Open access
Technical Papers
Jul 6, 2021

Performance-Based Wind Design of High-Rise Buildings Using Generated Time-History Wind Loads

Publication: Journal of Structural Engineering
Volume 147, Issue 9

Abstract

Performance-based wind design (PBWD) allowing inelastic behavior under extreme wind load has only been recently considered for design of high-rise buildings. Unlike performance-based seismic design, there are few guidelines and research on inelastic wind design. Time-history wind loads for PBWD rely on wind tunnel tests. However, conducting wind tunnel tests for preliminary structural concepts may not be practical due to changes in design and cost. To address these concerns, the following topics were studied: initial design using an RW factor to introduce inelastic behavior; generation of time-history wind loads from power spectral density (PSD) functions for inelastic analysis including vertical distribution and maximum directional load occurrence, and a case study for wind resistance performance evaluation of a reinforced concrete building design using an RW factor of 1, 2, and 3.

Introduction

For structural design, the main lateral loads considered are typically wind and seismic (Kang et al. 2013; Alinejad and Kang 2020). The structure is permitted to yield, in seismic design, to introduce inelastic behavior, and a response modification factor R (or RE) is employed. On the other hand, only elastic behavior is permitted in the strength design of conventional wind design codes. The following issues are commonly raised when the structure is designed to yield under wind loads: (1) relatively low level of loads compared to actual seismic loads, (2) fatigue failure by long duration of loading, (3) ratcheting by the mean component of wind load, and (4) fluid-structure-interaction by large deformation (i.e., aerodynamic instability problem). The difference in design philosophy causes complications in the design of high-rise buildings for wind and seismic loads.
Wind load can exceed that of seismic load, which is reduced by an RE factor, as the height of the building increases. Current elastic wind design requirements often lead to the need for excessive stiffness and strength of horizontal members such as beams, coupling beams, and braces. Due to overdesigned fuse elements, vertical members and joints under seismic load also increase.
To address this issue, researchers have begun to focus on performance-based wind design (PBWD). In the early stages of PBWD, the framework for evaluation of performance objectives at various hazard levels was developed. Paulotto et al. (2004) proposed a framework for PBWD for low and high performance levels. ACI (2006) Committee 375 published “Performance-Based Design of Concrete Building for Wind Loads (SP-240),” gathering six papers. It provides the differences in the design methodologies of wind and seismic design; the necessity of wind design for serviceability, strength, and stability level loads; modeling assumptions for wind design; and wind tunnel methods to overcome the limitation of code-based design. van de Lindt and Dao (2009) proposed PBWD for wood-frame buildings. Fragility analyses of wooden structures were carried out for occupant comport, continued occupancy, life safety, and structural integrity performances. Ciampoli et al. (2011) proposed a general procedure of PBWD, and carried out a risk assessment of a bridge by the proposed procedure. Probabilistic evaluation of collapse prevention, occupant safety, accessibility, and functionality, among others, was studied by Petrini and Ciampoli (2012). Nakai et al. (2013) studied aerodynamically unstable vibration, exterior claddings, and habitability level hazards for structures.
In recent research, PBWD has begun to extend into inelastic behavior. Aswegan et al. (2017) carried out a nonlinear analysis to verify PBWD of a case study building allowing yielding under extreme wind load. Mohammadi et al. (2019) carried out performance evaluation under extreme wind load by nonlinear time-history analysis (NTHA). El Damatty and Elezaby (2018) and Elezaby and El Damatty (2020) proposed a reduction of the resonant component of the design wind load by two to introduce inelastic behavior under the wind load. They adopted ductility-based design in wind engineering. The structural performance was verified by pushover analysis and resultant ductility demands. Verification by NTHA was not carried out. Bezabeh et al. (2019) studied an inelastic damage accumulation of single-degree-of-freedom (SDOF) bilinear and self-centering system by generated along-wind time-history load. Bezabeh et al. (2020) and Alinejad et al. (2020) proposed frameworks for an inelastic wind design including hazard scenarios of seismic and wind.
Unlike performance-based seismic design (PBSD), there are few guidelines for PBWD in practice. To meet the demand for PBWD, ASCE recently published a prestandard for PBWD (ASCE 2019). It permits inelastic behavior under wind load exceeding a mean recurrence interval (MRI) of 700 years. Inelastic behavior of lateral load resisting structural components or ductile actions, such as coupling beam, brace, and slender wall, is permitted while maintaining a gravity load resisting capacity, which is in line with the minimum performance requirement for gravity load resisting capacity under earthquakes in PBSD [ASCE 41-17 (ASCE 2017b)]. Therein, performance objectives, acceptance criteria, and performance evaluation methods including conventional time-history analysis and reliability assessment are presented. However, little information exists on how to introduce inelastic behavior in the initial design and acceptance criteria of inelastic performance for member levels.
Several researchers have studied NTHA for wind load (Aswegan et al. 2017; Mohammadi et al. 2019). Their research relied on wind tunnel tests to acquire time-history wind loads. The final verification by time-history wind loads from wind tunnel tests is inevitable. However, it is difficult to utilize wind tunnel test results for initial PBWD in practice due to frequent design changes and costs. Thus, a study on generation of time-history wind load from power spectral density (PSD) functions for incorporation into a design code and a PBWD case study of a building is conducted in this paper.

Significance of Research

Unlike PBSD, inelastic wind design has not been widely used in practice. There is still a lack of guidelines for initial design to introduce inelastic behavior. In addition, the unavailability of wind tunnel test data becomes an initial barrier of the preliminary design or PBWD of tall and/or irregular buildings, because inelastic design requires trial and error. To bridge this gap, an extensive nonlinear analysis study including the generation of 1,000 wind load time histories is conducted and detailed design procedures are developed. Particularly, the developed method of generating wind load time histories based on PSD functions can be very useful for the preliminary design or PBWD of tall and/or irregular buildings.

Case Study Model

Initial Design

A virtual 45-story reinforced concrete (RC) building, shown in Figs. 1 and 2, was used to conduct a case study. A building frame system with RC core walls and coupling beams, which are commonly found in lateral load–resisting systems for high-rise buildings, was selected. Building structural information, seismic load, and wind load conditions are summarized in Tables 14, respectively.
Fig. 1. Structural plan.
Fig. 2. Analysis model.
Table 1. Building information
Structural systemBuilding heightMaterialsNatural periods
ConcreteReinforcing bar (SD500)
Building frame system with ordinary shear walls180 m (story height of 4  m×45  stories)Strength fc: 40 MPaYield strength fy: 500 MPax-dir.: 4.431 s
Modulus of elasticity Ec: 30,008 MPaModulus of elasticity Es: 200,000 MPay-dir.: 4.456 s
  rz-dir.: 2.879 s
Table 2. Structural member information
Structural membersSizeEffective stiffness
Core wall(Thickness) main perimeter core wall: 900 mm(1st–6th stories)
Internal wall: 250 mmIn-plane stiffness: 1.0EA, 0.35EIga, 0.5GA
Out-of-plane stiffness: 0.25EIg
(7th–45th Stories)
In-plane stiffness: 1.0EA, 0.7EIg, 1.0GA
Out-of-plane stiffness: 0.25EIg
Coupling beam(width×height) 900×900  mm1.0EA, 0.15EIg, 1.0GA
Column(1st–24th stories) 1,000×1,000  mm1.0EA, 0.7EIgb, 1.0GA
(25th–45th stories) 800×800  mm
Beam(width×height) 600×800  mm1.0EA, 0.35EIg, 1.0GA
Slab(Thickness) 210 mmIn-plane stiffness: rigid diaphragm
Out-of-plane stiffness: 0.1EIgc

Note: EA = axial rigidity of gross section; EIg = flexural rigidity of gross section; and GA = shear rigidity of gross section.

a
The value of 0.35EIg is for elastic analysis. The effective stiffness is automatically calculated by fiber elements in nonlinear analysis.
b
The value of 0.7EIg is for elastic analysis. The effective stiffness is automatically calculated by fiber elements in nonlinear analysis.
c
Commonly used value of design practice in Korea.
Table 3. Seismic load conditions
ParametersContents
Effective ground acceleration parameter, S0.22g
Site classSC (more than 20-m depth to bedrock)
Response modification factor, RE5 (building frame system)
Importance factor, IE1.2
Load factor1.0
Damping ratio, ζE5%
Table 4. Wind load conditions
ParametersContents
Basic wind speed, V0 (100-year MRI)38  m/s
Surface roughness categoryB
Power law exponent of mean wind speed profile, α0.22
Topography factor, Kzt1.0
Importance factor, IW1.0
Load factor1.3
Design wind speed, VH53.6  m/s
Damping ratio, ζW1.2% [ISO 4354 (ISO 2009)]
Response modification factor, RW1, 2, 3
To determine the inelastic analysis model, initial design is performed per Korea Building Code (MOLIT 2016). Because MOLIT (2016) uses a wind load factor of 1.3 and wind directionality factor of 1.0, the 100-year MRI wind speed in MOLIT (2016) corresponds to that of 1,700-year MRI in ASCE 7-16 (ASCE 2017a). The basic wind speed of 38  m/s used in this paper is also a wind speed for one of the strong wind hazard regions in Korea. Low-to-intermediate seismic and strong wind hazards are assumed, and accordingly, ordinary shear walls and nonconforming transverse reinforcement for beams and coupling beams are used.
To introduce yielding of members under wind loads, reduction of design force is required. Bezabeh et al. (2020) used a load reduction factor for both background and resonant components to compare the damage accumulation of structures with tuned mass damper (TMD) and self-centering systems under along-wind load. However, applying the reduction factor to the resonant component only was recommended for conventional buildings due to large ductility demands. If sufficient postyield stiffness is secured in the conventional buildings, the background component can be reduced. In contrast, El Damatty and Elezaby (2018) proposed the response modification factor (RW) for resonant component only. Because it is difficult to estimate the postyield stiffness of structure in the early stage of design, this approach is employed in this study. The general form of design wind load is per the following equation:
W=Wmean+Wbackground+(Wresonant/RW)
(1)
where W = design wind load of along-, across-, or torsional-wind load; Wmean = mean component (considered in along-wind load only); Wbackground = fluctuating of wind load itself; and Wresonant = wind load induced by the resonant effect.
For application, a range for the RW factor is needed. Design results using an RW factor of 1, 2, and 3 are compared in this study. RW factors larger than 3 are not effective for a reduction of the initial design load. Design loads depending on the RW factor are shown in Fig. 3. An RW factor of 1 is for conventional elastic wind design. The structural design results of members are shown in the Appendix.
Fig. 3. Design seismic load and wind loads depending on RW factors.
To compare with the seismic load by a response spectrum analysis, the load combination factor of 1.3 was considered for wind loads. The governing lateral design load for high-rise buildings is typically across-wind load due to its large resonant component (Marukawa et al. 1992; Tamura et al. 1996; Ha et al. 2007; Kang et al. 2019). Because large portions of elastic across- and torsional-wind loads are composed of resonant components, those with an RW factor decrease drastically, whereas the reduction of along-wind is limited due to a large portion of its mean and background components. The reduction of lateral loads significantly affects member forces of coupling beams that mainly resist lateral loads, while the member force reduction of vertical members is insignificant.

Inelastic Modeling

There are no current guidelines for inelastic modeling under wind loads. Unlike seismic design, there are few inelastic cyclic loading tests of structural members considering extreme wind loads. Abdullah et al. (2020) studied the inelastic performance of RC coupling beams under extreme wind loads. With hundreds of cyclic loads, RC coupling beams showed a larger pinching effect compared to that by seismic loads, but the overall difference is small. In this study, inelastic behavior of RC structures under seismic and wind loads is assumed to be similar, and PBSD guidelines such as those of the AIK (2019), the Tall Buildings Initiative (TBI 2017), and PEER/ATC 72-1 (PEER/ATC 2010) are employed for inelastic modeling. ETABS version 17 was used for the nonlinear analyses.

Material Models for Fiber Elements

Fiber elements were employed for the inelastic modeling of columns and shear walls (Kang et al. 2009). Expected strengths, expected moduli of elasticity, and backbone curves were determined by AIK (2019) (Fig. 4). The hysteretic behaviors of concrete and reinforcing bar (rebar) fibers were verified using the experimental result of the RW2 specimen by Thomsen and Wallace (1995). Verification is shown in Fig. 5.
Fig. 4. Material models for fiber elements: (a) concrete model (C40); and (b) rebar model (SD500).
Fig. 5. Verification of hysteresis model for fiber elements.

Concentrated Plastic Hinges

Concentrated moment hinges at each end of the beams and coupling beams were employed to model inelastic behavior. Backbone curves and acceptance criteria were determined by AIK (2019), which is identical to ASCE 41-17 (ASCE 2017b).
Takeda model (Takeda et al. 1970) and pivot model (Dowell et al. 1998) were employed for modeling the hysteretic behavior of beams and coupling beams, respectively. The Takeda model was used to describe the hysteretic behavior of conventional RC members without variables, and the variables of pinching point and unloading stiffness degradation used for the pivot model were determined by the experimental result of the HB3-10L-T50 specimen by Xiao et al. (1999), which has the same span-to-depth ratio of the analysis model. All beams and coupling beams were designed on the basis of flexure control and nonconforming transverse reinforcement.
Although the hysteretic behavior of RC members depends on many variables such as rebar ratio and arrangement, span-to-depth ratio, confinement, etc., the hysteresis model shown in Fig. 6 was used in this study for simplicity.
Fig. 6. Verification of hysteresis model for coupling beams.

Gravity Load

An expected gravity load combination, 1.0D+0.25L+1.0E, is used in PBSD. However, for the performance objective of continuous occupancy, the ASCE prestandard for PBWD (ASCE 2019) suggests use of the same gravity load combination that is used for elastic design, 1.2D+1.0L+1.0W. Because the ASCE prestandard for PBWD (ASCE 2019) is based on ASCE 7-16 (ASCE 2017a), the wind load factor is 1.0. However, the wind load factor of 1.3 noted in MOLIT (2016) is used in this study.

Acceptance Criteria

ASCE prestandard PBWD (ASCE 2019) permits inelastic behavior under 700–3,000-year MRI wind loads. The performance objective is defined as continuous occupancy.
Acceptance criteria for force-controlled actions are defined as a demand-capacity ratio not exceeding 1.0. The capacity for force-controlled action is the design strength per ACI 318-19 (ACI 2019) with application of appropriate strength reduction factor ϕ.
Acceptance criteria for deformation-controlled actions are defined for linear analysis as a demand-capacity ratio not exceeding 1.25. The capacity for deformation-controlled action is expected strength with application of a ϕ factor of 1.0. There are no distinct inelastic behavior capacity criteria of each member for NTHA.
Alinejad et al. (2020) suggested structural performance levels for wind and seismic hazard scenarios based on ASCE 41-17 (ASCE 2017b). Structural performance level up to damage control (DC), which is the midpoint of immediate occupancy (IO) and life safety (LS), was suggested for PBWD. For an upcoming earthquake event, a small RW factor for PBWD was recommended.
This approach can be employed based on a designer’s judgement for design of buildings—including that for supertall buildings, greater than 200 m per MOLIT (2016), subject to extreme wind hazard and enhanced criteria such as IO.

Generation of Time-History Wind Load

Code-based wind load is derived by a frequency domain analysis. However, a frequency domain analysis is invalid in an inelastic system. To conduct inelastic PBWD and verification, a time-history wind load is required. Wind tunnel testing is recognized as a reasonable means for determining wind loads. However, it has disadvantages in that it is expensive and not adaptable to design changes. Shinozuka and Deodatis (1991) suggested a generating time history from a PSD function. Hwang et al. (2015) studied generating time-history wind loads from PSDs for an evaluation of habitability. Because the structure remained in the elastic range under habitability evaluation load (1-year return period), directional loads were considered separately and correlation of directional loads could be considered in generation for inelastic analysis. In this study, a time-history load generation from MOLIT (2016) PSD functions for inelastic analysis was carried out.

Along-Wind Load

Along-wind load of KBC is based on the gust load factor (GLF) method suggested by Davenport (1967). The maximum equivalent load is determined by GLF and the standard deviation of the load. The standard deviation can be calculated by the square root of the integration of the PSD function. To calculate the standard deviation of along-wind load, the PSD of wind speed and aerodynamic admittance function were presented. MOLIT (2016) employed the PSD wind speed suggested by Karman (1948). The normalized form of the PSD is shown in Table 5.
Table 5. PSD of along-wind speed
ParametersEquations
Normalized PSD of along-wind speedfSv(f)σv2=4fLH/VH{1+71(fLH/VH)2}5/6
Standard deviation of along-wind speedσv=IHVH
Turbulence intensityIH=0.1(HZg)α0.05
Turbulence lengthLH=100(H30)0.5

Note: Sv(f) = PSD of along-wind speed; f = frequency; VH = design wind speed at the reference height of H; H = reference building height (m); α = power law exponent of mean wind speed profile; and Zg = height of the atmospheric boundary layer from the ground surface (m).

To convert wind speed PSD into wind load PSD, aerodynamic admittance needs to be considered. MOLIT (2016) presents the conversion by
|χ(f)|2=0.84{1+2.1(fH/VH)}{1+2.1(fB/VH)}
(2)
By adopting the quasi-steady assumption, the fluctuating component and PSD of along-wind force at the reference height can be expressed by Eqs. (3) and (4), respectively
p(t)=ρCD*AVHv(t)
(3)
Sp(f)=(ρCD*AVH)2Sv(f)|χ(f)|2
(4)
where t = time; ρ = air density (1.22  kg/m3); A = projection area; v(t) = fluctuating component of wind speed; and CD* = wind force coefficient for fluctuating wind force and can be expressed by (Ha 2017)
CD*0.75CD
(5)
where CD = wind force coefficient for mean along-wind force.
Because an enclosed building was assumed, CD is the difference between the external pressure coefficient of the windward and leeward walls as expressed in Table 6.
Table 6. External pressure coefficients Cpe
ParametersConditionsEquations
Windward wall, Cpe1Cpe1=0.8kz+0.03(D/B)
Leeward wall, Cpe2D/B1Cpe2=0.5
D/B>1Cpe2=0.5+0.25ln(D/B)0.8
Pressure distribution coefficient for vertical profile, kzzzbkz=(zb/H)2α
zb<z<0.8Hkz=(z/H)2α
z0.8Hkz=0.82α

Note: B = breadth of building (perpendicular to along-wind direction) (m); D = depth of building (perpendicular to across-wind direction) (m); z = height from the ground (m); and zb = height above ground surface starting atmospheric boundary layer (m).

Across- and Torsional-Wind Loads

The across- and torsional-wind loads in MOLIT (2016) were determined by wind tunnel tests. Empirical equations for PSDs of across- and torsional-wind loads in MOLIT (2016) are shown in Tables 7 and 8, respectively. The PSD of across wind is a form of normalized PSD of overturning moment by the across-wind load, and the PSD of torsional wind is a form of base moment by torsional-wind load.
Table 7. PSD of across wind
ParametersConditionsEquations
Normalized PSD of across-wind overturning momentfSL(f)σM,L2=j=1m4kj(1+0.5βj)βjπ(f/npj)2[{1(f/npj)2}2+4βj2(f/npj)2]
Standard deviation of across-wind overturning momentσM,L=CM,LqHBH2
Across-wind fluctuating overturning moment factorCM,L=0.0073(D/B)30.0629(D/B)2+0.1959(D/B)
Number of peaks of across-wind spectrumm={1(D/B<3)2(D/B3)
Amplification factorFirst peakk1=1
Second peakk2=0.01
Normalized bandwidth factorFirst peakβ1=17.25(D/B)432.99(D/B)3+25.37(D/B)29.39(D/B)+1.4119.68(D/B)424.49(D/B)3+13.02(D/B)24.77(D/B)+1
Second peakβ2=0.31(D/B)0.41
Normalized frequencyFirst peaknp1=0.11{1+0.39(D/B)2}0.8VHB
Second peaknp2=0.61(D/B)0.89VHB

Note: SL(f) = PSD of across-wind overturning moment. Application conditions for first and second peaks are 0.2D/B5 and 3D/B5, respectively.

Table 8. PSD of torsional wind
ParametersEquations
Normalized PSD of torsional-wind momentfST(f)σT2={0.14KT2(VT*)2βTπD(B2+D2)2L2B3(VT*4.5or6VT*10)ST,4.5exp[3.5ln(ST,6ST,4.5)ln(VT*4.5)](4.5<VT*<6)
Standard deviation of torsional-wind momentσT=CTqHB2H
Torsional-wind fluctuating moment factorCT=[0.0066+0.015(D/B)2]0.78
Normalized wind speed for torsional windVT*=VHfBD
Amplification factor for torsional vibration spectrumKT={1.1(D/B)+0.97(D/B)2+0.85(D/B)+3.3+0.17(VT*4.5)0.077(D/B)0.16(D/B)20.96(D/B)+0.42+0.35(D/B)+0.095(6VT*10)
Exponential factor for normalized wind speedβT={(D/B)+3.6(D/B)25.1(D/B)+9.1+0.14(D/B)+0.14(VT*4.5)0.44(D/B)20.0064(D/B)40.26(D/B)2+0.1+0.2(6VT*10)

Note: L = larger of B and D; ST(f) = PSD of torsional-wind moment; and ST,4.5, ST,6 = normalized PSDs of torsional-wind moment when VT* are 4.5 and 6, respectively.

As shown in Table 8, application of torsional-wind PSD is limited by the range of normalized wind speed, VT*. The range limit of VT* is assumed sufficient for an estimation of a resonant component with a range of natural frequency common to high-rise buildings. The standard deviation for background and resonant components σT is calculated from the CT factor rather than the square root of integration of PSD function. But a PSD with sufficient frequency range is required in generation of time-history load. Because the value of VH used in this study is large, the PSD with VT* exceeding 10 is assumed in Eq. (6). For the low-frequency range (Fig. 7), it was determined to have a PSD shape similar to the proposed normalized PSD by Kil et al. (2007). Applicable to only certain conditions of the wind speed and size of structure used in this case study is Eq. (6)
fST(f)σT2=1.45VT*0.64(VT*>10)
(6)
Fig. 7. Normalized torsional PSDs of torsional wind.

Generation of Time-History Loads for Nonlinear Analysis

Time-history loads are generated from the PSD function as (Hwang et al. 2015)
X(t)=i=1n2S(fi)Δfcos(2πfit+θi)
(7)
where S(f) = PSD function; Δf = interval of frequency; and θ = randomly generated phase angle set (02π).
The along-wind time-history load is composed of the mean and background components. Because vertical distribution of the along-wind force is addressed in the wind force coefficient in MOLIT (2016), the along-wind time-history load can be expressed as
FD(z,t)=XP(z,t)+12ρVH2CD(z)A
(8)
where z = height from the ground; and XP = generated time history from along-wind force PSD.
The PSDs of across- and torsional-wind loads are moments at the base of the structure. The generated time histories from the base moment PSDs are needed to be converted to the story forces and story torsional moments. In an equivalent static analysis, a vertical load distribution based on a linear mode shape appears appropriate, because across- and torsional-wind loads are governed by the resonant component, i.e., linear first mode shape. However, in the case of time-history analysis, only the mean and background components are applied, and the resonant force is inherently induced during the analysis. For this study, the vertical profiles of fluctuating across- and torsional-wind loads are assumed to be uniform (Ryu et al. 2019). By assuming uniform distribution along the height, across-wind story forces and torsional-wind moments can be expressed by Eqs. (9) and (10), respectively
FL(zi,t)=2hiH2XL(t)
(9)
MT(zi,t)=hiHXT(t)
(10)
where zi and hi = height from ground and story height of the ith floor, respectively; and XL and XT = generated time histories from across- and torsional-wind PSDs, respectively.
Generated time-history loads have values at the initial start. If the loads were applied directly to a structure, the dynamic response would be overestimated due to sudden loading. Thus, a filter function, F(t), for gradual loading is required. As shown in Fig. 8, an additional 100 s of gradual loading time for the start and end of the wind load duration of 600 s was considered in this study.
Fig. 8. Filter function F(t).
The time-history load for each direction can be expressed by
W(z,t)=F(t)F(t)
(11)
where F(t) can be FD(z,t), FL(z,t), and MT(z,t).
In this study, the time step of 0.05 s and a frequency range from 0.00125 (=1/totaltimeduration) to 10 Hz (Nyquist frequency corresponding to time step) was used.

Wind Load Combination

The possibility of simultaneous occurrence of maximum loads for along-, across-, and torsional-wind loads is low. Thus, wind load combinations for static analysis given in Tables 9 and 10 are presented in MOLIT (2016). The wind load combination in MOLIT (2016) is similar to that of the AIJ (2015). According to AIJ (2015) commentary, the response correlation is negligible between along-wind load and across-wind load, and between along-wind load and torsional-wind load. Meanwhile, the wind load combination factor for across- and torsional-wind loads has been suggested based on the correlation coefficient.
Table 9. Wind load combination
Load combinationAlong windAcross windTorsional wind
1WD0.4WL0.4WT
2WD(0.4+0.6GD)WLκWT
3WD(0.4+0.6GD)κWLWT

Note: WD, WL, WT = along-, across-, and torsional-wind loads, respectively; GD = gust effect factor for along-wind load; and κ = wind load combination factor for across- and torsional-wind loads (defined in Table 10).

Table 10. Wind load combination factor for across- and torsional-wind loads
D/Baf1B/VHaκ
0.50.10.55
0.20.65
0.60.80
10.10.55
0.30.55
0.60.65
2All values0.55

Note: f1 = smaller value of first mode natural frequencies of across- and torsional-wind directions.

a
Linear interpolation shall be used for D/B and f1B/VH.
Unlike equivalent static analysis in design codes, the load combination factor is not required if the time histories are appropriately generated so that maximum values of each directional load do not occur simultaneously. Because time histories for along-, across-, and torsional-wind loads are generated based on the random phase angle θ, proper random phase angle sets for each directional load are required to control the occurrence of the maximum load in each direction in the time domain. Because the correlation between along-wind and other direction loads is negligible according to AIJ (2015), a statistically independent random phase angle set for along-wind θalong was employed. If it is assumed that the trends of PSD shapes in each frequency for across- and torsional-wind loads are the same, and the same random phase angle set is used, the maximum load for each direction will occur at the same instant. If it is assumed that the trends of PSD shape for across- and torsional-wind load are similar, the magnitude of one directional wind load at the instant of the maximum of the other wind load can be controlled by using the same random phase angle set with an additional phase lag as in the following equation:
θtorsion=θacross+cos1(κ)
(12)
where θacross and θtorsion = sets of random phase angle for across- and torsional-wind load, respectively; and κ = wind load combination factor for equivalent static analysis in Table 10.
To compare instants of the maximum value occurrence depending on random phase angle sets, 1,000 wind loads were generated. Figs. 911 show ratios to the maximum value for other directional loads at the instant of the maximum of one directional wind. Because the mean values of across- and torsional loads are zero, the ratios are determined from absolute values. Fig. 9 represents the statistically independent random phase angle sets of along-, across-, and torsional-wind loads. Fig. 10 is for the same phase angle set of across- and torsional-wind loads, and Fig. 11 is the case if Eq. (12) is used. The expected ratio (average of 1,000 generated time histories) to the maximum value of along-wind load at the instants of maximum of across- and torsional-wind loads is about 0.7. It is much larger than those of across- and torsional-wind loads due to the mean component of along-wind load (Fig. 9). When statistically independent random phase angle sets are used, the expected ratio of across- and torsional-wind loads is about 0.24. Similar results were observed for the expected ratios of across- and torsional winds when along-wind load is the maximum. When the same phase angle set is used (Fig. 10), the expected ratio is 0.9. Using the proposed method, the expected ratio is 0.5, which is close to the value of wind load combination factor κ.
Fig. 9. Ratios to maximum wind loads at the instant of maximum of one directional load when θacross and θtorsion are independent of each other: (a) instant of maximum of across wind; and (b) instant of maximum of torsional wind.
Fig. 10. Ratios to maximum wind loads at the instant of maximum of one directional load when θacross and θtorsion are the same: (a) instant of maximum of across wind; and (b) instant of maximum of torsional wind.
Fig. 11. Ratios to maximum wind loads at the instant of maximum of one directional load when the phase lag in Eq. (12) is used: (a) instant of maximum of across wind; and (b) instant of maximum of torsional wind.
Even with the proposed method, the deviation of the ratios is still large due to the difference between PSD shapes of along- and torsional-wind loads. Thus, a desirable wind load set shown in Fig. 12 was selected. The load ratios at the instant of maximum of each directional load are given in Table 11. The maximum of each directional load is close to the MOLIT (2016) equivalent static load without resonant component, and is appropriate to be used for the early stage of PBWD. Note that the maximum value in the generated time history is obtained by the superposition of cosine functions with large range of frequencies. Stochastically, the maximum value of time history can be smaller or larger than the expected value (KBC load).
Fig. 12. Generated time-history wind loads: (a) time-history force at the top by along-wind load; (b) time-history base overturning moment by across-wind load; and (c) time-history base torsional moment by torsional-wind load.
Table 11. Load ratios of generated time-history wind loads
Directional load maximum instantAlong-wind load/absolute maximumAcross-wind/absolute maximumTorsional-wind/absolute maximum
Along wind10.3030.346
Across wind0.51810.491
Torsional wind0.6120.6401

Analysis Results

Nonlinear Static Analysis

Nonlinear static analysis or pushover analysis for each model was performed to verify the yield strength, yield displacement, overstrength factor (Ω), and ductility (μ) of the system. The vertical load distribution for pushover analysis was assumed to be the first mode shape. Fig. 13 shows load-displacement curves and seismic performance by AIK (2019) from the analysis. In the building frame system, the lateral load is mainly resisted by the core walls and coupling beams. Due to the relatively short length of the coupling beams, the coupling beams are subject to larger deformation than the shear walls (Lequesne et al. 2016). Thus, inelastic behavior is concentrated in coupling beams when a core wall system is used.
Fig. 13. Load-displacement curves and seismic performances in accordance with AIK (2019).
For whole models, yield and collapse of members first occurred at the coupling beams. In the inelastic modeling of coupling beams, ductility capacity depends on the applied shear ratio. If a coupling beam is designed to have a large moment capacity due to elastic wind design (RW=1), the corresponding applied shear force also increases, which results in a reduction of ductility capacity. Moreover, excessive shear forces can be applied to joints and core walls, which leads to a brittle system. To resist the maximum considered earthquake (MCE) with 2,475-year return period, a system with excessively large strength and low ductility may be inappropriate.
For the aforementioned scenario, improved seismic details such as conforming transverse reinforcement in beams and confinement in shear walls are required during a PBSD. On the other hand, buildings designed with RW factors of 2 and 3 show relatively low yield strengths and sufficient ductility capacity. The overstrength factor and ductility of each model by pushover analysis are summarized in Table 12.
Table 12. Overstrength factor and ductility of system
Design code and analysis modelOverstrength factor Ω (strength at collapse start/yield strength)Ductility μ (displacement at collapse start/yield displacement)
MOLIT (2016) (building frame system)2.55 (μ=RE for tall buildings)
RW=11.732.29
RW=22.334.76
RW=32.635.83
The overstrength factor for the structural system in elastic seismic design in MOLIT (2016) is 2.5. The ductility of a building with a long period is the same as the response modification factor RE (Chopra 2017). Considering assumed ductility of the structural system, the design seismic load is reduced by the RE factor in the design code. However, the design result of the wind load governing case (RW=1) deviates from the initial assumption. In other words, the seismic design for the wind load governing case cannot be guaranteed under MCE due to a lack of ductility. By using a larger RW factor, the design lateral load becomes close to the design seismic load, and the intended results appear. The wind resistance performance in light of ratcheting and low-cycle fatigue should be verified by NTHA.

Nonlinear Modal Time-History Analysis

The duration of wind load is much longer than that of the seismic load. For wind load, extremely large computation is required for NTHA. Thus, the nonlinear modal time-history analysis, called fast nonlinear analysis (FNA) (Wilson 2002), was employed. Despite a deficiency that geometric nonlinearity cannot be considered, FNA was computationally effective compared to the conventional time integration method.
FNA is applicable primarily to a system with small to moderate inelastic deformation, which is suitable for a PBWD. Unlike seismic design with a large reduction of design force by the RE factor, the RW factor is limited due to mean and background components, which results in a limited inelastic behavior. For verification, a modal analysis up to 50 modes by Ritz vector was carried out. The accumulated modal participating mass ratios of 99% for x-, y-, and rz-directions were achieved. A modal damping ratio of 1.2% was used for the NTHA.
Analysis results of each model are shown in Figs. 1416. The model designed by the RW factor of 1.0 remained elastic. All vertical members remained elastic in all analysis models. In the models with RW factors of 2 and 3, plastic hinges occurred in the majority of the coupling beams and several beams for the across-wind direction. Due to mean and background components, the reduction of design member forces by along-wind load was limited, and no inelastic behavior was observed for the along-wind direction. Despite a large reduction in design across-wind load, plastic rotation of coupling beams and beams was negligible (Fig. 14) when compared with acceptance criteria (plastic rotation of 0.5%–0.6%) for IO in PBSD. Because of system redundancy, design strength reduction by the RW factor of 3 resulted in a propagation of yielded members due to force redistribution rather than an increase in plastic rotations.
Fig. 14. Maximum plastic rotations: (a) coupling beams; and (b) moment frame beams.
Fig. 15. Story forces at the maximum overturning moment by NTHA: (a) along-wind governing; and (b) across-wind governing.
Fig. 16. Maximum story displacement and drift results by NTHA: (a) story displacement; and (b) story drift.
Fig. 15 shows story forces at the maximum overturning moment for x- and y-directions. Although the same time-history loads are used for all models, the instants of maximum overturning moment occurrence are different. Vertical load distributions of governing directions in NTHA are similar to those of code-based loads regardless of occurrence instants, because the maximum loads are governed by the first mode shape. However, as shown in Fig. 15(a), those in the perpendicular direction in NTHA are difficult to predict. The gradients of vertical load distributions of along-wind governing cases in NTHA are steeper than that of code-based load due to the resonant component of the first mode shape, while the along-wind story forces in NTHA are quite smaller than those of the code-based force. This appears to be the difference in application of the simplified equation for calculating the background component of along-wind load in MOLIT (2016).
In contrast, across-wind story forces of NTHA and elastic design were close. The difference of the lower part of across wind was due to uniform load distribution of the background component and influence of high-order modes.
The maximum story displacement and drift are shown in Fig. 16. The model with the RW factor of 1 shows smaller responses than the elastic design model. The smaller response is attributed to the expected strength of concrete and the resultant expected modulus of elasticity in the NTHA model.
The models designed by RW factors of 2 and 3 show larger responses than the model with the RW factor of 1 due to yielding. Like the results of the plastic rotation of hinges, displacement and drift of the model with RW factors of 2 and 3 were almost the same. All of the nonlinear models satisfied the acceptance criterion of peak drift at the top of H/300 (=600  mm) in the ASCE prestandard (ASCE 2019).
The prestandard for PBWD (ASCE 2019) mentions that in the performance evaluation of a building, it shall be safe to ratcheting and low-cycle fatigue. The ratcheting is defined as a progressive accumulation of plastic deformation leading to P-delta instability.
Bezabeh et al. (2019) introduced inelastic behavior by reducing the wind force so that the mean wind load would be 0.3 and 0.6 of yield strength. The approach does not consider a quasi-static load of the background component, and may result in underestimation of yield strength. When this approach is used for the initial design, performance evaluation for the damage accumulation by along-wind load is required.
In this study, only the resonant component was reduced, and no inelastic behavior by along-wind was observed due to the redundancy of the system. Unlike an idealized SDOF system, which is commonly used in research, real structures in design practice have large postyield stiffness and overstrength. No significant damage accumulation was expected when using the RW factor in this design approach.
Fatigue failure can occur with a small plastic deformation due to the long duration of wind load. Low-cycle fatigue depends on the number of cycles and extent of plastic deformation. Considering the fundamental natural frequency and time duration of 600 s, the expected total number of cycles is about 135 times and the corresponding strain limit of rebar is about 0.01 (Alinejad et al. 2020). Given that rebar strain is close to its yield strain of 0.00295, plastic deformation is presumed negligible. Thus, sufficient performance is perceived with regard to low-cycle fatigue.
Performance evaluation using generated time-history wind load can be used for a preliminary PBWD. Also, time-history wind loads from wind tunnel tests can be used for verification, if necessary.

Conclusion

Preliminary PBWD of the case study RC building was carried out using time-history wind load generated from PSD functions. The resonant component was reduced by the RW factor to introduce inelastic behavior during the initial elastic design. Performance evaluations of the RW factor were carried out, and findings are summarized as follows:
1.
Use of the RW factor can effectively reduce across- and torsional-wind loads. As a result, design forces on horizontal members, especially coupling beams, are reduced significantly. The RW factor is determined such that the seismic load reduced by RE should be smaller than the along-wind load reduced by RW. In the design building case study, this was the case for all RW factors of 1, 2, and 3, in part due to the relatively small demand of seismic load.
2.
Preliminary PBWD can be carried out by time-history wind load generated from PSD functions. For an NTHA, time-history wind loads need to be generated considering gradual loading/unloading, vertical distribution of mean and background rather than resonant component’s mode shape, and maximum load occurrence.
3.
Generation of wind load time histories is cost-effective; nonetheless, final performance verification by wind tunnel test results is recommended due to the following limitations: (1) lack of precise PSD functions for various shapes of structures and wind directions, (2) uncertainties in the local wind pressure, and (3) aerodynamic instability (an aeroelastic wind tunnel test is occasionally required).
4.
Using pushover analysis, the ductility of the system increased when RW factors of 2 and 3 were employed. By reducing the excessive strength of fuse elements such as coupling beams, seismic performance under MCE in PBSD can be improved.
5.
Coupling beams were vulnerable under time-history wind loads, but inelastic behavior is limited due to strength reduction factor ϕ, the detailing requirement of minimum rebar ratio and maximum spacing, rounding up in member design, and expected material strength in nonlinear analysis.
6.
Buildings designed using RW factors of 2 and 3 satisfy the IO performance objective and drift limit in ASCE (2019). No ratcheting and low-cycle fatigue failure is expected due to limited inelastic deformation.

Appendix. Design Results of Main Structural Members

A summary of design results by RW factors of 1, 2, and 3 are shown in Tables 1322. IDs of members are shown in Fig. 1. The design results may vary by engineers, and there is room for optimization.
Table 13. Design results of moment frame beams B1
StoryReinforcing bar ratio (%)
RW=1RW=2RW=3
ρtopρbottomρvρtopρbottomρvρtopρbottomρv
43–450.430.230.080.34 (20.8%)0.23 (0%)0.08 (0%)0.34 (20.8%)0.23 (0%)0.08 (0%)
40–420.400.200.080.34 (14.9%)0.17 (16.3%)0.08 (0%)0.34 (14.9%)0.14 (30.0%)0.08 (0%)
37–390.430.260.080.34 (20.8%)0.17 (33.3%)0.08 (0%)0.34 (20.8%)0.14 (44.3%)0.08 (0%)
34–360.460.260.080.34 (25.9%)0.17 (33.3%)0.08 (0%)0.34 (25.9%)0.14 (44.3%)0.08 (0%)
31–330.490.280.080.34 (30.4%)0.17 (40.1%)0.08 (0%)0.34 (30.4%)0.14 (50.0%)0.08 (0%)
28–300.490.280.080.34 (30.4%)0.20 (29.9%)0.08 (0%)0.34 (30.4%)0.14 (50.0%)0.08 (0%)
25–270.520.320.080.34 (34.4%)0.20 (36.8%)0.08 (0%)0.34 (34.4%)0.17 (46.0%)0.08 (0%)
22–240.520.340.080.34 (34.4%)0.20 (42.0%)0.08 (0%)0.34 (34.4%)0.17 (50.4%)0.08 (0%)
19–210.520.340.080.34 (34.4%)0.20 (42.0%)0.08 (0%)0.34 (34.4%)0.17 (50.4%)0.08 (0%)
16–180.520.340.080.34 (34.4%)0.20 (42.0%)0.08 (0%)0.34 (34.4%)0.17 (50.4%)0.08 (0%)
13–150.490.320.080.34 (30.4%)0.20 (36.8%)0.08 (0%)0.34 (30.4%)0.17 (46.0%)0.08 (0%)
10–120.460.280.080.34 (25.9%)0.20 (29.9%)0.08 (0%)0.34 (25.9%)0.17 (40.1%)0.08 (0%)
7–90.430.260.080.34 (20.8%)0.17 (33.3%)0.08 (0%)0.34 (20.8%)0.14 (44.3%)0.08 (0%)
4–60.370.230.080.34 (8.0%)0.14 (37.4%)0.08 (0%)0.34 (8.0%)0.11 (49.8%)0.08 (0%)
1–30.340.170.080.32 (8.2%)0.11 (32.9%)0.08 (0%)0.32 (8.2%)0.11 (32.9%)0.08 (0%)

Note: Values in parentheses are reduced percentages in comparison to the results of RW factor of 1.

Table 14. Design results of moment frame beams B2
StoryReinforcing bar ratio (%)
RW=1RW=2RW=3
ρtopρbottomρvρtopρbottomρvρtopρbottomρv
43–450.320.140.080.26 (19.0%)0.11 (19.7%)0.08 (0%)0.26 (19.0%)0.11 (19.7%)0.08 (0%)
40–420.340.170.080.26 (25.7%)0.11 (32.9%)0.08 (0%)0.23 (33.8%)0.11 (32.9%)0.08 (0%)
37–390.340.170.080.26 (25.7%)0.11 (32.9%)0.08 (0%)0.23 (33.8%)0.11 (32.9%)0.08 (0%)
34–360.340.200.080.28 (17.2%)0.11 (42.7%)0.08 (0%)0.23 (33.8%)0.11 (42.7%)0.08 (0%)
31–330.340.210.080.28 (17.2%)0.11 (44.9%)0.08 (0%)0.26 (25.7%)0.11 (44.9%)0.08 (0%)
28–300.340.260.080.32 (8.2%)0.14 (44.3%)0.08 (0%)0.26 (25.7%)0.11 (55.3%)0.08 (0%)
25–270.340.320.080.32 (8.2%)0.17 (46.0%)0.08 (0%)0.26 (25.7%)0.11 (63.8%)0.08 (0%)
22–240.370.340.080.34 (8.0%)0.17 (50.4%)0.08 (0%)0.28 (23.9%)0.14 (58.6%)0.08 (0%)
19–210.370.340.080.34 (8.0%)0.17 (50.4%)0.08 (0%)0.28 (23.9%)0.14 (58.6%)0.08 (0%)
16–180.370.340.080.34 (8.0%)0.17 (50.4%)0.08 (0%)0.28 (23.9%)0.14 (58.6%)0.08 (0%)
13–150.370.340.080.34 (8.0%)0.17 (50.4%)0.08 (0%)0.28 (23.9%)0.14 (58.6%)0.08 (0%)
10–120.340.290.080.32 (8.2%)0.17 (41.6%)0.08 (0%)0.28 (17.2%)0.14 (51.2%)0.08 (0%)
7–90.340.280.080.28 (17.2%)0.14 (50.0%)0.08 (0%)0.26 (25.7%)0.11 (59.9%)0.08 (0%)
4–60.340.230.080.26 (25.7%)0.11 (49.8%)0.08 (0%)0.23 (33.8%)0.11 (49.8%)0.08 (0%)
1–30.320.140.080.20 (36.8%)0.11 (19.7%)0.08 (0%)0.20 (36.8%)0.11 (19.7%)0.08 (0%)

Note: Values in parentheses are reduced percentages in comparison to the results of RW factor of 1.

Table 15. Design results of moment frame beams B3
StoryReinforcing bar ratio (%)
RW=1RW=2RW=3
ρtopρbottomρvρtopρbottomρvρtopρbottomρv
43–450.460.260.080.34 (25.9%)0.20 (22.0%)0.08 (0%)0.34 (25.9%)0.20 (22.0%)0.08 (0%)
40–420.430.280.080.34 (20.8%)0.17 (40.1%)0.08 (0%)0.34 (20.8%)0.14 (50.0%)0.08 (0%)
37–390.460.280.080.34 (25.9%)0.17 (40.1%)0.08 (0%)0.34 (25.9%)0.14 (50.0%)0.08 (0%)
34–360.490.320.080.34 (30.4%)0.20 (36.8%)0.08 (0%)0.34 (30.4%)0.14 (54.9%)0.08 (0%)
31–330.490.320.080.34 (30.4%)0.20 (36.8%)0.08 (0%)0.34 (30.4%)0.17 (46.0%)0.08 (0%)
28–300.520.340.080.34 (34.4%)0.20 (42.0%)0.08 (0%)0.34 (34.4%)0.17 (50.4%)0.08 (0%)
25–270.520.340.080.34 (34.4%)0.23 (33.8%)0.08 (0%)0.34 (34.4%)0.17 (50.4%)0.08 (0%)
22–240.550.340.080.34 (37.9%)0.23 (33.8%)0.08 (0%)0.34 (37.9%)0.17 (50.4%)0.08 (0%)
19–210.550.340.080.34 (37.9%)0.23 (33.8%)0.08 (0%)0.34 (37.9%)0.17 (50.4%)0.08 (0%)
16–180.520.340.080.34 (34.4%)0.23 (33.8%)0.08 (0%)0.34 (34.4%)0.17 (50.4%)0.08 (0%)
13–150.520.340.080.34 (34.4%)0.20 (42.0%)0.08 (0%)0.34 (34.4%)0.17 (50.4%)0.08 (0%)
10–120.490.320.080.34 (30.4%)0.20 (36.8%)0.08 (0%)0.34 (30.4%)0.17 (46.0%)0.08 (0%)
7–90.430.280.080.34 (20.8%)0.17 (40.1%)0.08 (0%)0.34 (20.8%)0.14 (50.0%)0.08 (0%)
4–60.370.230.080.34 (8.0%)0.14 (37.4%)0.08 (0%)0.34 (8.0%)0.14 (37.4%)0.08 (0%)
1–30.340.170.080.32 (8.2%)0.11 (32.9%)0.08 (0%)0.28 (17.2%)0.11 (32.9%)0.08 (0%)

Note: Values in parentheses are reduced percentages in comparison to the results of RW factor of 1.

Table 16. Design results of moment frame beams B4
StoryReinforcing bar ratio (%)
RW=1RW=2RW=3
ρtopρbottomρvρtopρbottomρvρtopρbottomρv
43–451.440.700.241.26 (12.5%)0.52 (24.9%)0.24 (0%)1.16 (19.4%)0.52 (24.9%)0.24 (0%)
40–421.260.700.121.06 (15.9%)0.39 (44.6%)0.08 (33.3%)0.99 (21.1%)0.39 (44.6%)0.08 (33.3%)
37–391.260.700.121.06 (15.9%)0.39 (44.6%)0.08 (33.3%)0.99 (21.1%)0.39 (44.6%)0.08 (33.3%)
34–361.260.700.121.06 (15.9%)0.39 (44.6%)0.08 (33.3%)0.99 (21.1%)0.39 (44.6%)0.08 (33.3%)
31–331.260.700.121.06 (15.9%)0.39 (44.6%)0.08 (33.3%)0.99 (21.1%)0.39 (44.6%)0.08 (33.3%)
28–301.260.700.120.99 (21.1%)0.39 (44.6%)0.08 (33.3%)0.93 (26.5%)0.39 (44.6%)0.08 (33.3%)
25–271.260.700.120.99 (21.1%)0.39 (44.6%)0.08 (33.3%)0.93 (26.5%)0.39 (44.6%)0.08 (33.3%)
22–241.260.700.120.99 (21.1%)0.39 (44.6%)0.08 (33.3%)0.93 (26.5%)0.39 (44.6%)0.08 (33.3%)
19–211.260.700.120.99 (21.1%)0.39 (44.6%)0.08 (33.3%)0.86 (32.0%)0.39 (44.6%)0.08 (33.3%)
16–181.160.700.120.93 (20.2%)0.39 (44.6%)0.08 (33.3%)0.83 (28.9%)0.36 (48.9%)0.08 (33.3%)
13–151.160.700.120.86 (26.1%)0.39 (44.6%)0.08 (33.3%)0.78 (32.9%)0.36 (48.9%)0.08 (33.3%)
10–121.060.520.080.83 (22.2%)0.36 (30.8%)0.08 (0%)0.73 (31.0%)0.36 (30.8%)0.08 (0%)
7–90.930.520.080.73 (21.1%)0.36 (30.8%)0.08 (0%)0.64 (31.2%)0.36 (30.8%)0.08 (0%)
4–60.780.400.080.59 (24.2%)0.27 (33.4%)0.08 (0%)0.55 (29.0%)0.26 (36.4%)0.08 (0%)
1–30.550.280.080.46 (16.1%)0.23 (20.1%)0.08 (0%)0.43 (21.6%)0.23 (20.1%)0.08 (0%)

Note: Values in parentheses are reduced percentages in comparison to the results of RW factor of 1.

Table 17. Design results of moment frame beams B5
StoryReinforcing bar ratio (%)
RW=1RW=2RW=3
ρtopρbottomρvρtopρbottomρvρtopρbottomρv
43–451.350.700.241.16 (14.1%)0.52 (24.9%)0.24 (0%)1.16 (14.1%)0.52 (24.9%)0.24 (0%)
40–421.160.700.080.99 (14.3%)0.39 (44.6%)0.08 (0%)0.93 (20.2%)0.39 (44.6%)0.08 (0%)
37–391.160.700.080.99 (14.3%)0.39 (44.6%)0.08 (0%)0.93 (20.2%)0.39 (44.6%)0.08 (0%)
34–361.160.700.080.99 (14.3%)0.39 (44.6%)0.08 (0%)0.93 (20.2%)0.39 (44.6%)0.08 (0%)
31–331.160.700.080.99 (14.3%)0.39 (44.6%)0.08 (0%)0.93 (20.2%)0.39 (44.6%)0.08 (0%)
28–301.160.700.080.93 (20.2%)0.39 (44.6%)0.08 (0%)0.86 (26.1%)0.39 (44.6%)0.08 (0%)
25–271.160.700.080.93 (20.2%)0.39 (44.6%)0.08 (0%)0.86 (26.1%)0.39 (44.6%)0.08 (0%)
22–241.160.700.080.93 (20.2%)0.39 (44.6%)0.08 (0%)0.83 (28.9%)0.36 (48.9%)0.08 (0%)
19–211.160.700.080.86 (26.1%)0.39 (44.6%)0.08 (0%)0.83 (28.9%)0.36 (48.9%)0.08 (0%)
16–181.070.700.080.83 (22.9%)0.36 (48.9%)0.08 (0%)0.78 (27.3%)0.36 (48.9%)0.08 (0%)
13–151.060.520.080.83 (22.2%)0.36 (30.8%)0.08 (0%)0.73 (31.0%)0.36 (30.8%)0.08 (0%)
10–120.990.520.080.73 (26.5%)0.36 (30.8%)0.08 (0%)0.68 (31.2%)0.36 (30.8%)0.08 (0%)
7–90.860.520.080.68 (20.2%)0.36 (30.8%)0.08 (0%)0.64 (25.7%)0.36 (30.8%)0.08 (0%)
4–60.730.400.080.58 (20.4%)0.28 (29.2%)0.08 (0%)0.55 (24.5%)0.26 (36.4%)0.08 (0%)
1–30.550.260.080.46 (16.1%)0.23 (11.0%)0.08 (0%)0.43 (21.6%)0.23 (11.0%)0.08 (0%)

Note: Values in parentheses are reduced percentages in comparison to the results of RW factor of 1.

Table 18. Design results of moment frame column C1
StoryReinforcing bar ratio (%)
RW=1RW=2RW=3
ρρvρρvρρv
43–450.710.090.55 (22.5%)0.09 (0%)0.55 (22.5%)0.09 (0%)
40–420.710.090.71 (0%)0.09 (0%)0.71 (0%)0.09 (0%)
37–390.710.090.71 (0%)0.09 (0%)0.71 (0%)0.09 (0%)
34–360.710.090.71 (0%)0.09 (0%)0.71 (0%)0.09 (0%)
31–330.710.090.71 (0%)0.09 (0%)0.71 (0%)0.09 (0%)
28–300.710.090.71 (0%)0.09 (0%)0.71 (0%)0.09 (0%)
25–270.710.090.71 (0%)0.09 (0%)0.71 (0%)0.09 (0%)
22–240.610.070.61 (0%)0.07 (0%)0.61 (0%)0.07 (0%)
19–210.610.070.61 (0%)0.07 (0%)0.61 (0%)0.07 (0%)
16–180.610.070.61 (0%)0.07 (0%)0.61 (0%)0.07 (0%)
13–150.610.070.61 (0%)0.07 (0%)0.61 (0%)0.07 (0%)
10–120.610.070.61 (0%)0.07 (0%)0.61 (0%)0.07 (0%)
7–91.700.070.61 (64.1%)0.07 (0%)0.61 (64.1%)0.07 (0%)
4–62.230.071.38 (38.1%)0.07 (0%)1.03 (53.8%)0.07 (0%)
1–32.570.071.70 (33.9%)0.07 (0%)1.70 (33.9%)0.07 (0%)

Note: Values in parentheses are reduced percentages in comparison to the results of RW factor of 1.

Table 19. Design results of moment frame column C2
StoryReinforcing bar ratio (%)
RW=1RW=2RW=3
ρρvρρvρρv
43–450.710.090.71 (0%)0.09 (0%)0.71 (0%)0.09 (0%)
40–420.710.090.71 (0%)0.09 (0%)0.71 (0%)0.09 (0%)
37–390.710.090.71 (0%)0.09 (0%)0.71 (0%)0.09 (0%)
34–360.710.090.71 (0%)0.09 (0%)0.71 (0%)0.09 (0%)
31–330.710.090.71 (0%)0.09 (0%)0.71 (0%)0.09 (0%)
28–300.710.090.71 (0%)0.09 (0%)0.71 (0%)0.09 (0%)
25–270.710.090.71 (0%)0.09 (0%)0.71 (0%)0.09 (0%)
22–240.610.070.61 (0%)0.07 (0%)0.61 (0%)0.07 (0%)
19–210.610.070.61 (0%)0.07 (0%)0.61 (0%)0.07 (0%)
16–180.610.070.61 (0%)0.07 (0%)0.61 (0%)0.07 (0%)
13–150.610.070.61 (0%)0.07 (0%)0.61 (0%)0.07 (0%)
10–121.380.070.61 (55.8%)0.07 (0%)0.61 (55.8%)0.07 (0%)
7–92.230.071.03 (53.8%)0.07 (0%)0.61 (72.6%)0.07 (0%)
4–62.570.072.23 (13.2%)0.07 (0%)1.70 (33.9%)0.07 (0%)
1–32.570.072.57 (0%)0.07 (0%)2.23 (13.2%)0.07 (0%)

Note: Values in parentheses are reduced percentages in comparison to the results of RW factor of 1.

Table 20. Design results of moment frame column C3
StoryReinforcing bar ratio (%)
RW=1RW=2RW=3
ρρvρρvρρv
43–450.550.090.55 (0%)0.09 (0%)0.55 (0%)0.09 (0%)
40–420.710.090.71 (0%)0.09 (0%)0.71 (0%)0.09 (0%)
37–390.710.090.71 (0%)0.09 (0%)0.71 (0%)0.09 (0%)
34–360.710.090.71 (0%)0.09 (0%)0.71 (0%)0.09 (0%)
31–330.710.090.71 (0%)0.09 (0%)0.71 (0%)0.09 (0%)
28–300.710.090.71 (0%)0.09 (0%)0.71 (0%)0.09 (0%)
25–270.710.090.71 (0%)0.09 (0%)0.71 (0%)0.09 (0%)
22–240.610.070.61 (0%)0.07 (0%)0.61 (0%)0.07 (0%)
19–210.610.070.61 (0%)0.07 (0%)0.61 (0%)0.07 (0%)
16–180.610.070.61 (0%)0.07 (0%)0.61 (0%)0.07 (0%)
13–150.610.070.61 (0%)0.07 (0%)0.61 (0%)0.07 (0%)
10–121.030.070.61 (40.8%)0.07 (0%)0.61 (40.8%)0.07 (0%)
7–92.230.070.61 (72.6%)0.07 (0%)0.61 (72.6%)0.07 (0%)
4–62.570.071.70 (33.9%)0.07 (0%)1.38 (46.3%)0.07 (0%)
1–32.570.072.57 (0%)0.07 (0%)2.23 (13.2%)0.07 (0%)

Note: Values in parentheses are reduced percentages in comparison to the results of RW factor of 1.

Table 21. Design results of coupling beam CB
StoryReinforcing bar ratio (%)
RW=1RW=2RW=3
ρtop=ρbottomρvρtop=ρbottomρvρtop=ρbottomρv
43–451.020.440.56 (45.1%)0.11 (75.0%)0.43 (57.8%)0.08 (81.8%)
40–421.110.440.61 (45.0%)0.11 (75.0%)0.46 (58.6%)0.08 (81.8%)
37–391.220.660.67 (45.1%)0.28 (57.6%)0.51 (58.2%)0.16 (75.8%)
34–361.390.660.76 (45.3%)0.28 (57.6%)0.56 (59.7%)0.16 (75.8%)
31–331.540.660.82 (46.8%)0.29 (56.1%)0.62 (59.7%)0.19 (71.2%)
28–301.690.880.90 (46.7%)0.44 (50.0%)0.67 (60.4%)0.28 (68.2%)
25–271.780.880.98 (44.9%)0.44 (50.0%)0.73 (59.0%)0.28 (68.2%)
22–241.990.881.02 (48.7%)0.44 (50.0%)0.77 (61.3%)0.28 (68.2%)
19–211.990.881.02 (48.7%)0.44 (50.0%)0.82 (58.8%)0.29 (67.0%)
16–181.990.881.02 (48.7%)0.44 (50.0%)0.82 (58.8%)0.29 (67.0%)
13–151.990.881.02 (48.7%)0.44 (50.0%)0.82 (58.8%)0.29 (67.0%)
10–121.780.880.94 (47.2%)0.44 (50.0%)0.77 (56.7%)0.28 (68.2%)
7–91.540.660.82 (46.8%)0.29 (56.1%)0.70 (54.5%)0.28 (57.6%)
4–61.170.660.61 (47.9%)0.16 (75.8%)0.53 (54.7%)0.16 (75.8%)
1–30.640.280.36 (43.8%)0.08 (71.4%)0.36 (43.8%)0.08 (71.4%)

Note: Values in parentheses are reduced percentages in comparison to the results of RW factor of 1.

Table 22. Design results of main core walls MCW
StoryReinforcing bar ratio (%)
RW=1RW=2RW=3
ρverticalρhorizontalρverticalρhorizontalρverticalρhorizontal
43–450.160.280.16 (0%)0.28 (0%)0.16 (0%)0.28 (0%)
40–420.160.280.16 (0%)0.28 (0%)0.16 (0%)0.28 (0%)
37–390.160.280.16 (0%)0.28 (0%)0.16 (0%)0.28 (0%)
34–360.280.280.16 (42.9%)0.28 (0%)0.16 (42.9%)0.28 (0%)
31–330.280.280.16 (42.9%)0.28 (0%)0.16 (42.9%)0.28 (0%)
28–300.280.280.16 (42.9%)0.28 (0%)0.16 (42.9%)0.28 (0%)
25–270.280.280.16 (42.9%)0.28 (0%)0.16 (42.9%)0.28 (0%)
22–240.280.280.16 (42.9%)0.28 (0%)0.16 (42.9%)0.28 (0%)
19–210.280.280.16 (42.9%)0.28 (0%)0.16 (42.9%)0.28 (0%)
16–180.280.280.16 (42.9%)0.28 (0%)0.16 (42.9%)0.28 (0%)
13–150.280.280.16 (42.9%)0.28 (0%)0.16 (42.9%)0.28 (0%)
10–120.280.280.16 (42.9%)0.28 (0%)0.16 (42.9%)0.28 (0%)
7–90.630.280.16 (74.6%)0.28 (0%)0.16 (74.6%)0.28 (0%)
4–61.740.280.16 (90.8%)0.28 (0%)0.16 (90.8%)0.28 (0%)
1–32.500.280.28 (88.8%)0.28 (0%)0.16 (93.6%)0.28 (0%)

Note: Values in parentheses are reduced percentages in comparison with to the results of RW factor of 1.

Data Availability Statement

All data, models, and code generated or used during the study appear in the published article.

Acknowledgments

This work was supported by the Korea Agency for Infrastructure Technology Advancement (KAIA) grant funded by the Ministry of Land, Infrastructure and Transport (Grant No. 21CTAP-C151831-03). Also, the support by the Institute of Construction and Environmental Engineering at Seoul National University is acknowledged.

References

Abdullah, S. A., K. Aswegan, S. Jaberansari, R. Klemencic, and J. W. Wallace. 2020. “Performance of reinforced concrete coupling beams subjected to simulated wind loading.” ACI Struct. J. 117 (3): 283–295.
ACI (American Concrete Institute). 2006. Performance-based design of concrete building for wind loads (SP-240). Farmington Hills, MI: ACI.
ACI (American Concrete Institute). 2019. Building code requirements for structural concrete. Farmington Hills, MI: ACI.
AIJ (Architectural Institute of Japan). 2015. AIJ recommendations for loads on buildings. Tokyo: AIJ.
AIK (Architectural Institute of Korea). 2019. Guidelines for performance-based seismic design of residential buildings, 64. Seoul, South Korea: AIK.
Alinejad, H., S. Y. Jeong, and T. H.-K. Kang. 2020. “Performance-based design of tall buildings for wind load and application of response modification factor.” Wind Struct. 31 (2): 153–164.
Alinejad, H., and T. H.-K. Kang. 2020. “Engineering review of ASCE 7-16 wind-load provisions and wind effect on tall concrete-frame buildings.” J. Struct. Eng. 146 (6): 04020100. https://doi.org/10.1061/(ASCE)ST.1943-541X.0002622.
ASCE. 2017a. Minimum design loads and associated criteria for buildings and other structures. ASCE 7-16. Reston, VA: ASCE.
ASCE. 2017b. Seismic evaluation and retrofit of existing buildings. ASCE 41-17. Reston, VA: ASCE.
ASCE. 2019. Prestandard for performance-based wind design, 127. Reston, VA: ASCE.
Aswegan, K., R. Larsen, R. Klemencic, J. Hooper, and J. Hasselbauer. 2017. “Performance-based wind and seismic engineering: Benefits of considering multiple hazards.” In Proc., Structures Congress 2017, 473–484. Reston, VA: ASCE.
Bezabeh, M., G. Bitsuamlak, and S. Tesfamariam. 2019. “The first step towards the performance-based design of wind-excited buildings: A critical examination of ductility-demand, hysteretic energy, and control of damage accumulation.” In Proc., 15th Int. Conf. on Wind Engineering, 1–4. Atsugi, Japan: International Association for Wind Engineering.
Bezabeh, M. A., G. T. Bitsuamlak, and S. Tesfamariam. 2020. “Performance-based wind design of tall buildings: Concepts, frameworks, and opportunities.” Wind Struct. 31 (2): 103–142.
Chopra, A. K. 2017. Dynamics of structures: Theory and applications to earthquake engineering. London: Pearson.
Ciampoli, M., F. Petrini, and G. Augusti. 2011. “Performance-based wind engineering: Towards a general procedure.” Struct. Saf. 33 (6): 367–378. https://doi.org/10.1016/j.strusafe.2011.07.001.
Davenport, A. G. 1967. “Gust loading factors.” J. Struct. Div. 93 (3): 11–34. https://doi.org/10.1061/JSDEAG.0001692.
Dowell, R. K., F. Seible, and E. L. Wilson. 1998. “Pivot hysteresis model for reinforced concrete members.” ACI Struct. J. 95 (5): 607–617.
El Damatty, A. A., and F. Y. Elezaby. 2018. “The integration of wind and structural engineering.” In Proc., The 2018 World Congress on Advances in Civil, Environmental, and Materials Research (ACEM18), 27–31. Daejeon, South Korea: International Association of Structural Engineering and Mechanics.
Elezaby, F., and A. El Damatty. 2020. “Ductility-based design approach of tall buildings under wind loads.” Wind Struct. 31 (2): 143–152.
Ha, Y.-C. 2017. Evaluation of the along-wind load and response of high-rise buildings. [In Korean.]. Seongnam, South Korea: MIDAS Information Technology.
Ha, Y.-C., Y.-S. Kil, S.-K. Hwang, and D.-W. Kim. 2007. “Simplified formulas for estimating the across-wind induced response of rectangular tall buildings.” [In Korean.] J. Archit. Inst. Korea Struct. Constr. 23 (6): 39–48.
Hwang, J. S., S. H. Lee, and Y. C. Ha. 2015. “Time history analysis of building structure using load generation of design wind load spectrum.” [In Korean.] J. Wind Eng. Inst. Korea 19 (2): 43–49.
ISO. 2009. Wind actions on structures. ISO 4354. Geneva: ISO.
Kang, T. H.-K., S. Y. Jeong, and H. Alinejad. 2019. “Understanding of wind load determination according to KBC 2016 and its application to high-rise buildings.” [In Korean.] J. Wind Eng. Inst. Korea 23 (2): 83–89.
Kang, T. H.-K., R. D. Martin, H.-G. Park, R. Wilkerson, and N. Youssef. 2013. “Tall building with steel plate shear walls subject to load reversal.” Struct. Des. Tall Spec. Build. 22 (6): 500–520. https://doi.org/10.1002/tal.700.
Kang, T. H.-K., J. W. Wallace, and K. J. Elwood. 2009. “Nonlinear modeling of flat-plate systems.” J. Struct. Eng. 135 (2): 147–158. https://doi.org/10.1061/(ASCE)0733-9445(2009)135:2(147).
Karman, T. 1948. “Progress in the statistical theory of turbulence.” Proc. Natl. Acad. Sci. 34 (11): 530–539. https://doi.org/10.1073/pnas.34.11.530.
Kil, Y.-S., D.-W. Kim, H.-R. Kim, and Y.-C. Ha. 2007. “Torsional-wind fluctuating moment coefficient and power spectral density coefficient for estimating torsional-wind load of tall buildings.” [In Korean.] J. Wind Eng. Inst. Korea 11 (2): 187–194.
Lequesne, R. D., G. J. Parra-Montesinos, and J. K. Wight. 2016. “Seismic response of fiber-reinforced concrete coupled walls.” ACI Struct. J. 113 (3): 435–445.
Marukawa, H., T. Ohkuma, and Y. Momomura. 1992. “Across-wind and torsional acceleration of prismatic high rise buildings.” J. Wind Eng. Ind. Aerodyn. 42 (1–3): 1139–1150. https://doi.org/10.1016/0167-6105(92)90121-P.
Mohammadi, A., A. Azizinamini, L. Griffis, and P. Irwind. 2019. “Performance assessment of an existing 47-story high-rise building under extreme wind loads.” J. Struct. Eng. 145 (1): 04018232. https://doi.org/10.1061/(ASCE)ST.1943-541X.0002239.
MOLIT (Ministry of Land, Infrastructure and Transport). 2016. Korean building code. KBC 2016. Sejong, South Korea: MOLIT (Korea).
Nakai, M., K. Hirakawa, M. Yamanaka, H. Okuda, and A. Konishi. 2013. “Performance-based wind-resistant design for high-rise structures in Japan.” Int. J. High-Rise Build. 2 (3): 271–283.
Paulotto, C., M. Ciampoli, and G. Augusti. 2004. “Some proposals for a first step towards a performance based wind engineering.” In Proc., of the IFED-Int. Forum in Engineering Decision Making. Stoos, Switzerland: International Forum on Engineering Decision Making.
PEER/ATC (Pacific Earthquake Engineering Research Center/Applied Technology Council). 2010. Modeling and acceptance criteria for seismic design and analysis of tall buildings. PEER/ATC 72-1. Redwood City, CA: PEER/ATC.
Petrini, F., and M. Ciampoli. 2012. “Performance-based wind design of tall buildings.” Struct. Infrastruct. Eng. 8 (10): 954–966.
Ryu, H.-J., D.-H. Shin, and Y.-C. Ha. 2019. “The wind load evaluation on building considering vertical profile of fluctuating wind force.” J. Archit. Inst. Korea Struct. Constr. 35 (7): 157–164.
Shinozuka, M., and G. Deodatis. 1991. “Simulation of stochastic processes by spectral representation.” Appl. Mech. Rev. 44 (4): 191–204. https://doi.org/10.1115/1.3119501.
Takeda, T., M. A. Sozen, and N. N. Nielson. 1970. “Reinforced concrete response to simulated earthquakes.” J. Struct. Div. 96 (12): 2557–2573. https://doi.org/10.1061/JSDEAG.0002765.
Tamura, Y., H. Kawai, Y. Uematsu, H. Marukawa, K. Fujii, and Y. Taniike. 1996. “Wind load and wind-induced response estimations in the recommendations for loads on buildings, AIJ 1993.” Eng. Struct. 18 (6): 399–411. https://doi.org/10.1016/0141-0296(95)00121-2.
TBI (Tall Buildings Initiative). 2017. Guidelines for performance-based seismic design of tall buildings. Berkeley, CA: Pacific Earthquake Engineering Research Center, Univ. of California.
Thomsen, J. H., IV, and J. W. Wallace. 1995. Displacement-based design of reinforced concrete structural walls: An experimental investigation of walls with rectangular and T-shaped cross-sections. Potsdam, NY: Dept. of Civil and Environmental Engineering, Clarkson Univ.
van de Lindt, J. W., and T. N. Dao. 2009. “Performance-based wind engineering for wood-frame buildings.” J. Struct. Eng. 135 (2): 169–177. https://doi.org/10.1061/(ASCE)0733-9445(2009)135:2(169).
Wilson, E. L. 2002. Three-dimensional static and dynamic analysis of structures. Berkeley, CA: Computers and Structures.
Xiao, Y., A. Esmaeily-Ghasemabadi, and H. Wu. 1999. “High-strength concrete short beams subjected to cyclic shear.” ACI Struct. J. 96 (3): 392–400.

Information & Authors

Information

Published In

Go to Journal of Structural Engineering
Journal of Structural Engineering
Volume 147Issue 9September 2021

History

Received: Sep 12, 2020
Accepted: Mar 22, 2021
Published online: Jul 6, 2021
Published in print: Sep 1, 2021
Discussion open until: Dec 6, 2021

Authors

Affiliations

Seung Yong Jeong [email protected]
Ph.D. Student, Dept. of Architecture and Architectural Engineering, Seoul National Univ., 1 Gwanak-ro, Gwanak-gu, Seoul 08826, Korea. Email: [email protected]
Hamidreza Alinejad [email protected]
Ph.D. Student, Dept. of Architecture and Architectural Engineering, Seoul National Univ., 1 Gwanak-ro, Gwanak-gu, Seoul 08826, Korea. Email: [email protected]
Thomas H.-K. Kang, M.ASCE [email protected]
Professor, Dept. of Architecture and Architectural Engineering, Seoul National Univ., 1 Gwanak-ro, Gwanak-gu, Seoul 08826, Korea (corresponding author). Email: [email protected]

Metrics & Citations

Metrics

Citations

Download citation

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. Simply select your manager software from the list below and click Download.

Cited by

  • Performance-Based Wind Design of Tall Buildings Considering Corner Modification and Inelastic Behavior, Journal of Structural Engineering, 10.1061/JSENDH.STENG-13100, 150, 7, (2024).
  • Structural Performance of Precast Concrete Column Joint with Clamped Headed Bar during Construction, ACI Structural Journal, 10.14359/51738667, 120, 3, (2023).
  • Improved Method of Determination of Basic Wind Speed with Terrain Effects Using Graph Neural Network, Journal of Structural Engineering, 10.1061/JSENDH.STENG-11367, 149, 1, (2023).
  • The Influence of Wind Direction on the Inelastic Responses of a Base-Isolated Square Section High-Rise Building, Buildings, 10.3390/buildings12081208, 12, 8, (1208), (2022).
  • Post-Tensioned Self-Centering System Efficiency against Extreme Wind Loads, ACI Structural Journal, 10.14359/51734900, 119, 5, (2022).
  • Nonlinear Wind and Earthquake Loads on Tall Steel-Braced Frame Buildings, Journal of Structural Engineering, 10.1061/(ASCE)ST.1943-541X.0003375, 148, 8, (2022).
  • Consideration of terrain features from satellite imagery in machine learning of basic wind speed, Building and Environment, 10.1016/j.buildenv.2022.108866, 213, (108866), (2022).

View Options

Media

Figures

Other

Tables

Share

Share

Copy the content Link

Share with email

Email a colleague

Share