section is assembled from contributions of individual layers for which inelastic cyclic
material constitutive relationships are applied. The cubic
elasto-plastic elements are combined with material models for concrete, which account for active confinement
and reinforcing steel with nonlinear hardening. On the
concrete side, the uniaxial constant confinement concrete
model, Martinez-Rueda and Elnashai [19], has been
chosen. For steel, the advanced multisurface steel model
for cyclic plasticity, which defines the stressstrain
response of steel in terms of a series of cubic polynomials, Elnashai and Izzuddin [20], is utilised. The
parameters used in the material models are the mean
values.

4. Load pattern and seismic action


According to the data used in the design [12], a live
load Q=2.0 kN/m2 is considered to calculate the total
gravity loads on the frames, which is applied as point
loads at nodes. Using the appropriate coefficients from
EC8, the vertical loads are combined with seismic
actions in a combination of 1.0G+0.15Q+EL for all stories except the top floor, where it was taken equal to

1.0G+0.30Q+EL. To account for inertia effects during


dynamic analysis, masses are calculated in a manner
consistent with the gravity loading combinations and are
represented by lumped 2D mass elements.
Due to the fact that the lateral force profiles in static
pushover analyses will influence the structural response,
three different load patterns have been utilised to represent the distribution of inertia forces imposed on the
building. The first shape is calculated as SRSS combinations (for the first three modes) of the load distributions obtained from modal analyses of the buildings.
The choice of this load shape is made to take into consideration the anticipated effect of higher modes of
vibrations for moderate long period and irregular structures (the 12-storey and the 8-storey frame buildings),
as well as for buildings with hybrid lateral resistance
systems (the 8-storey frame-core structures). The design
code lateral load pattern and a uniform load distribution
shape have been also utilised. The latter represents the
lateral forces that are proportional to the vertical distribution of the mass at various levels. On the other hand,
the code lateral load shape represents the forces obtained
from the predominant mode of vibration. The use of the
uniform load shape may be justified in the light of a
possible soft storey mechanism of the 8-storey irregular
buildings. If this mechanism occurs the response will be
controlled by a large drift in the first storey. Therefore,
this load distribution may give better predictions of the
overall response. The inverted triangular (code) and the
rectangular (uniform) load shapes also represent the
extreme cases from the linear distribution point of view.
The shape of the lateral load should be selected on
the light of anticipated changes in inertia forces as the
structure moves from the elastic to the plastic phases.
Ideally, this shape should be modified with the changes
in inertia forces during the actual earthquake. These
changes mainly depend on the characteristics of both the

412

A.M. Mwafy, A.S. Elnashai / Engineering Structures 23 (2001) 407424

record and the structure. Several trials [2,3] have been


made to permit of changes in inertia forces with the level
of inelasticity through the use of adaptive load patterns.
The underlying approach of this technique is to redistribute the lateral load shape with the extent of inelastic
deformations. The load shape is suggested to be redistributed according to the global displacement shape, the
level of storey shear demands or a combination of mode
shapes obtained from secant stiffnesses. This redistribution is performed at each time step, which leads to a
substantial increase in the computational effort. Moreover, the pushover analysis has not been widely established as yet in the design office environment. Therefore,
for common types of building the need for more taxing
approaches is by no mean fully established. The variable
load distribution option may be appropriate for special
and long period structures, despite that eminence of this
technique has not been confirmed yet [9,10]. On this
basis, the aforementioned fixed load distribution shapes
have been utilised for the current study. It is also worth
mentioning that the NEHRP (FEMA 273) guidelines recommend utilising fixed load patterns with at least two
load profiles. The first shape should be the uniform load
distribution and the other is the code profile or the load
shape obtained from multimodal analyses. The code lateral load distribution is allowed if more than 75% of the
total mass participates in the predominant mode.
Timehistory analyses employ four artificially generated 10-s duration acceleration records, referred to as
Art-rec1 to Art-rec4, as well as two natural records. The
artificial accelerograms were generated to fit the Eurocode 8 elastic response spectrum for medium soil class
as shown in Fig. 4 for a PGA=0.3 g. The use of the
artificial accelerograms is in order to allow effective
comparisons and calibrations with the design code.
Moreover, the effect of the vertical component of the
seismic excitation is worthy of consideration [21], particularly for the irregular frame structures where the

Fig. 4. Acceleration spectra for the artificial accelerograms (5%


damping).

planted columns are supported by long span beams.


Towards this end, two natural ground motions have been
selected in terms of the V/H ratio (peak vertical-to-horizontal acceleration). The Kobe (Hyogo-ken Nanbu at
Kobe University, Japan, 1995) and the Loma Prieta
(Northern California at Saratoga Aloha Ave, USA,
1989) earthquakes are employed and applied with and
without the vertical components, giving two analyses for
each record. However, for the sake of brevity, results of
the effect of vertical ground motion on the seismic
response are not presented herein. Comprehensive
results of this study are given elsewhere [22]. Characteristics of the records that have been used are given in
Table 2, while their acceleration response spectra for 5%
damping are shown in Fig. 5.
In order to apply the outlined procedure for the evaluation of dynamic collapse envelopes, scaling of the records utilised is frequently required. The technique of scaling earthquake records to possess equal values of
spectrum intensity was based on a proposal by Housner
[23]. The spectrum intensity is defined as the area under
the pseudo-velocity spectrum between certain period
limits. It is suggested in the current study to modify the
limits employed in the original method (between 0.1 and
2.5 s) to be between 0.8 Ty and 1.2 T2D, where Ty and
T2D are the inelastic periods of the structure at global
yielding and at twice the design ground acceleration,
respectively. This follows the proposal of MartinezRueda [24,25], modified for the range used here. Therefore, the normalisation factor for an accelerogram (n) is
equal to the ratio SIc/SIn. Where, SIc and SIn are the
areas under the code velocity spectrum and the velocity
spectrum of the scaled accelerogram, respectively. SIc
and SIn are calculated between periods of 0.8 Ty and 1.2
T2D, as explained above.
It is also worth mentioning that there is no need to
use the aforementioned scaling method with the artificial
accelerograms since they are already spectrum-compatible. Hence, the four artificial records are scaled according to their PGA. The buildings are analysed first under
the artificial records at different PGA levels and the
recorded top response time history is utilised to obtain
the inelastic periods Ty and T2D of each building from
Fourier analyses (average for four artificial records). The
scaling factors are then calculated for the longitudinal
component of natural records and used for scaling the
accelerograms up to collapse. The factors used to scale
the longitudinal earthquake component are also used to
scale the vertical component of the motions, when
employed, to keep the V/H ratio constant. Table 3 shows
the average normalisation factors to ground acceleration
0.30 g for each of the three groups of buildings. Finally,
it should be noted that the quoted values of PGA are not
of the natural or scaled records but rather multiples of
the design ground acceleration.

A.M. Mwafy, A.S. Elnashai / Engineering Structures 23 (2001) 407424

413

Table 2
Characteristics of records used in analysis
Earthquake

Date

Ms

Kobe (Japan)
17/01/95
Loma Prieta (USA)
18/10/89
Artificial Records Art-rec1, Art-rec2, Art-rec3,
and Art-rec4

7.20
7.17

Table 3
Normalisation factors for ground acceleration 0.30 g
IF-buildings RF-buildings FW-buildings

Artificial
0.30
Kobe (KBU) 0.54
Loma Prieta
1.15
(SAR)

KBU
SAR

PGA (g)
Horiz.

Vert.

0.276
0.319

0.431
0.349

V/H

No. of input runs

1.56
1.09

2
2
4

Fig. 5. Elastic spectra for the long. component of the natural records
(5% damping).

Earthquake

Station

Average

0.30
0.61

0.30
0.56

0.30
0.57

1.25

1.32

1.24

5. Collapse criteria and incremental dynamic


collapse dynamic pushover results
Three types of analyses have been performed using
the structural models described earlier. Eigenvalue
analyses are conducted to determine the elastic periods
and the mode shapes of the buildings needed for calculating the first lateral load profile of the static pushover
analysis (combination of loads from modal shapes).
Inelastic static pushover and dynamic analyses are then
performed using the calculated lateral load shapes and
the seismic actions with increasing severity. The analyses are progressed until all the predefined collapse limits
are exceeded. In both static and dynamic analyses, permanent loads are first applied and iteration to equilibrium is performed. This is followed by applying the horizontal action (loads or ground acceleration). The analysis

is both inelastic and geometrically nonlinear. The large


displacement formulation is an updated Lagrangian
form, where convected member axes are used to derive
member deformations.
The criteria used for defining collapse are classified
into two groups; local and global criteria. Two failure
criteria on the member-level are applied: the ultimate
curvature, which is normally controlled by the maximum
compression strain at the extreme fibre of the confined
concrete and shear failure in any structural member. An
empirical axial load-sensitive shear model capable of
providing an experimentally verifiable estimate of shear
supply in RC members was proposed by Priestley et al.
[26] and has been utilised in this study [27]. The code
shear supply model has also been employed after eliminating the design safety factors. On the structure level,
three collapse criteria are chosen: a limit corresponds to
a maximum inter-storey drift of 3% of the storey height,
formation of a sidesway mechanism and reduction in lateral resistance by considering the loaddisplacement
curve of the structure. Additionally, the criterion used to
define global yield threshold, which is essential for the
proposed scaling method of the records, is selected as
the yield displacement of the equivalent elasto-plastic
system with reduced stiffness evaluated as the secant
stiffness at 75% of the ultimate load of the real system.
The utilised shear models are implemented with other
collapse and yield criteria in a post-processing program
connected to ADAPTIC [28]. This post-processsor traces
the shear supplydemand situation at each time step at
both ends of all members. It also performs the appropriate calculations to evaluate the local and global response
parameters of the structure and directly apply the selected criteria.
The results of more than 1300 inelastic timehistory
analyses were employed to perform regression analyses
to obtain the dynamic pushover (ideal) envelope for each
of the twelve examined buildings. Figs. 68 depict
dynamic response points and the fitted regression equations of the response of the buildings subjected to the
eight seismic actions considered for all limit states. The
fitted envelopes for the upper and lower response points,
the number of analyses carried out, the design base shear
and the correlation coefficient for each case are also

414

A.M. Mwafy, A.S. Elnashai / Engineering Structures 23 (2001) 407424

Fig. 6.

Dynamic collapse analysis results for the irregular frame structures.

shown. The actual response of the 8-storey irregular


frame structures illustrated in Fig. 6 show how the
results of the eight seismic actions follow the same trend
and shape the pushover envelopes without the need to
apply curve fitting. This is clear from the correlation
coefficient values, which are almost invariably above
0.9. It is worth mentioning that the number of timehistory analyses shown on each graph varies according to
the number of trials needed to identify the collapse and
the yield limits, as discussed above.
Concerning the 12-storey regular frame structures,
Fig. 7 shows a higher scatter in the dynamic analysis
results of different ground motions than the results of
the 8-storey irregular frame buildings. Moreover, the
scatter for the two buildings designed for the higher
design ground acceleration gives the impression of being
higher than the other pair of results. The low correlation
of the former and the high correlation of the latter are
reflected in the correlation values which are equal to 0.69
and 0.66 for the first pair and 0.93 and 0.88 for the
second one. It should be noted that the main difference
between the two pairs is in the longitudinal and transverse reinforcement of the structural members, while the

dimensions of the cross-sections for this group are the


same except a slight changes, mainly in the beams crosssection width (from 0.35 m for the first pair to 0.30 m
for the second one). In spite of the aforementioned
observations, the difference in scatter between the higher
and the lower design ground acceleration pair diminishes
when calculating the difference between the lower and
the upper response envelope for each case (quotient of
minimum and maximum strength for the eight records).
This value is equal to 0.69 and 0.72 for the 0.30 g design
ground motion pair and 0.76 and 0.74 for the other pair.
This is more consistent since the calculated inelastic periods, which are the main cause of the different response,
are very close for the four buildings, as subsequently
discussed. Finally, the difference in the correlation
values between the two pair of buildings can only be
justified in the light of the lower number of runs (or
response points) needed to achieve yield and collapse
for each pair. This number is equal to 97 and 105 for
the first pair and 60 and 73 for the second one.
The high sensitivity to changes in the input motion
observed in the 12-storey frame buildings are also
reflected, to a lesser extent, in the 8-storey frame-wall

A.M. Mwafy, A.S. Elnashai / Engineering Structures 23 (2001) 407424

Fig. 7.

415

Dynamic collapse analysis results for the regular frame structures.

group. Fig. 8 illustrates the results of the timehistory


analyses for this group. At collapse limit state, a scatter
is observed for values of VMin/VMax shown in Table 4.
A higher hardening stiffness is also observed for the 0.30
g design ground acceleration pair compared to the other
pair. This is not observed in the other two groups of
building. Previous analytical investigations [29,30] have
indicated that base shear demands of wall structures are
sensitive to higher mode effects. Once a plastic hinge
has formed at the base of the wall, higher mode effects
can considerably amplify the base shear as well as the
shear at each storey level. The results shown in Fig. 8
confirm that such amplification may occur and could be
large. It is also worth mentioning that the thickness of
the core-walls for the higher design ground acceleration
pair is 0.35 m, compared to 0.25 m for the other pair.
This causes an increase in the mass at each storey level
for the former, hence higher amplification of base shear
demand. Alongside the high initial stiffness of this pair,
the difference in response between the two pairs of
building shown in Fig. 8 can be explained.

6. Contribution of the elongated period to the


seismic response
The scatter observed for some buildings is mainly in
the post-elastic range, and is associated with the spread
of yielding and member failure throughout the structure.
Subsequently, the stiffness of the structure decreases, the
fundamental period elongates and the distribution of the
inertia forces along the building undergoes continuous
change. To provide insight into the response of the
investigated buildings, extensive analyses in the frequency domain (Fourier analyses) of the acceleration
response at the top have been conducted to identify the
predominant inelastic period of each building under consideration. Fig. 9 illustrates the calculated periods
(average for the eight seismic actions) at the design and
twice the design ground acceleration, along with the
elastic period for each building calculated from eigenvalue analyses.
It is clear to what extent the fundamental periods of
the buildings are elongated as a result of the spread of

416

A.M. Mwafy, A.S. Elnashai / Engineering Structures 23 (2001) 407424

Fig. 8.

Dynamic collapse analysis results for the frame-wall structures.

Table 4
Observed response at global collapse for the eight records
Group

Reference name

IF-H030
IF-M030
IF-M015
IF-L015
RF-H030
RF-M030
RF-M015
RF-L015
FW-H030
FW-M030
FW-M015
FW-L015

Min/Max

Roof disp. (mm)


Max

Min

Mean

613
635
492
590
690
796
735
785
643
660
643
652

503
500
381
380
580
611
630
607
599
576
590
598

542
570
449
465
625
684
681
694
631
625
621
626

0.82
0.79
0.77
0.64
0.84
0.77
0.86
0.77
0.93
0.87
0.92
0.92

Base shear (kN)


VMax

VMin

Mean

11,614
13,930
7699
9229
15,647
16,278
9743
12,735
20,821
23,300
12,724
16,153

9918
12,713
6663
8102
11,568
12,076
9234
11,009
15,520
18,123
8769
11,604

10,567
13,146
7123
8685
13,689
13,990
9453
11,972
17,849
20,738
10,642
13,425

VMin/VMax

Storeya

0.85
0.91
0.87
0.88
0.74
0.74
0.95
0.86
0.75
0.78
0.69
0.72

1,
1,
1,
1,
2,
4,
5,
4,
2,
2,
2,
2,

2,
2,
4,
2,
4,
5,
6,
5,
3,
5,
3
3,

3,
4,
5
4,
8,
8,
8
9
6,
7

5
5
5
9
9

Location where interstorey drift collapse criterion is observed for the eight ground motions.

cracks and yielding. The average elastic periods for the


three groups of building are 0.69, 0.90, and 0.56 s,
respectively. On the other hand, the calculated inelastic
periods at the design and twice the design ground acceleration are (1.301.46), (1.651.80), and (0.811.00) s,

respectively. It is observed that the average percentage


of elongation in the period is (100%), (90%), and (60%).
The percentage increase is clearly related to the overall
stiffness of the structural system of the building. The
maximum calculated elongation is recorded in the most

A.M. Mwafy, A.S. Elnashai / Engineering Structures 23 (2001) 407424

417

Fig. 9. Elastic and inelastic (at the design and twice the design ground acceleration) predominant response periods of the buildings average
for the eight seismic actions.

flexible system, where the first storey can be considered


as a soft storey; whereas the minimum elongation is
observed in the stiff frame-wall structural system. The
results point towards an important conclusion, employment of elastic periods of vibration in estimating design
forces leads to high levels of overstrength (ratio of
actual-to-required strength). Moreover, they lead to nonuniform safety margins for different structural systems.
To facilitate the comparison between the input accelerograms utilised in this study in terms of the frequency
content, the records are scaled to ground acceleration
equal to 0.30 g and used to obtain the Fourier spectrum
for each record. The normalisation factors used are the
average of the scaling factors utilised to perform the collapse analysis for the three groups of buildings, as shown
in Table 3. The Fourier amplitude spectra for the acceleration records (one of the utilised artificial records, Artrec1, and the horizontal component of the two natural
records) are shown in Fig. 10. The average inelastic period for each group of building is also shown on the
graphs. It is clear that the input motions, with the exception of the Kobe (KBU) record, have high amplitude that
may amplify the effect of the second mode of vibration
for structures with period between 0.35 and 0.50 s. The
amplitude in this period range is higher than the amplitude corresponding to the fundamental period of the
three groups of structures. This is one of the reasons for
the scatter in the results of the 12-storey regular frame
and the 8-storey frame-wall structures. It is verified that
the high response points in Figs. 7 and 8 are for the
artificial and Loma Prieta (SAR) records, while the low
response is for the Kobe (KBU) record. This may also
justify obtaining a higher maximum base shear corresponding to almost an identical top deflection when
applying the same ground motion with higher PGA. Furthermore, the Fourier spectral ordinate corresponding to
the inelastic fundamental period of the buildings can also
be utilised to justify the scatter in the results of the
second and the third group of buildings. For the 12-storey buildings, the average inelastic fundamental period

is about 1.75 s, which corresponds to high amplification


in the Loma Prieta (SAR) record only. This also
accounts for the high response of the 12-storey frame
structures when subjected to the latter record. The same
applies, to a lesser extent, to the artificial and Loma Prieta (SAR) records when imposed on the frame-wall
structures (inelastic period 0.91 s), compared to the Kobe
(KBU) record. For this reason the observed scatter for
this group is less than the 12-storey buildings. On the
other hand, the ordinates of the spectra correspond to the
inelastic period of the 8-storey irregular buildings are
equivalent, hence the high correlation for this group.

7. Inelastic static-to-collapse static pushover


analyses
Following the success in obtaining the incremental
dynamic response envelopes for the twelve buildings
under investigation, inelastic static pushover analyses are
performed to assess the applicability of the technique
(for different load distributions) in predicting the overall
dynamic response of structures. Figs. 1113 illustrate the
base shear vs top displacement plots for the three lateral
load profiles utilised along with the incremental dynamic
envelopes for the twelve buildings. The dynamic pushover curve for each case is shown in the form of the
upper and lower response envelope as well as the best
fit of the timehistory analysis results. The global yield
and collapse thresholds are also shown. It should be
pointed out that it was decided to choose only one global
yield limit from the limits obtained from the four pushover envelopes (the three static and the dynamic one).
This is due to the need to unify and simplify obtaining
this limit, which is necessary for the suggested method
of scaling the input seismic actions explained earlier.
The yield limit state obtained from the static pushover
analysis using the code lateral load shape is selected for
this purpose. For collapse, the observed upper and lower
global collapse limits from the eight earthquake records

418

A.M. Mwafy, A.S. Elnashai / Engineering Structures 23 (2001) 407424

Fig. 10. Fourier amplitude spectra for the input accelerograms (scaled to 0.30 g) and the average inelastic period of the buildings.

Fig. 11.

Static and dynamic pushover analysis results for the irregular frame structures.

A.M. Mwafy, A.S. Elnashai / Engineering Structures 23 (2001) 407424

419

Fig. 12. Static and dynamic pushover analysis results for the regular frame structures.

as well as from the three pushover analyses are


presented. It is important, however, to note that the static
pushover and the incremental dynamic collapse analyses
are continued beyond all the predefined collapse thresholds. This is to ensure that all collapse states are bounded
by the dynamic analysis.
The static pushover method is rarely used to predict
seismic demands when a particular ground motion is
imposed on a structure [9,10]. If this is needed, the top
target displacement expected when this ground motion
is imposed on the building should be estimated. It is
beyond the scope of this study to address the approaches
of estimating the target displacement. A review of these
methods was given in the latter two references. Since
the main application of the static pushover analysis is to
estimate the seismic capacity of structures, the following
observations are driven by this requirement. In this
application of the procedure, the analysis is usually continued until any of the predefined collapse criteria is
exceeded.
In general, it is clear in all cases that the response of
the buildings is sensitive to the shape of the lateral load
distribution. This is particularly true when moving from

the code and the multimodal load patterns to the uniform


load shape. It is also noticeable that the difference
between load shape A (the design code load pattern,
which is almost an inverted triangle) and load shape B
(load shape from multimodal analysis) is very small.
Although higher mode effects are confirmed in the
response of the second and the third group of buildings,
as explained earlier, the multimodal analysis load pattern
did not show an enhanced capability to predict these
effects. This is due to the fact that this load shape represents the distribution of inertia forces in the elastic
range only, while the amplification of higher mode
effects are observed in the post-elastic phase. Table 5
presents the results at global collapse limit state for the
three load shapes. In terms of the predicted ultimate
strength and drift at collapse, the differences between
load A and B are less than 4%, for the twelve buildings.
As a general trend, the collapse is observed earlier
when applying the uniform load than the triangular load.
Collapse is observed slightly earlier than the triangular
distribution when imposing the multimodal load. In Figs.
1113 the lower collapse limits from static analyses are
always from the uniform load and the upper limits are

420

A.M. Mwafy, A.S. Elnashai / Engineering Structures 23 (2001) 407424

Fig. 13. Static and dynamic pushover analysis results for the frame-wall structures.

Table 5
Results at global collapse limit state for the three load patterns
Group

Reference name

IF-H030
IF-M030
IF-M015
IF-L015
RF-H030
RF-M030
RF-M015
RF-L015
FW-H030
FW-M030
FW-M015
FW-L015

a
b
c
d

Roof disp. (mm)

Storeyd

Base shear (kN)

Aa

Bb

Cc

Aa

Bb

Cc

Aa

Bb

Cc

534
552
474
516
648
712
656
688
570
580
575
590

528
534
462
498
624
688
640
664
560
570
565
580

508
480
432
450
552
568
600
592
535
545
530
545

10,091
12,690
6652
8253
12,135
13,083
7332
9817
13,243
16,671
7880
10,001

10,446
13,056
6914
8508
12,499
13,444
7554
10,136
13,796
17,241
7988
10,119

11,592
14,219
7620
9147
14,650
15,748
9235
12,175
16,425
20,754
9843
12,490

4th
3rd
3rd
3rd
5th
5th
5th
5th
3rd
3rd
3rd
3rd

3rd
3rd
3rd
2nd
4th
5th
5th
5th
3rd
3rd
3rd
3rd

3rd
2nd
2nd
2nd
3rd
3rd
4th
4th
3rd
3rd
3rd
2nd

Triangular load.
Multimodal load.
Uniform load.
Storey at which collapse is observed.

A.M. Mwafy, A.S. Elnashai / Engineering Structures 23 (2001) 407424

from the triangular load. Moreover, the maximum interstorey drift collapse limit when employing the uniform
load pattern is observed in lower storeys than the
recorded collapse from the triangular or the multimodal
loads. This observation is also recorded in dynamic
analyses, where collapse is observed in lower storeys for
records that impose higher base shear. Fig. 14 depicts
the relationship between collapse limit states of the three
load shapes for one of the investigated buildings
(building RF-L015). It should be emphasised that the
loaddeformation envelope is for global response, which
is a function of the point of application of resultant force.
The uniformly distributed load gives the lowest point;
hence the maximum strength and earlier global yield and
collapse limit states. On the other hand, resultant in the
triangular load case is applied at a higher point; consequently lower strength and delayed global yield and collapse are observed in all cases.
Despite the fact that all load shapes do not represent
the actual distribution of relative inertia forces during
the dynamic analysis, almost an identical response is
observed in the first group of buildings between the
dynamic analysis best-fit envelopes and the static
response obtained from the triangular and the multimodal distributions. On the other hand, the uniform load
overestimates the initial stiffness and the maximum base
shear in the four buildings. Table 6 illustrates graphically
the differences between the results of the static pushover
analysis for the triangular and the uniform load patterns
on one side, and the incremental dynamic analysis
(average for eight records) on the other, at global collapse limit state. Since the triangular load shape is simple
and show very close results with the multimodal load
pattern, it was decided to exclude the latter from this
comparison. It is clear that the uniformly distributed load
is unconservative in predicting collapse limit states
(underestimates the drift and overestimates the strength).
The overall prediction of collapse using the triangular
load is significantly better. Although it slightly under-

Fig. 14.

421

estimates the average drift of dynamic analysis in two


buildings, it is between the upper and lower drift limits
obtained from the eight records, as shown in Fig. 11.
Higher modes effect and concentration of inelastic
deformations are expected to be significant in the first
group of structures where the buildings exhibit two
sources of irregularity and weak storey. Notwithstanding, four static pushover analyses using the simple triangular load pattern have succeeded in predicting the
average results of more than 600 inelastic timehistory
analyses. It is also important to note that the good design
of these buildings and the high overstrength associated
with structural elements, particularly the columns, prevented any undesirable mode of failure. The results show
that utilising the triangular load shape only to predict the
global response of low rise frames as well as well
designed irregular frame structures is adequate.
In contrast to the first group of buildings, the results
of the static pushover of the 12-storey group, illustrated
in Fig. 12, show discrepancies with the dynamic
response envelope in the post-elastic range. While the
static pushover results of the triangular and the multimodal load pattern show a good agreement with the
dynamic results best fit in the elastic range, both give a
conservative prediction of the maximum lateral strength,
as also shown in Table 6 for the triangular load. However, in the four buildings the triangular load response
is higher than the lower limit envelopes obtained from
dynamic analyses employing natural and artificial records. On the other hand, the capacity curve obtained
from the uniformly distributed load overestimates the
response in the elastic range. However, it gives better
prediction of the ultimate strength. It is also clear from
Fig. 12 that the triangular load shape gives good prediction of the deformation at collapse, while the uniform
load underestimates the collapse limit state in the four
buildings. It is concluded for this group of buildings that
the triangular distribution is again the most suitable load
pattern given that the uniform load, which is rec-

Differences between the three lateral load patterns.

422

A.M. Mwafy, A.S. Elnashai / Engineering Structures 23 (2001) 407424

Table 6
Differences between static and dynamic pushover analysis at global collapse limit state
Group

Difference in drift
Triangular

Difference in strength
Uniform

Triangular

Uniform

ommended by the NEHRP guidelines (FEMA 273), is


unconservative in predicting the response and the drift
at collapse.
Concerning the frame-wall group of buildings, the differences between the static pushover and the dynamic
analysis results are larger than for the other two groups.
In terms of the predicted elastic response and initial stiffness, the triangular and the multimodal load shapes show
good correlation with the dynamic analysis best fit for
three buildings and a conservative prediction for the
fourth (FW-M030). In the post-elastic range, the two
load shapes underestimate the lateral strength obtained
from the timehistory analyses. Table 6 shows that the
triangular load prediction of strength at global collapse
is between 20 and 26% less than the average results of
dynamic analyses. Similar to the 12-storey buildings, the
uniform load overestimates the elastic response but gives
better prediction of the lateral strength at collapse for
this group of buildings. Moreover, none of the investigated load patterns give reasonable prediction of the high
hardening stiffness obtained from dynamic analysis for
the higher design ground motion pair. With regard to
predicting the drift at collapse, both the triangular and
the uniform load patterns are unconservative. This is
clear from Fig. 13 and Table 6.
The comparison between static and dynamic pushover
analysis for this group of buildings shows more discrepancies than the second group, especially for the 0.30 g
design ground motion pair. As explained earlier, these
differences are mainly due to higher mode effects, which
amplify the base shear following formation of first plastic hinge at the base of the wall. In pushover analysis,
once the wall attained its ultimate lateral strength, it will
deform by plastic hinging at the base [31]. Clearly, for
this type of structure (shear frame response plus cantil-

ever wall response) the amplification of the base shear


during the dynamic analysis is difficult to predict by
pushover analysis. However, the triangular load profile
shows good correlation in the elastic range, conservative
predictions of the ultimate strength and reasonable estimations of the collapse limit state (underestimates the
drift by about 8%). Hence, it may be employed for estimating the seismic capacity and collapse limit state.
Finally, if the static pushover analysis is utilised as a
tool for predicting seismic demands instead of estimating
capacities, the analysis is usually performed until the
roof drift corresponding to the design ground acceleration is attained. Table 7 presents the average for eight
ground motions of the maximum top displacement
observed from timehistory analyses at the design
ground acceleration. Clearly, the target displacement is
almost always below the global yield limit state. The
comparison between the static pushover and the dynamic
analysis discussed above show that the triangular load
gives better estimation of the response in the elastic
range. In few buildings, however, it underestimates the
initial stiffness. In contrast, in the same range the uniform load shape overestimates the stiffness and the base
shear in all cases. From the design point of view, the
uniformly distributed load is conservative for the twelve
buildings investigated. It is concluded that the use of two
load distributions is needed for estimating the seismic
demand. The simple triangular or the multimodal shape,
which correlate well with dynamic analysis results and
the uniform load pattern which shows a conservative
prediction of demands in almost all cases considered.
This conclusion is supported by the observations
obtained from the results of the frame-wall group of
buildings. The uniform load pattern can provide a conservative estimation of shear demand below collapse

A.M. Mwafy, A.S. Elnashai / Engineering Structures 23 (2001) 407424

423

Table 7
Observed maximum roof displacement at the design ground acceleration (average for eight ground motions)
Group 1

Roof disp. (mm)

Group 2

Roof disp. (mm)

Group 3

Roof disp. (mm)

IF-H030
IF-M030
IF-M015
IF-L015

285
202
138
136

RF-H030
RF-M030
RF-M015
RF-L015

290
304
146
163

FW-H030
FW-M030
FW-M015
FW-L015

115
120
56
67

limit states; hence it can support preventing undesirable


shear failure. Moreover, it shows reasonable estimation
of shear at global collapse limit state (at 3% interstorey
drift). It is worth mentioning that, according to the
SEAOC (Vision 2000), complete collapse is considered
once the interstorey drift exceeds 2.5%. Utilising this
definition of collapse leads to obtaining conservative
prediction of shear demand in all buildings investigated
when employing the uniformly distributed load.

8. Conclusions
The applicability and accuracy of inelastic static pushover analysis in predicting the seismic response of RC
buildings are investigated. Twelve RC buildings with
various characteristics, incremental dynamic analysis
employing eight natural and artificial records, static
pushover analysis using three lateral load distributions
and local and global limit state criteria are utilised.
Based on the large amount of information obtained,
which is nonetheless far from comprehensive, the following conclusions are drawn:
Subject to adequate modelling of the structure, careful
selection of the lateral load distribution and articulate
interpretation of the results, pushover analysis can
provide insight into the elastic as well as the inelastic
response of buildings when subjected to earthquake
ground motions.
Static pushover analysis is more appropriate for low
rise and short period frame structures. For welldesigned buildings but with structural irregularities,
the results of the procedure also show good correlation with the dynamic analysis. In this study,
response obtained for a group of four 8-storey irregular frame buildings using an inverted triangular lateral
load distribution is identical to inelastic timehistory analysis.
The experience gained from previous studies can help
to eliminate the discrepancies between static and
dynamic analysis results for special and long period
buildings. These differences are mainly due to the
limited capability of the fixed load distribution to predict higher mode effects in the post-elastic range. To
overcome this problem, more than one load pattern

should be selected to guarantee providing an accurate


or slightly conservative prediction of capacities and
demands.
The investigation carried out on two sets of four 12storey frame buildings and four 8-storey frame-wall
structures show that a conservative prediction of
capacity and a reasonable estimation of deformation
is obtained using the simple triangular or the multimodal load distribution. The same load patterns
slightly underestimate the demand of some buildings
in the elastic range. On the other hand, the uniform
load provides a conservative prediction of seismic
demands in the range before first collapse. It also
yields an acceptable estimation of shear demands at
the collapse limit state.
Comparison between the triangular and the multimodal distribution results show differences less than 4%,
for the twelve buildings, since the former captures the
characteristics of the most important mode of
vibration. The load distribution from multimodal
analysis only represents the distribution of inertia
forces in the elastic range; hence higher mode effects
are not entirely accounted for in the post-elastic
domain.
The elongation in the fundamental period of structures
due to extensive yielding and cracking during earthquakes depends on the overall stiffness of the structural system of the building. In the current study, the
observed elongation ranges between 100% for the
most flexible irregular frame system and 60% for the
stiff frame-wall structural system. Employment of
elastic periods in seismic code does not therefore provide uniform levels of safety for different structural
systems.
The results of the dynamic collapse analysis show
clearly that each earthquake record exhibits its own
peculiarities, dictated by frequency content, duration,
sequence of peaks and their amplitude. The dispersion
in the results of different ground motions depends on
the characteristics of both the structure and the record.
The Fourier spectral analysis is an important tool to
investigate the observed variability of the results and
to identify the elongated inelastic periods of the structure.
The importance of pushover analysis as an assessment
and design tool warrants much needed further devel-

424

A.M. Mwafy, A.S. Elnashai / Engineering Structures 23 (2001) 407424

opments. These may be classified as tools and


behaviour. There is considerable scope for development of tools for more efficient and versatile pushover
analysis techniques. One such development would be
the continuous assessment of the effect of inelasticity
on the load distribution used, taking into account the
shape of the spectrum. This would enable the accurate
and realistic analysis of highly irregular structures.
With regard to behaviour, analysis of a larger sample of buildings that includes high-rise structures and
structures with heavily irregular strength distribution
is needed.
To close, it is emphasised that, notwithstanding the range
of structures analysed, the number of records employed
and the rigour of the limit state criteria monitored, the
conclusions are, strictly speaking, applicable to the range
investigated. However, some generality may be claimed
by noting that every effort has been made to select distinct structural systems, comprehensive limit states and
verified investigation tools.

[12]

[13]

[14]

[15]

[16]

[17]

[18]

[19]

References

[20]

[1] Saiidi M, Sozen MA. Simple nonlinear seismic analysis of R/C


structures. Journal of the Structural Division, ASCE
1981;107(ST5):93751.
[2] Fajfar P, Gaspersic P. The N2 method for the seismic damage
analysis of RC buildings. Earthquake Engineering and Structural
Dynamics 1996;25:3146.
[3] Bracci JM, Kunnath SK, Reinhorn AM. Seismic performance and
retrofit evaluation of reinforced concrete structures. Journal of
Structural Engineering, ASCE 1997;123(1):310.
[4] FEMA. NEHRP guidelines for the seismic rehabilitation of buildings. FEMA 273, Federal Emergency Management Agency,
1996.
[5] SEAOC. Performance based seismic engineering of buildings.
Vision 2000 Committee, Structural Engineers Association of California, Sacramento, CA, 1995.
[6] Lawson RS, Vance V, Krawinkler H. Nonlinear static push-over
analysis why, when, and how? In: Proceedings 5th US NCEE,
vol. 1. IL, USA: Chicago, 1994:28392.
[7] Mitchell D, Paultre P. Ductility and overstrength in seismic
design of reinforced concrete structures. Canadian Journal of
Civil Engineering 1994;21:104960.
[8] Faella G, Kilar V. Asymmetric multistorey R/C frame structures:
push-over versus nonlinear dynamic analysis. In: Proceedings
11th ECEE, Paris, 1996, CD-Rom.
[9] Krawinkler H, Seneviratna GDPK. Pros and cons of a pushover
analysis of seismic performance evaluation. Engineering Structures 1998;20(4-6):45264.
[10] Tso WK, Moghadam AS. Pushover procedure for seismic analysis of buildings. Progress in Structural Engineering and Materials
1998;1(3):33744.
[11] Eurocode 8. Design provisions for earthquake resistance of structures. Part 1-1, 1-2 and 1-3, Comite Europeen de Normalisation,

[21]

[22]
[23]

[24]

[25]

[26]

[27]

[28]

[29]

[30]
[31]

European Pre-standard ENV 1998-1-1, 1-2, and 1-3, Bruxelles,


1994.
Fardis MN. Analysis and design of reinforced concrete buildings
according to Eurocodes 2 and 8. Configuration 3, 5 and 6, Reports
on Prenormative Research in Support of Eurocode 8, 1994.
Izzuddin BA, Elnashai AS. ADAPTIC a program for adaptive
large displacement elastoplastic dynamic analysis of steel, concrete and composite frames. ESEE Report No. 89/7, Imperial College, London, 1989.
Elnashai AS, Elghazouli AY. Performance of composite
steel/concrete members under earthquake loading, Part I: Analytical model. Earthquake Engineering and Structural Dynamics
1993;22(4):31545.
Broderick BM, Elnashai AS. Seismic resistance of composite
beamcolumns in multi-storey structures, Part 2: Analytical
model and discussion of results. Journal of Construction Steel
Research 1994;30(3):23158.
Elnashai AS, Pinho R. Repair and retrofitting of RC walls using
selective techniques. Journal of Earthquake Engineering
1998;2(4):52568.
Salvitti LM, Elnashai AS. Evaluation of behaviour factors for RC
buildings by non-linear dynamic analysis. In: Proceedings 11th
WCEE, Acapulco, Mexico, 1996, CD-Rom, Paper No. 1820.
Panagiotakos TB, Fardis MN. Effect of column capacity design
on earthquake response of reinforced concrete buildings. Journal
of Earthquake Engineering 1998;2(1):11345.
Martinez-Rueda JE, Elnashai AS. Confined concrete model under
cyclic load. Materials and Structures 1997;30(197):13947.
Elnashai AS, Izzuddin BA. Modelling of material nonlinearities
in steel structures subjected to transient dynamic loading. Earthquake Engineering and Structural Dynamics 1993;22:50932.
Papazoglou AJ, Elnashai AS. Analytical and field evidence of the
damaging effect of vertical earthquake ground motion. Earthquake Engineering and Structural Dynamics 1996;25:110937.
Mwafy AM. Seismic performance of code-designed RC buildings. PhD thesis, Imperial College, University of London, 2000.
Housner G. Spectrum intensities of strong-motion earthquakes.
In: Proceedings of the Symposium on Earthquake and Blast
Effects on Structures. CA, USA: Los Angeles, 1952:2036.
Martinez-Rueda JE. Energy dissipation devices for seismic
upgrading of RC structures. PhD thesis, Imperial College, University of London, 1997.
Martinez-Rueda JE. Scaling procedure for natural accelerograms
based on a system of spectrum intensity scales. Earthquake Spectra 1998;14(1):13552.
Priestley MJN, Verma R, Xiao Y. Seismic shear strength of
reinforced concrete columns. Journal of Structural Engineering,
ASCE 1994;120(8):231029.
Elnashai AS, Mwafy AM, Lee D. Collapse analysis of RC structures including shear. ASCE Structures Congress, New Orleans, 1999.
Mwafy AM. Seismic performance of RC buildings under multiaxial earthquake loading. PhD transfer report, Imperial College,
London, 1998.
Ghosh SK. Required shear strength of earthquake resistant
reinforced concrete shear walls. In: Krawinkler H, Fajar P, editors. Nonlinear Seismic Analysis and Design of Reinforced Concrete Buildings. Oxford: Elsevier Science, 1992.
Seneviratna GDPK. Evaluation of inelastic MDOF effects for
seismic design. PhD thesis, Stanford University, 1995.
Krawinkler H. New trends in seismic design methodology. In:
Proceedings 10th ECEE. The Netherlands: Rotterdam,
1995:82130.

Vous aimerez peut-être aussi