(2024) Crosswind Response of Base Isolated Tall Buidlings

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

Engineering Structures 305 (2024) 117722

Contents lists available at ScienceDirect

Engineering Structures
journal homepage: www.elsevier.com/locate/engstruct

Predicting crosswind response of tall buildings: Base isolation and


nonlinear aeroelastic effects
Guoqing Huang a, Yuhang Fan a, Xinzhong Chen b, *, Xudong Yang a, Zhihao Li c
a
School of Civil Engineering, Chongqing University, Chongqing, China
b
National Wind Institute, Department of Civil, Environmental and Construction Engineering, Texas Tech University, Lubbock, USA
c
Institute for Smart City of Chongqing University in Liyang, Chongqing University, Jiangsu, China

A R T I C L E I N F O A B S T R A C T

Keywords: This study introduces an analysis framework to evaluate the crosswind response of tall buildings under the in­
Base-isolated tall buildings fluence of base isolation and nonlinear aeroelastic effect near the vortex lock-in wind speed. The motion-induced
Vortex-induced vibration and buffeting forces are determined based on those in a building with a fixed base. The primary focus is on
Aerodynamic damping
modeling the motion-induced force in tall buildings with base isolation, where no pre-existing model exists. The
Motion-induced wind force
fundamental modal displacement is used to describe the building motion relative to the base. The Bouc-Wen
model is employed to capture the hysteresis in the relation of shear force and displacement of the isolation
system. Through response history analysis, the response statistics of a 65-story building with a square cross-
section are determined and compared with those of a fixed-base building. An investigation is conducted to
examine the impact of different isolation layer parameters on building response. Additionally, this study explores
how the motion-induced force model influences the building response. The findings indicate that base isolation
effectively mitigates nonlinear aeroelastic effects, leading to a significant reduction in crosswind response.

1. Introduction be expressed as a polynomial function of vibration amplitude. Addi­


tionally, it can be modeled as a polynomial function of time-varying
Tall and flexible buildings are susceptible to crosswind excitation, vibration displacement and/or velocity through the application of the
particularly near the vortex lock-in wind speed. The aeroelastic effect harmonic balance technique [4]. The stochastic narrow-band building
caused by motion-induced force can generate negative aerodynamic response can be computed utilizing the equivalent nonlinear equation
damping that is dependent of vibration amplitude and wind speed (e.g., method [3–5], which gives analytical solutions of response probability
[22,23,30]). Aerodynamic damping can be extracted from wind-induced distribution and statistics. Feng and Chen [12] extended their analysis
response data. Hao et al. [15] pointed out that the random decrement approach to include the hysteresis in the restoring force characteristics
technique (e.g., [8,17,28]) cannot be used for nonlinear damping and associated with inelastic building response.
presented an approach of damping identification based on response Tall buildings incorporating base isolation systems are being
statistics measured in wind tunnel at different structural damping ratios. increasingly constructed, primarily to enhance their resilience against
Wu and Chen [34] established an approach using an unscented Kalman seismic loading (e.g., [9,27]). Their performance under wind excitation
filter technique to derive nonlinear aerodynamic damping from sto­ has been investigated in literature (e.g., [16,21,24]). The soft base layer
chastic response time history. Wu et al. [35] further presented an results in a decrease in structural frequency thus an increase in
approach to determine nonlinear aerodynamic damping from the wind-induced response when the base isolation system is within linear
probability density distribution of building response. The elastic. Conversely, the hysteretic damping generated by the isolation
motion-induced force on a vibrating building model can also be directly system’s inelastic response to strong winds diminishes building response
measured via force balance or surface pressure scanning (e.g., [10,31]; (e.g., [14] and [13]). Katagiri et al. [20] showed that base-isolated tall
Katagiri et al., [18]; [19]). The aerodynamic damping corresponds to the buildings with light weight can have unstable crosswind and torsional
motion-induced force component in phase of vibration velocity. responses. Tanaka and Ohtake [29] confirmed the influence of negative
The aerodynamic damping ratio at a specific reduced wind speed can aerodynamic damping on base-isolated tall building using wind tunnel

* Corresponding author.
E-mail address: xinzhong.chen@ttu.edu (X. Chen).

https://doi.org/10.1016/j.engstruct.2024.117722
Received 29 October 2023; Received in revised form 17 January 2024; Accepted 19 February 2024
Available online 24 February 2024
0141-0296/© 2024 Elsevier Ltd. All rights reserved.
G. Huang et al. Engineering Structures 305 (2024) 117722

isolation system is modeled in two degrees of freedom (2DOFs) system,


with the following equations of motion [13,14]
Q(t)
Ls ẍb + q¨1 + 2ω01 ξ01 q̇1 + ω201 q1 = (1)
Ms

Ls Ms F(t)
ẍb + q¨1 + 2ωb ξb ẋb + αω2b xb + (1 − α)ω2b zb = (2)
mb + m mb + m

żb = Aẋb − β|ẋb ||zb |n− 1 zb − γẋb |zb |n (3)

where Ms = ϕT Mϕ is generalized mass; ϕ is building mode shape; M =


diag(m1 , m2 …, mN ) is N × N building mass matrix (N is total number of
stories); Ls = ϕT Mr/Ms ; r = [1, 1, …, 1]T ; q1 is generalized displace­
ment, i.e., building top displacement relative to the base; x =
[x1 , x2 , …, xN ]T = ϕq1 is building displacement vector relative to the
base; xb is base slab displacement relative to the ground; ω01 and
ξ01 are modal frequency and damping ratio of the building with a fixed
base; mb is mass of base slab; m = rT Mr is entire mass of the building
without base isolation system; kb and cb are stiffness and damping
Fig. 1. Analysis model of a base-isolated tall building. coefficient of the isolation system; ωb =
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
kb /(mb + m)and ξb = cb /[2ωb (mb +m)] are frequency and damping
response data. ratio of the isolation system calculated using the initial stiffness and
In this research, a framework is presented for calculating crosswind rigid upper structure assumption; α is second stiffness ratio; zb is hys­
response of base-isolated buildings considering the nonlinear aero­ teretic displacement; and A, β, γ and n are Bouc-Wen model parameters,
1
dynamic damping effect. The building motion relative to the base is and A = 1, β = γ; xy = [A/(β + γ)]n is yield displacement;
represented in fundamental modal displacement. The shear force and Q(t) = ϕT p(t) = Qb (t) +Qse (t) and F(t) = rT p(t) = Fb (t) +Fse (t) are
displacement relation of the isolation system is described by Bouc-Wen generalized modal force and base shear force, which include both buf­
hysteresis model [33]. The wind loading on the building is represented feting and motion-induced (self-excited) components, denoted by the
by generalized modal force and base shear force. Their buffeting force subscripts “b” and “se”, respectively; p(t) = [p1 (t), p2 (t), …, pN (t) ]T is N ×
components are same as those of the building with a fixed base. The 1 story wind loading vector. When the base isolation system is within
primary focus is on modeling the motion-induced force components
linear elastic, we have zb = xb , and the first modal frequency of the base-
using the force information of the building with a fixed base, where no isolated building is denoted as ωs that is slightly lower than ω01 . When
pre-existing model exists. With the proposed framework, the response
the base displacement xb is set to be zero, Eq.(1) reduces to the equation
characteristics of a 65-story building having a square cross-section are of motion of the building with a fixed base.
quantified. The influences of different isolation layer parameters on the
building response are explored. The effect of motion-induced force
model is also discussed. 2.2. Modeling of buffeting wind loads

2. Response analysis approach The cross power spectral density (CPSD) function between i-th and j-
th story forces is described as (e.g., [4])
2.1. Equations of dynamic building motion ( ⃒ ⃒)
( z )αs ( z )αs kz fH ⃒zi − zj ⃒
Spi pj (f ) = Sp0 (f )
i j
exp − (4)
The response of a base-isolated building under crosswind excitation H H UH H
is linear elastic and the relative displacement to the base is represented /⃒
( )2 ⃒
by the fundamental modal displacement (Fig. 1). The hysteresis rela­ 1 2 ⃒
Spo (f ) = ρU H BH0 SCM (f ) ⃒Jz (f )|2 (5)
tionship between shear force and displacement of the isolation system is 2 ⃒
described by Bouc-Wen model. Therefore, the building with base

Fig. 2. PSDs and coherence of the generalized wind force and base shear force coefficients.

2
G. Huang et al. Engineering Structures 305 (2024) 117722

Fig. 3. Aerodynamic damping ratio for a building with a square cross-section (ρB2 /ms = 1/172).

Fig. 4. Time history samples of buffeting forces of the base-isolated building (U/fB= 12.9).

⃒ ⃒)
(
H0
)2 ∑ ∑N ( zi )αs +1 ( zj )αs +1 (
kz fH ⃒zi − zj ⃒ 2.3. Modeling of motion-induced wind loads
(6)
2 N
|Jz (f ) | = exp −
H i=1 j=1 H H UH H
2.3.1. Building with a fixed base
where ρ is air density; UH is wind speed at the building top; B, H and The motion-induced (self-excited) load can be divided into two
H0 are building width, height, and story height; zi is i-th floor elevation; components: one is in phase of building displacement, and another is in
αs is exponent of wind load profile; kz is decay factor; SCM (f) is PSD of phase of building velocity. The component in phase of building
base bending moment coefficient; and f is frequency in Hz. displacement contributes to aerodynamic stiffness, whose effect can be
Accordingly, the power spectra matrix of the buffeting components neglected as compared to structural stiffness. The component in phase of
of generalized modal force and shear force, Qb (t) and Fb (t), is calculated building velocity causes aerodynamic damping and can have significant
as effect on building response.
[
SQb Qb (f ) SQb Fb (f )
] [ T]
ϕ
][ ] The motion-induced wind load at i-th story under a given wind speed
= [Spi pj (f ) ϕ r (7) near the vortex lock-in wind speed is expressed as follows [2,4,11,26].
SFb Qb (f ) SFb Fb (f ) rT ( )
1 ϕ q˙1
pise (t) = ρU 2H BH0 kH ∗01i (k) i (8)
2 UH

3
G. Huang et al. Engineering Structures 305 (2024) 117722

Fig. 5. Samples of time-varying crosswind displacements of the base-isolated building (ξ01 = 1%).

⃒ ⃒ ( )2
⃒ϕ q 1 ⃒ ϕ q1
H ∗01i (k) = B10 (k) + B20 (k)⃒⃒ i ⃒⃒ + B30 (k) i (9) B1 (k) = Γ 2 B10 (k); B2 (k) = Γ3 B20 (k); B3 (k) = Γ 4 B30 (k) (12)
B B
( ) N
H0 ∑
where k = ωs B/UH is reduced frequency; ωs = 2πfs is building vibration Γn = (ϕ )n (13)
H i=1 i
frequency; B10 , B20 and B30 are model coefficients and depend on
reduced frequency; and ϕi is the mode shape coordinate at the i-th story. Obviously, Γn ≈ 1/(n +1) for a linear building mode shape.
It is noted that one additional term proportional to the absolute value of Substituting Eqs. (10) and (11) into Eq.(1) and by setting xb = 0, the
vibration displacement with the model coefficient B20 (k) is added in aerodynamic damping ratio of the fixed-base building with single vi­
the derivative H∗01i (k) as compared to the widely used formulation bration mode is calculated as [4].
introduced in Basu and Vickery [2], Ehsan and Scanlan [11] and Simiu ( )
and Scanlan [26]. Higher polynomial terms of vibration displacement ξa = −
1 ρB2 H
H ∗1 (k) (14)
can also be added if necessary [4]. 4 Ms
The motion-induced component of generalized modal force is then The above aerodynamic damping model is identical to the following
calculated as model as a function of vibration amplitude Aq1 , which is derived from
[ ( )]
∑N 1 2 q̇1 the harmonic balance technique (e.g., [4])
Qse (t) = ϕi p ∗
ise (t) = ρU H BH kH 1 (k) (10)
i=1 2 UH ( ) ⃒ ⃒ ( ) ( )2
4 ⃒ Aq ⃒ 1 Aq 1
H ∗1 (k) = B1 (k) + B2 (k)⃒⃒ 1 ⃒⃒ + B3 (k) (15)
⃒q ⃒ (q )2 3π B 4 B
⃒ 1⃒ 1
(11)
H ∗1 (k) = B1 (k) + B2 (k)⃒ ⃒ + B3 (k)
B B
where Aq1 is amplitude of q1 (t).

4
G. Huang et al. Engineering Structures 305 (2024) 117722

Fig. 6. RMS values and peak factors of displacements of the base-isolated building (w/ aerodynamic damping).

The motion-induced generalized force model, Eqs. (10) and (11), can ( )
1 ẋb + ϕi q̇1
be directly determined using the forced vibration building model test piseb (t) = ρU 2H BH0 kH ∗01ib (k) (16)
2 UH
with various vibration amplitudes. When the base bending moment is
directly measured and utilized to forecast the generalized modal force, ⃒ ⃒ ( )2
⃒xb + ϕi q1 ⃒
an empirical correction is necessary for a nonlinear variation in modal ∗
H01ib (k) = B10 (k) + B20 (k)⃒⃒ ⃒ + B30 (k) xb + ϕi q1
⃒ (17)
shape along the building’s height (e.g., [7,36]). B B

2.4. Building with a base isolation It is noted that same model coefficients B10, B20 and B30 are used for
both base-isolated and fixed building with same geometric
The motion-induced wind load or aerodynamic damping model for a configuration.
base-isolated building currently does not exist. In the following, a new Accordingly, the motion-induced components of generalized modal
model is developed following the model for the building with a fixed force and shear force are calculated as
base. Akin to Eqs. (8) and (9), the motion-induced i-th story wind load of [ ( ) ( )]
∑N 1 2 ẋb q̇1
the building with base isolation can be defined as: Qse (t) = ϕ p iseb (t) = ρU BH kH ∗
(k) + kH ∗
(k)
i=1 i 2 H 1b
UH 1s
UH
(18)

5
G. Huang et al. Engineering Structures 305 (2024) 117722

Fig. 7. RMS values and peak factors of displacements of the base-isolated building (w/o aerodynamic damping).

6
G. Huang et al. Engineering Structures 305 (2024) 117722

Fig. 8. RMS value of building top acceleration of the base-isolated building.

∑N [ ( ) ( )] ⎡ ⎤ ⎡ ⎤ ⎡ ⎤
1 2 ∗ ẋb ∗ q̇1 qx q̇x 0
Fse (t) = piseb (t) = ρ U BH kH (k) + kH (k)
i=1 2 H 1Fb
UH 1b
UH ⎢ ⎥ ⎢ − 1 − 1 − 1 ⎥ ⎢ −1 ⎥
v=⎢ ⎥ ⎢ ⎥ ⎢
⎣ q̇x ⎦;g(v) = ⎣ − M Cq̇x − M K 1 qx − M K 2 zb ⎦;D = ⎣ M D0 ⎦

(19)
n− 1 n
zb Aẋb − β|ẋb ||zb | zb − γẋb |zb | [0,0]
[ ⃒x ⃒ (⃒q ⃒) ]
∗ ⃒ b⃒ ⃒ 1⃒ (27)
H1b (k) = B10 (k)Γ1 + B20 (k) Γ1 ⃒ ⃒ + Γ2 ⃒ ⃒ #
B B
[ ( ) (x ) (q ) (q )2 ] (20) ⎡ ⎤
xb 2 b 1 1 [ ]
+B30 (k) Γ1 + 2Γ2 + Γ3 q1 (t) 1 Ls
B B B B ⎢ ⎥
qx (t) = ; M = ⎣ Ls Ms ⎦
[ ⃒x ⃒ (⃒q ⃒) ] xb (t) 1
∗ ⃒ b⃒ ⃒ 1⃒ mb + m
H1s (k) = B10 (k)Γ2 + B20 (k) Γ2 ⃒ ⃒ + Γ3 ⃒ ⃒ # [ ]
B B
[ ( ) (x ) (q ) (q )2 ] (21) 2(ω01 ξ01 + ωs ξas ) 2ωs ξab
xb 2 b 1 1 C=
+B30 (k) Γ2 + 2Γ3 + Γ4 2ωs ξaFs 2(ωb ξb + ωs ξaFb )
B B B B (28)
[ ⃒x ⃒ (⃒q ⃒) ] ( ) [ ]
2 T
∗ ⃒ b⃒ ⃒ 1⃒ K 1 = diag ω201 , αω2b ; K 2 = 0, (1 − α)ωb ;
H1Fb (k) = B10 (k)Γ0 + B20 (k) Γ0 ⃒ ⃒ + Γ1 ⃒ ⃒ #
B B
[ ( ) (x ) (q ) (q )2 ] (22) [ ]
xb 2 b 1 1 ( )
+B30 (k) Γ0 + 2Γ1 + Γ2 1 1 Qb (t)
B B B B D0 = diag , ; R(t) =
Ms mb + m Fb (t)
The equations of motion of the base-isolated building, i.e., Eqs. (1)
and (2), can then be rewritten as follows by applying Eqs. (18) and (19): The spectral representation method [6,25] is used to generate the
time histories of the generalized modal force and base shear force from
Qb (t) their PSD matrix. The building response history is then computed using
Ls ẍb + q̈1 + 2ωs ξab ẋb + 2(ω01 ξ01 + ωs ξas )q̇1 + ω201 q1 = (23)
Ms the Runge-Kutta method.
Ls Ms
ẍb + q̈ + 2(ωb ξb + ωs ξaFb )ẋb + 2ωs ξaFs q̇1 3. Crosswind response of a 65-story building
mb + m 1
(24)
Fb (t) 3.1. Building model
+αω2b xb + (1 − α)ω2b zb =
mb + m
( ) ( ) As an illustrative example, a 65-story building featuring a square
1 ρB2 H 1 ρB2 H
ξab = − ∗
H1b ; ξas = − ∗
H1s ; cross-section situated within an urban area is being examined. The
4 Ms 4 Ms
building height, width and story height are H = 260 m, B = 30 m and H0
( ) ( ) ( )
1 ρ B2 H ∗ 1 ρB2 H ∗ Ms = 4 m. The density of the building is 192 kg/m3. The mass of each story
ξaFb =− H1Fb ; ξaFs = − H1b = ξ
4 mb + m 4 mb + m mb + m ab throughout the building’s height is considered to be constant. For the
(25) building with a fixed base, the fundamental modal frequency and
damping ratio are f01 = 0.16 Hz and ξ01 = 1%. The mode shape exhibits
where ξab , ξas , ξaFb and ξaFs are aerodynamic damping ratios. a linear variation along the building’s height.
The equations of motion can be represented in the state-space For the base isolation system, the shear force and displacement have
equations: a bilinear hysteretic relation described by the Bouc-Wen model with the
modal parameters: A = 1,n = 6,β = γ = 0.5/xny . The yield displace­
v̇ = g(v) + DR(t) (26)
ment is xy = 0.025 m. The initial stiffness and mass are kb = 402 MN/m
and mb = 4.08 × 105 kg. The ratio of second stiffness to first stiffness is
α = 0.12. The damping is ignored, i.e., ξb = 0, while the influence of

7
G. Huang et al. Engineering Structures 305 (2024) 117722

Fig. 9. RMS value and peak factor of top displacement of the fixed-base building (w/ aerodynamic damping).

value of ξb on building response will be discussed later. When the base shear force coefficients are defined as
building is treated as a rigid body, the frequency of system is fb = /( ) /( )
CQb (t) = Qb (t) 0.5ρU 2H BH ; CFb (t) = Fb (t) 0.5ρU 2H BH (29)
ωb /2π = 0.47 Hz. The isolation system reduces fundamental building
modal frequency from f01 = 0.16 Hz to fs = 0.15 Hz, calculated by Fig. 2(a) portrays their PSDs where their STDs are σCQb = 0.1462 and
modal analysis using the initial stiffness of the base isolation system. The σ CFb = 0.2531. In Fig. 2(b) their coherence is presented. The normalized
Strouhal number for the building is set as St = 0.090 in accordance with PSDs exhibit close similarity, and there is a strong coherence between
the guidelines from the Architectural Institute of Japan (AIJ) [1]. As a them.
result, the critical vortex lock-in wind speed is calculated as Ucr = The motion-induced force of the building with a fixed base having a
fs B/St , yielding a value of 51.6 m/s. linear mode shape is calculated using the empirical model by Watanabe
The PSDs of story wind forces are defined by Eq.(4)-(6) with kz = 5 et al. [32], developed based on forces vibration wind tunnel data. The
and αs = 0.11. The PSD of the base bending moment coefficient, as aerodynamic damping ratio at a specific reduced wind speed is repre­
recommended by AIJ [1], is applied with its standard deviation (STD) sented as follows (e.g., [4]):
σ CM = 0.1572 as depicted in Fig. 2(a). The generalized modal force and

Fig. 10. RMS value and peak factor of top displacement of the fixed-base building (w/o aerodynamic damping).

8
G. Huang et al. Engineering Structures 305 (2024) 117722

Fig. 11. RMS value of top acceleration of the fixed-base building.

Fig. 12. RMS value of building top displacement vs. building damping ratio (w/ aerodynamic damping).

( )( 2 ) ( ) ( 2)
1 ρB H 3 ρB
ξaeq = − H1 ∗ (k) = − H1 ∗ (k) = a1 (k) + a2 (k)σq + a3 (k)σ2q (30)
4 Ms 4 ms

where ms = ϕT Mϕ/ϕT ϕ is effective mass per unit building height; and vortex lock-in speed. The polynomial model of H∗1 (k), unaffected by the
σ q = σ q1 /B = Aq1 /(√2B), σ q1 and Aq1 are RMS or STD and amplitude building mass ratio, is employed with ρB2 /ms = 1/149 to determine the
of displacement at building top. For the linear mode shape normalized aerodynamic damping ratio of the building under current investigation.
with a unit coordinate at the building top, H/ϕT ϕ = 3.
Fig. 3 illustrates the comparison between the aerodynamic damping 3.2. Crosswind response characteristics
ratio calculated using the empirical model and that obtained from the
fitted polynomial function model. Here, σ q = σ q1 /B ranges from 1.7% to 100 response history samples are simulated at each mean wind speed
14.2% and ρB2 /ms = 1/172 [32]. The aerodynamic damping undergoes to calculate the response statistics including the RMS (STD), mean
a rapid transition from positive to negative values near the critical extreme (peak), peak factor and kurtosis. The time duration of each
sample is 13 min with a time interval of 0.04 s. The first 3-min response

9
G. Huang et al. Engineering Structures 305 (2024) 117722

Fig. 13. Influence of damping ratio of isolation layer.

is excluded to eliminate transient effect. The response without consid­ damping leads to a substantial increase in building response near the
eration of aerodynamic damping and that of fixed-base building are also vortex lock-in wind speed. Moreover, the response displays a non-
calculated to examine the base isolation and aeroelastic effects. Gaussian probability distribution characterized by decreased kurtosis
Fig. 4 displays the time history samples of the generalized modal and peak factor. For brevity, the response kurtosis is not presented here.
force and shear force coefficients at the reduced wind speed UH /fs B = U/ For buildings with higher structural damping, the effect of aerodynamic
fB = 12.9. Fig. 5 illustrates the time histories of building top displace­ damping diminishes, and the significance of external buffeting force
ment relative to the base and the base displacement relative to the increases.
ground with structural damping ratio ξ01 = 1% at UH /fB= 4.3, 8.6 and In contract, the base-isolated building experiences a notable reduc­
12.9, respectively. Figs. 6 and 7 portray the RMS values and peak factors tion in response due to the hysteresis damping resulting from the in­
of top and base displacements of the base-isolated building under elastic isolation layer response. This concurrently diminishes the
varying wind speed and structural damping ratios. Both scenarios, with influences of aerodynamic and structural damping. The response ex­
and without accounting for aerodynamic damping, are explored. Fig. 8 hibits comparable characteristics to that of a fixed-base building with a
illustrates the RMS value of the absolute acceleration of the building’s higher level of structural damping. The insignificant effect of aero­
top. To provide a comparison, the response of the fixed-base building is dynamic damping on the response is further illustrated by comparing
also computed, and the results are presented in Figs. 9–11. the responses with and without accounting for aerodynamic damping.
In the case of the fixed-base building, the negative aerodynamic At the wind speed range U/fB= 5–11, the base layer has intermediate

10
G. Huang et al. Engineering Structures 305 (2024) 117722

Fig. 14. Influence of initial stiffness of isolation layer.

level of yielding and exhibits infrequent drift as shown in Fig. 5(d). The base isolation system, akin to the behavior observed in the fixed-base
probability distribution of base displacement is characterized by a building with higher structural damping.
softening non-Gaussian distribution with heightened peak factor and
kurtosis compared to a Gaussian response process. In contrast, the 4. Influence of base isolation parameters
building top displacement exhibits reduced peak factor and kurtosis
indicative of hardening non-Gaussian distribution. Additionally, the The effects of various isolation layer parameters on building response
response of a traditional building with a fixed base also demonstrates a are examined. These parameters include damping ratio ξb , initial stiff­
hardening non-Gaussian characteristic within the wind speed range ness kb , yield displacement xy , and second stiffness ratio α. In each
where the influence of aerodynamic damping is notable. instance, only a single parameter varies from the baseline case, where
In Fig. 12, the influence of structural damping on the RMS value of ξ01 = 1%, ξb = 0%, kb = 402 MN/m, xy = 0.025 m, and α = 0.12.
building top displacement is illustrated for both fixed-base and base- The influence of damping ratio ξb is presented in Fig. 13. The base
isolated buildings at U/fB = 12, 14 and 16. The augmentation of displacement decreases with increase of ξb . The building top displace­
structural damping can significantly mitigate the aeroelastic effect on ment diminishes with the rising ξb at lower wind speeds, where the base
the building response in the case of a fixed base. Nonetheless, it exerts a layer behaves linearly elastic. However, at higher wind speeds, the in­
comparatively lesser influence on the response of base-isolated building cremental effect of ξb reduces the yield level of the base layer, leading to
because of the existence of the large hysteresis damping caused by the a reduction in hysteresis damping and ultimately causing an elevation in

11
G. Huang et al. Engineering Structures 305 (2024) 117722

Fig. 15. Influence of yield displacement of isolation layer.

the building top displacement. It is worth noting that the peak value of the second stiffness ratio α.
the top displacement varies with wind speed differently from its RMS
value, as the peak factor is also influenced by base layer’s yield level. 5. Influence of motion-induced force model
Evidently, there exists an optimal ξb value to attain the lowest building
top displacement. The influence of motion-induced force model is investigated. In the
Fig. 14 illustrates the impact of initial stiffness kb . As kb rises, the base above calculation, both motion-induced generalized force and base
displacement decreases. Simultaneously, the building top displacement shear are considered as functions of total vibrating displacement and
shows a decrease with increasing kb value at lower wind speeds but velocity of the building. This model is denoted as Model 1. This model is
increases at higher wind speeds. further simplified by consideration of relative displacement and velocity
The influences of yield displacement xy and second stiffness ratio α of building instead of total displacement and velocity, referred to as
are depicted in Figs. 15 and 16. At lower wind speeds, no effect is Model 2. In Model 2, the contribution of base layer vibration on motion-
observed as the base isolation response is linear elastic. However, at induced force thus aerodynamic damping is neglected. Accordingly, the
higher wind speeds, the response of base layer increases as xy and α motion-induced generalized modal force and base shear force given in
decrease, leading to a decrease in the building top displacement due to Eqs.(18) and (19) become
increased hysteresis damping. It is noteworthy that both the RMS and ∑N [ ( )]
1 2 q̇1
peak values of top displacement exhibit higher sensitivity to variation in Qse (t) = ϕi p ∗
iseb (t) = ρUH BH kH1s (k) (31)
i=1 2 UH

12
G. Huang et al. Engineering Structures 305 (2024) 117722

Fig. 16. Influence of second stiffness of isolation layer.

∑N 1 2
[ ( )]
q̇1 high degree of similarity at wind speeds below the vortex lock-in speed
Fse (t) = p ∗
iseb (t) = ρU H BH kH 1b (k) (32) where the aerodynamic damping effect is negligible. The difference can
i=1 2 UH
be found at higher wind speeds under lower structural damping where
(⃒q ⃒ )
⃒ 1⃒
(q )2 the modeling of negative aerodynamic damping is important. The
(33)
1
H ∗1s (k) = B10 (k)Γ2 + B20 (k)Γ3 ⃒ ⃒ + B30 (k)Γ4 = H ∗1 (k) response calculated by Model 3 is the minimum, followed by Model 2
B B
and Model 1. At ξ01 = 0.5% and U/fB = 12.27, the underestimations of
(⃒q ⃒ ) (q )2
⃒ 1⃒ RMS value of building top displacement by Models 2 and 3 as compared
(34)
1
H ∗1b (k) = B10 (k)Γ1 + B20 (k)Γ2 ⃒ ⃒ + B30 (k)Γ3
B B to Model 1 reach the maximum, which are 5.30% and 8.44%. The un­
The Model 2 can be further simplified by neglecting the effect of derestimation of RMS value of base displacement between Model 2 and
motion-induced base shear, i.e., Fse (t) = 0, referred to as Model 3, where Model 1 reaches the maximum of 9.45%. At ξ01 = 0.5% and
both fixed-base and base-isolated buildings have same aerodynamic U/fB= 11.63, the underestimation of RMS value of base displacement by
damping. Model 3 reaches the maximum of 12.74%. This result indicates that the
Fig. 17 compares the estimated RMS values of building base and top motion-induced shear force and the contribution of motion of base
displacements from these three models. The result without consider­ isolation system to the motion-induced force cannot be disregarded in
ation of aerodynamic damping, denoted as “no self-excitation”, is also the response prediction.
presented. The RMS responses computed using three models exhibit a

13
G. Huang et al. Engineering Structures 305 (2024) 117722

Fig. 17. Influence of motion-induced force model on building response.

6. Conclusion positive structural damping. The base isolation proves effective in


reducing the vortex-induced crosswind building response. Notably, the
This study introduced an analysis framework for predicting response base displacement at the intermediate level of yielding demonstrates a
of tall buildings subjected to crosswind excitation influenced by base softening non-Gaussian characteristic with an elevated peak factor,
isolation and nonlinear aerodynamic damping. Special attention was leading to the building top displacement exhibiting a hardening non-
given to the modeling of nonlinear motion-induced force or aero­ Gaussian nature with a reduced peak factor.
dynamic damping of base-isolated building from the force information The rise in inelastic base displacement results in an increase of
of the building with a fixed base. The coefficient of this force model, hysteresis damping, thereby assisting in mitigating the impact of nega­
expressed in terms of aerodynamic derivative, was represented as a tive aerodynamic damping on the building response. It also reduces the
polynomial function of building displacements. Additionally, this force influence of building damping ratio on building response. An optimal
model enables the computation of the motion-induced shear force damping ratio of base layer exists, which minimizes building response
required for base-isolated buildings. and enhances building performance.
As an illustrative example, the crosswind response of a 65-story It was observed that, although the effect of aerodynamic damping is
building with a square cross-section was investigated. The findings somewhat less sensitive to response of a base-isolated building, accurate
revealed that the hysteresis damping arising from the yielding of the modeling remains crucial at higher wind speeds when the structural
base layer results in a notable reduction in the building response and damping and the yield level of base layer are low. In these cases, it is
mitigates the influence of both negative aerodynamic damping and important to consider both the motion-induced shear force and the role

14
G. Huang et al. Engineering Structures 305 (2024) 117722

of motion of the base isolation system in predicting the motion-induced [11] Ehsan F, Scanlan RH. Vortex-induced vibration of flexible bridges. J Eng Mech
1990;1392:1392–411. https://doi.org/10.1061/(ASCE)0733-9399(1990)116:6.
force for response analysis.
[12] Feng C, Chen X. Crosswind response of tall buildings with nonlinear aerodynamic
The motion-induced force model introduced in this study will un­ damping and hysteretic restoring force character. J Wind Eng Ind Aerodyn 2017;
dergo validation through a wind tunnel investigation of an elastic base- 167:62–74.
isolated building model. The outcomes of this validation will be docu­ [13] Feng C, Chen X. Evaluation and characterization of probabilistic alongwind and
crosswind responses of base-isolated tall buildings. J Eng Mech 2019;145(12):
mented in subsequent studies. 4019097.
[14] Feng C, Chen X. Estimation of inelastic crosswind response of base-isolated tall
buildings: performance of statistical linearization approaches. J Struct Eng 2019;
CRediT authorship contribution statement
145(12):4019161.
[15] Hao W, Chen X, Yang Q. Extraction of nonlinear aerodynamic damping of
Quoqing Huang: Formal analysis, Data curation. Xudong Yang: crosswind-excited tall buildings from aeroelastic model tests. J Eng Mech 2020;146
Writing – original draft, Formal analysis, Data curation. Zhihao Li: Data (3):04020006.
[16] Kareem A. Modelling of base-isolated buildings with passive dampers under winds.
curation. Yuhang Fan: Writing – original draft, Formal analysis, Data J Wind Eng Ind Aerodyn 1997;72(1):323–33.
curation. Xinzhong Chen: Writing – review & editing, Supervision, [17] Kareem A, Gurley K. Damping in structures: its evaluation and treatment of
Project administration. uncertainty. J Wind Eng Ind Aerodyn 1996;59(2-3):131–57.
[18] Katagiri J, Ohkuma T, Marikawa H. Motion-induced wind forces acting on
rectangular high-rise buildings with side ratio of 2. J Wind Eng Ind Aerodyn 2001;
89(14-15):1421–32.
Declaration of Competing Interest [19] Katagiri J, Ohkuma T, Marukawa H. Analytical method for coupled across-wind
and torsional wind responses with motion-induced wind forces. J Wind Eng Ind
The authors declare that they have no known competing financial Aerodyn 2002;90(12-15):1795–805.
[20] Katagiri J, Ohkuma T, Marukawa H, Yasui H. Unstable aerodynamic responses of
interests or personal relationships that could have appeared to influence
base-isolated high-rise buildings. J Struct Constr Eng 2012;77(681):1637–44.
the work reported in this paper. [21] Katagiri, J., Ohkuma, T., Marukawa, H., Tsurumi, T., 2014. Study of wind response
of isolated layer of a base-isolated high-rise building using rain-flow method. In:
Proceedings of the 23rd National Symposium on Wind Engineering,Tokyo.,
Data availability
307–312(in Japanese).
[22] Kawai H. Vortex induced vibration of tall buildings. J Wind Eng Ind Aerodyn 1992;
Data will be made available on request. 41(1-3):117–28.
[23] Kwok K, Melbourne WH. Wind-induced lock-in excitation of tall structures. J Struct
Div 1981;107(1):57–72.
Acknowledgment [24] Liang B, Xiong S, Tang J. Wind effects on habitability of base-isolated buildings.
J Wind Eng Ind Aerodyn 2002;90(12–15):1951–8.
[25] Shinozuka M, Jan CM. Digital simulation of random processes and its applications.
The support by the National Natural Science Foundation of China
J Sound Vib 1972;25(1):111–28.
(No. 51978107 and 52178456) is greatly acknowledged. [26] Simiu E, Scanlan RH. Wind Effects on Structures: Fundamentals and Applications
to Design. third ed. New York: Wiley,; 1996.
[27] Takewaki I, Fujita K. Earthquake input energy to tall and base-isolated buildings in
References time and frequency dual domains. Struct Des Tall Spec Build 2009;18(6):589–606.
[28] Tamura Y, Suganuma S-Y. Evaluation of amplitude-dependent damping and
[1] Architectural Institute of Japan. Guidelines for Wind-resistant Design of Base- natural frequency of buildings during strong winds. J Wind Eng Ind Aerodyn 1996;
isolated Buildings. Tokyo, Japan: JSSI,; 2012. 59(2-3):115–30.
[2] Basu RI, Vickery BJ. Across-wind vibrations of structures of circular cross-section. [29] Tanaka H, Ohtake K. Study on aerodynamic vibration for super high-rise building.
Part II: development of a mathematical model for full-scale application. J Wind Eng Wind Eng, JAWE 2016;41(1):16–22 (in Japanese).
Ind Aerodyn 1983;12(1):75–97. [30] Tanaka H, Tamura Y, Ohtake K, Nakai M, Kim YC. Experimental investigation of
[3] Caughey T. On the response of non-linear oscillators to stochastic excitation. aerodynamic forces and wind pressures acting on tall buildings with various
Probab Engineering Mech 1986;1(1):2–4. unconventional configurations. J Wind Eng Ind Aerodyn 2012;107:179–91.
[4] Chen X. Estimation of stochastic crosswind response of wind-excited tall buildings [31] Vickery JB, Steckley A. Aerodynamic damping and vortex excitation on an
with nonlinear aerodynamic damping. Eng Struct 2013;56(nov.):766–78. oscillating prism in turbulent shear flow. J Wind Eng Ind Aerodyn 1993;49:
[5] Chen X. Analysis of crosswind fatigue of wind-excited structures with nonlinear 121–40.
aerodynamic damping. Eng Struct 2014;74:145–56. [32] Watanabe Y, Isyumov N, Davenport AG. Empirical aerodynamic damping function
[6] Chen X, Kareem A. Proper orthogonal decomposition-based modeling, analysis and for tall buildings. J Wind Eng Ind Aerodyn: J Int Assoc Wind Eng 1997;72:313–21.
simulation of wind loads and their effects. J Eng Mech, ASCE 2005;131(4):325–39. [33] Wen Y-K. Method for random vibration of hysteretic systems. J Eng Mech Div
[7] Chen X, Kwon D-K, Kareem A. High-frequency force balance technique for tall 1976;102(2):249–63.
buildings: a critical review and some new insights. Wind Struct 2014;18(4): [34] Wu Y, Chen X. Identification of nonlinear aerodynamic damping from stochastic
391–422. crosswind response of tall buildings using unscented Kalman filter technique. Eng
[8] Cheng CM, Lu PC, Tsai MS. Acrosswind aerodynamic damping of isolated square- Struct 2020;220:110791.
shaped buildings. J Wind Eng Ind Aerodyn 2002;90(12–15):1743–56. [35] Wu Y, Chen X, Wang Y. Identification of nonlinear aerodynamic damping of wind-
[9] Chimamphant S, Kasai K. Comparative response and performance of base-isolated excited flexible structures by curve-fitting non-Gaussian response probability
and fixed-base structures. Earthq Eng Struct Dyn 2016;45(1):5–27. density function. J Wind Eng Ind Aerodyn 2020;206:104311.
[10] Cooper K, Nakayama M, Sasaki Y, Fediw A, Resende-Ide S, Zan S. Unsteady [36] Xu YL, Kwok KCS. Mode shape corrections for wind tunnel tests of tall buildings.
aerodynamic force measurements on a super-tall building with a tapered cross Eng Struct 1993;15:618–35.
section. J Wind Eng Ind Aerodyn 1997;72:199–212.

15

You might also like