Sie sind auf Seite 1von 25

Bulletin of Earthquake Engineering

https://doi.org/10.1007/s10518-019-00570-y

ORIGINAL RESEARCH

Nonlinear dynamic response of a tall building to near‑fault


pulse‑like ground motions

Necmettin Güneş1 · Zülfü Çınar Ulucan2

Received: 29 May 2018 / Accepted: 29 January 2019


© Springer Nature B.V. 2019

Abstract
The objective of this study is to determine the effects of near-fault ground motions with
large, medium and small pulse periods on a core-wall tall building. Three near-fault and
one far-field ground motion sets, each with 11 different records, are rotated to the max-
imum pseudo-velocity spectrum directions. The mean transition periods of each set are
obtained by the smoothed tripartite spectrum, which illustrates that the acceleration–sensi-
tive region extends with increasing pulse duration. A 40-story core-wall building is sub-
jected to 44 different record pairs, and the analysis results show that the ratio of pulse dura-
tion to first mode period (­Tp/T1), governs the nonlinear response. It is demonstrated that
higher mode effects increase while the first mode sensitive region moves away from the
second mode sensitive region. The distribution profiles of the story drift, tension strain, and
beam rotation demands along the height of the building are substantially different depend-
ing on the ­Tp/T1 ratio. In the large and medium pulses records, the core-wall reaches the
flexural yielding capacity above the podium level, and this phenomenon increases the story
drift ratio and flexural beam rotation demands depending on the displacement demands at
the end of the acceleration-sensitive region. On the contrary, near-fault with small pulses
and far-field ground motions induce yielding of core-wall at upper stories due to higher
mode effects. Thus, post-yield story shear force distributions significantly change com-
pared to the large and medium pulses ground motions.

Keywords  Near-fault ground motions · Pulse period · Nonlinear dynamic analysis · Higher
mode effects · Tall buildings

* Necmettin Güneş
ngunes@cumhuriyet.edu.tr
Zülfü Çınar Ulucan
zculucan@firat.edu.tr
1
Department of Architecture, Cumhuriyet University, 58140 Sivas, Turkey
2
Department of Civil Engineering, Fırat University, 23000 Elazığ, Turkey

13
Vol.:(0123456789)
Bulletin of Earthquake Engineering

1 Introduction

The performance-based design has been proposed in tall building guidelines as an alterna-
tive to code-based elastic design approaches to determine the structural demand parameters
with acceptable accuracy (PEER/TBI2010; LATBSDC 2017; TEC 2018). The require-
ments in these guidelines generally focus on three topics, that is, limiting the use of the
elastic method to the Service Level Earthquakes (SLE) shaking and the preliminary design
phase, suggesting to incorporate the realistic simulation of structural members behavior
into mathematical model, and checking the performance goals in the levels of SLE and
Maximum Considered Earthquake (MCE).
Response spectrum analysis (RSA) has been widely used to design tall buildings due to
its simple application properties and software tool varieties. However, many studies have
shown that the RSA cannot capture the shear force demands of the structural walls under
the higher mode effects with acceptable accuracy (Vuran and Aydınoğlu 2016; Dezhdar
and Adebar 2012; Mehmood et  al. 2017; Rejec et  al. 2013; Rutenberg 2013; Kazaz and
Gülkan 2016). Moreover, Rad and Adebar (2008) compared the results of RSA and Lin-
ear Time History analyses (LTHA) for an 84  m high cantilever wall, and they showed
that envelop moments obtained by these two methods were about 20% different along the
height of the wall. Therefore, many seismic design codes take shear magnification factors
into account since the wall shear forces cannot be determined realistically with response
spectrum analysis (CEN 2004; TEC 2018; NZS 2006). For the preliminary design to be
more effective, it is recommended that the response spectrum analysis results should be
amplified by the shear force magnification factor (Vuran and Aydınoğlu 2016; Dezhdar and
Adebar 2012).
The incorporation of cyclic deterioration of Reinforced Concrete (RC) structural mem-
bers into nonlinear dynamic analysis models with acceptable precision has been recom-
mended (PEER/ATC 2010; NIST 2017). Haselton et al. (2008) proposed a simple proce-
dure to modify the monotonic loading curve of RC beams and columns to capture cyclic
envelope curve by using 255 test results. Thomsen and Wallace (2004) tested the RW2
shear wall specimen, following, Orakçal and Wallace (2006) processed the test result to
discard the shear deformations and calibrated the constitutive parameters. Naish et  al.
(2009) carried out the testing of eight coupling beams. Modeling capabilities of softwares
are also crucial to obtain accurate results and to simulate test results in the analysis model
(Görgülü and Taskin 2015; Carvalho et al. 2013; Basone et al. 2017). It is shown that Per-
form-3D software is capable of simulating the tested behavior of shear walls and coupling
beams (PEER/ATC 2010; Wallace et al. 2009).
Rupture mechanism affects the ground motions recorded in near-fault, and the dislo-
cation of rupture toward a record station with a velocity similar to shear wave leads to a
single distinct pulse at the beginning of the record in fault-normal direction. Sometimes
this pulse could be observed in fault-parallel direction due to the permanent displacement
of rupture (Somerville et al. 1997; Bray and Rodriguez-Marek 2004). The amplitude and
duration of this pulse increase the spectral displacement and interstory drift demands.
The ratio of pulse duration to first mode period (­Tp/T1) is a critical parameter to spec-
ify the structural response of pulse-like ground motions (Alavi and Krawinkler 2004;
Baker and Cornell 2006; Akkar et al. 2005; Champion and Liel 2012). While the records
with ­Tp < T1 create the higher mode effects and maximum response at upper stories, in
the records with T­ p > T1, the first mode governs the response, and the maximum demands
occur at the lower stories (Baker and Cornell 2006; Champion and Liel 2012; Alavi and

13
Bulletin of Earthquake Engineering

Krawinkler 2001, 2004). When the T ­ p ≈ T1, the maximum elastic response demands are
obtained (Akkar et al.2005). However, with nonlinear behavior, the periods elongate and
response demands decrease.
The near-fault ground motions have wider acceleration and narrower velocity-sensitive
regions than ordinary ones (Chopra and Chintanapakdee 2001). Malhotra (1999) studied
the tripartite spectra of three near-fault and one far-field ground motions and demonstrated,
by an elastic shear beam model, that wide acceleration sensitive region decreased higher
mode contribution to first story drift. Chopra and Chintanapakdee (2001) showed that
when the periods were normalized to the acceleration-sensitive region transition period
­(Tc), the response of sdof systems became similar for both near-fault and far-field ground
motions.
In the present paper, a 40-story tall building was subjected to two types of strong motion
records; a) near-fault ground motion with varying pulse duration and b) ordinary far-field
records. The comparison results for story drift ratios, wall shear stresses, the axial strains
of core-wall critical outer edges, overturning moments, flexural and coupling beam rota-
tions, etc. were given. It was shown that the effects of the T ­ p/T1 ratios on these structural
demand parameters could be classified by smoothed tripartite spectra which were good
indicators to project the higher mode effects and first-mode dominant response. In practice,
it is prevalent to reduce the thickness of the shear wall at upper stories. The present study
revealed that when the shear wall thicknesses were reduced at the upper stories, due to
the higher mode effects, core-wall may have reached elastic capacity at those levels firstly.
Thus, the post-yield redistribution of story shear forces significantly changed below the
yielding level.

2 Model description and elastic design

The 40-story RC core-wall model, including four parking stories below ground, one
podium and mechanical stories and 34 residential stories, had a height of 141.5 m. Three-
dimensional view and typical plan layouts below and above ground are given in Figs.  1
and 2. Basement walls thickness was 0.4  m, and all slab thicknesses were 0.2  m, except
ground floor at which slab thickness was taken as 0.3 m to provide the backstay diaphragm
effects. All coupling beams were 2.5 m in length, and the other model properties are given
in Table 1. Gravity loads are presented in Table 2. The elastic modulus of concrete (­ Ec) was
obtained as given in ACI 363R-92 and expected elastic modulus (­ Eexp) was taken as 1.14 E ­c
for each concrete class. In the Design Base Earthquake (DBE) and Maximum Considered
Earthquake (MCE) levels, different stiffness modifiers were assumed as shown in Table 3.
The elastic modeling and analysis were carried out using Etabs software with the DBE
level stiffness assumption in Table  3 (CSI 2016a). The first three vibration mode peri-
ods and their participation mass ratios (PM) in X, Y, and torsional directions are given in
Table 4 for DBE and MCE levels, separately. In the MCE level model, the expected elastic
moduli were used. Therefore the MCE level mode periods were obtained smaller than DBE
level ones. The geometric means of short and long period spectral acceleration for DBE
(475 years return period) level were 1.8 g and 0.72 g, and ones for MCE (2475 years return
period) level were 2.7 g and 1.08 g. The elastic design was conducted in accordance with
the Turkish Earthquake Code (TEC 2018), and the 5% damped response spectra are given
in Fig.  3. Modal responses were combined by using Complete Quadratic Combination

13
Bulletin of Earthquake Engineering

Fig. 1  Three-dimensional view
of Etabs model

(CQC) method. According to TEC, the minimum base shear ratio was 0.056. However,
design base shear ratios were taken as 0.09 and 0.072 in X and Y directions. The design
story shear forces and drift ratios for DBE level are given in Fig. 4.

3 Ground motions data sets and rotation

The studies on directivity effects of ground motions show that the maximum direction is
oriented toward the fault-normal direction in the areas within faults distance closer than
5 km, whereas that is random in the farther regions (Huang et al. 2008; Stewart et al. 2011;
Watson-Lamprey and Boore 2007; Beyer and Bommer 2006; Shahi and Baker 2014a).
Therefore, ASCE 7-16 states that the ground motions recorded in a site closer than 5 km
from faults must be rotated to fault-normal and parallel directions. Nevertheless, the rota-
tion of each ground motion to fault-normal and parallel directions is difficult and time-con-
suming to achieve due to the complexity of geographical fault position, and to overcome
this difficulty, maximum velocity direction is proposed and used in attenuation relations
(Akkar and Gülkan 2003). Pseudo-velocity spectrum (PSV) is also a good indicator due to
the including clues about the pulse period and structural demand (Akkar and Gülkan 2003;
Rupakhety et al. 2011).
In the present study, the maximum PSV direction was used instead of fault-normal direc-
tion. In order to show the effectiveness of maximum PSV direction, Kocaeli (Yarımca),
Erzincan (Erzincan) and Kobe (Kjm) records were selected. Their polar diagrams of nor-
malized maximum PSV values and the velocity histories of maximum PSV directions are
given in Figs.  5 and 6, respectively. The maximum PSV direction was aligned with the

13
Bulletin of Earthquake Engineering

Fig. 2  a Typical plan for ground and below. b Typical plan for above ground

13
Bulletin of Earthquake Engineering

Table 1  General model
properties Story heights 3.5 m parking floors
4.5 m ground floor
3.5 m above ground floor
Column dimensions 1.1 × 1.1 m from 1st to 13th
1.0 × 1.0 m from 14th to 26th
0.90 × 0.90 m from 27th to 39th
0.55 × 0.55 m other parking area columns
Beam dimensions 0.55 × 0.80 m ground floor and below
0.80 × 1.0 m from 5th to13th
0.65 × 1.0 m from 14th to 26th
0.50 × 1.0 m from 27th to 39th
Coupling beams 0.80 × 1.2 m from 1st to 13th
0.65 × 1.2 m from 14th to 26th
0.50 × 1.2 m from 27th to 39th
Slab 0.30 m ground floor
0.20 m all other floors
Core wall thicknesses 0.80 m from 1st to 13th
0.65 m from 14th to 26th
0.50 m from 27th to 40th
Specified concrete strength 60 Mpa from 1st to 13th
50 Mpa from 14th to 26th
40 Mpa from 27th to 40th

Table 2  Gravity loads Superimposed dead load Live load


(kN/m2) (kN/m2)

Residential 1.5 2.0


Exit areas 1.5 3.5
Roof 1.5 3.5
Garage 1.5 5.0

fault-normal (FN) direction in the Erzincan and Kjm records both of which were influ-
enced by forward-directivity. However, in Yarımca record due to fling step, a pulse was
observed in the velocity records of fault-parallel (FP) direction. In both cases, maximum
PSV directions were consistent with the pulse contained components.
Table  5 shows the 44 ground motions compiled to represent near-fault and far-field
ground motions from PEER database (PEER 2017). For near-fault motions, records within
the 20 km distance to rupture plane were selected. To obtain different ­Tp/T1 ratios, these
records were grouped by duration of pulses as large, medium and small. In the first set, T­ p/
T1 ratios were bigger than 1 while in the second and third sets, these ratios were around
and below 1, respectively. The pulse period was obtained as the pulse duration in the veloc-
ity time history. The last set contained ordinary far-field ground motions whose recorded
distances ranged from 39 to 82  km. The present study is concerned with the effects of
the pulse durations on nonlinear behavior of a tall building. In order to obtain a sufficient
number of near-fault records with desired pulse duration, ground motions were necessarily
selected from different site classes.

13
Bulletin of Earthquake Engineering

Table 3  Stiffness assumptions Effective stiffness DBE level MCE level


(Ghodsi and Ruiz 2010; TEC
2018; LATBSDC 2017)
Core walls Flexural: 0.5Ec ­Ig Flexural: ­EexpIg*
Shear: 0.4EcAg Shear: 0.2EexpAg
Columns Flexural: 0.7EcIg Flexural: 0.7EexpIg
Shear: 0.4EcAg Shear: 0.4EexpAg
Beams Flexural: 0.35EcIg Flexural: 0.35EexpIg
Shear: 0.4EcAg Shear: 0.4EexpAg
Coupling beams Flexural: 0.2EcIg Flexural: 0.2EexpIg
Shear: 0.4EcAg Shear: 0.4EexpAg
Basement and ground Flexural:0.25EcIg Flexural: 0.25EexpIg
level slabs
Shear: 0.2EcAg Shear: 0.1EexpAg
Basement walls Flexural: 0.8EcIg Flexural: 0.8EexpIg
Shear: 0.2EcAg Shear: 0.2EexpAg

*Fiber model, E ­ exp was determined from confined Mander model


(Değer et al. 2014)

Components of each ground motion were rotated 180 degrees with one-degree incre-
ments, and PSV was obtained for each direction. The peak value of each velocity spectrum
was recorded to represent its own direction. Thus, direction-dependent PSV values were
determined as given in Fig. 5. The component in maximum PSV direction was taken as X,
and then Y component was defined as normal to maximum PSV direction. The resultant
spectrum of each record was scaled to the MCE level maximum direction target spectrum
with Mean Square Error (MSE) scaling procedure, in which the square of the logarithmic
difference of the target and recorded spectral acceleration was minimized in the 0.2T1 and
1.5T1 period range by considering adequate logarithmically equal spaced interval points as
given by Baker (2010) and PEER (2010).
Duration of the pulse in the near-fault records affects the spectral acceleration and dis-
placement demands (Alavi and Krawinkler 2004). Figure  7 shows the MCE level maxi-
mum direction target spectrum and scaled mean resultant spectra with scaling period range
(0.53 to 4.34  s). As illustrated in Fig.  7, in periods smaller than 1  s, the near-fault with
small pulses and far-field ground motions had significantly higher spectral acceleration
demands in comparison with other record sets. On the other hand, the near-fault ground
motions with the medium and large pulses had higher spectral acceleration demands in the
1–3 s period range and periods longer than 3 s, respectively.

4 Nonlinear modeling properties

The three-dimensional nonlinear modeling was carried out by Perform-3D software (CSI
2016b). Seismic mass of the stories was assumed to be lumped at the center of mass above
ground floors and distributed at the ground floor and below. Slabs were included in the
mathematical model with their reduced stiffness properties at the ground floor and below,
whereas they were modeled as rigid diaphragms above the ground level. Rayleigh damping
ratio was taken as 2.5% at the 0.2T1 and 1.5T1 periods, P-delta effect was considered, and

13

13
Table 4  Modal periods and participation mass ratios
Mode number DBE level MCE level

X direction Y direction Torsional direction X direction Y direction Torsional direction


TX (s) PM (%) TY (s) PM (%) TΦ (s) PM (%) TX (s) PM (%) TY (s) PM (%) TΦ (s) PM (%)

1 2.89 49.1 3.38 51.7 2.0 29.1 2.64 48.2 2.89 50.0 1.63 30.2
2 0.74 15.7 1.06 12.6 0.68 4.97 0.67 14.8 0.87 12.8 0.57 4.90
3 0.34 7.0 0.58 4.80 0.38 2.33 0.30 6.60 0.46 4.70 0.33 2.30
Bulletin of Earthquake Engineering
Bulletin of Earthquake Engineering

Fig. 3  DBE and MCE level geo- 3


metric mean response spectra DBE (475 years return period)

Spectral acceleration (g)


2.5
MCE (2475 years return period)
2

1.5

0.5

0
0 1 2 3 4 5 6
Period (sec)

40 40
X direction X direction
35 35
Y direction Y direction
30 30
25 25
Story level
Story level

20 20
15 15
10 10
5 5
0 0
0 20000 40000 60000 80000 0 0.001 0.002 0.003
Base shear (kN) Drift ratio
(a) (b)
Fig. 4  Design level a story shear forces b story drift ratios

the foundation was assumed to be rigid. Expected material strengths were used for concrete
and steel materials as f­ cc = 1.3fc and ­fse = 1.2fsy.

4.1 Core wall modeling

The core shear wall elements are illustrated in Fig.  8, and their longitudinal reinforce-
ment to cross-section area ratios (ρ) are given in Table  6. Core walls were modeled by
Perform-3D shear wall element, in which bending/axial behavior is modeled by vertical
nonlinear fiber elements. In each story, 110 concrete and reinforcement steel fiber ele-
ments were used in the mathematical model of the core-wall. Furthermore, in order to cap-
ture the inelastic behavior of boundary elements, the length of concrete fibers was grad-
ually decreased from web to boundary, and the outermost concrete fiber element length
was taken as 0.4 m for each wall element. In order to permit the plastic hinges formation
above the podium level, the height of fiber elements was taken as half of the corresponding
wall length. Mander’s confined concrete model was used to acquire concrete stress–strain
relationships (Mander et  al. 1988). The elastic moduli used in the nonlinear fiber model
were obtained from these relationships as shown in Fig. 9a, and the unconfined concrete
cover was neglected. Since the TEC requires special confinement for boundary elements,

13
Bulletin of Earthquake Engineering

90 Yarımca
1.5
120 60 Erzincan
Kjm
1

150 30

0.5
FP direction

180 0

210 330

240 300

270

FN direction

Fig. 5  Polar diagrams of normalized maximum PSV values for Kocaeli (Yarımca), Erzincan (Erzincan) and
Kobe (Kjm) records

100 100
Velocity (cm/sec)

Velocity (cm/sec)

50 50
0
0
-50
-50 -100
-100 -150
0 10 20 30 0 10 20
Time (sec) Time (sec)
(a) (b)
100
Velocity (cm/sec)

50
0
-50
-100
-150
0 10 20 30 40
Time (sec)
(c)
Fig. 6  Velocity time histories of max PSV directions for a Kocaeli (Yarımca) b Erzincan (Erzincan) c Kobe
(Kjm)

13
Bulletin of Earthquake Engineering

Table 5  Ground motions database


No Earthqauke Year Mw Station Rrup (km) Tp (s) Vs30 (m/s)

(a) Near-fault records (large pulse period)


1 Kocaeli 1999 7.4 Yarımca 4.80 4.62 297.0
2 Kocaeli 1999 7.4 Arçelik 13.50 5.85 523.0
3 Kocaeli 1999 7.4 Gebze 10.92 4.71 792.0
4 ChiChi 1999 7.6 TCU087 6.98 4.47 538.7
5 ChiChi 1999 7.6 TCU075 0.89 5.60 573.0
6 ChiChi 1999 7.6 TCU036 19.80 6.16 478.0
7 Imperial Valley 1979 6.5 El-Centro Array #4 7.05 5.15 208.9
8 TabaS 1978 7.3 Tabas 2.05 4.52 766.0
9 Darfield 2010 7.0 ROLC 1.54 6.50 298.0
10 Darfield 2010 7.0 LINC 7.11 5.15 263.2
11 El Mayor Cucapah 2010 7.2 WES 11.44 6.65 242.0
(b) Near-fault records (medium pulse period)
1 Erzincan 1992 6.7 Erzincan 4.40 2.13 352.1
2 Northridge 1992 6.7 Newhall-WPC Rd 5.50 2.36 286.0
3 Northridge 1992 6.7 Sylmar Olive view Md. 5.30 2.35 440.5
4 Superstition Hills 1987 6.5 Parachute Test Site 0.80 2.13 348.7
5 Northridge 1994 6.7 Jensen Filter Plant G.B. 5.43 2.46 525.8
6 Imperial Valley 1979 6.5 Agrarias 0.70 1.90 242.1
7 Irpinia/Italy 1980 6.9 Sturno 10.8 3.06 382.0
8 Imperial Valley 1979 6.5 El Centro Array #5 3.95 3.43 205.6
9 Loma Prieta 1992 6.9 Saratoga - Aloha Ave 8.50 2.34 380.0
10 Kobe 1995 6.9 Port Island 3.31 2.20 198.0
11 Superstition Hills 1987 6.5 Kornbloom Road 18.48 2.20 266.0
(c) Near-fault records (small pulse period)
1 Morgan Hill 1984 6.2 Coyote Lake Dam 0.53 0.86 561.4
2 Northridge 1994 6.7 Sepulveda Hospital 8.44 0.97 380.0
3 Northridge 1994 6.7 Pacoima Kagol Canyon 7.26 0.73 508.0
4 Parffield 2004 6.9 Zone 9 2.85 1.19 383.9
5 ChiChi 1999 7.6 CHY006 9.76 1.10 438.2
6 Cape Mendocino 1992 7.0 Cent. Beac Nav. Fac 18.31 1.70 459.0
7 Montenegro 1979 7.1 Bar Skupstina Opstina 6.98 1.60 462.2
8 Loma Prieta 1989 6.9 Gilroy Historic Bldg 10.97 1.35 308.0
9 Northridge 1994 6.7 Pardee sce. 7.46 1.12 325.7
10 Kobe 1995 6.9 KJMA 0.96 1.40 312.0
11 Parffield 2004 6.0 Cholama 3.0 1.22 326.6
(d) Far-field records
1 Kern County 1952 7.4 Santa Barbara Courthouse 82.19 – 515.0
2 Taiwan Smart1(45) 1986 7.3 Smart1E02 51.35 – 671.5
3 Imperial Valley-06 1979 6.5 Coachella Canal # 4 50.10 – 336.5
4 Colinga-01 1983 6.4 Parkfield-Vineyard Cany 32.17 – 308.9
5 Cape Mondenico 1992 7.0 Butler Valley Station 45.43 – 525.3
6 Iwate 2008 6.9 Rifu Town 57.78 – 520.7
7 Darfield 2010 7.0 CSHS 43.60 – 638.4

13
Bulletin of Earthquake Engineering

Table 5  (continued)
No Earthqauke Year Mw Station Rrup (km) Tp (s) Vs30 (m/s)

8 Northridge 1994 6.7 Manhattan-Beach 39.23 – 351.6


9 Taiwan Smart1(40) 1986 6.3 Smart1-O07 57.99 – 314.3
10 Northridge 1994 6.7 LA-Pico and Sentous 31.33 – 304.7
11 ChiChi-06 1999 6.3 CHY042 54.36 – 665.2

4
MCE Maximum direction
Scaling period range NF Large pulse
Spectral acceleration (g)

3 NF Medium pulse
NF Small pulse
FF
2

0
0 1 2 3 4 5 6
Period (sec)

Fig. 7  Scaled mean spectra of ground motion data sets and the MCE maximum direction target spectrum.
NF Near-fault ground motion; FF far-field ground motion

Fig. 8  The elements of core-wall

13
Bulletin of Earthquake Engineering

Table 6  Wall elements and Wall element Story level ρ (%)


longitudinal reinforcement ratios
W1 From 1st to 10th 0.8000
From 11th to 13th 0.6320
From 14th to 26th 0.6440
From 27th to 40th 0.7100
W2 From 1st to 10th 1.1156
From 11th to 13th 1.0125
From 14th to 26th 0.9877
From 27th to 40th 0.9900
W3 From 1st to 10th 1.1775
From 11th to 13th 1.1775
From 14th to 26th 0.9231
From 27th to 40th 1.0620
W4 From 1st to 10th 0.8770
From 11th to 13th 0.6570
From 14th to 26th 0.6580
From 27th to 40th 0.6440

70 700
60 600
50 500
Stress (Mpa)
Stress (Mpa)

40 400
30 300
Boundary, Mander Model
20 Boundary, Perform-3D 200
10 Web, Mander Model 100
Web, Perform-3D
0 0
0 0.005 0.01 0.015 0 0.05 0.1
Strain (mm/mm) Strain (mm/mm)
(a) (b)
Fig. 9  Material stress–strain relationships a specified concrete strength with 40 Mpa b reinforcing steel

the concrete stress–strain relationships varied considerably. Therefore, different concrete


stress–strain curves were used for the boundary and web fibers separately, and shear behav-
ior was modeled elastically.
RW2 specimen developed and tested by Thomsen and Wallace (2004) was modeled in
Perform-3D for adjusting cyclic degradation parameters. Orakçal and Wallace (2006) pro-
posed to modify reinforcing steel tensile stress and incorporate concrete tensile strength to
calibrate the analytical model. However, in the present study, the tensile strength of con-
crete was ignored, and the reinforcement steel behavior was assumed to be symmetrical
in tension and compression. The concrete stress–strain curves were obtained by Mander’s
confined model. The cyclic degradation parameters for concrete and reinforcement mate-
rials were used as proposed by Görgülü and Taskin (2015) and Ghodsi and Ruiz (2010),
respectively. However, the unloading stiffness factor for reinforcement material was taken

13
Bulletin of Earthquake Engineering

Fig. 10  Cyclic behavior of the 200


Perform-3D
RW2 test specimen and Perform- 150 Test
3D model results (test results
were provided by Dr. Orakcal) 100

Lateral load (kN)


50
0
-50
-100
-150
-200
-2.5 -1.5 -0.5 0.5 1.5 2.5
Lateral drift (%)

as − 0.75. Loading protocol of RW2 test was applied to Perform-3D model by cyclic load-
ing procedure as detailed in Orakçal and Wallace (2006). The comparison of the numerical
and test results is given in Fig. 10.

4.2 Modeling of moment frame beam and column

Columns and flexural beams were modeled with concentrated plastic hinge approach, in
which at each end, the concentrated plastic hinge was assumed with rigid end zone. The
part between the hinges was modeled elastically with reduced stiffness as given in Table 3.
The top reinforcement ratio of beam ends changed from 0.49% to 0.75%, whereas the
bottom reinforcement ratios were between 0.45% and 0.48%. Inelastic moment-rotation
curves of beams were calculated analytically by using expected material properties, and
then monotonic backbone curves were modified to incorporate cyclic degradation effects as
described in option 3 of PEER/ATC (2010) (Fig. 11).
Tower column longitudinal reinforcement ratios were taken as 2.1, 1.97 and 1.83%
depending on the cross-sectional dimensions with respect to the order given in Table  1.
Backbone curves of column plastic hinges were assumed to be elastic-perfectly plastic,
and interaction moment-axial force capacity diagrams were determined by using expected
material properties while the cyclic and strength deteriorations were disregarded.

Fig. 11  Modified backbone curve to account cyclic degradation effects (PEER/ATC 2010)

13
Bulletin of Earthquake Engineering

4.3 Coupling beam modeling

Diagonal reinforcement ratio of the coupling beams were around 1% and 2% in X and Y
directions, respectively. The coupling beams were modeled with a mid-length shear-dis-
placement hinge, and other parts were modeled elastically with their reduced stiffness.
Shear hinge modeling and cyclic degradation parameters were taken as suggested by Wal-
lace et  al. (2009). In Fig.  12, ­Vcode is the shear capacity of section given by codes, and
herein it was calculated according to TEC with expected material properties.

5 Response history analysis for MCE level

Response history analyses at MCE level were conducted at the nonlinear and linear time
history analyses stages. The performance evaluations were obtained by nonlinear time his-
tory analysis (NTHA) results, and the mean values of the shear stresses, drift ratios, flex-
ural and coupling beam rotations were checked with acceptance criteria of the collapse
prevention performance level for each ground motion set.
In order to figure out the energy dissipations and the post-yield story shear forces redis-
tribution, the MCE level elastic model was subjected to the ground motions in Table 5, and
the comparison results of NTHA and LTHA were obtained over the height.

5.1 Performance evaluation

Core wall shear stress demand is limited for tall buildings to obtain ductile behavior (TEC
2018; ASCE 318-14; PEER/ATC 2010). Therefore, TEC restricts the shear stresses for the
walls with and without coupling beams by 0.65√(fc′) and 0.85√(fc′), respectively. Under
all the ground motion sets, the mean shear stresses of one critical wall in X and Y direc-
tions were below the TEC allowable limits as shown in Figs. 13 and 14.
Story drift ratio is the most used demand parameter to assess the overall system perfor-
mance. Therefore the maximum mean drift ratio for a tall building is given as 0.03 (TBI
2010; LATBSDC 2017). Figure 15 shows that story drift ratios did not reach the accept-
able limit given by seismic design codes for all the ground motions sets. The mean strains
of all edges were obtained for the core-wall, and strains of two critical points are given in
Fig. 16. The maximum mean compression strain was 0.00155, which was under the limit of
the ACI 318-14 compression-controlled Sect. (0.002) and TEC uninterrupted performance

Fig. 12  Coupling beam force– 1.4


deformation curve (Wallace et al.
2009) 1.2
1
V/Vcode

0.8
0.6
0.4
0.2
0
0 2 4 6 8 10 12
Drift (% rotation)

13
Bulletin of Earthquake Engineering

40
NF Large pulse
35 NF Medium pulse
30 NF Small pulse
FF
Story level

25 0.85√fc'
20
15
10
5
0
-8000 -6000 -4000 -2000 0 2000 4000 6000 8000
Shear stress (kN/mm2)

Fig. 13  Shear stresses of a critical wall in the X direction

40
NF Large pulse
35 NF Medium pulse
NF Small pulse
30
FF
0.65√fc'
Story level

25
20
15
10
5
0
-6000 -4000 -2000 0 2000 4000 6000

Shear stress (kN/mm2)

Fig. 14  Shear stresses of a critical wall in the Y direction

40
NF Large pulse
35 NF Medium pulse
30 NF Small pulse
FF
Story level

25 0.03
20
15
10
5
0
-0.04 -0.03 -0.02 -0.01 0 0.01 0.02 0.03 0.04
Drift ratio

Fig. 15  Drift demand ratios

13
Bulletin of Earthquake Engineering

40 40
NF Large pulse NF Large pulse
35 NF Medium pulse 35 NF Medium pulse
NF Small pulse NF Small pulse
30 FF 30 FF
25
Story level

25

Story level
20 20
15 15
10 10
5 5
0 0
-0.003 0.002 0.007 0.012 0.017 -0.003 0.002 0.007 0.012
Compression (-) and tension (+) strain Compression (-) and tension (+) strain
(a) (b)
Fig. 16  Mean compression and tension strains a on gridlines 4-D b on gridlines E-6

level (0.0025). Tension strain demands were substantially different with respect to ground
motion sets. While the tension strains of the outer edges of the piers exceeded the yield
point at the podium level in ground motions with large and medium pulses, these strains
exceeded the yield point in small pulses and far-field ground motions at the 27th-floor
level. This phenomenon was in well agreement with overturning moment distributions.
As the flexural beams connected core-wall had considerably larger end rotations than
the periphery beams, their mean rotations only in X-direction are shown in Fig. 17a with
ASCE 41-13 collapse prevention limit. In the core-wall model, while the stiffness of the
shear walls reduced the story drift, they transferred significant moments to the beams; con-
sequently, beam end rotation demands increased. For this reason, the beam rotations were
close to ASCE 41-13 acceptance limit. Mean values of coupling beam rotations in X direc-
tion were below the ASCE 41-13 collapse prevention limit as illustrated in Fig. 17b.
Since the ductility behavior of the columns decreases due to the increase in the axial
load, it is desirable that the column axial forces at MCE level do not exceed 0.4Acfc (LATB-
SDC 2017). Despite the exceeding of this limit in all the ground motion sets as shown in
Fig. 18, the maximum average column rotation was 0.0028 rad, and this value was below
the immediate operation performance limit of ASCE 41-13, which is 0.003 rad. Therefore,
the inelastic behavior of columns did not significantly affect the overall performance.

5.2 Higher mode effects

The ground motions with ­Tp/T1 < 1 lead to higher mode effects and increase the structural
responses at the upper stories. On the other hand, in the ground motions with ­Tp/T1 > 1,
the first mode governs the structural responses which maximized at the lower stories
(Baker and Cornell 2007). In the acceleration-sensitive region, first-mode dominant behav-
ior governs the structural response, and this situation increases the story drift demands in
tall buildings. However, higher mode effects increase as the first mode period becomes
distant from acceleration–sensitive region (Malhotra 1999). The dominant peak of PSV
is highly correlated with pulse duration (Akkar et al. 2005; Alavi and Krawinkler 2000).
Thus, acceleration and velocity spectra are suggested to be considered to explain higher

13
Bulletin of Earthquake Engineering

40 NF Large pulse
35 NF Medium pulse
NF Small pulse
30 FF
25
Story level

20
15
10
5
0
-0.075 -0.05 -0.025 0 0.025 0.05 0.075
Rotation (rad)
(a)
40 NF Large pulse
35 NF Medium pulse
NF Small pulse
30 FF
ASCE 41-13 limit
25
Story level

20
15
10
5
0
-0.06 -0.03 0 0.03 0.06
Drift ratio (rad)
(b)
Fig. 17  a Flexural beam rotations b coupling beam rotations

40
NF Large pulse
35 NF Medium pulse
NF Small pulse
30 FF
0.4Acf'c
Story level

25
20
15
10
5
0
0.1 0 -0.1 -0.2 -0.3 -0.4 -0.5 -0.6
N/(Acf'c)

Fig. 18  Column normalized mean axial forces

13
Bulletin of Earthquake Engineering

40
NF Large pulse
35 NF Medium pulse
30 NF Small pulse
FF
25
Story level

20
15
10
5
0
-7500 -5000 -2500 0 2500 5000 7500
Moment (1000*kNm)

Fig. 19  Overturning moments in X-direction

mode effects, and although the mode periods lengthen during the nonlinear response, these
spectra give a primary aspect about dynamic behavior (Kalkan and Kunnath 2006).
As illustrated in Figs. 13 and 19, in the near-fault with small pulses and far-field ground
motion sets, relatively higher mean shear stresses and overturning moments were obtained
at upper stories. Conversely, in the ground motion sets with large and medium pulses, the
mean shear stresses were small at upper stories, and the mean overturning moments lin-
early decreased along with height above the podium level.
In the present study, to clarify the relation between ­ Tp/T1 ratio and higher mode
effects, smoothed tripartite (pseudo-spectral acceleration (PSA), pseudo-spectral velocity
(PSV) and pseudo-spectral displacement (PSD)) spectrum of each ground motion set was
obtained in maximum PSV direction as given in Fig. 20. These spectra revealed that large
pulses induced wide acceleration-sensitive regions.
The first three periods of the MCE level elastic model were 2.64, 0.67 and 0.30  s in
the X-direction, and related participation mass ratios were 48, 15, and 7%, respectively.
As shown in Fig. 20, near-fault with small pulses and far-field sets had short acceleration
sensitive regions (0.67 and 0.53 s), and large spectral acceleration demands at the second
and third mode periods gave rise to higher mode effects. The second mode period was very
close to acceleration-sensitive transition period at the small pulses set and in the velocity-
sensitive region at the far-field set, while the third mode period was in the acceleration-sen-
sitive region in both sets. As shown in Figs. 15, 16 and 17, the nonlinear behavior signifi-
cantly influenced the upper stories in these groups. The shifting of the second mode period
caused a decrease in shear force at upper stories, and it was unexpected that the elongated
third mode period exceeded the acceleration-sensitive region in both groups. Unlike the
second and third modes, the first mode period was in the displacement sensitive-region.
Thus, the first mode period elongation did not significantly increase the displacement
demands, which limited the drifts at the bottom and mid-height levels.
In the near-fault ground motions with medium pulses, the spectrum transition periods,
­Tc and ­Td were 1.2 and 3.17 s. As a result, the first mode period was in the velocity-sen-
sitive region, while the second and third mode periods were in the acceleration-sensitive
region. The core-wall outer edge strains exceeded the elastic limit at the podium level,

13
Bulletin of Earthquake Engineering

500

PS
100

10 1

D
(c
10

m
2

)
)
(g

10 0
A
PS

10
1
PSV (cm/sec)

10 -1
10

10
0
T 2 =0.67 sec

10 -2
10
-1

10 -3
1
10
-2

T 3 =0.3 sec T 1 =2.64 sec

0.1
0.01 0.1 1 10 50
T (sec)

Fig. 20  Smoothed tripartite spectra of four ground motion sets in maximum PSV directions

which excited the nonlinear response of the first mode and moved it toward the displace-
ment sensitive region. By the first mode period elongation, base shear decreased while
the second and third mode shear forces did not significantly change due to their behavior
remaining within elastic limits in the acceleration-sensitive region.
The near-fault ground motions with large pulses, whose transition periods were 4.34
­(Tc) and 5.45 (­Td) s, had widest acceleration-sensitive region. All the modes were in the
acceleration-sensitive region, and the first transition period was 1.64 times larger than the
first mode period. Thus, in contrast to the other ground motion sets, the shifting of the
first mode in the acceleration-sensitive region considerably increased the displacement
demands. As a result, the largest drift ratios, beam rotations at all stories and the highest
tension strains were obtained at the podium level.
When the core-wall flexural yielding capacity is exceeded at the upper stories, the
higher mode effects decrease (Değer et al. 2014; Panagiotou and Restrepo 2009). There-
fore, the drift demands at upper stories increased, and redistributions of the story shear
forces came into prominence. Elastic to inelastic story shear force ratios ­(Velas/Vine) in X
direction were obtained for all the ground motion sets to show the post-yield story shear
forces redistribution due to higher mode effects (Fig. 21).
The short acceleration-sensitive region in near-fault ground motions with small pulses
and far-field ground motions increased the higher mode effects, which caused the overall
moment to exceed the flexural yielding capacity at the upper stories. After yielding of the
core-wall, the shear forces decreased above the yielding level and increased at the mid-
height due to the post-yield redistribution. Therefore, the V ­ elas/Vine ratios were relatively
higher at upper than mid-height stories. In the large and medium pulse ground motions,
even though the beam rotations and drift ratios were maximized above the mid-height level

13
Bulletin of Earthquake Engineering

Fig. 21  Elastic to inelastic story 40


shear force ratio ­(Velas/Vine)
35
NF large pulse
30 NF medium pulse
NF small pulse
25

Story level
FF
20
15
10
5
0
0 1 2 3 4
Velas /Vine

of the tower, core-wall yielded at the podium level on account of the first mode effect.
Thus, the linear profile of the V
­ elas/Vine ratios was experienced along the height. However,
in the records with medium pulses, the largest V ­ elas/Vine ratios were obtained because of the
higher elastic story shear force demands.
The smoothed tripartite spectra showed that, when the first and second mode periods
were in the same or neighboring sensitive regions, the structural response based on the
first-mode dominant behavior as in the case of large and medium pulse ground motion sets.
Conversely, the higher mode effects increased while the sensitive region of the higher and
the first mode period moved away from each other. This situation was especially evident
when the second mode period was in the acceleration-sensitive region and the first mode
period was in the displacement-sensitive region as observed in the small pulses set.

5.3 Practical results for designers

The probability of the pulse occurrence and estimation of the pulse duration are two essen-
tial parameters which come into prominence for the design of structures in near-fault
zones. In some cases, even though the dislocation of rupture propagates toward a site, no
pulse takes place in velocity records (NIST 2011). Therefore, the percentage of pulse-like
motions have been investigated in near-fault zones (Shahi and Baker 2011; Hayden et al.
2012). In the literature, earthquake magnitude is taken into account as the critical param-
eter to predict the pulse period (Somerville 2003; Alavi and Krawinkler 2000; Bray and
Rodriguez-Marek 2004; Shahi and Baker 2014b; Cork et al. 2016).
The following two basic conclusions can be inferred from the estimation of pulse dura-
tion and the occurrence probability of pulse-like ground motions for designers. Firstly, the
likelihood of pulse-like motions occurrence decreases by moving away from the fault. Sec-
ondly, the pulse period increases with earthquake magnitude.
The pulse duration enabling the T ­ p/T1 < 1, ­Tp/T1 ≈ 1 and ­Tp/T1 > 1 conditions have
unique characteristics on the behavior of tall buildings (Baker and Cornell 2007) which
should be taken into account by designers. When the T ­ p/T1 ratio was below 1, the elon-
gated first mode period was usually in the displacement-sensitive region. Thus, as a result
of the higher mode effects, outer fiber elements of core-wall exceeded the yielding capacity
at the upper level, and this phenomenon was likely to increase by narrowing thicknesses
of shear walls and decreased longitudinal reinforcement ratios at the upper stories. For

13
Bulletin of Earthquake Engineering

these reasons, flexural and coupling beam rotations were more critical at the upper sto-
ries. PEER/ATC (2010) showed that in order to save analyses time, elastic modeling of the
region outside the plastic hinge zone, which was thought to be at the base or just above the
podium level, caused unrealistic story shear forces, especially at mid-height. Moreover, the
rotation demands of flexural and coupling beams increased above the yielding level. This
circumstance was also valid for far-field ground motion.
­ p/T1 ratio to 1, the higher mode effects in structural response
With the increasing of the T
decrease due to the first mode presence in the velocity-sensitive region (Malhotra 1999),
and the elastic response maximize as detailed by Akkar et  al. (2005). In this case, it is
useful for designers to keep in mind that the elastic analyses overestimated the story shear
forces compared to other records, and potential plastic hinge zone of core-wall was prob-
ably just above the podium level.
In the ground motion set with ­Tp/T1 > 1, first mode period most likely remained in
the acceleration-sensitive region, and the elongation of the first mode period gave rise to
increasing displacement demand in the acceleration-sensitive region. Thus, the highest ten-
sion strain, beam rotation, and story drift demands were observed at all stories, and the
plastic hinge region of the core-wall was just above the podium level.

6 Summary and conclusion

A core-wall tall building was designed and subjected to three sets of near-fault ground
motions with different T ­ p/T1 ratios and one set of far-field ground motions. All records
were rotated to the maximum PSV direction which was well-compatible with the pulse
direction. The T­ p/T1 ratios for the pulse-like ground motion sets were below, around and
bigger than 1, whereas the far-field records were obtained from 39 to 82  km distance to
fault rupture. The smoothed tripartite mean spectrum of pulse-like ground motion sets
highlighted that acceleration-sensitive region expanded with increasing pulse duration.
Although each record was scaled to the same spectrum, the results showed that all the
ground motion sets with different pulse periods had distinctive attributes on the core-wall
tall building, which must be considered during the design phase.
In the case of ­Tp/T1 < 1, the second and third mode periods were in the higher spectral
acceleration plateau compared to the first mode period. For this reason, the demands of
the story shear forces and overturning moments increased at the upper levels. When the
elastic capacity of core-wall was exceeded at the upper levels, the story shear forces at
the mid-height increased in consequence of post-yield redistribution, and thus the second
mode effect decreased. Although both of the near-fault ground motions with small pulses
and the far-field ground motions induced the higher mode effects, the impacts of the former
set were higher on the structural demand parameters such as drift ratios and core-wall ten-
sion strains.
In the case of ­Tp ≈ T1, the first mode period was in the velocity-sensitive region, while
the second and the third mode periods were in the acceleration-sensitive region. The dif-
ferences of spectral acceleration demands between these regions were not at the levels
enough to trigger the higher mode effects. Furthermore, the high elastic demands caused
the core-wall to reach the elastic capacity at the podium level firstly, and thus the first mode
behavior governed the nonlinear responses. After the first yielding, due to the fundamental
period elongation, the story shear force demands decreased rapidly, and the high V ­ elas/Vine
ratios were obtained as shown in Fig. 21.

13
Bulletin of Earthquake Engineering

In the near-fault ground motions with large pulses, the T ­ p/T1 > 1 caused the elongated
first mode period to remain in the acceleration-sensitive region. Although the large beam
rotations were observed at the mid and upper stories, the stiffness of the tower significantly
decreased in relation with the exceeding of the core-wall flexural yielding capacity at the
podium level. Therefore, the first mode period considerably shifted to longer periods in the
acceleration-sensitive region. As given in Fig. 20, the displacement demand of the accel-
eration-sensitive region of near-fault ground motion with large pulses set was significantly
larger than that of other sets. Consequently, considerably larger drift ratios and beam rota-
tion were obtained at all stories, and due to the low contribution of the higher mode effects
to structural response, linear profile of ­Velas/Vine ratios were observed along the height.

References
ACI 318-14 (2014) Building code requirements for structural concrete and commentary. American Concrete
Institute, Farmington Hills
Akkar S, Gülkan P (2003) A Near-fault design spectrum and its drift limits. In: Fourth international confer-
ence on seismology and earthquake engineering. International Institute of Earthquake Engineering and
Seismology, Tehran
Akkar S, Yazgan U, Gülkan P (2005) Drift estimates in frame buildings subjected to near-fault ground
motions. ASCE J Struct Eng 131(7):1014–1024
Alavi B, Krawinkler H (2000) Consideration of near-fault ground motion effects in seismic design. In: 12th
world conference on earthquake engineering, Auckland, New Zealand
Alavi B, Krawinkler H (2001) Effects of near-fault ground motions on frame structures. John A. Blume
Earthquake Engineering Center Stanford. Report No: 138
Alavi B, Krawinkler H (2004) Behavior of moment-resisting frame structures subjected to near-fault ground
motions. Earthq Eng Struct Dyn 33:687–706
ASCE, SEI 7–10 (2010) Minimum design loads for buildings and other structures. American Society of
Civil Engineers, Reston
ASCE, SEI 41-13 (2013) Seismic Rehabilitation Standards Committee. Seismic Rehabilitation of Existing
Buildings. American Society of Civil Engineers, Reston
Baker JW (2010) Conditional mean spectrum: tool for ground-motion selection. J Struct Eng
137(3):322–331
Baker JW, Cornell CA (2006) Vector-Valued Ground Motion Intensity Measures for Probabilistic Seismic
Demand Analysis. PEER Report No: 08
Baker JW, Cornell CA (2007) A vector-valued measures for pulse-like near-fault ground motions. Engineer-
ing Structure 30:1048–1057
Basone F, Cavaleri L, Di Trapani F, Muscolino G (2017) Incremental dynamic based fragility assessment
of reinforced concrete structures: stationary versus non-stationary artificial ground motions. Soil Dyn
Earthq Eng 103:105–117
Beyer K, Bommer JJ (2006) Relationships between median values and between aleatory variabilities for dif-
ferent definitions of the horizontal component of motion. Bull Seismol Soc Am 96:1512–1522
Bray JD, Rodriguez-Marek A (2004) Characterization of forward-directivity ground motions in the near-
fault region. Soil Dyn Earthq Eng 24:815–828
Carvalho G, Bento R, Bhatt C (2013) Nonlinear static and dynamics analyses of reinforced concrete build-
ings-comparison of different modeling approaches. Earthq Struct 4(5):451–470
CEN (2004) Eurocode 8–Earthquake resistant design of structures. General rules–specific rules for various
materials and elements. Brussels, Belgium
Champion C, Liel A (2012) The effect of near-fault directivity on building seismic collapse risk. Earthq Eng
Struct Dyn 41(10):1391–1409
Chopra AK, Chintanapakdee C (2001) Comparing response of SDF systems to near-fault and far-fault earth-
quake motions in the context of spectral regions. Earthq Eng Struct Dyn 30:1769–1789
Cork TG, Kim JH, Mavroeidis GP, Kim JK, Halldorsson B, Papageorgiou AS (2016) Effects of tectonic
regime and soil conditions on the pulse period of near-fault ground motions. Soil Dyn Earthq Eng
80:102–118
CSI (2016a) ETABS, Extended 3D Analysis of Building Systems Software, Nonlinear Version 9.7.1. Com-
puters and Structures, Inc, Berkeley, CA

13
Bulletin of Earthquake Engineering

CSI (2016b) PERFORM-3D, nonlinear analysis and performance assessment for 3D structures, version
6. Computers and Structures Inc, Berkeley
Değer ZT, Yang TY, Wallace JW, Moehle J (2014) Seismic performance of reinforced concrete core wall
buildings with and without moment resisting frames. Struct Des Tall Spec Build 24:477–490. https​
://doi.org/10.1002/tal.1175
Dezhdar E, Adebar P (2012) Estimating seismic demands on high-rise concrete shear wall buildings. In:
15th world conference on earthquake engineering, Lisbon
Ghodsi T, Ruiz JAF (2010) Pacific earthquake engineering research/seismic safety commission tall
building design case study 2. Struct Des Tall Spec Build 256:197–256
Görgülü O, Taskin B (2015) Numerical simulation of RC infill walls under cyclic loading and calibration
with widely used hysteretic models and experiments. Bull Earthq Eng 13:2591–2610
Haselton CB, Liel AB, Taylor Lange S, Deierlein GG (2008) Beam-column element model calibrated
for predicting flexural response leading to global collapse of RC frame buildings, PEER Report
2007/03, Pacific Earthquake Engineering Research Center, University of California, Berkeley,
California
Hayden C, Bray J, Abrahamson N, Acevedo-Cabrera AL (2012) Selection of near-fault pulse motions for
use in design. In: 5th international world conference on earthquake engineering, Lisboa
Huang YN, Whittaker AS, Luco N (2008) Maximum spectral demands in the near-fault region. Earthq
Spectra 24:319–341
Kalkan E, Kunnath SK (2006) Effects of fling step and forward directivity on seismic response of build-
ing. Earthq Spectra 22:367–390
Kazaz İ, Gülkan P (2016) Dynamic shear force amplification in regular frame–wall systems. Struct Des
Tall Spec Build 25:112–135. https​://doi.org/10.1002/tal
LATBSDC (2017) An alternative procedure for seismic analysis and design of tall buildings located in
the Los Angeles region. Los Angeles Tall Buildings Structural Design Council, Los Angeles
Malhotra PK (1999) Response of buildings to near-field pulse-like ground motions. Earthq Eng Struct
Dyn 28:1309–1326
Mander JB, Priestley MJN, Park R (1988) Theoretical stress–strain model for confined concrete. ASCE J
Struct Eng 114:1804–1826
Mehmood T, Warnitchal P, Ahmed M, Qureshi MI (2017) Alternative approach to compute shear ampli-
fication in high-rise reinforced concrete core wall buildings using uncoupled modal response his-
tory analysis procedure. Struct Des Tall Spec Build 26:1–18
Naish D, Wallace JW, Fry JA, Klemencic R (2009) Experimental evaluation and analytical modeling of
ACI 318-05/08 reinforced concrete coupling beams subjected to reversed cyclic loading. UCLA –
SGEL 6
NIST (2011) Selecting and scaling earthquake ground motions for performing response-history analyses.
NEHRP Consultants Joint Venture for NIST, Gaithersburg
NIST (2017) Guidelines for nonlinear structural analysis for design of buildings, part IIb—concrete moment
frames, under preparation by Applied Technology Council, ATC Project 114, pending publication
NZS (2006) NZS 3101: part1-concrete structures standard, part- 2 commentary on the design of con-
crete structures. New Zealand Standards, Wellington
Orakçal K, Wallace JW (2006) Flexural modeling of reinforced concrete walls-experimental verification.
ACI Struct J 103(2):196–206
Panagiotou M, Restrepo JI (2009) Dual-plastic hinge design concept for reducing higher-mode effects on
high-rise cantilever wall buildings. Earthq Eng Struct Dyn 38:1359–1380
PEER (2010) Technical report for the PEER ground motions database web application
PEER Ground Motion Database (2017) Pacific Earthquake Engineering Research Center. University of
California, California
PEER/ATC (2010) Modeling and acceptance criteria for seismic design and analysis of tall buildings
(PEER/ATC 72-1). Applied Technology Council, Redwood City- Pacific Earthquake Engineering
Center, Berkeley
PEER/TBI (2010) Guidelines for performance-based seismic design of tall buildings. PEER Report
2010/05. The TBI Guidelines Working Group, Pacific Earthquake Engineering Research Center,
University of California, Berkeley, CA
Rad BR, Adebar P (2008) Dynamic shear amplification. In: High-rise concrete walls: effect of multiple flex-
ural hinges and shear cracking. The 14 the world conference on earthquake engineering, Beijing, China
Rejec K, Isokovic T, Fischiner M (2013) Seismic shear force magnification in RC cantilever structural
walls, designed according to Eurocode 8. Bull Earthq Eng 10:567–586
Rupakhety R, Sigurdsson SU, Papageorgiou AS (2011) Quantification of ground-motion parameters and
response spectrum in the near-fault region. Bull Earthq Eng 9:893–930

13
Bulletin of Earthquake Engineering

Rutenberg A (2013) Seismic shear forces on RC walls: review and bibliography. Bull Earthq Eng
11:1727–1751. https​://doi.org/10.1007/s1051​8-013-9464-1
Shahi SK, Baker JW (2011) Regression models for predicting the probability of near-fault earthquake
ground motion pulses, and their period. In: 11th International conference on applications of statistics
and probability in Civil Engineering, Zurich, Switzerland
Shahi SK, Baker JW (2014a) NGA-West2 models for ground motion directionality. Earthq Spectra
30:1285–1300
Shahi S, Baker JW (2014b) An empirically calibrated framework for including the effects of near-fault
directivity in probabilistic seismic hazard analysis. Bull Seismol Soc Am 101:742–755
Somerville PG (2003) Magnitude scaling of the near-fault rupture directivity pulse. Phys Earth Planet Inter
137:201–212
Somerville PG, Smith N, Graves R, Abrahamson N (1997) Modification of empirical strong ground motion
attenuation relations to include the amplitude and duration effects of rupture directivity. Seismol Res
Lett 68(1):199–222. https​://doi.org/10.1785/gssrl​.68.1.199
Stewart JP, Abrahamson NA, Atkinson GM, Baker J, Boore DM, Bozorgnia Y, Campbell KW, Comartin
CD, Idriss IM, Lew M, Mehrain M, Moehle JP, Naeim F, Sabol TA (2011) Representation of bi-direc-
tional ground motions for design spectra in building codes. Earthq Spectra 27:927–937
TEC (2018) Turkish Earthquake Code. Disaster and Emergency Management Authority, Ankara
Thomsen JH, Wallace JW (2004) Displacement-based design of slender reinforced concrete structural
walls—experimental verification. ASCE J Struct Eng 130(4):618–630
Vuran E, Aydınoğlu MN (2016) Capacity and ductility demand estimation procedures for preliminary design
of coupled core wall systems of tall buildings. Bull Earthq Eng 14:721–745. https​://doi.org/10.1007/
s1051​8-015-9853-8
Wallace JW, Naish D, Fry JA, Klemencic R (2009) RC core walls–testing and modeling of coupling beams.
PEER Center Annual Meeting
Watson-Lamprey J, Boore DM (2007) Beyond SaGMRotI: conversion to SaArb, SaSN, and SaMaxRot. Bull
Seismol Soc Am 97(5):1511–1524

Publisher’s Note  Springer Nature remains neutral with regard to jurisdictional claims in published maps and
institutional affiliations.

13

Das könnte Ihnen auch gefallen