Sie sind auf Seite 1von 22

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/257355422

Optimum lateral load pattern for seismic


design of elastic shear-buildings incorporating
soil–structure interaction effects

Article in Earthquake Engineering & Structural Dynamics · May 2013


DOI: 10.1002/eqe.2252

CITATIONS READS

12 161

2 authors:

Behnoud Ganjavi Hong Hao


University of Mazandaran Curtin University
22 PUBLICATIONS 99 CITATIONS 553 PUBLICATIONS 7,856 CITATIONS

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Dynamic behavior of precast segmental bridge piers under vehicle collision View project

See the Project Goal...... View project

All content following this page was uploaded by Behnoud Ganjavi on 25 December 2017.

The user has requested enhancement of the downloaded file.


EARTHQUAKE ENGINEERING & STRUCTURAL DYNAMICS
Earthquake Engng Struct. Dyn. 2013; 42:913–933
Published online 7 September 2012 in Wiley Online Library (wileyonlinelibrary.com). DOI: 10.1002/eqe.2252

Optimum lateral load pattern for seismic design of elastic


shear-buildings incorporating soil–structure interaction effects

Behnoud Ganjavi*,† and Hong Hao


School of Civil and Resource Engineering, The University of Western Australia, Crawley, WA 6009, Australia

SUMMARY
Recently, several new optimum loading patterns have been proposed by researchers for fixed-base systems
while their adequacy for soil–structure systems has not been evaluated yet. Through intensive dynamic
analyses of multistory shear-building models with soil–structure interaction subjected to a group of 21 artificial
earthquakes adjusted to soft soil design spectrum, the adequacy of these optimum patterns is investigated. It is
concluded that using these patterns the structures generally achieve near optimum performance in some range of
periods. However, their efficiency reduces as soil flexibility increases especially when soil–structure interaction
effects are significant. In the present paper, using the uniform distribution of damage over the height of
structures, as the criterion, an optimization algorithm for seismic design of elastic soil–structure systems
is developed. The effects of fundamental period, number of stories, earthquake excitation, soil flexibility,
building aspect ratio, damping ratio and damping model on optimum distribution pattern are investigated.
On the basis of 30,240 optimum load patterns derived from numerical simulations and nonlinear statistical
regression analyses, a new lateral load pattern for elastic soil–structure systems is proposed. It is a function
of the fundamental period of the structure, soil flexibility and structural slenderness ratio. It is shown that
the seismic performance of such a structure is superior to those designed by code-compliant or recently
proposed patterns by researchers for fixed-base structures. Using the proposed load pattern in this study,
the designed structures experience up to 40% less structural weight as compared with the code-compliant
or optimum patterns developed based on fixed-base structures. Copyright © 2012 John Wiley & Sons, Ltd.

Received 7 January 2012; Revised 31 July 2012; Accepted 5 August 2012

KEY WORDS: soil–structure interaction; optimum strength distribution; uniform damage distribution;
elastic behavior; seismic code

1. INTRODUCTION

It is known that structural configuration in terms of stiffness and strength distributions have a key role in
the seismic response and behavior of structures. The current seismic design codes generally regard the
seismic effects as lateral inertia forces, which is called a force-based design procedure. Therefore, the
distribution of story stiffness and strength over the height of the structures are designed primarily
based on these static forces. The shape of these lateral load patterns along the height of structures from
various standards such as EuroCode-8 [1], Mexico City Building Code [2], Uniform Building Code
[3], NEHRP 2003 [4], ASCE/SEI 7–05 [5], Australian Seismic code [6] and IBC 2009 [7] depends on
the fundamental period of the structures and their mass. They are derived primarily based on elastic
dynamic analysis of the corresponding fixed-base structures without considering soil–structure
interaction (SSI) effect. In the United States, the current code-specified seismic design procedures are
mainly based on the NEHRP Recommended Provisions published in 2003 [4]. It should be mentioned

*Correspondence to: Behnoud Ganjavi, School of Civil and Resource Engineering, The University of Western Australia,
35 Stirling Highway, Crawley, WA 6009, Australia.

E-mail: ganjavi@civil.uwa.edu.au, b.ganjavi@gmail.com

Copyright © 2012 John Wiley & Sons, Ltd.


914 B. GANJAVI AND H. HAO

that the seismic design criteria in ASCE/SEI 7–05 [5], exclusively based on the NEHRP 2003 [4], is also
adopted in IBC 2009 [7] for minimum design load criteria. The reliability of using the code-specified
lateral load patterns for fixed-base building structures have been investigated during the past two
decades [8–13]. Leelataviwat et al. [10] evaluated the seismic demands of midrise moment-resisting
frames designed in accordance to UBC 94 [14]. They proposed improved load patterns using the
concept of energy balance applied to moment-resisting frames with a preselected yield mechanism.
Lee and Goel [11] also proposed new seismic lateral load patterns by using high-rise moment-resisting
frames up to 20-story with the same concept that Leelataviwat et al. [10] used. However, they used
SDOF response modification factor and structural ductility factors and dealt with a limited number of
ground motions. Their proposed load pattern fundamentally follows the shape of the lateral load
pattern in the code provisions (i.e., UBC 1994, 1997) and is a function of mass and the fundamental
period of the structure. In a more comprehensive research, Mohammadi et al. [12] investigated the
effect of lateral load patterns specified by the United States seismic codes on drift and ductility
demands of fixed-base shear building structures under 21 earthquake ground motions, and found that
using the code-specified design lateral load patterns do not lead to a uniform distribution and
minimum ductility demands. Ganjavi et al. [13] investigated the effect of equivalent static and spectral
dynamic lateral load patterns specified by the major seismic codes on height-wise distribution of drift,
hysteretic energy and damage subjected to severe earthquakes in fixed-base reinforced concrete
buildings. They concluded that in strong ground motions, none of the lateral loading patterns will lead
to uniform distribution of drift, hysteretic energy and damage, and an intense concentration of the
values of these parameters can be observed in one or two stories especially in equivalent static
method. More recently, several studies have been conducted by researchers to evaluate and improve
the code-specified design lateral load patterns [15–18]. However, all researches have been
concentrated on the different types of structures with rigid foundation, that is, without considering SSI
effects. SSI is one of the important factors that can significantly affect the seismic responses of
structures located on alluvium and soft soils by altering the overall stiffness and energy dissipation
mechanism of the systems. In fact, a soil–structure system behaves as a different system having longer
period and generally higher damping because of energy dissipation by hysteretic behavior and wave
radiation in the soil [19].
On the other hand, recently many studies have been conducted by researchers to investigate the
effects of SSI on elastic and inelastic response of SDOF and MDOF systems [20–26]. However, the
adequacy of code-compliant lateral loading patterns, which is fundamentally based on fixed-base
structures, have not been investigated for soil–structure systems in detail. For the first time, Ganjavi
and Hao [27] through performing intensive analyses of 5760 shear-buildings with SSI subjected to a
group of 30 earthquakes recorded on alluvium and soft soils investigated the adequacy of IBC-2009
code-complaint lateral loading patterns for elastic and inelastic soil–structure systems. They
concluded that using the code-specified load pattern leads to nearly uniform (optimal) ductility
demands distribution for structures having short periods and within the elastic range of response.
For structures with longer periods, however, it loses its efficiency as the number of stories and soil
flexibility increase especially for the cases of severe SSI effects. In the present study, the adequacy
of recently proposed lateral load patterns for fixed-base buildings, which will be explained in the
next section, is parametrically investigated for elastic soil–structure systems. Optimization algorithm
proposed by Hajirasouliha and Moghaddam [17] for fixed-base buildings based on uniform
distribution of deformation over the height of the structure is developed for soil–structure systems.
By performing intensive numerical simulations of responses of elastic shear buildings with various
dynamic characteristics and SSI parameters, the effects of fundamental period of vibration,
earthquake excitation, damping ratio, damping model, the number of stories, soil flexibility and
structure aspect ratio (slenderness ratio) on the optimal lateral load pattern of soil–structure systems
are investigated. On the basis of the results of this study, a new lateral load pattern for seismic
resistant design of elastic shear-building structures supported on shallow foundation considering SSI
effect is proposed. The new lateral load pattern is based on the results of time-history dynamic
analysis subjected to 21 earthquake ground motions consistent to the IBC-2009 design spectrum of
soil type E, and is a function of the period of the structure, dimensionless frequency (an index for
the structure-to-soil stiffness ratio) and structure aspect ratio.

Copyright © 2012 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2013; 42:913–933
DOI: 10.1002/eqe
OPTIMUM LATERAL LOAD PATTERN INCORPORATING SSI EFFECT 915

2. SUPERSTRUCTURE MODELING AND GROUND MOTIONS

Because of its simplicity and capability of considering higher modes effects, the well-known shear-
beam model is one of the most frequently used models that facilitate performing a comprehensive
parametric study [12, 16, 17, 26, 27]. In the MDOF shear-building models utilized in the present
study, each floor is assumed as a lumped mass to be connected by elasto-plastic springs. Story
heights are 3 m and total structural mass is considered as uniformly distributed along the height of
the structure. In all MDOF models, lateral story stiffness is assumed as proportional to story shear
strength distributed over the height of the structure, which is obtained in accordance to the different
lateral load patterns [12, 17]. Five percent Rayleigh damping is assigned to the first mode and the
mode in which the cumulative mass participation is at least 95%. By varying the damping ratio, the
effects of different structural damping ratios and damping models on optimum lateral load patterns
are studied.
In this investigation, an ensemble of 21 earthquake ground motions with different characteristics
recorded on alluvium and soft soil deposits are compiled. The selected ground motions are
components of six earthquakes including Imperial Valley 1979, Morgan Hill 1984, Superstition
Hills 1987, Loma Prieta 1989, Northridge 1994 and Kobe 1995. All the selected ground motions are
obtained from earthquakes with magnitude greater than 6 having closest distance to fault rupture
more than 15 km without pulse type characteristics. To be consistent, using SeismoMatch software
[28] these seismic ground motions are adjusted to the elastic design response spectrum of IBC-2009
with soil type E. SeismoMatch is an application capable of adjusting earthquake accelerograms to
match a specific target response spectrum using wavelets algorithm. The ground motions utilized in
the present study have the predominant period ranging from 0.5 to 1.35 s, recorded on sites with
shear wave velocity from 90 to 350 m/s, which are approximately consistent with the IBC-2009
elastic response spectrum of soil type E. Therefore, they were grouped and adjusted to match the
design response spectrum of soil type E corresponding to IBC-2009. Figure 1 shows a comparison
of the 21 matched ground motion spectra with the target elastic design response spectrum of
IBC-2009 [7].

3. SOIL–STRUCTURE MODEL AND KEY PARAMETERS

3.1. Soil–structure model


Substructure method is used to model the soil–structure system. Using the substructure method, the soil
can be modeled separately and then combined to establish the soil–structure system. The soil-shallow-
foundation element, in which the kinematic interaction is zero, is modeled by an equivalent linear
discrete model based on the cone model with earthquake frequency-independent coefficients and
equivalent linear model [29, 30]. However, to consider the material damping viscous soil
1.6

1.4 Individual synthetic earthquakes


Pseudo Acceleration (g)

IBC-2009 (Soil type E)


1.2

0.8

0.6

0.4

0.2

0
0 0.5 1 1.5 2 2.5 3
Period (sec)

Figure 1. IBC-2009 (ASCE/SEI 7-05) design spectrum for soil type E and response spectra of 21 adjusted
earthquakes (5% damping) for selected ground motions.

Copyright © 2012 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2013; 42:913–933
DOI: 10.1002/eqe
916 B. GANJAVI AND H. HAO

impedances, that is, stiffness and damping coefficients are dependent on the natural frequency of the
system (i.e., soil-structure system) through an iterative method. Cone model based on the one-
dimensional wave propagation theory represents circular rigid foundation with mass mf and area
moment of inertia If resting on a homogeneous half-space. In lieu of the rigorous elastodynamic
approach, the simplified cone model can be used with sufficient accuracy in engineering practice
[30]. Typical 10-story shear building models of fixed-base and flexible-base systems used in this
study are shown in Figure 2. The sway and rocking degrees of freedom are defined as
representatives of translational and rocking motions of the shallow foundation, respectively,
disregarding the slight effect of vertical and torsional motion. The stiffness and energy dissipation of
the supporting soil are represented by springs and dashpot, respectively. In addition, while being
hysteretic inherently, soil material damping is assumed as commonly used viscous damping so
that more intricacies in time-domain analysis are avoided. The coefficients of springs and dashpots
for sway and rocking used to define the soil-shallow foundation model in Figure 2 are summarized
as follows:

8rv2s r 8rv2s r3
kh ¼ ; ch ¼ rvs Af ; k’ ¼ ; c’ ¼ rvp If (1)
2υ 3ð1  υÞ

where kh, ch, k’ and c’ are sway stiffness, sway viscous damping, rocking stiffness, and rocking
viscous damping, respectively. Equivalent radius and area of cylindrical foundation are denoted by r
and Af. Besides, r, υ, vp and vs are respectively the specific mass density, Poisson’s ratio, and
dilatational and shear wave velocity of soil. To consider the soil material damping, z0, in the soil-
foundation element, each spring and dashpot is respectively augmented with an additional parallel
connected dashpot and mass. Also, to modify the effect of soil incompressibility, an additional
mass moment of inertia ΔM’ equal to 0.3p(υ  1/3)rr5 can be added to the foundation for υ greater
than 1/3 [30]. Influence of soil nonlinearity on the soil-foundation element may be approximated
through conventional equivalent linear approach in which a degraded shear wave velocity,
compatible with the estimated strain level in soil, is utilized for the soil medium [29]. This is
currently used in the modern seismic provision such as NEHRP 2003 [4] and FEMA 440 [31]

(a) (b)

Figure 2. Typical 10-story shear-building models: (a) fixed-base model and (b) flexible-base model.

Copyright © 2012 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2013; 42:913–933
DOI: 10.1002/eqe
OPTIMUM LATERAL LOAD PATTERN INCORPORATING SSI EFFECT 917

where the strain level in soil is implicitly related to the PGA. In the present study, by considering a
range of reasonable values for dimensionless frequency, soil nonlinearity is approximately modeled.

3.2. Key parameters


For a specific earthquake ground motion, the dynamic response of the structure can be interpreted
based on the properties of the superstructure relative to the soil beneath it. It has been shown that
the effect of these factors can be best described by the following dimensionless parameters [19, 32, 33].
• A dimensionless frequency as an index for the structure-to-soil stiffness ratio defined as

ofix H
a0 ¼ (2)
vs
where ofix is the natural frequency of the fixed-base structure. It can be shown that the practical
range of a0 for conventional building structures is from zero for the fixed-base structure to about
 which is the effective height of structure
3 for the case with severe SSI effect [19]. Besides, H,
corresponding to the fundamental mode properties of the MDOF building, can be obtained from
the following equation:
" !#
X
n X
j
mj ’j1 hi
 ¼
H
j¼1 i¼1
(3)
X
n
mj ’j1
j¼1

where mj is the mass of the jth story; hi is the height from the base level to level j; and ’j1 is the
amplitude at jth story of the first mode.
• 
Aspect ratio of the building defined as H=r, where r is the equivalent foundation radius.
• 
Structure-to-soil mass ratio defined as m¼m tot =rr H where H is total height of the structure.
2

• Foundation-to-structure mass ratio mf /mtot and Poisson’s ratio of the soil denoted by υ.
• Material damping ratios of the soil z0 and the structure zS.
The first two factors, affecting the responses more prominently are usually considered as the key
parameters that define the main SSI effect. The third one controls the inelastic behavior of the
structure. The other parameters, having less importance, may be set to some typical values for
conventional buildings [19, 27, 30]. In the present study, the foundation mass is assumed to be 0.1
of the total mass of the MDOF building. The Poisson’s ratio is considered to be 0.4 for the alluvium
soil and 0.45 for the soft soil. Also, a damping ratio of 5% is assigned to the soil material.

4. CURRENT LATERAL LOADING PATTERNS FOR FIXED-BASE BUILDINGS

4.1. Code-specified seismic design lateral load patterns


The general formula of the lateral load pattern specified by the aforementioned seismic codes is
defined as:

wx hkx
Fx ¼ X
n :V (4)
k
wi hi
i¼1

where Fx and V are respectively the lateral load at level x and the total design lateral force; wi and wx
are the portion of the total gravity load of the structure located at the level i or x; hi and hx are the
height from the base to the level i or x; n is the number of stories; and k is an exponent that differs

Copyright © 2012 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2013; 42:913–933
DOI: 10.1002/eqe
918 B. GANJAVI AND H. HAO

from one seismic code to another. In IBC-2009 [7], k is related to the fundamental period of the
structure, which is equal to 1 or 2 for structures having a period of 0.5 s or less, and for structures
having a period of 2.5 s or more, respectively. For structures having a period between 0.5 and
2.5 s, k is computed by linear interpolation between 1 and 2. It should be mentioned that, the
distribution of lateral force based on IBC 2009 is identical to that of the NEHRP 2003 [4] and
ASCE/SEI 7–05 [5] provisions. Note that when k is equal to 1, the pattern corresponds to an
inverted triangular lateral load distribution and the response of building, thus, is assumed to be
controlled primarily by the first mode. While k equal to 2 corresponds to a parabolic lateral load
pattern with its vertex at the base in which the response is assumed to be influenced by higher
mode effects.

4.2. Lateral load pattern proposed by park and medina


Park and Medina [15], on the basis of the nonlinear dynamic analyses on fixed-base regular moment-
resisting frames subjected to 40 far-field ground motions from California earthquakes recorded on stiff-
soil sites, proposed a new lateral load pattern that is consistent, in format, with the lateral load
distributions of IBC 2009 and UBC 97 [3]. They concluded that the proposed approach provides, on
average, a more uniform distribution of story ductility ratios and story drift ratios, when compared
with the distributions obtained using current seismic code provisions, that is, the 2006 IBC which is
the same as the IBC 2009 [7]. Their proposed lateral load pattern is given by the following expression:

wx hkx  
Fx ¼ X
n : V  Ftop (5)
wi hki
i¼1

where Ftop and k are respectively consistent, in definition, with Ft and exponent k in UBC-97 and IBC
2009, which are defined as

Ftop
¼ 0:32  0:001H  0:13k; 22⩽H⩽66 and k ¼ 0:56  0:17mt ; 1⩽mt ⩽5 (6)
V

In Equation (6), Ftop is an additional force applied to the top (roof) story to incorporate higher mode
effects, and mt is the target global ductility demand of the structure.

4.3. Lateral load pattern proposed by Hajirasouliha and Moghaddam


On the basis of the research carried out by Mohammadi et al. [12], Hajirasouliha and Moghaddam [17]
with the nonlinear dynamic analyses on fixed-base shear building models subjected to 20 earthquake
ground motions recorded on alluvium soil, a new lateral load pattern as a function of the
fundamental period of the structure and target ductility was proposed, which is defined as

Fi ¼ ðai T þ bi Þmt ðci Tþdi Þ (7)

where Fi is the optimum load component at the ith story; T is the fundamental period of the structure; mt
is the target global ductility demand; and ai, bi, ci and di are constant coefficients at the ith story. These
coefficients can be obtained at each level of the structure by interpolating the values given in their
paper. As reviewed above, it should be noted that none of the above load patterns explicitly
considered the influences of SSI. In the present study, the adequacy of all the above-mentioned
lateral load patterns for the elastic response of soil–structure systems is investigated and discussed.

Copyright © 2012 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2013; 42:913–933
DOI: 10.1002/eqe
OPTIMUM LATERAL LOAD PATTERN INCORPORATING SSI EFFECT 919

5. OPTIMUM DISTRIBUTION OF ELASTIC DESIGN LATERAL FORCE FOR


SOIL–STRUCTURE SYSTEMS

5.1. Optimization procedure


As stated before, using code-compliant lateral load patterns does not lead to the optimum seismic
performance of structures. A recent study by Ganjavi and Hao [27] indicates that using code-
specified load pattern for soil–structure systems with severe SSI effect and high inelastic response
does not lead to uniform (optimal) ductility demand distribution over the height of structures. This
means that the deformation (ductility) demand in some stories of the building does not reach the
presumed target level of seismic capacity, implying that the structural material has not been entirely
exploited over the height of the building. This study deals with the development of the optimization
technique to distribute predefined structural damage in elastic response along the height of the
structure. In this regard, the optimization technique adopted by Mohammadi et al. [12] and
Hajirasouliha and Moghaddam [17] for fixed-base shear-building structures is used to develop the
optimal load pattern for elastic soil–structure systems. In this approach, the structural properties are
modified so that inefficient material is gradually shifted from strong to weak parts of the structure.
This process is continued until a state of uniform deformation is achieved [17]. In the present study,
the seismic demand parameter used to quantify the structural damage is the interstory displacement
ductility ratio (m ). It should be mentioned that although for the elastic response the ductility ratio is
limited to the value equal to one, this parameter can still be used as a damage index representing the
level of deformation with respect to the predefined target value. Considering the theory of uniform
deformation proposed by Mohammadi et al. [12] and Hajirasouliha and Moghaddam [17] for fixed-
base shear-building structures, the following step-by-step optimization algorithm is adopted and
proposed for elastic shear-buildings with SSI effects.
(1) Define an MDOF shear-building model depending on the prototype structure height and number
of stories.
(2) Assign an arbitrary value for total stiffness and strength and then distribute them along the
height of the structure based on the arbitrary lateral load pattern, for example, uniform pattern.
As mentioned earlier, the lateral story stiffness is assumed as proportional to the story shear
strength distributed over the height of the structure.
(3) Select an earthquake ground motion.
 and dimensionless frequency, a0, as the predefined
(4) Consider a presumed set of aspect ratio, H=r,
key parameters for SSI effects.
(5) Select the fundamental period of fixed-base structure and scale the total stiffness without altering
the stiffness distribution pattern such that the structure has a specified target fundamental period.
The following equation is used for scaling the stiffness to reach the target period by just one step:
  !
X
Þ Ti 2 X
n n
Kjþ1 ¼ : Kj (8)
i
Ttarget i¼1
iþ1 i

where Kj, Ti and Ttarget are story stiffness in the jth story, fixed-base period in the ith step and the target
fundamental fixed-base period, respectively. Refine H=r  based on the fundamental modal properties of
fixed-base MDOF structure (Equation (3)).
(6) Perform dynamic analysis for the soil–structure system subjected to the selected ground motion
and compute the total shear strength demand, (Vs)i. If the computed ductility ratio is equal to the
target value within the 0.5% of the accuracy, no iteration is necessary. Otherwise, total base
shear strength is scaled (by either increasing or decreasing) until the target ductility ratio is
achieved. To do this the following equation is proposed:

ðVs Þiþ1 ¼ ðVs Þi ðm max Þb (9)

where (Vs)i is the total base shear strength of MDOF system at ith iteration and mmax is the maximum
story ductility ratio among all stories. Parameter b is an iteration power, which is more than zero.

Copyright © 2012 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2013; 42:913–933
DOI: 10.1002/eqe
920 B. GANJAVI AND H. HAO

Results of this study indicate that b power for mt ⩽ 1(elastic state) can be taken as a constant value of
0.8 for all fixed-base and flexible base shear-building structures when subjected to any earthquake
excitation.
(7) Calculate the coefficient of variation (COV) of story ductility distribution along the height of the
structure and compare it with the target value of interest, which is set as 0.02 here. If the value of
COV is less than the presumed target value, the current pattern is regarded as optimum pattern.
Otherwise, the story shear strength and stiffness patterns are scaled until the COV decreases
below or equal to the target value.
(8) Stories in which the ductility demand is less than the presumed target values (i.e.mt = 1) are
identified and their shear strength and stiffness are reduced. To obtain the fast convergence in
numerical computations, the equation proposed by Hajirasouliha and Moghaddam [17] is
revised for elastic soil–structure systems as follows:

½Si qþ1 ¼ ½Si q :½mi a (10)

where [Si]q= shear strength of the ith floor at qth iteration, mi=story ductility ratio of the ith floor and
a = convergence parameter that has been considered equal to 0.1–0.2 as the acceptable range by
Hajirasouliha and Moghaddam [17] for elastic and inelastic fixed-base structures. The results of this
study show that for elastic fixed-base and soil structure systems a can be taken from 0.5 to 1 as the
best range. However, in most cases the value of 0.8 leads to the fastest convergence (i.e. less than 5
iterations). The effect of the convergence parameter on optimum elastic design of the shear-building
structures will be investigated in the next part. Now, a new pattern for lateral strength and stiffness
distributions is obtained.
(9) Control the current maximum story ductility ratio (mmax) and refine the total base shear strength
of soil–structure systems if mmax is not equal to the target value within the 0.5% of the accuracy
based on Equation (9) of step 6. Otherwise, go to the next step.
(10) Control the current fixed-base period and modify it if it is not equal to the target value within the
1% of the accuracy based on Equation (8) of step 5. Otherwise, go to the next step
 and refine it if the value is not equal to the previous value
(11) Control the current effective height (H)
within the 1% tolerance based on Equation (3). Otherwise, go to the next step
(12) Control the current Rayleigh-type damping coefficients and modify them if they are not equal to
the previous values within the 1% tolerance. Otherwise, go to the next step
(13) Convert the optimum shear strength pattern to the optimum lateral force pattern.
(14) Repeat steps 5–14 for different presumed target periods.
 and a0.
(15) Repeat steps 4–15 for different sets of H=r
(16) Repeat steps 3–16 for different earthquake ground motions.
(17) Repeat steps 1–17 for different number of stories.
To show the efficiency of the proposed method for optimum seismic design of soil–structure
systems in elastic range of response, the above algorithm is applied to a 10-story shear building with
 = 3, and a0 = 2 subjected to Kobe (Shin Osaka) simulated earthquake. Figure 3(a)
Tfix= 1.5 s, H=r
illustrates a comparison of IBC-2009 [7] load pattern with the optimum patterns of fixed-base and
soil–structure systems. As can be seen, there is a significant difference between the optimum pattern
of soil–structure systems and the other two patterns. These three patterns are applied to the same
10-story building with consideration of soil flexibility and then the height-wise distribution of story
ductility demand resulted from utilizing these lateral load patterns are computed and depicted in
Figure 3(b). In this figure beside the results of the aforementioned load pattern, the height-wise
distribution of story ductility demand resulted from utilizing modal response spectrum analysis with
consideration of SSI effect are also plotted for comparison. Modal response spectrum analysis is a
general procedure for linear analysis of the dynamic response of structures, which are allowed by
most current seismic codes particularly for tall buildings. In this case, the results of one and three
modes are shown. It can be seen that although using the SSI optimum pattern results in a completely
uniform distribution of the deformation, using other load patterns lead to a rather nonuniform
distribution of ductility demand along the height of the soil–structure systems in elastic range of

Copyright © 2012 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2013; 42:913–933
DOI: 10.1002/eqe
OPTIMUM LATERAL LOAD PATTERN INCORPORATING SSI EFFECT 921

10 10
9 (a) 9
Modal, SSI (1 mode) (b)
Modal, SSI (3 modes)
8 8 OPT Fixed-base
7 7 OPT SSI

Story
6 6 IBC
5 5
IBC
4 4
OPT SSI
3 3
2 OPT Fixed-base 2
1 1
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0 0.2 0.4 0.6 0.8 1 1.2
Lateral Force / Base Shear Ductility

Figure 3. Comparison of IBC-2009 with optimum designed models of fixed-base and soil–structure system:
 =3, Kobe
(a) lateral force distribution; (b) story ductility pattern, 10-story shear building with Tfix= 1.5 s, H=r
(Shin Osaka) simulated earthquake.

vibration. Moreover, for the case of modal analysis it is seen that considering only the first mode with
SSI effect does not have any significant preference over the code-compliant pattern. However,
considering three modes may lead to a more uniform pattern with respect to other patterns (i.e., code-
specified and optimum fixed-base patterns) but still far from the optimum one. It is noteworthy to
mention that although modal response spectrum analysis are allowed by most of the current seismic
codes, the preliminary design of most common buildings in force-base method is still based on
equivalent lateral force method specified by the current seismic codes. In fact, it can be much easier
for most of the structural engineers, even with low level of knowledge in earthquake engineering, to
use the equivalent lateral force method rather than modal analysis or time history analysis methods.
Hence, this study focuses on modifying the equivalent lateral force method to incorporate SSI effects
in elastic range of response. The COV of story ductility demand distributions resulted from applying
IBC pattern, the fixed-base optimum pattern, modal analysis with one mode, modal analysis with
three modes and SSI optimum pattern are 0.226, 0.196, 0.18, 0.104 and 0.003, respectively. This
indicates that SSI phenomenon through changing the dynamic characteristics of structures can
significantly affect damage distribution along the height of structures. Therefore, utilizing fixed-base
optimum load pattern may not result in an optimum seismic performance of soil–structure systems
and, thus, a more adequate load pattern accounting for SSI effects should be defined and proposed for
soil structure systems. This will be discussed more in the next sections.
In another point of view, it may be of interest to compare the required structural weight for buildings
that have been designed for different seismic load patterns. The efficiency of the selected load patterns,
then, can be further evaluated accordingly. In this regard, it is assumed that the weight of the lateral
load-resisting system at each story, WEi, is proportional to the story shear strength, Si [12].
Therefore, the total weight of the seismic resistant system, WE, can be calculated as

X
n X
n X
n
WE ¼ WEi ¼ l:Si ¼ l: Si ; (11)
i¼1 i¼1 i¼1

where l is the proportioning coefficient. According to Equation (11), normalized weight index (WI) of
the MDOF structure related to the presumed lateral load pattern can be defined as:

WI ¼ WE =W:PGA (12)

where W and PGA are total weight of the structure and peak ground acceleration, respectively. The
loading pattern that corresponds to the minimum weight index is considered as the most adequate
loading pattern. As discussed in Section 2, in the MDOF shear-building models utilized in the
present study, each floor is assumed as a lumped mass to be connected by elasto-plastic springs.
Therefore, in the typical shear-building model, mass and the lateral stiffness of each story can be
modeled by a lumped mass and horizontal spring representing lateral strength and stiffness. Thus,
through Equations (11) and (12) the structural weight in shear building can be directly related to the

Copyright © 2012 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2013; 42:913–933
DOI: 10.1002/eqe
922 B. GANJAVI AND H. HAO

lateral shear strength. However, this point is not correct for other structural systems such as moment-
resisting system and moment-resisting systems with shear wall. In fact, the increase in the weight of the
earthquake-resistant structural elements of these systems (because of beam and column or slab) is not
directly related to the total strength demands, and thus reduction of the weight index is usually
proportional to a percentage of the strength demands. This method has been widely used by
researchers to compare the efficiency of the different lateral load patterns, see Refs. [12, 16, 17]. As
an instance, the weight indexes corresponding to the aforementioned load patterns considered in the
previous example are calculated here to demonstrate the efficiency of the proposed algorithm. The
weight indexes computed from applying IBC-2009 pattern, the fixed-base optimum pattern, modal
analysis with one mode, modal analysis with three modes and SSI optimum pattern are respectively
8.68, 7.74, 7.36, 6.63 and 5.25. Therefore, the required structural weight value corresponding to the
proposed SSI optimum load pattern is respectively 39.5%, 32.2%, 31.1% and 20.8% less than that
of the IBC-2009, the fixed-base optimum patterns, modal analysis with one mode and modal
analysis with three modes, which means that utilizing the proposed algorithm can remarkably reduce
the required structural weight in elastic range of response. Therefore, to improve the seismic
performance of the structure under this specific earthquake the frame should be designed in
accordance to an equivalent lateral load pattern that is different from the conventional code-specified
and also the fixed-base optimum patterns.

5.2. Effect of convergence parameter


To examine the effect of power, a, defined in Equation (10) on convergence; the previous example again
is solved for different values of a. Figure 4 shows the variation of structural weight index corresponding
to different values of convergence powers of 0.1, 0.2, 0.5, 0.8, 1 and 1.5. It shows how the structural
weight index varies with the iteration step from a presumed initial load pattern, here the uniform load
pattern, toward the optimum pattern. Different from the finding by Hajirasouliha and Moghaddam [17]
who have proposed the best values of 0.1–0.2 for all ranges of ductility demands, that is, both elastic
and inelastic response, results of this study indicate that for elastic response, regardless of either being
fixed-base or flexible-base model, the values of 0.5–1 could be the best ranges for a. In addition, in
elastic response there is no fluctuation in convergence problem for the power of a ranging from 0.01
to 1 while the fluctuation happens for a greater than one as seen for the case of a = 1.5 in Figure 4.
On the basis of intensive analyses performed for both fixed-base and soil–structure systems in elastic
range of response, It can be concluded that, on average, the value of 0.8 could be a good value for a
to achieve the fastest convergence. As seen, using a = 0.8 the required numbers of iterations to reach
the optimum design are only 2 and 3 steps for respectively the fixed-base and soil–structure system.
However, it will be 32 and 16 steps in fixed-base systems and 32 and 19 steps in soil–structure
systems for a = 0.1 and 0.2, respectively. It is also interesting to note that after only one iteration, the
value of weight index reduces to less than 50% of its initial value, that is, from 14 to 6.66.

18 15
17 (a) Fixed base α= 0.1 α= 0.2 14 (b) SSI, 0= 2 α = 0.1 α = 0.2
16 α= 0.5 α= 0.8 13
α= 0.5 α = 0.8
15 12
14 α= 1 α= 1.5 11 α= 1 α = 1.5
13 10
WI

12 9
11 8
10 7
9 6
8 5
7 4
6 3
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30 32 0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30 32
Step Step

Figure 4. Variation of structural weight index for different values of convergence powers; 10-story
 = 3,a0=2, Kobe (Shin Osaka) simulated earthquake.
soil–structure system with Tfix= 1.5 s, H=r

Copyright © 2012 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2013; 42:913–933
DOI: 10.1002/eqe
OPTIMUM LATERAL LOAD PATTERN INCORPORATING SSI EFFECT 923

5.3. Effect of initial load pattern


Considering the proposed optimization algorithm for soil–structure systems, an initial strength and
stiffness distribution is required to reach the optimum answer. Moghaddam and Hajirasouliha [16]
concluded that for inelastic fixed-base shear buildings, the optimum lateral force pattern is not
dependent on the initial strength pattern, and to some extent it would affect the speed of convergence.
This point is investigated here for the case of elastic soil–structure system of the previous example by
considering the same initial patterns utilized by Moghaddam and Hajirasouliha [16], and the results are
depicted in Figures 5 and 6(a). As seen, the same results can be concluded for the shape of optimum
pattern for elastic soil–structure systems (see Figure 6(a)). However, the results of this study indicate
that using the proposed optimization algorithm for both elastic fixed-base and soil–structure systems the
initial strength pattern has no effect on the convergence speed. For example, using all assumed patterns
including rectangular, triangular, IBC-2009 and concentric patterns requires only three steps to converge
to the same optimum load pattern as shown in Figure 5 for both fixed-base and soil–structure systems.

6. EFFECT OF STRUCTURAL DYNAMIC CHARACTERISTICS AND SSI KEY PARAMETERS


ON OPTIMUM LATERAL FORCE PATTERN

Before proposing a general load pattern for optimum design of soil–structure systems in elastic range
of response, it is necessary to first investigate the effects of various parameters including structural
dynamic characteristics and SSI key parameters.

6.1. Effect of earthquake excitation


Effect of ground motion intensity on the optimum load profile of the elastic 10-story soil–structure

model with Tfix= 1.5, H=r= 3, and a0 = 2 subjected to Kobe (Shin Osaka) simulated earthquake with
the PGA multiplied by 0.5, 1, 2 and 3 factors are illustrated in Figure 6(b). The results indicate that
for a specific fundamental period, aspect ratio and dimensionless frequency, the optimum lateral
load pattern is completely independent of the ground motion intensity factor (SF).
To investigate the effect of varying earthquake ground motions on optimum lateral force pattern for
both fixed-base and flexible-base shear-building models in elastic range of response, individual results
of all 21 matched earthquake ground motions along with their average values for a 10-story shear
building are presented in Figure 7. The results are for systems with Tfix= 1.5 s for fixed-base system
(Figure 7(a)) and for soil–structure system with aspect ratio of 3, and dimensionless frequency of 2
(Figure 7(b)). As can be seen, it is obvious that the optimum lateral load pattern is sensitive to the
earthquake ground motion characteristics for both the fixed-base and especially the flexible-base
models. However, in most ground motions used in this study, there is not a big discrepancy in the

18 15
17 (a) Fixed base Rectangular 14 (b) SSI, 0= 2 Rectangular
16 13 Triangular
Triangular
15 12
IBC-2009 IBC-2009
14 11
Concentric Concentric
13 10
WI

12 9
11 8
10 7
9 6
8 5
7 4
6 3
0 1 2 3 4 0 1 2 3 4
Step Step

Figure 5. Effect of initial load pattern on optimization iteration steps; 10-story shear building. (a) Fixed-base
system and (b) soil–structure system with Tfix= 1.5 s, H=r  = 3, a0 = 2, Kobe (Shin Osaka) simulated
earthquake.

Copyright © 2012 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2013; 42:913–933
DOI: 10.1002/eqe
924 B. GANJAVI AND H. HAO

10 10
9 (a) 9 (b)
8 8
7 Concentric 7 SF= 0.5
6 6
Story
Triangular SF=1
5 5
Inverted Triangular SF=2
4 4
3 Rectangular 3 SF=3
2 2
1 1
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
Lateral Force / Base Shear Lateral Force / Base Shear

Figure 6. Effect of (a) initial load pattern and (b) ground motion intensity on optimum lateral force profile for

soil–structure systems with Tfix= 1.5 s, H=r=3 and a0= 2; Kobe (Shin Osaka) simulated earthquake.

Individual Average
10 10
9 9
8 8
7 7
6 6
Story

5 5
4 4
3 3
2 (a) 2 (b)
1 1
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 -0.05 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
Lateral Force / Base Shear Lateral Force / Base Shear

Figure 7. Optimum lateral force distribution for different earthquake excitations, 10-story building with

Tfix= 1.5 s: (a) fixed-base model; (b) soil–structure model with H=r=3 and a0= 2.

general pattern of optimum lateral load profile when compared with the corresponding averaged
pattern. Therefore, it is expected that utilizing the mean pattern will lead to acceptable designs
although some inevitable variation is not avoidable depending on the earthquake ground motion.
This will be demonstrated in the next part. It is also worth mentioning that, in general, the soil–
structure systems are more sensitive to the seismic excitation than fixed-base systems. As seen in
Figure 7(b), in some ground motions the sign of the lateral force corresponding to the SSI pattern is
negative for one or two stories. This phenomenon could be due to the effect of higher mode
contributions in soil–structure systems with severe SSI effects that are much more flexible than the
corresponding fixed-base building.

6.2. Effect of fundamental period


To study the effect of fundamental period on optimum lateral load pattern of elastic soil–structure
systems, the 15-story building models with H=r  = 3 and a0 = 2 having fixed-base fundamental
periods of 0.3, 0.6, 1, 2 and 3 s are considered. For each case, the optimum load patterns are derived
for the 21 matched earthquake ground motions and the average results are plotted in Figure 8(a). As
seen, the averaged optimum load pattern is significantly dependent on the fundamental period of
vibration such that increasing the fundamental period is generally accompanied by increasing the
lateral shear force at top stories, which can be interpreted as the effect of higher modes.

Copyright © 2012 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2013; 42:913–933
DOI: 10.1002/eqe
OPTIMUM LATERAL LOAD PATTERN INCORPORATING SSI EFFECT 925

15 1.0
14 (a) 0.9 (b) H/r= 3
13
0.8 N= 5
12
11 0.7

Relative Height
N= 7
10 T= 0.3 0.6
9 N= 10
Story

8 T= 0.6 0.5 N= 15
7 0.4
6
T= 1 N= 20
5 0.3
T= 2
4 0.2
3 T= 3 0.1
2
1 0.0
0 0.05 0.1 0.15 0.2 0.25 0.3 0 0.05 0.1 0.15 0.2
Lateral Force / Base Shear Normalized lateral Force / Base Shear

Figure 8. Effect of fundamental period (a) and the number of stories (b) on averaged optimum lateral force

profile for soil–structure systems with H=r=3 and a0= 2: Tfix= 1.5 s.

6.5. Effect of number of stories


To examine the effect of number of stories on the optimum distribution profile, the proposed
optimization algorithm is applied to 5-story, 7-story, 10-story, 15-story and 20-story soil–structure

models with Tfix= 1.5, H=r= 3 and a0 = 2 subjected to the 21 matched earthquake ground motions.
The average results are depicted in Figure 8(b). To compare the averaged optimum patterns
corresponding to different number of stories, the normalized lateral loads are plotted. In Figure 8(b),
the vertical and horizontal axes are relative height and normalized lateral load divided by base shear
strength, respectively. From this figure, it can be concluded that the optimum load patterns are
almost independent of the number of stories. This finding is also consistent with that by
Hajirasouliha and Moghaddam [17] for fixed-base shear-building structures.

6.6. Effect of dimensionless frequency


Figure 9 shows the effect of dimensionless frequency, a0 on averaged optimum load pattern of elastic
soil–structure systems subjected to 21 matched ground motions. As stated before, aspect ratio and
dimensionless frequency are two key parameters that can affect the response of the soil–structure
systems subjected to earthquake excitation. The results are plotted for the 10-story shear building
with two fundamental periods of 1 and 2 s, and H=r  =3 corresponding to three values of
dimensionless frequency (a0=1, 2, 3). It can be observed that dimensionless frequency can

Fixed-base 0 =1 0 =2 0 =3
1.0 1.0
(a) 0.9
(b)
0.9
0.8 0.8
Relative Height

0.7 0.7
0.6 0.6
0.5 0.5
0.4 0.4
0.3 0.3
0.2 0.2
0.1 0.1
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 -0.05 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
Lateral Force / Base Shear Lateral Force / Base Shear

Figure 9. Effect of dimensionless frequency on averaged optimum lateral force profile for 10-story soil–

structure systems with H=r=3: (a) Tfix= 1 s and (b) Tfix= 2 s.

Copyright © 2012 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2013; 42:913–933
DOI: 10.1002/eqe
926 B. GANJAVI AND H. HAO

significantly affect the averaged optimum load pattern such that increasing the value of dimensionless
frequency is accompanied by increasing the lateral load at bottom and top stories, and decreasing the
load in middle stories. This phenomenon could be again because of the effect of higher mode effect as a
result of increasing the fundamental period of the soil–structure systems.

6.7. Effect of aspect ratio


Figure 10 shows the effect of aspect ratio on averaged optimum load pattern of elastic soil–structure
systems. The results are for the 10-story shear building with Tfix=1.5 s, two dimensionless
frequencies (a0=1, 3), representing the insignificant and severe SSI effect, respectively and three

values of aspect ratio ( H=r= 1, 3, 5) representing respectively squat, average and slender buildings
subjected to 21 matched ground motions. As seen, for the case of insignificant SSI effect
(i.e. Figure 10(a)), increasing aspect ratio will not change the optimum load profile remarkably.
However, by increasing the dimensionless frequency and, therefore more significant SSI effect, the
aspect ratio will significantly affect the averaged optimum load pattern. The trend is to some extent
similar to that of the dimensionless frequency discussed in the previous section such that increasing
the value of aspect ratio is accompanied by increasing the lateral load at the bottom and especially
top stories, and decreasing the load in the middle stories, which is more pronounced for slender

buildings (i.e. H=r= 5).

6.8. Effect of structural damping ratio


The effect of structural damping ratio on optimum load pattern of elastic soil–structure systems is

illustrated in Figure 11(a) for a 10-story shear-building structure with Tfix= 1.5, H=r= 3 and a0 = 2
corresponding to four values of 0%, 2%, 5% and 10% of damping ratios subjected to matched Loma
Prieta earthquake (APEEL 2 - Redwood City). As can be seen, earthquake loads associated to the
top stories reduces by increasing the damping ratio which, in turn, reduces higher mode effects. The
phenomenon is more pronounced for the case of damping ratio of 10%. Therefore, one may
conclude that for the practical purpose, the optimum load pattern of elastic soil–structure systems
can be considered insensitive to the variation of damping ratio. The results are consistent with those
concluded for fixed-base systems by Hajirasouliha and Moghaddam [17].

6.9. Effect of structural damping model


To examine the effect of damping models on optimum load pattern of elastic soil–structure systems,
three conventional viscous damping models including stiffness-proportional damping, mass-
proportional damping and Rayleigh-type damping in which damping matrix is composed of the
superposition of a mass-proportional damping term and a stiffness-proportional damping term are
considered. In this case, optimum lateral load patterns of the 10-story soil–structure systems with

1.0 1.0
0.9 (a) 0= 1 0.9 (b) 0 =3
0.8 0.8
0.7 0.7
Relative Height

0.6 H/r= 1 0.6 H/r= 1


0.5 0.5
H/r= 3 H/r= 3
0.4 0.4
0.3 H/r = 5 0.3 H/r= 5
0.2 0.2
0.1 0.1
0.0 0.0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 -0.05 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
Lateral Force / Base Shear Lateral Force / Base Shear

Figure 10. Effect of aspect ratio on averaged optimum lateral force profile for a 10-story soil–structure
system with Tfix= 1.5 s.

Copyright © 2012 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2013; 42:913–933
DOI: 10.1002/eqe
OPTIMUM LATERAL LOAD PATTERN INCORPORATING SSI EFFECT 927

1.0 1.0
0.9 (a) 0.9 (b)
0.8 0.8

Relative Height
0.7 0.7
0.6 0.6
ξ = 0%
0.5 0.5
ξ = 2% Mass
0.4 0.4
ξ = 5% Rayleigh
0.3 0.3
Stiffness
0.2 ξ = 10% 0.2
0.1 0.1
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
Lateral Force / Base Shear Lateral Force / Base Shear

Figure 11. Effect of structural damping ratio (a) and damping model (b) on optimum lateral force profile;

10-story soil–structure system with H=r=3, a0= 2 and Tfix= 1.5 s; Loma Prieta (APEEL 2 - Redwood City)
earthquake.


Tfix= 1.5, H=r= 3 and a0 = 2 corresponding to the aforementioned damping models subjected to Loma
Prieta earthquake (APEEL 2 - Redwood City) are obtained and plotted in Figure 11(b). It can be seen
that while there is no significant difference between the results of mass-proportional and Rayleigh-type
damping models; the difference is pronounced when compared with that of the stiffness-proportional
damping model. The same result has been found for other fixed-base and soil–structure models
subjected to different seismic ground motions. This observation indicates that stiffness-proportional
damping model may lead to quite different predictions of structural responses as compared with the
Rayleigh damping model. It is, therefore, advisable to use Rayleigh-type damping model to better
incorporate the effect of higher modes.

7. NEW LATERAL LOAD PATTERN FOR ELASTIC SOIL–STRUCTURE SYSTEMS

To generalize the use of the proposed optimization algorithm for conceptual seismic design of elastic
soil–structure systems, it is necessary to develop statistical models for estimating the optimum design
lateral load pattern as a function of relevant structural and soil characteristics. Because of the variability
in ground motion characteristics, it is not straightforward to determine an equivalent lateral load pattern
to provide, on average, a uniform distribution of deformation along the height of the soil–structure
system when the system is subjected to earthquake excitations with different frequency contents.
Generally, it is believed that for design purposes, the design earthquake ground motion should be
classified for each structural performance and soil type category. More reliable load pattern, then,
can be obtained by computing the mean values of optimum patterns associated to the design
earthquakes compatible with each seismic design response spectrum. To be consistent, 21 seismic
ground motions compatible to the elastic design response spectrum of IBC-2009 [7] with soil type E
are selected. Numerous soil–structure systems including four different number of stories (i.e. 5-story,
10-story, 15-story and 20-story buildings), 30 fundamental periods ranging from 0.1 to 3 s with
 =1, 3, 5) representing respectively squat, average
intervals of 0.1, three values of aspect ratio ( H=r
and slender buildings, and four values of dimensionless frequency (a0=0, 1, 2, 3) are considered. It
should be noted that a0=0 and 3 correspond to the systems with fixed base and severe SSI effect,
respectively. Consequently, utilizing the proposed optimization algorithm, 30,240 optimum lateral
load patterns are derived for elastic soil–structure systems. For each fundamental period,
dimensionless frequency and aspect ratio, the mean optimum load pattern corresponding to 21
matched earthquake ground motions are obtained. It is expected that designs based on the mean
patterns would exhibit a more uniform damage along the height of soil–structure systems. On the
basis of the results of this study and nonlinear statistical regression analysis, the following
expression is proposed for optimum design of elastic shear-building considering SSI effect:

Copyright © 2012 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2013; 42:913–933
DOI: 10.1002/eqe
928 B. GANJAVI AND H. HAO

Fi ¼ ai þ bi Tfix 0:5 :LnðTfix Þ þ ci eð0:5Tfix Þ (13)


where Fi= optimum load component at the ith story; Tfix= fixed-base fundamental period; and ai, bi and
ci= constant coefficients of ith story, which are functions of aspect ratio and dimensionless frequency
and are given in Tables I to III for each level of structure. It should be noted that Fi is the optimum
lateral load component that must be scaled to the total load components at the end of calculation. In
addition, the optimum load patterns corresponding to values of the relative height, aspect ratio and
dimensionless frequency associated with the specified values in Tables I to III can be easily obtained
by linear interpolation of the two associated load patterns.
The efficiency of the proposed load pattern is investigated by comparing the structural weight
indexes resulted from using Equation (13) with code-compliant and recently proposed load patterns
for fixed-base structures. Accordingly, two aforementioned lateral load patterns by Park and Medina
[15] and Hajirasouliha and Moghaddam [17] and IBC-2009 [7] lateral load pattern are considered
here and average results of these patterns are parametrically compared with results of this study
(i.e. Equation (13)). For this purpose, the values of normalized WI of the 10-story shear buildings
designed based on the four above-load patterns for 30 fundamental periods ranging from 0.1 to 3 s,

three values of aspect ratio ( H=r=1, 3, 5) and two values of dimensionless frequency (a0=1, 3) are
calculated subjected to 21 matched earthquake ground motions. Then, the averaged values of
weight index for all the above-load patterns are computed and illustrated in Figure 12. On the basis
of the results presented in Figure 12, it can be observed that:
(1) For short periods of vibration, the structures designed in accordance with the optimum load
pattern proposed by Park and Medina [15] has the worst performance among all the load

Table I. Constant coefficient ai of Equation (13) as function of relative height.


 =1
H=r  =3
H=r  =5
H=r
Relative height a0=0 a0=1 a0=2 a0=3 a0=1 a0=2 a0=3 a0=1 a0=2 a0=3

0.05 29.25 17.74 11.83 19.92 31.52 38.37 15.05 43.98 34.60 50.53
0.10 21.00 13.75 20.06 15.80 32.31 23.73 18.58 48.27 22.64 34.09
0.20 9.50 0.67 10.22 22.26 29.33 39.63 3.73 38.72 20.95 19.01
0.30 11.21 19.42 39.53 51.49 0.60 2.62 13.99 0.34 1.97 2.87
0.40 29.20 41.46 60.68 25.50 60.89 30.51 15.25 57.49 19.99 34.07
0.50 45.45 50.00 81.38 34.53 81.51 84.61 11.72 100.88 42.95 20.39
0.60 64.20 61.09 61.49 56.48 76.02 80.59 36.88 90.28 52.20 12.30
0.70 66.77 51.69 54.20 57.88 66.84 69.10 45.41 69.24 36.57 4.41
0.80 87.47 104.28 78.30 86.03 96.71 80.04 119.67 81.91 82.42 87.42
0.90 118.23 105.59 88.62 103.89 97.41 122.26 164.23 100.49 145.46 182.15
1.00 156.27 117.54 95.89 123.76 130.43 154.95 238.68 130.73 239.11 328.05

Table II. Constant coefficient bi of Equation (13) as function of relative height.


 =1
H=r  =3
H=r  =5
H=r

Relative height a0=0 a0=1 a0=2 a0=3 a0=1 a0=2 a0=3 a0=1 a0=2 a0=3
0.05 14.42 11.66 11.52 19.81 18.84 33.19 37.12 26.00 36.68 18.20
0.10 14.02 12.37 17.01 17.92 22.67 28.86 39.29 32.41 36.39 26.58
0.20 14.80 8.88 13.76 0.26 25.62 38.22 28.60 33.74 41.52 32.80
0.30 8.16 1.51 11.24 14.55 12.23 13.12 21.69 14.90 27.10 28.50
0.40 0.33 7.76 19.14 5.54 18.35 8.30 5.19 17.38 5.74 7.84
0.50 8.42 9.90 29.40 17.97 27.54 45.31 23.89 42.70 37.57 27.24
0.60 13.88 12.15 21.60 25.45 23.29 42.44 32.83 36.88 39.70 19.17
0.70 9.14 2.93 8.54 13.58 13.69 23.23 17.79 17.60 16.45 0.80
0.80 9.94 18.04 3.49 7.03 14.90 7.19 25.51 8.53 11.22 15.19
0.90 6.16 1.51 16.10 15.01 7.65 4.94 10.15 9.33 4.96 19.70
1.00 13.77 35.03 62.89 63.45 35.05 42.51 14.38 41.38 8.72 22.17

Copyright © 2012 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2013; 42:913–933
DOI: 10.1002/eqe
OPTIMUM LATERAL LOAD PATTERN INCORPORATING SSI EFFECT 929

Table III. Constant coefficient ci of Equation (13) as function of relative height.


 =1
H=r  =3
H=r  =5
H=r
Relative height a0=0 a0=1 a0=2 a0=3 a0=1 a0=2 a0=3 a0=1 a0=2 a0=3

0.05 67.1 54.4 54.3 74.7 73 94.5 70.7 92.9 90.4 20.7
0.10 61.0 53.7 71.1 70.9 80.7 77.1 79.3 106.9 79.1 6.8
0.20 55.3 41.1 63.2 16.6 86.8 109.5 59.9 104.7 87.1 33.8
0.30 33.7 20.5 8.9 24.8 50.8 47.9 73.2 53.5 61. 56.1
0.40 13.9 5.5 35.3 16.5 36.6 6 69.5 31.4 18.5 92.4
0.50 3.5 9.8 61.8 2.5 60.7 77.3 23.8 94.1 24.1 57.6
0.60 22.3 17.6 26.8 24.1 44. 63.4 5 70.9 27.2 59.9
0.70 14.6 7.5 3.0 11.7 18.1 28.0 2.5 24.7 16. 72.1
0.80 31.2 58.8 21.1 35.1 45.6 21 82.2 23.8 25.9 36.1
0.90 54.7 36.8 11.0 34 19.7 52.8 112.9 23. 83.4 135.2
1.00 70.1 12.1 28.5 5.8 23.5 48.2 163.2 20 164.7 286

Park and Medina (Eqs. 5,6) IBC-2009 Hajirasouliha and Moghaddam (Eq. 7)
Proposed (Eq. 13)

1.5 1.5 1.5


0= 1 0= 1 0= 1
1.4 1.4 1.4
WIi/ WIOPT

1.3 1.3 1.3

1.2 1.2 1.2

1.1 1.1 1.1

1.0 1.0 1.0


0 1 2 3 0 1 2 3 0 1 2 3

2.0 2.0 2.0


0= 3 0= 3 0= 3
1.8 1.8 1.8
WIi/ WIOPT

1.6 1.6 1.6

1.4 1.4 1.4

1.2 1.2 1.2

1.0 1.0 1.0


0 1 2 3 0 1 2 3 0 1 2 3
Tfix Tfix Tfix

=1 =3 =5

Figure 12. The spectra of ratio of required to optimum structural weight for the 10-story soil–structure
systems designed according to different load patterns; average of 21 earthquakes.

patterns. This implies that this pattern loses its efficiency for this ranges of period even when SSI
effect is not significant (i.e. a0=1).
(2) For all ranges of period and SSI effects, the load pattern proposed in this study gives signifi-
cantly better results in comparison with the results of the other load patterns. The superiority
is more pronounced for the cases of longer periods with severe SSI effect. As seen, the ratios
of required to the optimum structural weight index for models designed with Equation (13)
are, on average, from 1.02 to 1.15, which can be considered as optimum for practical purposes.
(3) The loading patterns proposed by Hajirasouliha and Moghaddam [17] and IBC-2009 [7], on
average, give good results for structures with short periods and insignificant SSI effect; how-
ever, they lose their efficiency with increasing the dimensionless frequency and aspect ratio.
(4) Generally, by increasing the aspect ratio and dimensionless frequency (i.e. increasing SSI
effects) the two recently proposed optimum load patterns for fixed-base structures and code-
specified load pattern significantly lose their efficiency while the proposed load pattern in this

Copyright © 2012 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2013; 42:913–933
DOI: 10.1002/eqe
930 B. GANJAVI AND H. HAO

 = 5) with
study still display superior seismic performance, especially for slender buildings (H=r
predominant SSI effect (a0 =3). As an example, for the case of slender building with severe SSI
effect with Tfix= 1.5 s, the values of structural weight for structures designed with Equation (13),
Park and Medina [15], Hajirasouliha and Moghaddam [17] and seismic code are respectively
15%, 44%, 53% and 65% above the optimum weight. This implies that significant improvement
is achieved by utilizing the proposed load pattern of this study for sever soil–structure systems.
The average COV of story ductility ratio for the 10-story soil–structure systems designed according
to the different load patterns corresponding to the severe SSI effect a0= 3 with three values of aspect
ratios are plotted in Figure 13. As seen, the average values of COV corresponding to the Equation
(13) are less than those of the other patterns especially for the cases of the slender buildings with
long periods of vibration. As an instance, a comparison of the different lateral load patterns
corresponding to all forgoing patterns along with the resulted ductility demand distributions for the
 = 3 and a0 = 3 is provided in Figure 14. It is clear that there
10-story building with Tfix= 1.5 s, H=r
is a significant difference among the results obtained with different load patterns. Structures
designed based on the proposed load pattern in this study exhibit a much more uniform distribution
of damage along the height as compared with those designed according to other load patterns.
Figure 15 shows a comparison of the proposed equation in predicting the optimum load pattern with
the average numerical results. As shown, there is an excellent agreement between Equation (13) and

Park and Medina (Eqs. 5,6) IBC-2009 Hajirasouliha and Moghaddam (Eq. 7)
Proposed (Eq. 13)

0.5 0.5 0.5


H/r =1 H/r =3 H/r =5
0.4 0.4 0.4

0.3 0.3 0.3


COV

0.2 0.2 0.2

0.1 0.1 0.1

0.0 0.0 0.0


0 1 2 3 0 1 2 3 0 1 2 3
Tfix Tfix Tfix

Figure 13. The spectra of COV for the 10-story soil–structure systems designed according to different load
patterns; average of 21 earthquakes; a0= 3.

Park and Medina (Eqs. 5,6) IBC-2009 Hajirasouliha and Moghaddam (Eq. 7)
Proposed (Eq. 13)

10 10
9
(a) 9
(b)
8 8
7 7
6 6
Story

5 5
4 4
3 3
2 2
1 1
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.2 0.4 0.6 0.8 1 1.2
Lateral Force / Base Shear Ductility


Figure 14. Comparison of different load patterns for 10-story soil–structure systems with Tfix= 1.5 s, H=r=3
and a0= 3: (a) lateral force distribution and (b) story ductility pattern; average of 21 earthquakes.

Copyright © 2012 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2013; 42:913–933
DOI: 10.1002/eqe
OPTIMUM LATERAL LOAD PATTERN INCORPORATING SSI EFFECT 931

1.0 1.0
0.9 0.9
0.8 0.8
Relative Height

0.7 0.7
0.6 0.6
0.5 0.5
0.4 0.4
0.3 0.3
0.2 0.2
0.1 0.1
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45
Lateral Force / Base Shear Lateral Force / Base Shear

Figure 15. Correlation between Equation (13) and numerical results.

the average numerical results to estimate the optimum load pattern in elastic soil–structure systems
corresponding to different sets of structural and soil parameters.

8. CONCLUSIONS

This paper developed an optimization algorithm for optimum seismic design of elastic shear buildings
with soil–structure interaction effect for shallow foundation systems. On the basis of intensive
numerical analyses of structural models with different dynamic characteristics and SSI key
parameters subjected to 21 earthquake ground motions recorded on alluvium and soft soils with
shear wave velocity ranging from 90 to 350 m/s compatible with the IBC-2009 design spectrum, an
optimum lateral load pattern is proposed for the seismic-resistant design of shear buildings with
consideration of SSI effect. The present study provides a fundamental step towards the development
of the more rational seismic design methodology that explicitly account for the complex
phenomenon of soil–structure interaction in elastic range of response. The results of this study are
limited to the elastic shear-building structures, shallow foundation systems and selected earthquake
ground motions. Using the proposed equivalent lateral load pattern may not guarantee a uniform
distribution of deformation (ductility) along the height of the structure for the cases of structures
within inelastic range of response, embedded foundations, foundation supported by piles, earthquake
ground motions recorded on the extremely soft soil such as those recorded on lake zone of Mexico
City, and near-fault with forward directivity ground motions. This study, as the first effort on the
subject, could be a stepping stone for future research works. Therefore, more research works for
more complex structural configurations and behaviours especially for inelastic response, different
types of foundation and earthquake ground motions are deemed necessary. In addition, it is known
that the estimation of dynamic properties of soil–structure systems may involve significant
uncertainties. Therefore, it is necessary to investigate the effect of the uncertainty involved during
the determination of the dynamic properties of both structure and soil on the proposed optimum
lateral load pattern. It may be possible that under a context of large uncertainty, the use of the
proposed lateral load pattern that offers an optimum ductility distribution for systems exhibiting
important SSI effects may result in excessive ductility demands within a context of large
uncertainty, and thus that under these circumstances, another lateral load pattern may result in a
better seismic design. Although the subject of this uncertainty will need a separate study, which is
not in the scope of this paper, the optimization technique proposed in this study can be further

Copyright © 2012 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2013; 42:913–933
DOI: 10.1002/eqe
932 B. GANJAVI AND H. HAO

extended to derive the more rigorous optimum load patterns with consideration of possible
uncertainties in seismic resistant design of soil–structure systems. From the numerical results
obtained in this study, the following conclusions can be drawn:
• A value of 0.8 could be an appropriate value for a to achieve the fastest convergence in optimum
analysis.
• The optimum load pattern for elastic shear-building structures with SSI effect is highly dependent
on the fundamental period, dimensionless frequency, aspect ratio, seismic excitation and structural
damping model, but almost independent of the number of stories, structural damping ratio and
earthquake intensity.
• Although there is no significant difference between the results of mass-proportional and Rayleigh-
type damping models, the difference is pronounced when compared with that of the stiffness-
proportional damping model. The stiffness-proportional damping model may not lead to reliable
predictions of structural responses.
• The proposed load pattern, which is a function of fixed-base fundamental period, dimensionless
frequency and structure aspect ratio, gives significantly better structural design than the code-
compliant and recently proposed patterns by researchers for fixed-base structures.
• Generally, by increasing the aspect ratio and dimensionless frequency (i.e. increasing SSI effects)
the two load patterns by Park and Medina [15] and Hajirasouliha and Moghaddam [17] for fixed-
base structures and the code-specified load pattern significantly lose their efficiency while the
proposed load pattern of this study leads to better seismic performance (i.e. less structure weight
and more uniform damage distribution over height). Using the proposed load pattern in this
study leads up to 40% less structural weight for shear building structures as compared with the
code-compliant or optimum patterns developed based on fixed-base structures.

ACKNOWLEDGEMENTS
The first author gratefully acknowledges the Endeavour IPRS scholarship provided by The University of
Western Australia and Australian Government towards a successful completion of the present work.

REFERENCES
1. EuroCode-8. Final draft of EuroCode-8. Design of structure for earthquake resistance – Part 1: General rules for
buildings. European Committee for Standardization: Bruxelles, 2003.
2. Mexico City Building Code, 2003.
3. Uniform Building Code. International Conference of Building Officials. Whittier: California, UBC-1997.
4. Building Seismic Safety Council (BSSC). National Earthquake Hazard Reduction Program (NEHRP). Recommended
Provisions for Seismic Regulations for 348 New Buildings and Other Structures—Part 2: Commentary (FEMA
450–2). Federal Emergency Management Agency: Washington, D. C, 2003.
5. ASCE/SEI 7–05. Minimum Design Loads for Buildings and Other Structures. American Society of Civil Engineers:
Reston, VA, 2005.
6. Structural design actions. Earthquake actions in Australia. AS-1170.4; 2007.
7. IBC-2009. International Building Code. International Code Council, Country Club Hills: USA, 2009.
8. Anderson JC, Miranda E, Bertero VV. Evaluation of the seismic performance of a thirty-story RC building. UCB/
EERC-91/16, Earthquake Engineering Research Centre. University of California: Berkeley, 1991.
9. Gilmore TA, Bertero VV. Seismic performance of a 30-story building located on soft soil and designed according to
UBC 1991. UCB/EERC-93/04. Earthquake Engineering Research centre. University of California: Berkeley, 1991.
10. Leelataviwat S, Goel SC, Stojadinovic´ B. Toward performance-based seismic design of structures. Earthquake
Spectra 1999; 15:435–461.
11. Lee SS, Goel SC. Performance-Based Design of Steel Moment Frames Using Target Drift and Yield Mechanism.
Report No. UMCEE 01–17, Department of Civil and Environmental Engineering, University of Michigan: Ann
Arbor, 2001.
12. Mohammadi RK, El Naggar MH, Moghaddam H. Optimum strength distribution for seismic resistant shear
buildings. International Journal of Solids and Structures 2004; 41(22):6597–6612.
13. Ganjavi B, Vaseghi Amiri J, Ghodrati Amiri G, Yahyazadeh Ahmadi Q. Distribution of Drift, Hysteretic Energy and
Damage in Reinforced Concrete Buildings with Uniform Strength Ratio. The 14th World Conf. on Earthquake En-
gineering. Mira Digital Publishing: 3800 Park Avenue, St. Louis, MO 63110, USA, October 12–17, 2008.
14. Uniform Building Code. International Conference of Building Officials. Whittier: California, UBC-1994.
15. Park K, Medina RA. Conceptual seismic design of regular frames based on the concept of uniform damage. Journal
of Structural Engineering (ASCE) 2007; 133(7):945–955.

Copyright © 2012 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2013; 42:913–933
DOI: 10.1002/eqe
OPTIMUM LATERAL LOAD PATTERN INCORPORATING SSI EFFECT 933

16. Moghaddam H, Hajirasouliha I. Optimum strength distribution for seismic design of tall buildings. The Structural
Design of Tall and Special Buildings 2008; 17:331–349.
17. Hajirasouliha I, Moghaddam H. New lateral force distribution for seismic design of structures. Journal of Structural
Engineering (ASCE) 2009; 135(8):906–915.
18. Goel SC Liao WC, Bayat MR, Chao SH. Performance-Based Plastic Design Method for Earthquake-Resistant Struc-
tures: An Overview. The Structural Design of Tall Special Buildings 2010; 19:115–137.
19. Mahsuli M, Ghannad MA. The effect of foundation embedment on inelastic response of structures. Earthquake
Engineering and Structural Dynamics 2009; 38(4):423–437.
20. Chouw N, Hao H. Significance of SSI and non-uniform near-fault ground motions in bridge response I: effect on re-
sponse with conventional expansion joint. Engineering Structures 2008; 30(1):141–153.
21. Barcena A, Esteva L. Influence of dynamic soil–structure interaction on the nonlinear response and seismic reliability
of multistory systems. Earthquake Engineering and Structural Dynamics 2007; 36(3):327–346.
22. Rodriguez ME, Montes R. Seismic response and damage analysis of buildings supported on flexible soils.
Earthquake Engineering and Structural Dynamics 2000; 29(5):647–665.
23. Aviles J, Perez-Rocha L. Soil–structure interaction in yielding systems. Earthquake Engineering and Structural
Dynamics 2003; 32(11):1749–1771.
24. Aviles J, Perez-Rocha JL. Influence of foundation flexibility on Rm and Cm factors. Journal of Structural Engineering
(ASCE) 2005; 131(2):221–230.
25. Aviles J, Perez-Rocha JL. Use of global ductility for design of structure–foundation systems. Soil Dynamics and
Earthquake Engineering 2011; 31(7):1018–1026.
26. Ganjavi B, Hao H. Elastic and Inelastic Response of Single- and Multi-Degree-of-Freedom Systems Considering Soil
Structure Interaction Effects. Australian Earthquake Engineering Society Conference. Elsevier: Barossa Valley,
South Australia, 18–20 November, 2011.
27. Ganjavi B, Hao H. A parametric study on the evaluation of ductility demand distribution in multi-degree-of-freedom
systems considering soil–structure interaction effects. Engineering Structures 2012; 43:88–104.
28. SeismoMatch. A computer program for adjusting earthquake records to match a specific target response spectrum.
Available from: http://www.seismosoft.com, 2011 [June 2011].
29. Moghaddasi M, Cubrinovski M, Chase JG, Pampanin S, Carr A. Probabilistic evaluation of soil–foundation–
structure interaction effects on seismic structural response. Earthquake Engineering and Structural Dynamics
2011; 40(2):135–154.
30. Wolf JP. Foundation Vibration Analysis using Simple Physical Models. Prentice-Hall: Englewood Cliffs, NJ, 1994.
31. FEMA 440. Improvement of nonlinear static seismic analysis procedures. Report No. FEMA 440, Federal Emergency
Management Agency, prepared by Applied Technology Council, 2005.
32. Veletsos AS. Dynamics of structure–foundation systems. In Structural and Geotechnical Mechanics, Hall WJ (ed.),
A Volume Honoring N.M. Newmark. Prentice-Hall: Englewood Cliffs, NJ, 1977; 333–361.
33. Veletsos AS, Meek JW. Dynamic behavior of building–foundation system. Earthquake Engineering and Structural
Dynamics 1974; 3(2):121–138.

Copyright © 2012 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2013; 42:913–933
DOI: 10.1002/eqe

View publication stats

Das könnte Ihnen auch gefallen