Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

International Journal of Civil Engineering (2020) 18:1209–1228

https://doi.org/10.1007/s40999-020-00529-0 (0123456789().,-volV)(0123456789().
,- volV)

RESEARCH PAPER

The Effect of Nonlinear Soil–Structure Interaction on the Ductility


and Strength Demands of Vertically Irregular Structures
Hamid Asadi-Ghoozhdi1 • Reza Attarnejad1

Received: 5 January 2020 / Revised: 22 May 2020 / Accepted: 25 May 2020 / Published online: 8 June 2020
Ó Iran University of Science and Technology 2020

Abstract
Soil–structure interaction (SSI) is key to the elastic and inelastic seismic responses of buildings. In this research, the
ductility and strength demands of vertically irregular MDOF buildings, considering nonlinear SSI effects, are paramet-
rically investigated. The superstructure is modeled as a nonlinear multi-story shear building, and the beam on nonlinear
Winkler foundation (BNWF) concept is employed to simulate the response of shallow foundations. Specifically, combined
stiffness–strength irregularities are introduced by reducing lateral properties at a specific story of the regular (reference)
models. Soil–structure systems with 5, 10, and 15 stories are analyzed under three sets of earthquake records corresponding
to different soil classes. A wide range of key parameters including number of stories, fundamental period, level of
inelasticity, aspect ratio, and site class are scrutinized through nonlinear time history analyses. The results reveal that SSI
can reduce the strength and ductility demands of the vertically irregular structures, especially those with short periods. This
beneficial effect becomes even more significant for systems with low ductility ratios. It is also concluded that the median
ductility demand increases in the modified story owing to the softness/weakness of the first story. Furthermore, this
increase due to the code strength regularity limit reached up to 78% and 36% in fixed-base and flexible-base conditions,
respectively. Finally, simplified equations are proposed to estimate the maximum ductility demands of regular and irregular
structures with flexible-base conditions.

Keywords Vertical irregularity  Soil–structure interaction  Ductility demand distribution  Nonlinear dynamic analysis 
Site class  Strength demand

1 Introduction 1989 Loma Prieta, the 1994 Northridge, the 1995 Kobe,
and the 1999 Izmit disasters as well as the recent earth-
Functional or architectural limitations may cause structural quakes in Kumamoto, Japan and the Iran–Iraq border.
irregularities, among which the non-geometric vertical The effects of vertical irregularities on the linear and
irregularities generally occur due to the non-uniform dis- nonlinear responses of structures have been considerably
tribution of mass, stiffness, or strength along the height of a studied, with results showing that these irregularities
building. Such irregularities happen particularly in base- increase the seismic demand of structures [1–5]. The
ments, where large open spaces are allocated to parking, irregularity limits imposed by the 1994 Uniform Building
lobbies, or shops. The soft/weak story mechanism can lead Code (UBC) in terms of structural stiffness, strength, and
to severe damage and collapse, as evidenced by experi- mass as well as their effects on the maximum ductility
ences and observations during past earthquakes, such as the demand of irregular structures was evaluated [1]. Al-Ali
and Krawinkler [2] investigated the influence of vertical
irregularities on height-wise variations in seismic demand
& Reza Attarnejad
attarnjd@ut.ac.ir and found that mass irregularity effects are less severe than
those of vertical stiffness and strength irregularities. The
Hamid Asadi-Ghoozhdi
h_asadi@ut.ac.ir effects of vertical irregularities on seismic demands of a
frame designed in accordance with the philosophy of
1
School of Civil Engineering, College of Engineering, strong column-weak beam were determined [3]. Under this
University of Tehran, Tehran, Iran

123
1210 International Journal of Civil Engineering (2020) 18:1209–1228

configuration, plastic hinges are permitted to develop only over the height of structures [21–24]. The influence of
at the beam ends and the base of the columns. Introducing a peripheral wall openings in the basement and the number
soft and/or weak story increases drift demands in modified of basement floors on the seismic response of buildings
and neighboring stories but decreases drift demands in with braced framed system was studied by considering SSI
other stories. The effects of mass, stiffness, strength, and [25].
coupled stiffness-strength irregularities under different The effects of SSI on the strength reduction factor (SRF,
intensities of ground motions were studied [4]. By denoted as Rl) have been extensively investigated over the
expressing all limit state capacities using a common last decade. A parametric study was carried out to deter-
intensity measure, the reference and modified structures mine the SRF of elasto-plastic SDOF systems, considering
were subjected to a set of earthquake records scaled to both site and interaction effects [26]. The results showed
several intensity levels. Simple equations were proposed that SSI generally decreases the SRF. Rl for MDOF soil–
for rapidly estimating variations in critical inter-story drift structure systems were estimated using proposed simplified
demands due to coupled stiffness–strength irregularities equations that are a function of fixed-base fundamental
[6]. The modification of the ductility reduction factor for periods, ductility ratios, number of stories, and other
vertically irregular structures under pulse-like ground important SSI parameters [27]. Lu et al. [28] also studied
motions was suggested [7], which significantly reduces the the performance-based design of MDOF shear buildings
aforementioned factor, thereby enabling control over the with consideration for SSI effects and used the results as
ductility demands of structures with vertical irregularities basis in introducing two design factors that capture strength
[8]. Researchers have also evaluated the reliability of demands of SDOF and MDOF systems. It should be noted
concrete moment frames with vertical stiffness or strength that foundation uplift and soil yielding were not considered
irregularities using probabilistic methods [9, 10]. Numeri- in the above studies.
cal investigations under seismic base excitation have been The effects of foundation uplift on the linear and non-
devoted to the elastic and inelastic responses of a class of linear responses of SDOF systems were evaluated [29, 30],
nonlinear oscillators that represent structural systems with and the results showed a decrease in base shear, drift, and
relatively rigid superstructures on a soft ground story ductility demand as well as an increase in top displace-
[11, 12]. The concept of direct displacement-based seismic ment. Based on the seismic responses of low-rise steel
design (DDBD) has been applied to several vertically frames to which nonlinear SSI was incorporated using the
irregular reinforced concrete (RC) frames [13]. Moreover, beam on nonlinear Winkler foundation (BNWF) approach,
a parametric study of simplified two-DOF models was it was ascertained that foundation uplifting and soil yield-
carried out to investigate the seismic responses of struc- ing reduces the ductility demands of buildings [31]. The
tures with soft or weak stories [14]. Despite the value BNWF concept was also employed in a parametric study
presented by these studies, however, they were conducted regarding the influence of nonlinear SSI on the displace-
under the assumption that structures are built on strong ment amplification factor (C1) [32]. The results indicated
rock foundations (fixed-base conditions). that the equations suggested in current rehabilitation codes
In reality, the flexibility of soil beneath a structure and underestimate top displacement demands in the presence of
the soil–foundation–structure interaction can affect the foundation rocking and uplifting.
seismic behaviors of buildings. Soil–structure interaction Relatively few studies have been dedicated to the
(SSI) initially changes the seismic input motion applied to influence of SSI on vertically irregular buildings. An
a system—a phenomenon known as kinematic interaction. assessment was conducted on the two-story, two-bay
Second, the dynamic characteristics of a soil–structure asymmetric frame by considering the effects of material
system are changed, with such an alteration always (soil and superstructure) yielding and geometric (uplifting
accompanied by an increase in natural period. In addition, and P-D) nonlinearities [33]. It was concluded that foun-
the energy radiation of waves propagating away from a dation uplift and soil yielding strongly reduce inter-story
structure in semi-infinite soil media generally increases drift and increase total displacements of soil–structure
soil–structure damping [15]. Most parametric studies on systems. Nonlinear analysis of a three-dimensional model
inelastic soil–structure systems focused on single-degree- was performed for evaluating the seismic responses of
of-freedom (SDOF) structures (e.g., [16–19]). The seismic setback steel buildings in a probabilistic framework [34]. It
behavior of a reference wind turbine was evaluated con- was observed that SSI reduces seismic capacity and
sidering SSI effects [20]. In recent years, a number of amplifies the drift demands of flexible-base setback
studies have been conducted to investigate the drift and buildings. An investigation was also carried out on the
ductility demands of multiple-degree-of-freedom (MDOF) probabilistic seismic performance of non-geometric verti-
structures, taking into account soil flexibility generally on cally irregular steel buildings, with SSI effects taken into
the basis of a regular distribution of stiffness and strength account [35], and a more recent study clarified the

123
International Journal of Civil Engineering (2020) 18:1209–1228 1211

influence of linear SSI on the ductility demands of verti- Force


cally irregular structures [36]. α K0
Most of the above-mentioned examinations centered on Vy
case studies or regular structures and inadequately eluci-
dated the effects of SSI on the seismic demands of verti- Kμ =K0.μβ
cally irregular systems. Parametric studies that account for
K0
the effects of vertical irregularities and nonlinear SSI are
Deformation
also rare. The present study aims to assess the influence of dy dm
SSI on the ductility and strength demands of vertically
irregular structures. To this end, shear building structures
Strain hardening ratio (α)
are subjected to three sets of ground motions for site Unloading stiffness parameter (β)
classes C, D, and E in fixed-base and flexible-base condi- -Vy
tions. Effects of number of stories, design ductility, fun-
damental period, soil flexibility, site class, aspect ratio, and
the irregularity ratio on ductility demand distribution are Fig. 1 Takeda hysteretic model
extensively evaluated. For practical applications, simplified
expressions are proposed to compute maximum ductility have similar patterns as regards the vertical distribution of
demands of regular and vertically irregular buildings with lateral seismic forces. In general terms, the seismic design
flexible bases. force (Fx) applied to each level x, due to response in the
first mode of vibration can be expressed as:
w x hk
2 Soil–Structure Systems Fx ¼ Pn x k V ð1Þ
i¼1 wi hi

2.1 Reference Structural Models where V is the total base shear of an MDOF structure; wi
and hi are the effective weight imposed on level i and the
Contrary to discussions regarding the limitations of well- height of the floor at level i from the base, respectively.
known shear beam models, these representations have been Parameter n denotes the number of stories, and the k-ex-
widely used in seismic analyses of multi-story regular (e.g., ponent is a function of the fundamental period of a building
[27, 28, 37]) and irregular [6, 7] buildings, because they (T1). Using the Rayleigh method, 5% damping is assigned
effectively capture both nonlinear behaviors and higher to the first mode of vibration and the vibration mode with at
mode effects and are therefore suitable for extensive least a 90% cumulative mass participation.
parametric studies due to simplicity and low computational
cost. The structures assessed in this study are 5-, 10-, and 2.2 Vertically Irregular Models
15-story shear-type models, with mass assumed to be
concentrated at each floor level. Story height is set at A structural irregularity is defined as a change in the
3.5 m, and the total structural mass is considered uniformly geometric characteristics or parameters of mass, stiffness,
distributed along the height of a structure. The force–dis- and strength. The limits on vertical stiffness and strength
placement relationships in shear beams are specified using irregularities in the Iranian seismic code [38] and ASCE
the Takeda hysteretic rule with a strain hardening of 5% 7-10 [39] are similar. These limits are described as follows:
(Fig. 1). This model represents the cyclic behaviors of non-
deteriorating (ductile) RC members. 1. Lateral stiffness irregularity: If the lateral stiffness of a
All the reference models (fixed-base regular structures) story is less than 70% of that in the story above it or
are first designed in accordance with the Iranian seismic less than 80% of the average stiffness of the three
code of practice (i.e., Standard No. 2800) [38] and ASCE stories above it, this level is referred to as a ‘‘soft
7-10 [39]. The lateral stiffness of each story is assumed story.’’ If these values decrease to 60% and 70%,
proportional to its shear strength, which is distributed over respectively, then the story is described as an ‘‘ex-
height on the basis of the code-specified lateral load pat- tremely soft story.’’
tern. The total stiffness of a structure is calculated in a way 2. Lateral strength irregularity: A story in a building is
that accorded a specified target fundamental period to the called a ‘‘weak story’’ if the lateral strength of a story is
building. The base shear strength of MDOF structures is less than 80% of that in the story above it. If the value
also determined in accordance with the presumed target decreases to 65%, then the story is described as an
ductility of the fixed-base models in a trial-and-error pro- ‘‘extremely weak story.’’
cess. Note that ASCE 7-10 and the Iranian seismic code

123
1212 International Journal of Civil Engineering (2020) 18:1209–1228

Given the fact that a change in stiffness is generally NEHRP classification, whose properties are presented in
associated with an alteration in strength in most practical Table 1. The rigid foundation was assumed to be rectan-
concrete frames, this study focuses on coupled stiffness– gular with foundation width B-to-length L (B/L) ratios of
strength irregularities, which are introduced separately in 0.4, 0.75, and 1.0 for the 5-, 10-, and 15-story structures,
two locations along the height of a building, namely the respectively. The soil beneath the foundation is replaced by
first story and the top story. Vertical irregularities are a set of spring-damper elements (Fig. 3). The horizontal
created by applying three modification factors. As shown dynamic stiffness of shallow foundation is modeled using a
in Fig. 2, irregular structures emerged from the reduction parallel spring-damper pair, whose coefficients are calcu-
in stiffness and strength parameters at the first story; that is, lated using the equations proposed by Gazetas [43]:
the parameters decreased by 90%, 80%, and 65% of those "  0:85 #
qVs2 L B
found on the upper (second). For irregularity at the roof kh ¼ 2 þ 2:5 ; ch ¼ qVs Ab ð2Þ
level, the stiffness and strength values were assumed to be 2m L
0.9, 0.8, and 0.65 times those of the reference regular
where q, t, and Vs represent the soil density, Poisson’s
structures. These modification factors are denoted by k.
ratio, and effective shear wave velocity of soil, respec-
Note that a fair comparison of the responses of regular and
tively, and Ab is the area of the foundation. Vertical soil
irregular structures necessitates equality in the main char-
springs were modeled through one-dimensional, zero-
acteristics of the buildings, such as fundamental periods,
length springs. The QzSimple2 material is used in the
base shear strengths, and damping properties [3]. Accord-
OpenSees (Open System for Earthquake Engineering
ing to Fig. 2, stiffness was uniformly scaled by coefficient
Simulation) software to define the nonlinear inelastic
gk for all the stories, thereby keeping the irregularity ratio
behaviors of the vertical springs. The behaviors of this
constant and maintaining the same period across the reg-
material were derived on the basis of the pile-calibrated
ular and irregular structures. Similarly, for a soft first story,
backbone curves developed by Boulanger et al. [44] and
the shear strengths were scaled uniformly to obtain an
were calibrated against a set of shallow foundation tests
irregular frame with the same yield base shear as that found
[41]. The initial vertical stiffness of the footing is derived
in a regular frame.
as follows [43]:
"  0:75 #
2.3 Soil–Foundation Model GL B
kv ¼ 0:73 þ 1:54 ð3Þ
1m L
The BNWF approach is used to model the behavior of
shallow foundation resting on a semi-infinite medium—a where G is the shear modulus of soil. As depicted in Fig. 3,
concept that has been widely implemented in previous the stiffness of the end region springs and end length ratio
studies (e.g., [40–42]). This method enables the simulta- are determined in such a way as to produce rotational
neous incorporation of geometric nonlinearity (foundation stiffness of the shallow foundations. The end length ratio is
uplift) and material nonlinearity (soil yielding) effects. The adapted from Harden et al. [40], who recommended this
selected soil types are classes C, D, and E according to the variable as a function of L/B. Generally, an increase in
spring stiffness Rk can be calculated as a function of
foundation end length ratio Re in the following manner:
 
12kh 3
kzi BL3  ð1  2Re Þ Lend
Rk;h ¼ 3
; Re ¼ ð4Þ
1  ð1  2Re Þ L

Equation (4) was derived by matching the moment


produced by the springs for a unit foundation rotation to
rotational stiffness kh. The stiffness and damping of the
surface foundation for rocking motion that are used to
define the soil–foundation model are described as follows:
 0:15
3qVs2 0:75 L
kh ¼ ðIb Þ ; ch ¼ qVa Ib ð5Þ
1m B
where Ib is the moment of inertia of the surface foundation
about the transverse axis. The coefficients of the vertical
Fig. 2 Introducing coupled stiffness–strength irregularity at the first damper elements as part of the q-z material can be evalu-
story ated thus:

123
International Journal of Civil Engineering (2020) 18:1209–1228 1213

Table 1 Soil properties used in this study


Soil type Mass density (q) Shear wave velocity (Vs0) Poisson’s ratio (t) Soil damping (ng)

C 1.9 t/m3 400 m/s 0.33 0.04


D 1.8 t/m3 200 m/s 0.4 0.07
E 1.7 t/m3 120 m/s 0.45 0.09

Center [47]. All the ground motions were obtained from


earthquakes with magnitudes greater than 6 and with the
m closest distance to fault ruptures exceeding 9 km. The main
characteristics of the selected ground motions are sum-
marized in Table 2. Three sets of ground motions corre-
sponding to site classes C, D, and E are used, and each set
m of the earthquakes consists of 20 seismic excitations. Fig-
Htot ×h ure 4 depicts the acceleration response spectra of 20
ground motions in each set in comparison with the corre-
sponding Iranian design spectrum.
All the ground motions in each set were scaled to the
m elastic design spectrum [design-basis earthquake (DBE)
level] in such a way as to guarantee that the spectral
Rigid elastic
element acceleration of an earthquake record matched the target
spectral acceleration at the fundamental period of a struc-
ture. Note that kinematic interaction effects can be
Zero-length
kend kmid kend Zero-length neglected for shallow foundations subjected to vertically
elements element
(QZSimple2 Lend Lmid Lend
(Equivalent propagating shear waves. Therefore, free-field ground
material) linear)
motions can serve as the foundation input motion.
L

4 Key Parameters
Fig. 3 Soil–MDOF structure model
The seismic responses of soil–structure systems depend on
ch
ciz ¼ PðS=2Þ ð6Þ the characteristics of the superstructure, soil, and input
2 i¼1 ðile Þ2 motion. The influence of these factors can be described
where le denotes the non-normalized spring spacing. A using the following key parameters:
minimum of 25 springs along the length of a footing is 1. Fixed-base period (T1): This is a fundamental period
suggested to ensure numerical stability and reasonable ranging from 0.3 to 1.5 s for 5-story structural models,
accuracy [45]. Note that the hysteretic damping of soil 0.5 to 2.5 s for 10-story structural models, and 1.0 to
material is explicitly included in the model. The vertical 3.0 s for 15-story structural models, with intervals of
bearing capacity of the underlying soil is calculated for 0.1 s implemented to cover actual ranges for practical
each system to acquire safety factors of 3 and 5 under structures.
gravity loads. In other words, the capacity of the BNWF 2. Design ductility of the structure (lt): Reference regular
springs is obtained with a presumed factor of safety against structures are designed on the basis of target ductility
vertical loading. levels of 1, 2, 4, and 6. Note that the ductility demand
of i-th story li is calculated as:
dmi
3 Selected Ground Motions li ¼ ð7Þ
dyi
A total of 60 earthquake ground motions with different where dmi and dyi are the i-th maximum inter-story
characteristics were chosen from FEMA-440 [46] for site displacement and yield displacement of the same story,
classes C, D, and E; these data were derived from the respectively.
database of the Pacific Earthquake Engineering Research

123
1214 International Journal of Civil Engineering (2020) 18:1209–1228

Table 2 Characteristics of the ground motions used in this study (a) Ground motions recorded on site class C (b) Ground motions recorded on
site class D (c) Ground motions recorded on site class E (soft soil)
Date Earthquake M Station Distance (km)* Component PGA (cm/s2)

(a)
10/15/79 Imperial Valley 6.8 El Centro, Parachute Test Facility 12.69 315 200.2
02/09/71 San Fernando 6.5 Pasadena, CIT Athenaeum 25.47 90 107.9
02/09/71 San Fernando 6.5 Pearblossom Pump 38.97 21 133.4
06/28/92 Landers 7.5 Yermo, Fire Station 23.62 0 167.8
10/17/89 Loma Prieta 7.1 APEEL 7, Pulgas 41.86 0 153.0
10/17/89 Loma Prieta 7.1 Gilroy#6, San Ysidro Microwave site 18.33 90 166.9
10/17/89 Loma Prieta 7.1 Saratoga, Aloha Ave. 8.50 0 494.5
10/17/89 Loma Prieta 7.1 Gilroy, Gavilon College Phys Sch Bldg 9.96 67 349.1
10/17/89 Loma Prieta 7.1 Santa Cruz, University of California 18.41 360 433.1
10/17/89 Loma Prieta 7.1 San Francisco, Diamond Heights 71.33 90 110.8
10/17/89 Loma Prieta 7.1 Fremont, Mission San Jose 39.51 0 121.6
10/17/89 Loma Prieta 7.1 Monterey, City Hall 44.35 0 71.6
10/17/89 Loma Prieta 7.1 Yerba Buena Island 75.17 90 66.7
10/17/89 Loma Prieta 7.1 Anderson Dam, Downstream 20.26 270 239.4
04/24/84 Morgan Hill 6.1 Gilroy, Gavilon College Phys Sci Bldg 14.84 67 95.0
04/24/84 Morgan Hill 6.1 Gilroy#6, San Ysidro Microwave Site 9.86 90 280.4
07/08/86 Palmsprings 6.0 Fun Valley 14.24 45 129.0
01/17/94 Northridge 6.8 Littlerock, Brainard Canyon 46.58 90 70.6
01/17/94 Northridge 6.8 Castaic, Old Ridge Route 20.72 360 504.2
01/17/94 Northridge 6.8 Lake Hughes#1, Fire station#78 35.81 0 84.9
(b)
06/28/92 Landers 7.5 Yermo, Fire Station 23.62 270 240.0
06/28/92 Landers 7.5 Palm Springs, Airport 36.15 90 87.2
06/28/92 Landers 7.5 Pomona, 4th and Locust, Free Field 117.5 0 65.5
01/17/94 Northridge 6.8 Los Angeles, Hollywood Storage Bldg. 24.03 360 381.4
01/17/94 Northridge 6.8 Santa Monica City Hall 26.45 90 866.2
01/17/94 Northridge 6.8 Los Angeles, N. Westmoreland 26.73 0 393.3
10/17/89 Loma Prieta 7.1 Gilroy 2, Hwy 101 Bolsa Road Motel 11.07 0 394.2
10/17/89 Loma Prieta 7.1 Gilroy 3, Sewage Treatment Plant 12.82 0 531.7
10/17/89 Loma Prieta 7.1 Hayward, John Muir School 52.68 0 166.5
10/17/89 Loma Prieta 7.1 Agnews, Agnews State Hospital 24.57 0 163.1
10/01/87 Whittier Narrows 6.1 Los Angeles, 116th St School 23.29 270 288.4
10/01/87 Whittier Narrows 6.1 Downey, County Maintenance Bldg. 20.82 180 193.2
10/15/79 Imperial Valley 6.8 El Centro#13, Strobel Residence 21.98 230 136.2
10/15/79 Imperial Valley 6.8 Calexico, Fire Station 10.45 225 269.6
04/24/84 Morgan Hill 6.1 Gilroy#4, 2905 Anderson Rd 11.54 360 341.4
04/24/84 Morgan Hill 6.1 Gilroy#7, Mantnilli Ranch, Jamison Rd 12.07 0 183.0
04/24/84 Morgan Hill 6.1 Gilroy#2, Keystone Rd. 13.69 90 207.9
04/24/84 Morgan Hill 6.1 Gilroy#3 Sewage Treatment Plant 13.02 90 189.8
02/09/71 San Fernando 6.5 Los Angeles, Hollywood Storage Bldg. 21.20 90 207.0
02/09/71 San Fernando 6.5 Vernon, Cmd Terminal Building 4814 Loma Vista 49.9 277 104.6
(c)
10/17/89 Loma Prieta 7.1 Foster City (APEEL 1; Redwood Shores) 43.80 90 277.6
10/17/89 Loma Prieta 7.1 Foster City (APEEL 1; Redwood Shores) 43.80 360 63.0
10/17/89 Loma Prieta 7.1 Larkspur Ferry Terminal 94.60 270 134.7
10/17/89 Loma Prieta 7.1 Larkspur Ferry Terminal 94.60 360 94.6
10/17/89 Loma Prieta 7.1 Redwood City (APEEL Array Stn. 2) 43.23 43 270.0

123
International Journal of Civil Engineering (2020) 18:1209–1228 1215

Table 2 (continued)
Date Earthquake M Station Distance (km)* Component PGA (cm/s2)

10/17/89 Loma Prieta 7.1 Redwood City (APEEL Array Stn. 2) 43.23 133 222.0
10/17/89 Loma Prieta 7.1 Treasure Island (Naval Base Fire Station) 77.42 0 112.0
10/17/89 Loma Prieta 7.1 Treasure Island (Naval Base Fire Station) 77.42 90 97.9
10/17/89 Loma Prieta 7.1 Emeryville, 6363 Christie Ave. 76.90 260 254.7
10/17/89 Loma Prieta 7.1 Emeryville, 6363 Christie Ave. 76.90 350 210.3
10/17/89 Loma Prieta 7.1 San Francisco, International Airport 58.65 0 231.5
10/17/89 Loma Prieta 7.1 San Francisco, International Airport 58.65 90 322.7
10/17/89 Loma Prieta 7.1 Oakland, Outer Harbor Wharf 74.26 35 281.4
10/17/89 Loma Prieta 7.1 Oakland, Outer Harbor Wharf 74.26 305 265.5
10/17/89 Loma Prieta 7.1 Oakland, Title & Trust Bldg. (2-story) 72.20 180 191.3
10/17/89 Loma Prieta 7.1 Oakland, Title & Trust Bldg. (2-story) 72.20 270 239.4
10/15/79 Imperial Valley 6.8 El Centro Array 3, Pine Union School 12.85 140 260.9
10/15/79 Imperial Valley 6.8 El Centro Array 3, Pine Union School 12.85 230 216.8
04/24/84 Morgan Hill 6.1 Foster City (APEEL 1; Redwood Shores) 53.89 40 45.1
04/24/84 Morgan Hill 6.1 Foster City (APEEL 1; Redwood Shores) 53.89 310 66.7
*
Closest distance to fault rupture

Individual ground motions Design Spectrum


3 3 3
Site Class C Site Class D Site Class E
Spectral Acceleration (g)

Spectral Acceleration (g)


Spectral Acceleration (g)

2.5 2.5 2.5

2 2 2

1.5 1.5 1.5

1 1 1

0.5 0.5 0.5

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

Fig. 4 Elastic response spectra of ground motions in each set with the corresponding design spectrum for site classes C, D, and E

3. Soil flexibility: Soil shear wave velocity (Vs0) is set at qult BL


FSv ¼ ð9Þ
120, 200, and 400 m/s, for site classes C, D, and E, Nu
respectively.
where qult denotes the ultimate vertical bearing
4. Aspect ratio of the building: This parameter is defined
capacity, and Nu is the applied gravity load. This factor
as:
was employed in modeling clearly different foundation
H behaviors, from uplifting-dominated (high FSv) to
s¼ ð8Þ
ðL=2Þ sinking-dominated (low FSv) responses.
where H  represents the effective height of the MDOF Other parameters are less important and can be assigned
structure and can be approximated as 70% of the total for conventional buildings using typical values [19, 21].
height of the structure, according to the provisions of One such parameter is the ratio of foundation mass to the
ATC3-06 [48] and ASCE 7-10. total mass of the structure, which is assumed to be 0.1.
5. Vertical factor of safety (FSv): The static factor of
safety against vertical loading is defined as:

123
1216 International Journal of Civil Engineering (2020) 18:1209–1228

5 Methodology and General Procedures 6. The soil–structure system with the same yield strength
as that calculated in Step 5 is subjected to the selected
Nonlinear dynamic analyses were conducted directly in the free-field ground motion, and ductility demand distri-
time domain using the OpenSees [49] software of the bution along the height of the system is computed.
Pacific Earthquake Engineering Research Center. The Note that the selected shear wave velocity and soil
dynamic analysis was performed under earthquake ground parameters correspond to the soil type chosen in step 4.
motion after the application of gravity load and initial 7. An initial irregular structure is created, with reductions
foundation settlement due to vertical loading. Soil–struc- in stiffness and strength parameters applied at the first
ture systems were subjected to 60 ground motions recorded story or roof level of a building (applying the k factor).
on site classes C, D, and E provided by FEMA-440 [46]. 8. The total stiffness and strength of the irregular
For each ground motion representing a specific site class, a structure are then scaled so that the fundamental period
family of 12,900 soil–structure systems covering 5-, 10-, and and base shear strength become similar to those of the
15-story models with different fixed-base periods and a wide corresponding reference model (through the coeffi-
range of predefined key parameters were considered. The cients gk and gs shown in Fig. 2).
systems included vertically irregular and reference regular 9. The final irregular structure with the base shear
structures in fixed-base and flexible-base conditions with strength obtained from step 5 is excited by the selected
aspect ratios of 2, 3, 5; FSv values of 3 and 5; design ductility earthquake record in fixed-base and flexible-base
ratios of 1, 2, 4, and 6; irregularity ratios of 0.9, 0.8, and 0.65; conditions.
two irregularity positions at the first story and roof level; and
three soil types C, D, and E. It should be mentioned that the
ranges of some key parameters, such as the fundamental 6 Effects of SSI on Ductility Demand
period, may be regarded as wider than their practical values to Distribution
cover all possible systems.
For each system, the yield strength demand of a regular This section comprehensively explains the effects of ver-
structure in the fixed-base state was first calculated for a tical irregularities and SSI on the ductility demand distri-
given design ductility under each earthquake ground bution of the MDOF structures. It also presents the separate
motion. For any given case, the base shear strength of evaluations of the influence of the predefined key
superstructure Vy was calculated via iteration to reach the parameters.
target ductility for the fixed-base models within an accu-
racy of 2%. Then, the ductility demand of each soil– 6.1 Effect of Irregularity Ratio
structure system was calculated, in which the yield strength
of the superstructure was similar to that of a fixed-base The effects of irregularity ratio k on the ductility demands
model. That is, the effects of irregularities and nonlinear of the 10-story models are illustrated for fixed-base and
SSI on ductility demand were evaluated. The procedure flexible-base shear buildings. As the irregularity ratio (k)
described above is summarized as follows (Fig. 5): decreases, the degree of irregularity increases, indicating
1. Shear building models are defined with consideration that the structures soften/weaken. Figure 6 shows the mean
of the number of stories. values obtained from 20 ground motions recorded on site
2. The fundamental period and design ductility of a fixed- class D. The results are derived for both fixed-base and
base regular system are selected. flexible-base models with soft/weak first stories, T1 = 1 s,
3. The stiffness and strength of the MDOF structure are s = 3, FSv = 3, and design ductility levels of 1, 4, and 6.
distributed along the height of the building in accor- The abscissa is the averaged ductility demand, and the
dance with the code-specified lateral load pattern. The vertical axis is the relative height of a structure. Figure 6
total stiffness and base shear strength of the structure indicates that as the irregularity ratio increases, median
must be scaled to achieve a target fundamental period ductility demand increases at the first story but decreases in
and target ductility ratio, respectively. the other stories for both fixed-base and flexible-base
4. An earthquake ground motion is chosen from three sets irregular buildings. This trend is intensified by increasing
of records corresponding to different soil classes the structural nonlinearity, so a target ductility of lt = 6
5. A nonlinear time history analysis is performed on causes the collapse of the first story. For the low ductility
reference structures in fixed-base conditions under the structures (elastic state), the ductility demand in a modified
given earthquake record, and Vy is calculated to story is similar to that in the regular cases, and the ductility
achieve the presumed design ductility with 2% demand in the upper stories decreases. In addition, non-
tolerance. linear SSI is beneficial for irregular structures and reduces

123
International Journal of Civil Engineering (2020) 18:1209–1228 1217

Select: n , T1 and μt Select: s , FSv and λ

Assign initial stiffness Ktot and strength Vy,tot , and


Calculate input parameters for the Winkler model
distribute them along the height of the structures

Scale total stiffness to achieve the target period T1 Analyze soil- structure system with the same yield
Adjust strength (Vy,tot) under the selected record and compute
Vy,tot the ductility demands of the soil–structure system
Select a ground motion from three sets of records
Create the initial irregular structure by reducing stiffness
Perform nonlinear time history analysis for fixed-base and strength parameters at the first or roof story
reference structures under the given earthquake (apply λ factor)

Calculate stories ductility ratio μi Scale the total stiffness and strength of the irregular
structure (by the coefficients ηk and ηs )

No | μma x -μt |≤ Yes Analyze the final irregular structure under the selected
Tolerance earthquake ground motion considering SSI.

Fig. 5 Flowchart of the general procedure for evaluation ductility demands of flexible-base MDOF buildings

Fig. 6 Effect of irregularity ratio on height-wise distribution of averaged ductility demand for 10-story structures with soft first story and
T1 = 1 s

the seismic demand of the modified story. As the irregu- generalized as long as the level of increase in ductility does
larity ratio and nonlinearity of a system increase, this not lead to collapse. Nonlinear SSI also leads to a more
positive effect is amplified in the modified story of the uniform distribution of ductility demand along the height
irregular structure. In other words, SSI effects diminish of the building.
structural ductility demand in a localized manner when the As previously stated, ASCE 7-10 [39] stipulates stiff-
level of inelasticity increases. This finding can be ness and strength regularity limits of 0.7 and 0.8,

123
1218 International Journal of Civil Engineering (2020) 18:1209–1228

respectively. Thus, based on the obtained results, a weak the case of irregularity at the first story, the ductility
story limit (k = 0.8) imposed on the first story in fixed-base demands of the upper stories decrease, as with the obser-
and flexible-base conditions corresponded to increases in vations discussed in the previous section. Nevertheless, a
ductility demands of the modified story up to 78% and weak top story exerts negligible effects on the seismic
36%, respectively. response of the other stories. SSI can reduce the ductility
demand of the irregular structures, especially in the mod-
6.2 Effect of Irregularity Position ified story, regardless of whether the irregularity is at the
first story or roof level. As a result, the ductility demand
To evaluate the effects of irregularity position on the distribution along the height of the soil–structure systems
height-wise distribution of ductility demand, 10-story shear becomes more uniform than the distribution in the corre-
building models exposed to earthquakes recorded on site sponding fixed-base model.
class D with and without SSI effects are considered. The
results correspond to three target ductility ratios (lt = 1, 2, 6.3 Effect of Fundamental Period
and 4), k = 0.8, T1 = 1 s, FSv = 3, and aspect ratio = 3
(Fig. 7). Two irregularities at the first story and roof level, Figure 8 depicts the effects of the fundamental period on
as well as the corresponding regular model, are examined. the averaged ductility demand of the 10-story vertically
As seen, the irregularity at the roof increases the demand of irregular systems. The conditions imposed on both fixed-
this story, whereas the responses of the other stories are the base and flexible-base structures are three target ductility
same as those observed in the regular cases. However, the demands (lt = 1, 2, and 6), four fundamental periods (0.5,
ductility demand distribution in this case is more non- 1, 1.5, and 2.5 s), k = 0.9 at the base, FSv = 3, and s = 3. It
uniform than for irregularity at the first story. Similar to the is observed that increasing the fundamental period results
case of irregularity at the first story, an increase in the level in a more non-uniform distribution of ductility demand
of inelasticity causes collapse mechanism at the modified along the height of the structures. The ductility demand of
story. Moreover, the influence of irregularity on ductility the soft first story owing to foundation uplift visibly
demands at stories further away from the irregular story decreases at T1 = 0.5 s. The reduction in such demand due
significantly depends on the location of the irregularity. In to soil flexibility is more prominent in the structures with

Fig. 7 Effect of irregularity position on height-wise distribution of averaged ductility demand for 10-story structures with T1 = 1 s

123
International Journal of Civil Engineering (2020) 18:1209–1228 1219

Fig. 8 Effect of fundamental period on height-wise distribution of averaged ductility demand for 10-story structures with k = 0.9 at the first story

short periods. This finding is attributed to the fact that a on the reduction in ductility demand for the 5-story soil–
soft superstructure and a high target ductility ratio acquire structure systems compared with the other structures. As
low SSI effects. In the elastic state, an almost uniform the number of stories increases, the height-wise distribution
distribution of ductility demand exhibits in the structures of ductility demand becomes more non-uniform, demon-
with short periods. By increasing the level of inelasticity, strating higher mode effects. These outcomes are more
ductility demand distribution for all ranges of periods pronounced for highly nonlinear structures. Note that the
becomes more non-uniform. Ductility demand distribution maximum ductility demand generally occurs at the top
does not significantly depend on periods for structures that story.
experience large inelastic deformations. The ductility
demand concentration at the first story due to vertical 6.5 Effect of Site Class
irregularities diminishes in significance for the long-period
structures. The influence of different site classes on the ductility
demand distribution along the height of fixed-base and
6.4 Effect of Number of Stories flexible-base structures is presented in Fig. 10. The results
are for the 10-story irregular buildings with three target
To compare the responses of the 5-, 10- and 15-story ductility ratios (lt = 1, 2, and 6), k = 0.9, T1 = 1 s, FSv-
structures, the story ductility demands of the fixed-base and = 3, and aspect ratio = 3. The averaged values for 20
flexible-base shear buildings are investigated. The results earthquakes corresponding to each site class are given. It
shown in Fig. 9 are for the structures with a fundamental can be seen that for the fixed-base structures, the ductility
period (T1 = 0.1n), three target ductility ratios (lt = 1, 2, demand is generally higher for site class E than for the
and 6), FSv = 3, and s = 3 under 20 ground motions other stiff soil profiles. By increasing inelastic behavior,
recorded on site class D. The dimensionless frequency is this trend is intensified at the first story but is less pro-
kept constant for different structures, and only the effects nounced at the upper stories. Therefore, the site effects can
of the number of stories are observed. As seen, SSI effects be evaluated in the fixed-base conditions. On the basis of
reduce the ductility demands of the shear buildings, espe- the site and SSI effects, the reduction in the ductility
cially those with low ductility ratios. Another important demand of the soil–structure systems is typically more
point is that soil flexibility exerts a more significant effect noticeable in soft soil conditions. Accordingly, despite the

123
1220 International Journal of Civil Engineering (2020) 18:1209–1228

Fig. 9 Effect of number of stories on height-wise distribution of averaged ductility demand for regular systems with T1 = 0.1n

Fig. 10 Effect of site class on height-wise distribution of averaged ductility demand for 10-story structures with T1 = 1 s

123
International Journal of Civil Engineering (2020) 18:1209–1228 1221

Fig. 11 Effect of aspect ratio and FSv on height-wise distribution of averaged ductility demand for 10-story structures with T1 = 1 s

greater ductility demand in the fixed-base structures loca- 7 Evaluation of Ductility Demand
ted on site class E, the SSI effects in this case yield the least Distribution Using Coefficient of Variation
ductility demand. Such demand for both fixed-base and
flexible-base buildings may be considered insensitive to For an improved assessment of the ductility demand dis-
soil type as inelasticity level increases. tribution along building height, the coefficient of variation
(COV) can be used, which is defined as the ratio of the
6.6 Effects of Factor of Safety and Aspect Ratio standard deviation of story ductility demand to the mean
ductility level of all stories. In this regard, the lower the
Figure 11 compares the ductility demand distributions of COV of story ductility demands, the greater the uniformity
the soil–structure systems for three aspect ratios and vari- in ductility demand distribution and the better the perfor-
ous combinations of FSv and target ductility ratios. The mance of a structure. The influence of irregularity ratios on
results are averaged values for the 10-story structures with the COV of story ductility demands in fixed-base and
k = 0.9, T1 = 1 s, and subjected to 20 ground motions flexible-base structures is depicted in Fig. 12. Mean COV
recorded on site class E. It is shown that the positive values are derived for the 10-story models with s = 3,
influence of SSI is more prominent as the aspect ratio FSv = 3, and three ductility ratios (1, 2, and 6) as well as 20
increases (i.e., slender building). The trend is less signifi- ground motions recorded on site class E. The abscissa in
cant in the structures suffering from large inelastic defor- Fig. 12 is the fundamental period of the fixed-base struc-
mations. This observation is compatible with the results of tures and the vertical axis is the averaged COV of story
Ganjavi and Hao [21]. As the nonlinearity of the system ductility demands. For both the fixed-base and flexible-
increases, further reduction in ductility demand occurs at base structures, the COV generally increases as structural
the first story. Moreover, the ductility demands of the soil– irregularity intensifies, especially for the buildings with
structure systems with low FSv are less than those of the short periods within the elastic range. In the inelastic
buildings with high FSv. This result can be ascribed to the range, two distinct segments are separated at the threshold
fact that decreasing FSv softens soil–foundation–structure period, which is related to the target ductility ratio. In the
system. first segment at a period of nearly less than 1.5 s, the

123
1222 International Journal of Civil Engineering (2020) 18:1209–1228

Fig. 12 Effect of irregularity ratio on averaged COV of story ductility demands for 10-story structures with s = 3

ductility demand distribution becomes non-uniform with PGA of each record. The horizontal axis is the fixed-base
increasing irregularity ratio. The results show a high COV fundamental period of a structure. In this case, the struc-
for the irregular structures with lt = 6 given the collapse of ture-to-soil stiffness ratio is kept constant for all periods
the first story, as described in Sect. 6.1. For periods longer and is defined as:
than this characteristic period, the COV of the story duc- xfix H 
tility demands in the irregular structures is lower than that a0 ¼ ð10Þ
Vs
of the demands in the regular ones. Moreover, the COV of
the story ductility demands for the long-period SSI systems where xfix is the natural frequency of the fixed-base
is greater than that of the structures characterized by short structure. It should be noted that the practical range of this
periods. On these bases, then, SSI effects more visibly parameter is from zero for the fixed-base structure to about
influence the decrease in COV for the short-period 2 for systems with dominant SSI effects. Accordingly, a0 is
structures. set at 0.75 and 1.5 for site classes D and E, respectively.
Figure 13 illustrates the results for the 10-story shear
buildings with fundamental periods ranging from 0.5 to
8 Effect of SSI on the Strength Demands 2.5 s; target ductility ratios of l, 2, and 6; FSv = 5; aspect
of Vertically Irregular Structures ratio = 3; and irregularity ratios of 0.9, 0.8, and 0.65, as
well as the corresponding regular model. As shown, non-
The effects of SSI on the elastic and inelastic strength linear SSI generally reduces the strength required to
demands of the MDOF structures are discussed in this acquire the target ductility ratio. However, the rate of this
section. Total shear strength is calculated in a way that the reduction is more remarkable for structures with elas-
maximum ductility demand among all stories reaches the tic behavior or low ductility ratios. This finding is consis-
desired target value for fixed-base and flexible-base tent with the results reported by Ghannad and Jahankhah
buildings. Figure 13 shows the strength demands of the [26]. In addition, strength demand decreases as the irreg-
regular and irregular 10-story structures under 20 ground ularity ratio increases, which is compatible with the results
motions recorded on site class D. The results are the presented in Sect. 6. It is also observed that the strength
average of the strength demands obtained under all ground required at each ductility level is inversely proportional to
motions and normalized by the total structural mass times the fundamental period of the structure.

123
International Journal of Civil Engineering (2020) 18:1209–1228 1223

Fig. 13 Effect of irregularity ratio on normalized strength demand for 10-story structures with s = 3 subjected to ground motions corresponding
to the site class D

Fig. 14 Normalized strength demand of 10-story structures for site class D and E with and without inclusion of SSI effect (solid line: Fixed-base,
dashed line: Flexible-base)

To investigate the effects of site class, the strength more apparent for the soil–structure systems with short
demands of the regular 10-story structures are determined periods and low ductility ratios. As a result, the strength
(Fig. 14), including the averaged strength demands of site reduction factor is lower in the soil–structure systems than
classes D and E. As seen, SSI exerts significant effects in in the fixed-base buildings. This finding indicates that using
site class E because of the greater a0 in this case. Therefore, the fixed-base strength reduction factor leads to an under-
the influence of SSI on the reduction in the strength estimation of the ductility demand of soil–structure sys-
demand is more noticeable in site class E. This tendency is tems and cannot generate the target ductility ratio.

123
1224 International Journal of Civil Engineering (2020) 18:1209–1228

9 Estimation of the Maximum Ductility corresponding to three target ductility demands and site
Demands of MDOF Soil–Structure Systems classes D and E. As seen, the expression can accurately
predict the ductility demands of the regular soil–structure
In this section, the simplified expressions are proposed for systems. The efficiency and accuracy of Eqs. (11) and (12)
estimating the maximum ductility demands of regular and in estimating the maximum ductility demands of irregular
irregular structures, with considering SSI effects. First, a structures with flexible base are also demonstrated in
parabolic function is developed to estimate the maximum Fig. 16.
ductility demand of the flexible-base regular systems
(lSSI
Reg ). Then, the ratio of the maximum ductility demand of
the flexible-base irregular structures (lSSI 10 Conclusion
Irg ) to that of the
regular ones (lSSI
Reg ) is regressed. The equations are given as: A comprehensive parametric study was performed to
lSSI
Reg ¼ 2
lt ðaTfix þ bTfix þ cÞ ð11Þ evaluate the effects of nonlinear SSI on the ductility and
strength demands of vertically irregular structures. For this
lSSI
Irg e purpose, a large number of soil–structure systems with a
¼ dTfix ð12Þ
lSSI
Reg wide range of key parameters were subjected to three sets
of ground motions corresponding to different soil classes.
where Tfix and lt are the fundamental period and target The results of this study are summarized as follows:
ductility ratio for fixed-base structures, respectively; a, b,
and c are the constant coefficients associated with influ- 1. The median ductility demand increases in the modified
encing parameters and can be obtained from Tables 3 and 4 story but decreases in the other stories due to the
for the regular buildings. Also, coefficients d and e in softness/weakness of the first story for both fixed-base
Eq. (12) are calculated from the following expressions: and flexible-base buildings. However, seismic demand
in the lower stories is insensitive to irregularity at the
d ¼ d1 þ d2 k þ d3 lt þ d4 s ð13Þ
roof level. It was shown that the code weak story limit
e ¼ e1 þ e2 k þ e3 lt þ e4 s ð14Þ (k = 0.8) imposed on the first story increases the
ductility demand of the modified story up to 78% and
The coefficients di and ei (i = 1, 2, 3, 4) in Eqs. (13) and
36% in fixed-base and flexible-base conditions,
(14) are found through a nonlinear regression analysis and
respectively.
are listed in Tables 5 and 6.
2. In general, SSI can considerably reduce the seismic
Figure 15 shows the comparison of the averaged duc-
demand of the regular and irregular structures, espe-
tility demand ratios from the numerical analyses with those
cially for systems with short periods. This beneficial
calculated using Eq. (11). The results are plotted for the
effect was further observed at the modified story with
10- and 15-story regular buildings with s = 3,
increasing the level of inelasticity. In other words, the

Table 3 Coefficients a, b and c to be used in Eq. (11) for 10-story structures


s=2 s=3 s=5
ai bi ci ai bi ci ai bi ci

Site class C
lt= 1 - 0.029 0.123 0.856 - 0.044 0.190 0.781 - 0.048 0.234 0.720
lt= 2 - 0.081 0.289 0.707 - 0.113 0.419 0.578 - 0.131 0.522 0.431
lt= 4, 6 - 0.125 0.43 0.61 - 0.121 0.42 0.61 - 0.130 0.467 0.553
Site class D
lt= 1 - 0.152 0.592 0.396 - 0.116 0.469 0.495 - 0.089 0.385 0.535
lt= 2 - 0.212 0.855 0.068 - 0.180 0.740 0.166 - 0.131 0.586 0.254
lt= 4, 6 - 0.248 1.019 - 0.070 - 0.232 0.917 0.093 - 0.151 0.705 0.146
Site class E
lt= 1 - 0.150 0.764 - 0.036 - 0.121 0.645 0.093 - 0.083 0.533 0.144
lt= 2 - 0.110 0.667 - 0.119 - 0.085 0.581 - 0.025 - 0.058 0.511 0.0
lt= 4, 6 - 0.078 0.607 - 0.108 - 0.127 0.743 - 0.145 - 0.090 0.652 - 0.132

123
International Journal of Civil Engineering (2020) 18:1209–1228 1225

Table 4 Coefficients a, b and c to be used in Eq. (11) for 15-story structures


s=3 s=5
2
ai bi ci R ai bi ci R2

Site class D
lt= 1 - 0.033 0.221 0.608 0.97 - 0.046 0.253 0.621 0.97
lt= 2 - 0.061 0.403 0.271 0.96 - 0.083 0.465 0.292 0.96
lt= 4, 6 - 0.091 0.548 0.135 0.94 - 0.116 0.618 0.142 0.90
Site class E
lt= 1 - 0.142 0.789 - 0.216 0.99 - 0.151 0.806 - 0.126 0.98
lt= 2 - 0.080 0.533 - 0.144 0.99 - 0.112 0.700 - 0.204 0.99
lt= 4, 6 - 0.078 0.626 - 0.348 0.97 - 0.145 0.909 - 0.454 0.98

Table 5 Coefficients di and ei in Eqs. (13) and (14) for 10-story nonlinear SSI effects indicated that the ductility
irregular structures demands of the soil–structure systems generally
i=1 i=2 i=3 i=4 decrease with the softening of soil.
4. For the fixed-base and flexible-base models, increasing
Site class C inelastic behavior leads to greater non-uniformity in
di 0.4293 0.5826 - 0.0248 - 0.0440 ductility demand distribution compared to the corre-
ei - 1.0114 0.9949 - 0.0728 0.1049 sponding elastic model. As expected, SSI effects
Site class D become less significant for highly nonlinear structures.
di 0.1973 0.8482 - 0.0489 - 0.0157 5. The ductility demands of the soil–structure systems
ei -0.4466 0.4831 - 0.0515 0.0279 with low FSv are generally smaller than those of the
Site class E buildings with high FSv. Soil flexibility and foundation
di 0.6201 0.5512 - 0.0208 - 0.0566 uplift exert more pronounced effects in the structures
ei - 1.1971 1.1514 - 0.1045 0.0688 with large aspect ratios (i.e., slender buildings).
6. For the structures with short periods, the averaged
COV of story ductility demands decreases as soil
flexibility increases, meaning a more uniform distribu-
Table 6 Coefficients di and ei in Eqs. (13) and (14) for 15-story tion of ductility demand than that found in the
irregular structures
corresponding fixed-base model. Increasing irregular-
i=1 i=2 i=3 i=4 ity ratio and inelasticity level (i.e., target ductility
Site class D
ratio) are normally accompanied by an increase in the
di 0.0020 1.0770 - 0.0497 - 0.0078
averaged COV of story ductility demands.
ei - 0.3009 0.3076 - 0.0232 0.0095
7. Nonlinear SSI reduces the strength required to achieve
the desired ductility ratio (elastic or inelastic). How-
Site class E
ever, this trend is less prominent as the inelasticity
di 0.4572 0.8078 - 0.0390 - 0.0667
level intensified. Therefore, SSI decreases the strength
ei - 0.8257 0.7483 - 0.0734 0.0610
reduction factor, and using the fixed-base strength
reduction factor leads to an underestimation of the
ductility demand of soil–structure systems. This find-
effects of SSI diminish structural ductility demand in a ing would be more critical for buildings located on soft
localized manner for the structures that undergo soil (e.g., site class E).
substantial inelastic deformations. Moreover, SSI 8. Simplified and efficient expressions were also provided
effects generate a more uniform distribution of ductil- to estimate the maximum ductility demands of regular
ity demand along building height in comparison with and irregular soil–structure systems through nonlinear
the corresponding fixed-base model. regression analyses.
3. The fixed-base structures experience greater ductility Note that the above-mentioned findings are based on a
demand under the ground motions recorded on site parametric study of shear-type MDOF buildings with
class E than in stiffer soil profiles. The site and coupled stiffness–strength irregularities located on shallow

123
1226 International Journal of Civil Engineering (2020) 18:1209–1228

n = 10, Data n = 10, Polynomial regression


n = 15, Data n = 15, Polynomial regression
1.5 2 4
μt =1 μt =2 μt =4
1.5 3

Site Class D
1
μ Reg
SSI

1 2
0.5
0.5 1

0 0 0
0 1 2 3 0 1 2 3 0 1 2 3
1.5 2 4
μt =1 μt =2 μt =4
1.5 3

Site Class E
1
μ Reg
SSI

1 2
0.5
0.5 1

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

Fig. 15 Comparison of the mean ductility demand ratios from numerical analyses with those calculated using Eq. (11) for regular soil–structure
systems with s = 3

n = 10, Data n = 10, Regressed


n = 15, Data n = 15, Regressed

3 3
s =3 s=5

Site Class D
2 2
SSI
µIrg

1 1

0 0
0 1 2 3 0 1 2 3
3 3
s =3 s=5
Site Class E

2 2
SSI
µ Irg

1 1

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

Fig. 16 Comparison of the mean ductility demand ratios from numerical analyses with those calculated using Eqs. (11) and (12) for irregular
buildings with k = 0.8 and lt = 4

123
International Journal of Civil Engineering (2020) 18:1209–1228 1227

foundations, which may not apply to other irregularity 145(9):04019087. https://doi.org/10.1061/(ASCE)ST.1943-541X.


types. In a general sense, the findings are limited by the 0002350
13. Giannakouras P, Zeris C (2019) Seismic performance of irregular
range of key parameters interrogated, such as superstruc- RC frames designed according to the DDBD approach. Eng
ture, soil, and input motion. Moreover, only inertial inter- Struct 182:427–445
action effects were considered in this paper, and kinematic 14. Tena-Colunga A, Hernández-Garcı́a DA (2020) Peak seismic
interaction was disregarded in the modeling. This demands on soft and weak stories models designed for required
code nominal strength. Soil Dyn Earthq Eng 129:105698. https://
assumption is commonly accepted for shallow foundations doi.org/10.1016/j.soildyn.2019.05.037
subjected to vertically propagating shear waves. Further 15. Wolf JP (1985) Dynamic soil–structure interaction. Prentice-Hall,
studies on RC frames should be performed to evaluate the Englewood Cliffs
following: (1) other types of vertical irregularities, (2) the 16. Bielak J (1978) Dynamic response of non-linear building–foun-
dation systems. Earthq Eng Struct Dyn 6(1):17–30. https://doi.
impact of material degradation, (3) structures with org/10.1002/eqe.4290060104
embedded foundations, and (4) the effects of pulse-like, 17. Müller FP, Keintzel E (1982) Ductility requirements for flexibly
near-fault ground motions. supported antiseismic structures. In: Proceedings of the seventh
european conference on earthquake engineering, Athens, Greece
18. Avilés J, Pérez-Rocha LE (2003) Soil–structure interaction in
yielding systems. Earthq Eng Struct Dyn 32(11):1749–1771.
https://doi.org/10.1002/eqe.300
References 19. Mahsuli M, Ghannad MA (2009) The effect of foundation
embedment on inelastic response of structures. Earthq Eng Struct
1. Valmundsson EV, Nau JM (1997) Seismic response of building Dyn 38(4):423–437. https://doi.org/10.1002/eqe.858
frames with vertical structural irregularities. J Struct Eng 20. Vatanchian M, Shooshtari A (2018) Investigation of soil–structure
123(1):30–41. https://doi.org/10.1061/(ASCE)0733-9445(1997) interaction effects on seismic response of a 5 MW wind turbine. Int J
123:1(30) Civil Eng 16(1):1–17. https://doi.org/10.1007/s40999-016-0059-5
2. Al-Ali AAK, Krawinkler H (1998) Effects of vertical irregularity 21. Ganjavi B, Hao H (2012) A parametric study on the evaluation of
on the seismic behaviour of building structures. Report No. 130, ductility demand distribution in multi-degree-of-freedom systems
John A. Blume Earthquake Engineering Center, Stanford considering soil–structure interaction effects. Eng Struct
University, Stanford 43:88–104. https://doi.org/10.1016/j.engstruct.2012.05.006
3. Chintanapakdee C, Chopra AK (2004) Seismic response of ver- 22. Khoshnoudian F, Attarnejad R, Paytam F, Ahmadi E (2015)
tically irregular frames: response history and modal pushover Effects of forward directivity on the response of soil–structure
analyses. J Struct Eng 130(8):1177–1185. https://doi.org/10.1061/ systems. Proc Inst Civil Eng Struct Build 168(9):664–679. https://
(ASCE)0733-9445(2004)130:8(1177) doi.org/10.1680/jstbu.13.00076
4. Fragiadakis M, Vamvatsikos D, Papadrakakis M (2006) Evalua- 23. Tomeo R, Pitilakis D, Bilotta A, Nigro E (2018) SSI effects on
tion of the influence of vertical irregularities on the seismic seismic demand of reinforced concrete moment resisting frames.
performance of a nine-storey steel frame. Earthq Eng Struct Dyn Eng Struct 173:559–572
35(12):1489–1509. https://doi.org/10.1002/eqe.591 24. Ganjavi B, Gholamrezatabar A, Hajirasouliha I (2019) Effects of
5. Habibi A, Asadi K (2017) Development of drift-based damage soil–structure interaction and lateral design load pattern on per-
index for reinforced concrete moment resisting frames with set- formance-based plastic design of steel moment resisting frames.
back. Int J Civil Eng 15(4):487–498. https://doi.org/10.1007/ Struct Design Tall Spec Build 28(11):e1624. https://doi.org/10.
s40999-016-0085-3 1002/tal.1624
6. Sadashiva VK, MacRae GA, Deam BL (2012) Seismic response 25. Tehranizadeh M, Barkhordari MS (2018) Effect of peripheral
of structures with coupled vertical stiffness–strength irregulari- wall openings in basement and number of basement floors on the
ties. Earthq Eng Struct Dyn 41(1):119–138. https://doi.org/10. base level of braced framed tube system. Int J Civil Eng
1002/eqe.1121 16(9):1157–1173. https://doi.org/10.1007/s40999-017-0270-z
7. Zhou J, Bu G, Wang H, Cai J (2013) Modification of ductility 26. Ghannad MA, Jahankhah H (2007) Site-dependent strength
reduction factor for vertically irregular structures subjected to reduction factors for soil–structure systems. Soil Dyn Earthq Eng
pulse-like ground motions. Adv Struct Eng 16(4):641–652. 27(2):99–110
https://doi.org/10.1260/1369-4332.16.4.641 27. Ganjavi B, Hao H (2014) Strength reduction factor for MDOF
8. CEN (2004) Eurocode 8: design of structures for earthquake soil-structure systems. Struct Design Tall Spec Build
resistance, part 1, general rules, seismic actions and rules for 23(3):161–180. https://doi.org/10.1002/tal.1022
buildings. European Committee for Standardizations, Brussels 28. Lu Y, Hajirasouliha I, Marshall AM (2016) Performance-based
9. Mondal G, Tesfamariam S (2014) Effects of vertical irregularity seismic design of flexible-base multi-storey buildings considering
and thickness of unreinforced masonry infill on the robustness of soil–structure interaction. Eng Struct 108:90–103. https://doi.org/
RC framed buildings. Earthq Eng Struct Dyn 43(2):205–223. 10.1016/j.engstruct.2015.11.031
https://doi.org/10.1002/eqe.2338 29. Yim C-S, Chopra AK (1984) Earthquake response of structures
10. Zhou J, Zhao W, Mao W (2015) Least favorable probability of with partial uplift on winkler foundation. Earthquake Eng Struct
failure for 5- and 10-story RC frame structures with vertical Dyn 12(2):263–281. https://doi.org/10.1002/eqe.4290120209
irregularities. J Earthquake Eng 19(7):1158–1180. https://doi.org/ 30. Ghannad MA, Jafarieh AH (2014) Inelastic displacement ratios
10.1080/13632469.2015.1025315 for soil–structure systems allowed to uplift. Earthquake Eng
11. Zeris C (2015) Seismic response of rocking oscillators on a soft Struct Dyn 43(9):1401–1421. https://doi.org/10.1002/eqe.2405
story: elastic response. J Struct Eng 141(8):04014196. https://doi. 31. Raychowdhury P (2011) Seismic response of low-rise steel
org/10.1061/(ASCE)ST.1943-541X.0001157 moment-resisting frame (SMRF) buildings incorporating nonlin-
12. Zeris C, Scodeggio A (2019) Seismic performance of rocking ear soil–structure interaction (SSI). Eng Struct 33(3):958–967.
oscillators on a soft story: inelastic response. J Struct Eng https://doi.org/10.1016/j.engstruct.2010.12.017

123
1228 International Journal of Civil Engineering (2020) 18:1209–1228

32. Khanmohammadi M, Mohsenzadeh V (2018) Effects of foun- 41. Raychowdhury P, Hutchinson TC (2009) Performance evaluation
dation rocking and uplifting on displacement amplification factor. of a nonlinear Winkler-based shallow foundation model using
Earthquake Eng Eng Vib 17(3):511–525. https://doi.org/10.1007/ centrifuge test results. Earthquake Eng Struct Dyn
s11803-018-0459-4 38(5):679–698. https://doi.org/10.1002/eqe.902
33. Anastasopoulos I, Gelagoti F, Spyridaki A, Sideri J, Gazetas G 42. Gajan S, Raychowdhury P, Hutchinson TC, Kutter BL, Stewart
(2014) Seismic rocking isolation of an asymmetric frame on JP (2010) Application and validation of practical tools for non-
spread footings. J Geotech Geoenviron Eng 140(1):133–151. linear soil–foundation interaction analysis. Earthquake Spectra
https://doi.org/10.1061/(ASCE)GT.1943-5606.0001012 26(1):111–129. https://doi.org/10.1193/1.3263242
34. Shakib H, Homaei F (2017) Probabilistic seismic performance 43. Gazetas G (1991) Formulas and charts for impedances of surface
assessment of the soil–structure interaction effect on seismic and embedded foundations. J Geotech Eng 117(9):1363–1381.
response of mid-rise setback steel buildings. Bull Earthq Eng https://doi.org/10.1061/(ASCE)0733-9410(1991)117:9(1363)
15(7):2827–2851. https://doi.org/10.1007/s10518-017-0087-9 44. Boulanger RW, Curras CJ, Kutter BL, Wilson DW, Abghari A
35. Homaei F, Shakib H, Soltani M (2017) Probabilistic seismic (1999) Seismic soil–pile–structure interaction experiments and
performance evaluation of vertically irregular steel building analyses. J Geotech Geoenviron Eng 125(9):750–759. https://doi.
considering soil–structure interaction. Int J Civil Eng org/10.1061/(ASCE)1090-0241(1999)125:9(750)
15(4):611–625. https://doi.org/10.1007/s40999-017-0165-z 45. Gajan S, Hutchinson TC, Kutter BL, Raychowdhury P, Ugalde
36. Asadi-Ghoozhdi H, Attarnejad R (2018) Evaluation of ductility JA, Stewart JP (2008) Numerical models for analysis and per-
demand distribution in vertically irregular frames allowed to formance-based design of shallow foundations subjected to
uplift. In: 16th european conference on earthquake engineering seismic loading. Report No. 2007/04, Pacific Earthquake Engi-
(16ECEE), Thessaloniki, Greece neering Research Center, University of California, Berkeley
37. Hajirasouliha I, Doostan A (2010) A simplified model for seismic 46. FEMA (2005) Improvement of nonlinear static seismic analysis
response prediction of concentrically braced frames. Adv Eng procedures. Report No. FEMA 440, Federal Emergency Man-
Softw 41(3):497–505. https://doi.org/10.1016/j.advengsoft.2009. agement Agency, prepared by Applied Technology Council,
10.008 Washington, DC
38. BHRC (2014) Iranian code of practice for seismic resistant design 47. PEER (2019) Ground motion database. Available: https://nga
of buildings (Standard 2800), 4th edn. Building and Housing west2.berkeley.edu/
Research Center, Tehran 48. ATC (1978) Tentative provisions for the development of seismic
39. ASCE (2010) Minimum design loads for buildings and other regulations for buildings. ATC 3-06, Applied Technology
structures: ASCE/SEI 7-10. American Society of Civil Engineers, Council, California
Reston 49. Mazzoni S, McKenna F, Scott M, Fenves G (2009) Open system
40. Harden CW, Hutchinson T, Martin GR, Kutter BL (2005) for earthquake engineering simulation (opensees) user command-
Numerical modeling of the nonlinear cyclic response of shallow language manual. Pacific Earthquake Engineering Research
foundations. Report No. 2005/04, Pacific Earthquake Engineering Center, University of California, Berkeley
Research Center (PEER), University of California, Berkeley

123

You might also like