Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Engineering Structures 266 (2022) 114666

Contents lists available at ScienceDirect

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

Structural control of high-rise buildings subjected to multi-hazard


excitations using inerter-based vibration absorbers
Haoshuai Qiao a, Peng Huang a, *, Dario De Domenico b, Qinhua Wang c
a
State Key Laboratory of Disaster Reduction in Civil Engineering, Tongji University, Shanghai 200092, PR China
b
Department of Engineering, University of Messina, Messina 98166, Italy
c
School of Civil and Architecture Engineering, Southwest University of Science and Technology, Mianyang 621010, PR China

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

Keywords: High-rise buildings located in strong wind and earthquake-prone regions are exposed to multiple hazards that
Inerter-based vibration absorber must be properly considered to prevent considerable socio-economic losses. In the present study, the Tuned Mass
Multiple-mode control Damper-Inerter (TMDI) is adopted for the structural control of high-rise buildings under wind and earthquake
Multi-hazard design
excitations by exploiting its inherent multiple-mode control effects. A generalized multi-degree-of-freedom
Parametric optimization
High-rise buildings
(MDOF) model of TMDI-controlled structure is established first and compared with the commonly-adopted
Structural control generalized single-degree-of-freedom (SDOF) model. Analytical investigation on the multiple-mode control ef­
fects of TMDI demonstrates that the damping effects weighted by mode coordinate differences at two terminals of
TMDI are mainly responsible for its multiple-mode control effects. Based on a real high-rise building project,
analysis errors induced by the generalized SDOF model against the more accurate generalized MDOF model are
quantified by a comprehensive Monte Carlo simulation, which suggests the necessity of adopting numerical
search for parametric design of TMDI under most circumstances. Therefore, the parameters of TMDI and its two
competitors, i.e., Multi-Tuned Mass Damper (MTMD) and Multi-Tuned Mass Damper-Inerter (MTMDI), are
optimized using a genetic algorithm, assuming both wind- and earthquake-induced responses as performance
objectives. Time and frequency domain analyses demonstrate the efficiency of utilizing the multiple-mode
control effects of a single TMDI for multi-hazard design. The trend of the equivalent damping ratio of TMDI-
controlled structure on an arbitrary mode is analyzed to identify the design parameters affecting most the
TMDI multiple-mode control effect at last.

1. Introduction hazard problem from the perspective of structural design, an intrinsic


dilemma exists for the structural design of high-rise buildings subjected
High-rise buildings are prone to excessive dynamic responses under to earthquakes and wind loads. Specifically, due to the differences in the
various natural hazards, e.g., earthquakes and strong winds. Severe frequency components of earthquakes and wind loads, structures with
natural disasters caused significant damage to the nonstructural com­ high stiffness have superior wind-resistant capacity at the expense of
ponents [1] and building envelopes [2] and threatened the serviceability poor seismic performance. In contrast, flexible structures with high
and survivability of the primary structure. Due to the low probability of ductility can exhibit good earthquake resistance but are vulnerable to
simultaneous occurrence of multiple dominant dynamic excitations [3], wind excitations [6]. Consequently, the application of additional energy
e.g., earthquakes and strong wind events, structural design is usually absorption and dissipation devices could be a promising method to
governed by a single hazard with the highest design load intensity [4]. achieve desirable vibration mitigation under multiple hazards.
However, structures constructed in both earthquake- and typhoon-prone With the increasing development of structural control techniques,
regions inevitably suffer from both hazards during their service life, various vibration control devices, have been studied and employed as
wherein particular concerns should be paid to the multi-hazard design of supplemental energy dissipation systems to mitigate the excessive re­
structures to prevent critical damage and increase structural resilience. sponses excited by multiple hazards, primarily fluid viscous dampers
Despite that some attempts [3,5] were made to solve the multi- [6], semi-active control systems [7] and multiple tuned mass dampers

* Corresponding author.
E-mail address: huangtju@tongji.edu.cn (P. Huang).

https://doi.org/10.1016/j.engstruct.2022.114666
Received 5 March 2022; Received in revised form 11 June 2022; Accepted 5 July 2022
Available online 15 July 2022
0141-0296/© 2022 Elsevier Ltd. All rights reserved.
H. Qiao et al. Engineering Structures 266 (2022) 114666

(MTMDs) [8]. A common feature of the above-introduced control stra­ performance of IVA-equipped structures under wind and earthquake
tegies is that all these systems perform well over a wide excitation excitations (multi-hazard scenario). More specifically, the multi-mode
bandwidth, i.e., they can simultaneously control multiple vibration control effects of TMDI, TMD, and TID are analytically studied first
modes or be tuned to a specific mode in real-time, which is the crucial based on the generalized MDOF model. Subsequently, based on a real
feature for structural control strategies designed under multiple hazards. high-rise building project subjected to both wind and seismic excita­
In the recent two decades, inerter-based control systems have drawn tions, errors induced by adopting the generalized SDOF model are
great interest from researchers in the field of vibration control benefit­ quantitatively evaluated through a comprehensive Monte Carlo simu­
ting from their peculiar characteristics of lightness, effectiveness, small lation. Further, the structural control performance of optimally-
stroke, and sound robustness. Although the idea of utilizing inertial designed TMDI, MTMD, and MTMDI in time and frequency domains
features could be traced back to the liquid mass pump concept proposed are competitively assessed. Finally, the trend of the equivalent damping
in the 1970s [9], the significant impulse to the development of inerter- ratio of TMDI-controlled structure on an arbitrary mode is critically
based control strategies was given in the early 2000s, when Smith [10] analyzed in order to identify the design parameters affecting most the
introduced the concept of inerter by using a formal analogy between TMDI multiple-mode control effect.
mechanical and electrical circuits. An ideal inerter is a massless two-
terminal element that can generate output force proportional to its 2. Dynamic model and theoretical fundamentals
inertance (or apparent mass) and the relative acceleration at its two
terminals. The inertance of the inerter can be realized via various 2.1. Mathematical model of TMDI-controlled MDOF structure
mechanisms, e.g., rack-and-pinion and ball-screw [11], or fluid mech­
anism [12,13], and can reach very high values, even thousands of times As the conventional TMD and the novel TID can be retrieved by
higher than the physical mass of the inerter device [14]. By introducing removing the inerter elements and auxiliary mass from a TMDI,
the inerter to conventional vibration absorbers, various inerter-based respectively, the TMDI-controlled MDOF structure (having N DOFs) is
vibration absorbers (IVAs) have been proposed, e.g., tuned inerter adopted to establish a more general mathematical model. Following this
damper (TID) [15], tuned mass damper inerter (TMDI) [16,17], multi idea, a two-dimensional (2D) TMDI-equipped lumped mass model sub­
TMDI (MTMDI) [18], tuned liquid inerter system (TLIS) [19], tuned jected to distributed forces simulating wind loads, or base acceleration
liquid column damper inerter (TLCDI) [20,21], etc. The results showed simulating earthquake excitation is shown in Fig. 1. Following previous
that properly designed IVAs generally exhibit superior efficiency in suggestions from the literature [25], to realize a larger modal difference
different usage scenarios compared to conventional vibration absorbers. between the two terminals of TMDI, a pendulum layout with openings
Besides the considerable control performance achieved around the on slabs of consecutive floors is considered. It is noted that taking an
oscillation frequency of the IVAs, another unique and convenient feature interstory layout of TMDI with installation-floor softening [27] or
of the IVA has attracted the attention of the researchers, i.e., the introducing TMDI to a mega brace [28] may be also applicable.
multiple-mode control effect of a single passive IVA that can simulta­ Based on structural dynamics concepts, the motion equations of the
neously control multiple modes of a multi-degree-of-freedom (MDOF) TMDI-controlled MDOF structure can be formulated in the
structure [15,22–24]. Different from conventional passive vibration matrix–vector format as follows.
absorbers, e.g., tuned mass dampers (TMDs), which have narrow
Mẍ(t) + Cẋ(t) + Kx(t) = p(t) + 1i Fi (t) − 1a Fa (t) (1)
effective bandwidth and could only control a single mode in most cases,
the aforementioned multiple-mode control effect endows passive IVAs where M, C, and K are the mass, viscous damping, and stiffness matrices
the potential to become competitive alternatives to multiple (e.g.,
of the primary structure, respectively. The vector x(t) =
MTMD [8]), semi-active, active, and hybrid control devices for vibration
control against multiple hazards. {x1 (t), ⋯, xa (t), ⋯, xi (t), ⋯, xN (t)}T collects the relative displacement of
Dealing with the challenging multi-hazard design by exploiting the each DOF of the primary structure at arbitrary time t, and ẋ and ẍ
multiple-mode vibration control potential of IVAs requires a clear un­ represent the corresponding velocity and acceleration response vectors.
derstanding of the essential factors that are responsible of this capa­ 1k (k = 1,…,N) is the N-by-1 location index vector whose kth element is 1
bility. In previous studies, the generalized single-degree-of-freedom and all the other elements are zeros. The term p(t) represents the exci­
(SDOF) model was commonly adopted to simplify a more realistic MDOF tation vector exerted on each DOF at time t and can be expressed in Eq.
model and facilitate analyzing the dynamic behaviors of IVA-equipped (2) for the two representative excitations, i.e., wind loads and seismic
systems. Lazar et al. [15] suggested that the TID behaves like a inputs,

damper at higher frequencies, which could be the reason for its multiple- ⎨
mode control effects. Giaralis and Petrini [25] pointed out that the {p1 (t), ⋯, pN (t)}T , wind loads
p(t) = (2)
‘higher-mode-damping effect’ of the TMDI is related to nondiagonal ⎩ − MJüg (t), seismic inputs
elements in the mass matrix of TMDI-equipped structures. Further, Dai
[23] and Xu et al. [26] performed numerical analyses on the higher- where J is the N-by-1 influence vector whose elements are all equal to 1.
mode damping effect of TMDI and concluded that this feature is üg (t) is the horizontal acceleration of the earthquake ground motion at
induced by the relatively large modal coordinate difference between its time t.
two terminals at higher modes. The control forces generated by TMDI system at its two terminals
However, the generalized SDOF model could be inaccurate for acting on the DOFs to which its upper and lower terminals attached can
analyzing IVA-equipped structures, especially for IVAs experiencing be calculated as.
large modal coordinate difference at the two terminals (e.g., for an IVA [ ]
having multistory topology or being tuned to a high order mode). Fi (t) = kt [y(t) − xi (t)] + ct ẏ(t) − ẋi (t) (3-1)
Therefore, to investigate the multiple-mode control effect of IVAs for
multi-hazard design, the present work extends the generalized SDOF
model to a more precise generalized MDOF model to investigate the

2
H. Qiao et al. Engineering Structures 266 (2022) 114666

Fig. 1. TMDI-controlled MDOF structure under earthquake and wind loads.

[ ]
expression is shown below.
Fa = b ẍa (t) − ÿ(t) (3-2)
− ω2 bXa (ω) + (kt + iωct )Xi (ω) − mt Ü g (ω)
Y(ω) = (6)
where the xa , xi , and y are the displacement of the ath DOF, ith DOF and − ω2 (mt + b) + kt + iωct
the auxiliary mass related to the ground, respectively. Their corre­ By substituting Eq. (6) into Eqs. (4-1) and (4-2), we obtain Eqs. (7-1)
sponding expressions in frequency domain can be obtained as. and (7-2), respectively.
Fi (ω) = (kt + iωct )[Y(ω) − Xi (ω)] (4-1) Fi (ω) = A0 (ω)Xa (ω) − [A2 (ω) + A0 (ω) ]Xi (ω) + B1 (ω)Ü g (ω) (7-1)

Fa (ω) = − ω2 b[Xa (ω) − Y(ω)] (4-2) Fa (ω) = [A1 (ω) + A0 (ω) ]Xa (ω) − A0 (ω)Xi (ω) + B2 (ω)Ü g (ω) (7-2)
Following the D’ Alembert principle, the dynamic equilibrium
where
equation of TMDI system in both time and frequency domains are
expressed in Eqs. (5-1) and (5-2), respectively. − ω2 b(kt + iωct ) ω4 mt b
A0 (ω) = , A1 (ω) = ,
[ ] − ω2 (mt + b) + kt + iωct − ω2 (mt + b) + kt + iωct
(mt + b)ÿ(t) − bẍa (t) + kt [y(t) − xi (t) ] + ct ẏ(t) − ẋi (t) = − mt üg (t) (5-1) − ω2 mt (kt + iωct )
A2 (ω) = ,
− ω2 (mt + b) + kt + iωct
− ω2 (mt + b)Y(ω) + ω2 bXa (ω) + (kt + iωct )[Y(ω) − Xi (ω)] = − mt Ü g (ω) (8-1)
(5-2) − mt (kt + iωct ) − ω2 bmt
B1 (ω) = , B2 (ω) = .
where − mt üg (t) is the inertia force only existing for TMDI and TMD − ω t + b) + kt + iωct
2 (m − ω (mt + b) + kt + iωct
2

under seismic excitations. (8-2)


Further, the response of the TMDI in frequency domain, i.e., Y(ω), It is noted that the control force of TID and TMD in the frequency
can be represented by Xa (ω) and Xi (ω) by solving Eq. (5-2), and the domain can be easily expressed using Eq. (7) by setting mt = 0 and b = 0,

3
H. Qiao et al. Engineering Structures 266 (2022) 114666

respectively.
Thus, we can reformulate the Fourier transform of the vector of

⎧ ⎫ ⎧ ⎫
⎡ ⎤⎪ ⎪ ⎪

⎪ X1 (ω) ⎪
⎪ ⎪ 0 ⎪

⎪ ⎪ ⎪ ⎪
⎢0 ⎥⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪

⎢ ⋯ 0 ⋯ 0 ⋯ 0 ⎥⎪
⎪ ⋮ ⎪
⎪ ⎪
⎪ ⋮ ⎪

⎢⋮ ⋱ ⋮ ⋱ ⋮ ⋱ ⋮ ⎥⎪











⎢ ⎥⎪ ⎪ ⎪ ⎪
⎢ ⎥⎪

⎪ Xa (ω) ⎪



⎪ B2 ⎪

⎢0 ⋯ − [A1 (ω) + A0 (ω) ] ⋯ A0 (ω) ⋯ 0 ⎥⎨ ⎬ ⎪ ⎨ ⎪

⎢ ⎥
1i Fi (ω) − 1a Fa (ω) = ⎢ ⋮ ⋱ ⋮ ⋱ ⋮ ⋱ ⋮ ⎥ ⋮ + ⋮ Ü g (ω) = TX(ω) + B(ω)Ü g (ω). (9)
⎢ ⎥⎪ ⎪
⎪ ⎪ ⎪
⎢0 ⋯ A0 (ω) ⋯ − [A2 (ω) + A0 (ω) ] ⋯ 0 ⎥⎪

⎪ Xi (ω) ⎪
⎪ ⎪

⎪ B1



⎢ ⎥⎪ ⎪ ⎪ ⎪
⎢⋮ ⋱ ⋮ ⋱ ⋮ ⋱ ⋮ ⎥⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪

⎢ ⎥⎪
⎪ ⎪
⎪ ⎪ ⎪ ⎪

⎣0 ⋯ 0 ⋯ 0 ⋯ 0 ⎦⎪

⎪ ⋮ ⎪




⎪ ⋮ ⎪



⎪ ⎪
⎪ ⎪ ⎪ ⎪


⎩ ⎪ ⎩ ⎭
XN (ω) ⎭ 0

control forces in Eq. (1) into matrix form as (⃒ ⃒)


⃒xk − xk− 1⃒
Δmax = max ⃒⃒ ⃒
⃒ (k = 1, …, n) (13-1)
hk
By performing Fourier transform on Eq. (1) and substituting Eq. (9)
⃒ ⃒)
into it, we obtain ẍamax = max(⃒ẍ + Jüg ⃒ (13-2)
( )
− ω2 M + iωC + K − T X(ω) = P(ω) + B(ω)Ü g (ω) (10) where the displacement time history of the kth DOF (denoted as xk) and
the acceleration time histories of all DOFs are calculated by using a step-
where P(ω) denotes the Fourier transform of the excitation vector p(t).
by-step numerical integration, i.e., the Newmark-β method [29]. At last,
Subsequently, the matrix and the vector of displacement frequency
the four tuning parameters of TMDI in dimensionless format, i.e., mass
response functions (FRFs) for TMDI-controlled MDOF structure sub­
ratio μt, inertance ratio βt, frequency ratio υt, and damping ratio ζt, are
jected to wind and seismic excitations can be correspondingly derived as
defined as follows:
X(ω) [ ]− √̅̅̅̅̅̅̅̅̅̅̅̅̅
(11-1)
1
HWind (ω) = = − ω2 M + iωC + K − T mt b ωt kt ct
P(ω) μt = ,β = , νt = = /ω1 and ζt = (14)
Mtot t Mtot ω1 mt + b 2(mt + b)ωt
X(ω) [ ]− 1
HSeismic (ω) = = − ω2 M + iωC + K − T { − MJ + B(ω) }
Üg (ω) where Mtot and ω1 are the total mass and circular frequency of the
(11-2) fundamental mode of the structure, and ωt is the circular frequency of
the TMDI.
FRFs for wind-induced acceleration and seismically-induced inter­
story drift ratio and absolute acceleration can be accordingly derived
based on Eq. (11) following concepts of stochastic dynamic analysis and 2.2. Generalized MDOF model vs generalized SDOF model
are not shown here for simplicity.
Having established the dynamic model of the TMDI-installed MDOF In this section, the generalized MDOF model, whose diagonal ele­
system, the responses taken as performance indexes are introduced. ments are accurate enough to represent corresponding modes, is estab­
As the wind loads at both along- and across-wind directions are lished to analytically illustrate the error embedded in the commonly
usually regarded as weakly stationary processes, the covariance matrix adopted generalized SDOF model for multiple-mode control design
of the fluctuating components of wind-induced displacement responses [26,30], and subsequently, to prove the necessity of adopting the pre­
can be calculated as follows: sent generalized MDOF model for multi-hazard design.
( 2)
∫ +∞ By substituting X(ω) = ΦQ(ω) into Eq. (10) and premultiplying each
σ x l,m = [Sx (ω) ]l,m dω (12-1)
0
term with ΦT , we obtain.
( )
− ω2 ΦT MΦ + iωΦT CΦ + ΦT KΦ − ΦT TΦ Q(ω)
Sx (ω) = HWind (ω)SAF (ω)HHWind (ω) (12-2)
= ΦT P(ω) + ΦT B(ω)Ü g (ω) (15)
where SAF (ω) is the single-side power spectral density (PSD) matrix of
( )
the centralized wind loads. When l = m (l, m ≤ N), the σ 2x l,m term where the mode shape matrix Φ can be expanded as.
represents the variance of the displacement response corresponding to ⎡ ⎤
φ ⋯ φN,1
the lth (or the mth) DOF. As the passive control device has no control ⎢ ⋮1,1 ⋱
Φ = [φ1 , ⋯, φN ] = ⎣ ⋮ ⎥ ⎦ (16)
effects against static loads, the mean wind loads are not considered. φ1,N ⋯ φN,N
Differently, considering the strong nonstationary characteristics of
earthquakes and the fact that the maximum responses may occur at the Further, Eq. (15) can be expressed as.
beginning of the earthquake events, the maximum values, i.e., the ( )
maximum interstory drift ratio Δmax (denoting hk the interstory height) − ω2 M* + iωC* + K * − ΦT TΦ Q(ω) = P* (ω) + ΦT B(ω)Ü g (ω) (17)
and maximum absolute acceleration ẍamax , are considered for seismic
In Eq. (17), M* , C* , and K * are the diagonal matrices of generalized
design, which can be evaluated as follows:
mass, damping, stiffness. P* (ω) is the generalized force vector at fre­
quency ω, respectively, while ΦT TΦ is a full matrix whose elements can
be expressed in a general form as (p, q = 1,…,N).

4
H. Qiao et al. Engineering Structures 266 (2022) 114666

( ) ( )( )
− ΦT TΦ p,q = φp,a − φp,i φq,a − φq,i A0 (ω) + φp,a φq,a A1 (ω) + φp,i φq,i A2 (ω)
mt b
A1 (ω)|ω≫ωt = (iω)2 (23-2)
(18) mt + b
For TMDI whose auxiliary mass is limited at a low value, the inertia mt
term B(ω) is neglected, which results in Eq. (19). A2 (ω)|ω≫ωt = iω ct (23-3)
mt + b
( )
− ω2 M* + iωC* + K * − ΦT TΦ Q(ω) = HQ (ω)− 1 Q(ω) = P* (ω) (19) From Eqs. (23), A0(ω) and A2(ω) tend to be damping effects at higher
frequencies, while A1(ω) behaves like an additional mass and shows
Despite the non-diagonal elements of HQ (ω) are non-zero, it has been
increasing negative stiffness effects.
numerically proven that the influences of non-diagonal elements are
By substituting Eqs. (23) into Eq. (18), the following equation is
minor [23], which indicates that each diagonal element in HQ (ω) can
obtained for the TMDI.
represent a certain mode of the coupled system, i.e., the FRF of the pth
mode can be formulated as. ( ) [( )( )
− ΦT TΦ p,q = iω φp,a − φp,i φq,a − φq,i b + φp,i φq,i mt
] ct
[( mt + b
)− 1 ]
HQ,p (ω) = − ω2 M* + iωC* + K * − ΦT TΦ (20) mt b
p,p + (iω)2 φp,a φq,a (24)
mt + b
However, to decouple Eq. (19), most previous research works
Corresponding equations for TID and TMD can be retrieved by sub­
[15,23,26,27,30,31] assumed that a certain mode dominates the total
mitting Eq. (23-1) and Eq. (23-3) into Eq. (18), respectively. By
responses, i.e., X(ω) = ΦQ(ω) ≈ φp Qp (ω)(p = 1, ⋯, or n), and conse­
combining Eq. (24) and Eq. (20), it can be inferred that the multiple-
quently, established a generalized SDOF model to represent the con­
mode control effects of TMDI are attributed to both damping and
cerned mode as follows.
negative stiffness effects weighted by modal coordinate difference and
( )
− ω2 m*p + iωc*p + kp* − φTp Tφp Qp (ω) = P*p (ω) (21) modal coordinates at its two terminals of multiple modes while not a
single mode. Between TMDI and TID, besides the damping effects
Further, the FRF of the pth mode is. contributed by A0(ω), the additional damping term A2(ω) and negative
( )− 1 stiffness term A1(ω) theoretically enhance the multiple-mode control
HQ,p (ω) = − ω2 m*p + iωc*p + kp* − φTp Tφp effect of TMDI in comparison to that of TID. As for TMD (b = 0), the
[ much smaller ct,TMD (βt of IVA is usually hundreds to thousand times μt of
( ) ]− 1
= − ω2 m*p + iωc*p + kp* − ΦT TΦ p,p (22) TMD, and the optimal ζt of IVA is also larger than that of TMD) leads to a
much narrower effective bandwidth.
where m*p , c*p , and k*p are the generalized mass, damping, stiffness of the Particularly, when the second terminal of TMDI is linked to the
ground, Eq. (24) degenerates into.
pth mode.
( )
By comparing Eqs. (20) and (22), it is noticed that an error is − ΦT TΦ p,q = iωφp,i φq,i ct (25)
introduced as all non-diagonal elements in ΦT TΦ are ignored when
performing the inversion operation. From the perspective of dynamics, Under this case, this grounded TMDI behaves like a non-
this error is introduced as the minor responses of all other modes could conventional TMD having an excessive effective mass ratio μeff = μt +
( ) βt, and it can still achieve considerable damping effects due to the large
be amplified by φq,a − φq,i , φq,a , and φq,i (q ∕ = p), and further, have
significant influences on the concerned pth mode. By observing Eq. (18), μeff, and correspondingly, large ct,TMDI.
such error becomes unacceptable for the MDOF system controlled by a The main concepts behind the above-explained mechanism of
( ) multiple-mode control effect of TMDI, TID, and TMD are summarized in
well-tuned TMDI which has large φq,a − φq,i , φq,a , and φq,i (q ∕ = p), e.g.,
Fig. 2.
having a multistory topology. Therefore, previous generalized SDOF
model-based methods [26,30,31] that seem to be suitable for the present
3. Case study
work are unreliable, and a numerical search should be implemented
based on the MDOF model to consider modal interactions. The quanti­
3.1. Building information
tative evaluation of errors of the generalized SDOF model will be per­
formed in Section 3.3.
The high-rise building taken as the host structure is located on the
southeast coast of China, which is both an earthquake- and typhoon-
prone zone, and thus, both earthquakes and strong wind events are
2.3. Mechanism of multiple-mode control effect of TMDI
included in the design stage. The building has 64 stories with a height of
300 m. Two facades of the building and the planar view of a standard
The multiple-mode control effects of TMDI endow it with the po­
floor for 6–9 stories are displayed in Fig. 3 (a) and (b), respectively. As
tential to realize multi-hazard design, which is analytically studied in
shown in Fig. 3 (b), the building has an almost symmetric planar layout
this section. Although some explanations for this feature have been re­
of structural components with negligible eccentrical distance on each
ported in the literature [23,26], the adopted generalized SDOF model
floor. Thus, the three-dimensional (3D) structure can be decomposed
makes them questionable.
into two independent 2D sub-structures representative of the dynamic
The mechanism of multiple-mode control effect of TMDI is investi­
behavior along the x- and the y-axis. As later explained in section 3.2.1,
gated through Eq. (18), by focusing on the variations of A0 (ω), A1 (ω),
the dynamic characteristics of the structure along the y-axis are indi­
and A2 (ω) in the T matrix (cf. Eq. (9)) along with frequency. Supposing
vidually considered in the present study to establish the lumped mass
that the control device is tuned to control the fundamental mode of a
model introduced in section 2.1.
flexible structure, when the natural frequencies of concerned higher
Based on the finite element model of the structure, the lumped mass
modes are significantly larger than that of the fundamental mode, the
of each floor is calculated by adding dead load to 0.5 times live load, and
limit cases of A0 (ω), A1 (ω), and A2 (ω) at higher frequencies can be
the y-axis lateral stiffness coefficients between two consecutive floors
expressed below.
are calculated by dividing the interstory shear force by interstory drift.
A0 (ω)|ω≫ωt = iω
b
ct (23-1) The total mass of the structure is 288363.6 tonnes. The distributions of
mt + b lumped masses and lateral stiffness along floors are shown in Fig. 4 (a),
where the floors having excessive lumped mass are refuge floors marked

5
H. Qiao et al. Engineering Structures 266 (2022) 114666

Fig. 2. Mechanism of multiple-mode control effect of TMDI, TID, and TMD.

Fig. 3. (a) West (left) and south (right) facades of the building and (b) plan view of standard floor for 6–9 stories.

in Fig. 3 (a). The first six mode shapes of the simplified model are loads are measured from 0◦ to 350◦ at an interval of 10◦ , as shown in
depicted in Fig. 4 (b), where the bending-shear deformation of the host Fig. 5 (b). Particularly, due to the coastal construction location of the
structure can be observed. The corresponding natural frequencies and structure, the wind field type for approaching wind from directions of
other key dynamic characteristics are listed in Table 1. According to the 100◦ to 220◦ is A-type [35]. In contrast, the wind field type is C-type for
preliminary assessments of the dynamic responses of the uncontrolled the other wind directions, which reflects the wind field characteristics of
structure under both hazards, the structure is assumed to remain elastic urban areas [35].
in the present work since the responses are generally below the Based on the preliminary wind-induced vibration calculations, it is
threshold of elastic limit. This assumption may overestimate the per­ found that the maximum standard derivations (STD) of displacement
formance of control devices [32–34] under the seismic excitations and acceleration responses occur on the top floor along the y-axis at
having higher intensities. wind directions of 80◦ and 100◦ , respectively, which are mainly induced
by vortex shedding effects. Thus, only the motion along the y-axis is
3.2. Excitations considered for the primary 3D structure, and the wind-induced vibra­
tions at these two directions are selected as the control objectives for the
This section introduces the detailed information of excitations parametric optimization of TMDI. Two segments of time histories of the
adopted in the latter parametric optimization and performance aerodynamic force on the 64th floor at wind direction of 90◦ (across-
assessment. wind) and 180◦ (along-wind) are shown in Fig. 6 (a) and (b), and the
PSDs of the corresponding zero-centered wind loads are shown in Fig. 6
3.2.1. Wind loads (c) and (d).
The wind loads acting on the host structure are measured and In Fig. 6 (a), the wind loads at the across-wind direction are nearly
calculated based on the synchronous multi-point pressure tests per­ zero-mean. A predominant frequency around ω = 0.8 can be observed in
formed on a scaled rigid building model carried out in the TJ-3 wind Fig. 6 (c), which is attributed to the vortex shedding. Differently, the
tunnel of Tongji University as shown in Fig. 5 (a). The aerodynamic wind loads at the along-wind direction are mainly governed by the

6
H. Qiao et al. Engineering Structures 266 (2022) 114666

Fig. 4. Distribution of (a) lumped mass and lateral stiffness along floors and (b) the first six order mode shapes.

motion time histories from the FEMA P695 record set [36] are adopted
Table 1
as the simulated seismic inputs. These time histories are two perpen­
Dynamic characteristics of the lumped mass model.
dicular components of tens of earthquake records with various charac­
Mode number 1 2 3 4 5 6 teristics, e.g., peak ground acceleration (PGA) and predominant
Natural frequency (rad/s) 1.14 2.81 4.51 6.14 7.81 9.47 frequency recorded by different stations. Detailed seismologic infor­
Damping ratio 2% 2% 2% 4% 4% 4% mation on the 100 records can be found in the reference [37].
Mode participation factor 1.477 0.807 0.605 0.505 0.432 0.408
To ensure consistency between the design code [38] and the simu­
lated seismic inputs, the PGAs of the 100 records are artificially scaled to
incoming turbulence and have an obvious non-zero mean value as 0.1 g following the seismic hazard map of the installation site of the
shown in Fig. 6 (b). No obvious predominant frequency is observed in high-rise building. The mean displacement and acceleration response
Fig. 6 (d). spectra (ζ = 0.02) are shown in Fig. 7 (a) and (b). Two time–frequency
spectra of one far-field record and one near-field earthquake record are
3.2.2. Earthquake ground motions shown in Fig. 7 (c) and (d) as examples to demonstrate the strong non-
To draw more general results when assessing the TMDI performance stationary property of the seismic inputs.
in seismic response reduction and to take into consideration the record- Both Fig. 7 (a) and (b) indicate that flexible structures having long
to-record variability for the seismic inputs, 44 far-field and 56 near-field periods are more susceptible to near-field pulse-type earthquakes. Be­
(including both pulse-type and non-pulse-type) earthquake ground sides the strong non-stationarity shown in Fig. 7 (c) and (d), it can be
observed that the energy of earthquake excitations usually distributes at

Fig. 5. (a) Photograph and (b) coordinates of the wind tunnel test.

7
H. Qiao et al. Engineering Structures 266 (2022) 114666

Fig. 6. Time history segments of wind loads at wind directions of (a) 90◦ and (b) 180◦ ; PSDs of zero-centered wind loads at wind directions of (c) 90◦ and (d) 180◦ .

Fig. 7. Mean (a) displacement and (b) acceleration response spectra of the 100 scaled earthquake records; time–frequency spectra of the (c) No. 21 (far-filed) and (d)
No. 70 (near-field pulse-type) earthquake records.

8
H. Qiao et al. Engineering Structures 266 (2022) 114666

higher frequencies in comparison to wind loads.


‖HQ,pGSDOF ‖2 or ∞ − ‖HQ,pGMDOF ‖2 or ∞
Relative error = (p = 1 or 3) (26-1)
3.3. Error induced by the generalized SDOF model ‖HQ,pGMDOF ‖2 or ∞

‖HQ,pGMDOF ,Controlled ‖2
To prove the analytical results obtained in Section 2.2, the FRFs of Reductionratio = 1 − (p = 1 or 3) (26-2)
modal displacements of the first and third modes are calculated in terms ‖HQ,pGMDOF ,Uncontrolled ‖2
of both Eq. (20) and Eq. (22) as examples. Four representative control
where the subscripts, i.e., GSDOF and GMDOF, indicate that the FRF is
devices are considered to draw more general conclusions, whose pa­
calculated based on the generalized SDOF and MDOF models, respec­
rameters are listed as follows:
tively. Errors are evaluated considering two types of norms, namely the
H2 norm of the FRF representing the STD of the modal responses under
• Conventional TMD (TMDC): μt = 0.5%;
white noise excitation having S0 = 1, and the H∞ norm that is crucial to
• Non-conventional TMD (TMDNC): μt = 10%;
optimization strategies based on fixed-point theory.
• Interstory TMDI (tp = 1): μt = 0.5%, βt = 0.2:0.2:1.0;
In Fig. 8, the relative errors for TMDI generally increase with a larger
• TMDI with multistory topologies (tp = 4): μt = 0.5%, βt = 0.2:0.2:1.0.
reduction ratio, which indicates that a more effective TMDI is more
prone to suffer from misevaluation of control effects based on the
A Monte-Carlo (MC) simulation is conducted to study the analysis
generalized SDOF model. For the four vibration absorbers shown in
errors under different scenarios. The considered frequency ratios for the
Fig. 8 (a) and (b), the percentages of relative errors in H2 norm within
devices tuned to the first and the third modes are uniformly distributed
the tolerance of [-5%, 5%] are 100.0 %, 93.6 %, 99.8 %, and 74.9 %,
within [0.8, 1.2] and [3.8, 4.2] (ω3/ω1 = 3.97), respectively. The
respectively. The corresponding percentages for H∞ norm are 99.6 %,
damping ratios are set to be 0:0.1:1. The installation location i is con­
72.0 %, 98.7 %, and 55.7 %. Differently, when the four control devices
strained within 2:2:64 (TMDI having a 4-story topology is grounded
are tuned to the third mode, the corresponding percentages of relative
when it is installed on the 2nd and the 4th floor). Following these con­
errors in H2 norm decrease to 99.6 %, 53.6 %, 60.5 %, and 19.1 %, and
figurations, by evaluating the FRFs of 61,440 sets of control devices, the
those in H∞ norm decrease to 96.1 %, 35.6 %, 42.1 %, 10.8 %. These
distribution of errors in the H2 and H∞ norms of FRFs are shown in
results numerically prove the unreliability of the generalized SDOF
Fig. 8, where the relative errors and reduction ratios under white-noise
model for multistory TMDI design especially when tuned to higher
excitation are defined as follows:
modes. The generalized SDOF model is feasible to design interstory

Fig. 8. Relative errors of the (a) H2 and (b) H∞ norms of FRFs of the first mode controlled by vibration absorbers tuned to the first mode; relative errors of the (c) H2
and (d) H∞ norms of FRFs of the third mode controlled by vibration absorbers tuned to the third mode.

9
H. Qiao et al. Engineering Structures 266 (2022) 114666

Fig. 9. Modal displacement FRFs of the first mode controlled by (a) TMDIQ1,tp=1 installed on the 22nd floor and (b) TMDIQ1,tp=4 installed on the 36th floor; modal
displacement FRFs of the third mode controlled by (c) TMDIQ3,tp=1 installed on the 58th floor and (d) TMDIQ3,tp=4 installed on the 58th floor.

TMDI targeting at fundamental mode control but suffers from severe directions of 80◦ and 100◦ (σx,64 and σ ẍ,64 , respectively) and the mean
reduction in accuracy when designing interstory TMDI tuned to the third earthquake-induced peak interstory drift ratio and absolute floor ac­
mode. Differently, the errors in calculating FRFs of TMDC remain minor, celeration among all floors triggered by ground motions simulated by
whereas the errors of TMDNC are relatively unacceptable. 100 records (termed Δmax and ẍamax , respectively). These four indexes
To explicitly show the error of the FRFs calculated based on the correspond to survivability and serviceability in wind-resistance design
generalized SDOF model, the four most effective TMDIs in the MC and seismic demands for structural and non-structural components. As
simulation are selected as representatives (marked in Fig. 8). FRFs of the expressed in Eqs. (27), the IVA system comprises n individual IVAs,
first and third modes controlled by these four TMDIs calculated based on where the kth IVA has six parameters to be optimized, i.e., mass ratio μt,k,
the two models are shown in Fig. 9. inertance ratio βt,k, frequency ratio υt,k, damping ratio ζt,k, installation
Fig. 9 demonstrates that the FRFs calculated based on the general­ floor ik (shown in Fig. 1), and the downward multistory topology
ized SDOF model can be significantly different from the generalized number tpk = ik-ak.
MDOF model except for the fundamental mode control scenario in Fig. 9 ⎧ ({ } { } { } { } )
(a). Furthermore, by taking the wind loads at wind direction of 80◦ as ⎪
⎪ minimize f1 μt,k , βt,k , υt,k , ζt,k , {ik }, {tpk } = σx,64 (k = 1, ..., n)


excitations, where the first mode dominates the total response, the ⎪

⎪ ({ } { } { } { } )
⎪ f2 μt,k , βt,k , υt,k , ζt,k , {ik }, {tpk } = σẍ,64 (k = 1, ..., n)
maximum relative error on the STD of the 1st order modal displacement ⎪


⎪ ({ } { } { } { } )

response of the generalized SDOF model could reach 39.35 % for a TMDI ⎪

⎪ f3 μt,k , βt,k , υt,k , ζt,k , {ik }, {tpk } = Δmax (k = 1, ..., n)

having tp = 4. ({ } { } { } { } )
f4 μt,k , βt,k , υt,k , ζt,k , {ik }, {tpk } = ẍamax (k = 1, ..., n)
The results above numerically validate the reliability of the previous ⎪


⎪ s.t. μt,k ∈ [0, 0.5%], βt,k ∈ [0, 1], υt,k ∈ [0, 20],
optimization strategies for fundamental mode control of interstory ⎪




TMDI [35] and TID [26], in which case the modal coordinate difference ⎪

⎪ ζt,k ∈ [0, 1.0], ik ∈ [1, 64], tpk ∈ [1, 4] (k = 1, ..., n)

is usually small, but strongly suggest adopting the generalized MDOF ⎪


∑n
⎩ μ ≤ 0.5%.
model for multiple-mode control design of IVA. k=1 t,k

(27)
3.4. Parametric optimizations of TMDI and its competitors Following the expression in Eq. (27), the MOP for parameters of
TMDI can be defined by setting n = 1. Noteworthy, the lower bounds for
By referring to the optimization problems (OPs) defined for IVAs in μt,k and βt,k are zeros, which indicates that TID and TMD are simulta­
the previous researches [17,18,39], a general multi-objective optimi­ neously considered in the optimization for TMDI. To evaluate the con­
zation problem (MOP) for the parameters of the IVA system can be trol performance of a single TMDI with other multiple-mode control
defined. Based on Eqs. (12) and (13), four response indexes are strategies, other two MOPs for parameters of MTMDI and MTMD can be
considered to be objectives, i.e., the STD of the wind-induced displace­ defined by modifying Eq. (27) accordingly. To be specific, having in
ment and acceleration response on the top floor excited by wind loads at

10
H. Qiao et al. Engineering Structures 266 (2022) 114666

mind that only the first few modes dominate the total responses of Table 2
flexible structure, n is set to be six with the upper bounds of tpk (k = 1, Optimal parameters of representative TMDI, MTMD, and MTMDI.
…,6) being decreased to 1 to represent an MTMDI system consisting of Control device μt βt υt ζt i tp
six interstory TMDIs. As for MTMD, n is also set to be six, and the upper
TMDI 0.32% 0.91 1.07 0.29 55 4
bound of βt,k (k = 1,…,6) is set to zero to eliminate the inerter elements. MTMD TMD-1 0.08% \ 1.01 0.03 43 \
As the responses of the controlled MDOF system cannot be expressed TMD-2 0.09% \ 4.19 0.05 26 \
explicitly based on the more accurate generalized MDOF model and are TMD-3 0.07% 7.11 0.05 37
probably affected by multiple modes under the multi-hazard scenario, it TMD-4 0.09% \ 8.04 0.02 48 \
TMD-5 0.08% \ 8.58 0.01 40 \
is rationally deduced that the implicit objective functions in MOPs are TMD-6 0.09% \ 9.51 0.09 41 \
non-convex, where the conventional gradient-based algorithms are not MTMDI TMDI-1 0.08% 0.95 1.04 0.04 22 1
applicable to realize global optimizations. Herein, a metaheuristic TMDI-2 0.11% 0.79 7.01 0.65 33 1
optimization algorithm, i.e., the built-in MATLAB© function ‘gamul­ TMDI-3 0.05% 0.80 8.63 0.61 19 1
TMDI-4 0.07% 0.71 9.04 0.48 34 1
tiobj’, which is coded based on the nondominated sorting genetic al­
TMDI-5 0.06% 0.67 12.21 0.59 17 1
gorithm II (NSGA-II) [40], is adopted. Apart from the population size TMDI-6 0.09% 0.51 13.17 0.65 20 1
being enlarged to 600 to facilitate searching for the globally optimal
individual, other configurations of the optimization algorithm remain at
the default settings.
Table 3
Towards an MOP, an optimal set of design parameters which can
Responses of the uncontrolled, TMDI-installed, MTMD-installed, and MTMDI-
simultaneously reach the minimum values on all four objectives does not installed structures.
exist. In this manner, design problems mathematically described by the
Control strategy Δmax (%) ẍamax (m/s2) σx,64 (m) σẍ,64 (m/s2)
MOPs are transformed to identify the Pareto optimal solutions (also
known as Pareto Set, PS). The objective vectors of the Pareto optimal Uncontrolled 0.261 2.078 0.095 0.051
solutions are also named as a Pareto Front (PF) which is located in a TMDI 0.218 (16.5)* 1.627 (21.7) 0.072 (23.9) 0.039 (24.9)
MTMD 0.255 (2.4) 1.997 (3.9) 0.088 (8.4) 0.047 (7.9)
four-dimensional space in the present study. Each solution on the PF is
MTMDI 0.210 (19.7) 1.498 (27.9) 0.083 (12.5) 0.045 (12.4)
Pareto optimal as no other individual could achieve further decrease on
*
a certain objective without increases on the other three objectives. : The values in parentheses are reduction ratios in %.
Finally, by defining a control performance index as ro = fo/fo,uncontrolled
(o = 1,2,3, and 4 for σ x,64, σ ẍ,64 , Δmax , and ẍamax ), three representative differences in wind-resistance and seismic design. Specifically speaking,
optimal solutions are selected from the three PSs in terms of a good two wind-induced responses could be mitigated once the fundamental
trade-off on four performance indexes as shown in Fig. 10. mode is controlled, while higher modes contribute more to the seismic
In Fig. 10, the ro of three PSs are enveloped by corresponding responses.
negative (NIS) and positive ideal solutions (PIS), which are depicted by The optimal parameters and objective values of the selected TMDI,
connecting the maximum and the minimum ro of four responses, MTMD, and MTMDI are listed in Table 2 and Table 3, respectively.
respectively. The PIS of TMDI has minor ro (o = 1,2,3,4) than those of the In Table 2, the optimal TMDI has a frequency ratio slightly larger
other two control devices, which demonstrates the potentials of TMDI than 1 and is installed on the 55th floor with a tp = 4 topology, which
with multiple topologies in vibration control against both hazards. indicates that large multistory topologies could enhance the perfor­
Constrained by the practical device mass, the MTMD can hardly achieve mance of TMDI. The TMDs and TMDIs in MTMD and MTMDI systems
considerable reductions on all four objectives, especially on two seismic installed along the elevation of the structure are tuned to distinct fre­
responses. As for the MTMDI, despite that the considerable inertance quency ratios to pursue multiple-mode control effects. Further para­
could enhance its control performance on all four responses in com­ metric analyses indicate that the control effects of multistory TMDI on
(
parison to MTMD, the interstory layout strictly limits φp,a − φp,i of
) two seismic responses are almost equally impacted by variations in
frequency and damping ratios, which is different from the trends around
each TMDI in the MTMDI system, and thus, makes its overall control
its oscillation frequency widely reported in literature, i.e., being more
performance worse than that of the single multistory TMDI. Besides, an
sensitive to variations in the frequency domain.
interesting phenomenon can be observed in Fig. 10 (a) and (c) that the
Correspondingly, objective values of the three control devices are
lines between r1 and r2 are almost horizontal to each other, while
listed in Table 3.
crossings exist between r1 and r4 (also r2 and r3), which represents the
In Table 3, the responses of the controlled structure prove that the

Fig. 10. Pareto optimal solutions of (a) TMDI, (b) MTMD, and (c) MTMDI.

11
H. Qiao et al. Engineering Structures 266 (2022) 114666

Fig. 11. Variations of STDs of (a) displacement and (b) acceleration on the top floor at 36 wind directions; maximum (c) interstory drift ratio and (d) absolute
acceleration along 100 earthquake records.

TMDI and MTMDI have overall better performance than MTMD in mitigation effects at all wind directions can be assured by applying
controlling the vibrations triggered by both hazards, especially on either a single TMDI or an MTMDI system. Differently, the control effects
decreasing the two seismically-induced responses. Between TMDI and of TMDI and MTMDI on two earthquake-induced responses vary from
MTMDI, the TMDI outperforms MTMDI in wind-induced vibration record to record, which can be attributed to both the diversities of fre­
control as TMDI has a multistory topology, which further enhances the quency components and the strong non-stationarity of earthquake
control performance around its oscillation frequency in comparison to ground motions. Particularly, it can be observed in Fig. 11 (c) and (d)
TMDI-1 in the MTMDI system. In contrast, the MTMDI achieves more that the near-field pulse-type ground motions generally trigger excessive
reduction on two earthquake-induced responses, which benefits from responses of the structure, which is in line with the response spectra
the TMDIs tuned to higher modes in the MTMDI system. However, it is shown in Fig. 7 (a). Apart from the maximum values usually occurring at
noteworthy that the overall performance of a single multistory TMDI the beginning of an earthquake event, the mean maximum root-mean-
could be further enhanced by increasing the floors it spans, if necessary square (RMS) values under 100 earthquake records are also calculated
and deemed a feasible solution. as an indicator of the vibration mitigation effect along the entire time
history of the response. TMDI can averagely suppress RMS values of
interstory drift ratio and absolute acceleration to 79.4 % and 74.2 %,
3.5. Assessment on the control effects of IVAs respectively. The corresponding reductions achieved by MTMDI are
87.5% and 72.1%.
Considering that the MTMD is less competitive than the two IVAs, Further, the distribution of mean responses along floors are shown in
the detailed analyses on control effects are only carried out for TMDI and Fig. 12.
MTMDI in this section. The responses of uncontrolled, TMDI-controlled, In Fig. 12 (a) and (b), it is obvious that the curvilinear shapes of both
and MTMDI-controlled structures under excitations of wind loads at 36 mean displacement and acceleration responses profiles are highly
wind directions and 100 earthquake records are shown in Fig. 11. similar to the first-order mode shape shown in Fig. 4 (b), which dem­
Fig. 11 shows that TMDI performs better than MTMDI on mitigating onstrates that the first mode dominates the total wind-induced re­
the wind-induced displacement and acceleration responses. The average sponses. Similar to the results reported in Fig. 11, TMDI outperforms
reduction ratios on STDs of wind-induced displacement and acceleration MTMDI at all floors on wind-induced vibration control. Differently, the
responses achieved by TMDI are 21.3 % and 25.4 %, respectively. The profiles of Δmax are heavily influenced by the distribution of the lumped
corresponding reduction ratios for MTMDI are 11.0 % and 13.2 %. mass as multiple fallbacks corresponding to refuge floors having
Despite the wind loads at different wind directions consisting of excessive lumped mass can be observed. The installation of TMDI
different energy components contributed by incoming turbulence and spanning from the 51st floor to the 55th floor effectively decrease Δmax
vortex shedding effects, the energy still mainly distributes at low fre­ on these floors. Despite the MTMDI system can suppress the maximum
quencies, which makes the fundamental mode dominate the total re­
Δmax on the 58th floor of uncontrolled structure from 0.225 % to
sponses. Thus, once the fundamental mode is controlled, stable

12
H. Qiao et al. Engineering Structures 266 (2022) 114666

Fig. 12. Mean profiles of STDs of wind-induced (a) displacement, (b) acceleration, seismically-induced maximum (c) interstory drift ratio and (d) absolute
acceleration.

0.183%, several undesirable peaks on the installation floors of TMDI-2, and 14.8 % on maximum values, respectively, corresponding to 31.0 %
-3, -4, -5, -6 are observed. Towards mitigation effects onẍamax , MTMDI and 9.5 % decrease in RMS values.
outperforms TMDI as ẍamax from the 15th floor to the top can be effec­ Besides the analyses in the time domain, it is more straightforward to
assess the multiple-mode control effects of TMDI and MTMDI in the
tively controlled, while TMDI can only suppress ẍamax above the 45th
frequency domain. The moduli of FRFs for four responses are shown in
floor.
Fig. 14.
After analyses on mean values, detailed assessments are carried out
In Fig. 14, both TMDI and MTMDI can effectively decrease the values
on the four most unfavorable load cases. Firstly, the segments of
of the first peaks of FRFs, and TMDI performs slightly better. On the
response time histories are shown in Fig. 13.
second peaks representing the dynamic behavior of the second mode,
In Fig. 13, both TMDI and MTMDI can achieve certain control effects
MTMDI barely suppresses the peak values while a significant decrease is
on the wind-induced vibrations, especially when the structure is excited
achieved by TMDI. In contrast, MTMDI outperforms TMDI on the third
to relatively large responses, e.g., from 1220 s to 1300 s. In Fig. 13 (c),
and the fourth peaks. In general, the curves of FRFs demonstrate that the
both devices failed to suppress the interstory drift at the beginning of the
multiple-mode control effects of a single TMDI are competitive with that
earthquake, but TMDI can efficiently dissipate the energy afterward. The
of MTMDI consisting of six interstory TMDIs.
reduction ratios achieved by TMDI and MTMDI on maximum interstory
Further, considering the differences in stationarity of wind loads and
drift ratios are 18.8 % and 2.1 %, respectively. The corresponding
earthquake excitations, the PSDs of time histories in Fig. 13 (a) and (b)
reduction ratios on RMS values are 27.3 % and 4.4 %. As for the absolute
and time–frequency spectra of time histories in Fig. 13 (c) and (d) are
acceleration, TMDI and MTMDI can achieve reduction ratios of 25.0 %

13
H. Qiao et al. Engineering Structures 266 (2022) 114666

Fig. 13. Segments of time histories of (a) wind-induced displacement responses at 80◦ , (b) acceleration responses at 100◦ on the top floor, (c) interstory drift ratio on
the 49th floor under No. 70 earthquake record and (d) absolute acceleration on the 64th floor under No. 70 earthquake record.

Fig. 14. Moduli of FRFs of wind-induced (a) displacement and (b) acceleration responses on the top floor under the excitation acting on the top floor, (c) interstory
drift ratio on the 58th floor, and (d) absolute acceleration on the top floor.

14
H. Qiao et al. Engineering Structures 266 (2022) 114666

Fig. 15. PSDs of (a) wind-induced displacement responses at 80◦ , (b) acceleration responses at 100◦ on the top floor; time–frequency spectra of (c) interstory drift
ratio on the 49th floor under No. 70 earthquake record and (d) absolute acceleration on the 64th floor under No. 70 earthquake record.

⃒ ⃒
Fig. 16. Distribution of (a) ζeq and (b) ⃒φp,i − φp,i− tp ⃒ (p = 1,…,6) for the first six modes.

shown below. (b). The control effect on the first peak in FRFs is further enhanced in
In Fig. 15 (a) and (b), the reductions on the first peaks of PSDs of PSDs as the predominant frequencies of wind loads are even lower than
wind-induced responses (the vertical axes in subplots are linear) corre­ ω1 of the structure. Fig. 15 (c) and (d) demonstrate the strong non-
spond well with reductions on the first peaks of FRFs in Fig. 14 (a) and stationarity of seismic responses in comparison to the weak-

15
H. Qiao et al. Engineering Structures 266 (2022) 114666

stationarity of wind-induced vibration. In the strong-motion duration consisting of six interstory TMDIs on wind-induced vibration but has
from about 40 s to 60 s, the first two modes contribute most to the slightly worse mitigation effects under seismic excitations than
interstory drift ratio responses, while the second and the third modes MTMDI. Both IVAs show moderate performance on decreasing peak
dominate the absolute acceleration. The reductions on different energy values under earthquake excitations but perform better in sup­
components are also in line with FRFs in Fig. 14 (c) and (d). pressing RMS values. Considering all performance indexes, TMDI
and MTMDI show superiority over MTMD comprising six TMDs. To
3.6. Trend of the equivalent damping ratio achieve better control performance in multi-hazard design, it is
recommended to design one multistory TMDI accompanied by a few
As the combination of wind and earthquake excitations is only a more interstory dampers (e.g., viscous dampers or TMDIs) or span­
particular case of multi-hazard scenario, the multiple-mode control ef­ ning more floors to further enhance its multiple-mode control effects.
fects of TMDI are further discussed by eliminating the excitation- • Analysis error on the estimate of the H2 and H∞ norms of displace­
dependency by evaluating the additional equivalent damping ratios ment FRF for structures controlled by conventional TMD (TMDC),
(denoted as ζeq) of the TMDI-controlled structure. Noteworthy, the nonconventional (large mass ratio) TMD (TMDNC), interstory TMDI,
structure-dependency is difficult to be removed as the mode shapes have and TMDI with multistory topology indicate that the generalized
significant impacts on the control performance of IVAs but vary among SDOF system is only reliable for designing TMDC and interstory
different buildings. However, certain trends identified for the present TMDI tuned to the fundamental mode of vibration. As for TMDNC and
case-study structure may be useful to identify the factors that affect most TMDI tuned to higher modes or having a multistory topology, it is
the multiple-mode control effect of the TMDI and to draw some design recommended to take the generalized MDOF model for a more reli­
guidelines that may be applicable to other structures. able vibration control performance evaluation and parametric
By minimizing the differences of modal responses between the un­ design.
controlled structure with assumed equivalent damping ratios and the
TMDI-controlled structure, ζeq on the first six modes of the structure Funding
introduced by TMDI installation (whose parameters are listed in Table 2)
with all possible installation layouts (i = 1:64, tp = 1:4) are shown in This research was funded by State Key Laboratory of Disaster
Fig. 16 (a). Reduction in Civil Engineering (SLDRCE21-06). This support is grate­
In Fig. 16, it is evident that ζeq has a positive nonlinear correlation to fully acknowledged.
⃒ ⃒
p,i− tp (p = 1,…,6), especially for higher modes. The high values
⃒φ − φ ⃒
p,i
⃒ ⃒
of ζeq in Fig. 16 (a) correspond to minor ⃒φ − φ
p,i
⃒ at higher modes in
p,i− tp
CRediT authorship contribution statement
Fig. 16 (b), which are contradictory with the conclusions obtained from
the generalized SDOF system [26] and may be attributed to the influence Haoshuai Qiao: Conceptualization, Methodology, Software, Data
of ζeq on the first mode. Further numerical analyses demonstrate that curation, Writing – original draft, Visualization. Peng Huang: Concep­
decreasing ct by taking a minor βt, υt, or ζt (cf. Eq. (14)) will lead to minor tualization, Resources, Writing – review & editing, Supervision. Dario
ζeq regardless of TMDI installation layout. Particularly, the decreasing υt De Domenico: Methodology, Writing – review & editing, Visualization.
will significantly decrease ζeq on the first mode due to the mistuning of Qinhua Wang: Conceptualization, Writing – review & editing, Funding
TMDI. Overall, the positive correlation between ζeq and ⃒φ − φ
⃒ ⃒
⃒ acquisition.
p,i p,i− tp
suggests that the optimal installation location of TMDI designed for
multiple-mode control could be preliminarily determined as between
⃒ ⃒ Declaration of Competing Interest
two DOFs having large ⃒φp,i − φp,i− tp ⃒ on multiple modes. As for the
determination of tuning parameters, it is recommended to use an The authors declare that they have no known competing financial
available optimization strategy based on the MDOF model and not the interests or personal relationships that could have appeared to influence
generalized SDOF model in Eq. (22). the work reported in this paper.

4. Conclusions
References

In the present research, the structural control of high-rise buildings [1] Perrone D, Calvi PM, Nascimbene R, Fischer EC, Magliulo G. Seismic performance
subjected to a multi-hazard scenario characterized by wind loads and of non-structural elements during the 2016 Central Italy earthquake. Bull Earthq
earthquake excitations is investigated by using TMDI, MTMD, and Eng 2019;17(10):5655–77.
[2] Zhicheng O, Spence Seymour MJ. A performance-based damage estimation
MTMDI. The mechanism of multiple mode control of TMDI, which is framework for the building envelope of wind-excited engineered structures. J Wind
beneficial for vibration control in multi-hazard design, is analytically Eng Ind Aerodyn 2019;186:139–54. https://doi.org/10.1016/j.jweia.2019.01.001.
studied and numerically verified in the context of a real case-study [3] Potra FA, Simiu E. Multihazard design: structural optimization approach. J Optim
Theory Appl 2010;144(1):120–36. https://doi.org/10.1007/s10957-009-9586-4.
building. The main contents and findings of this study are summarized [4] Wen YK. Minimum lifecycle cost design under multiple hazards. Reliab Eng Syst
as follows. Saf 2001;73(3):223–31. https://doi.org/10.1016/s0951-8320(01)00047-3.
[5] Potra FA, Simiu E. Optimization and multihazard structural design. J Eng Mech-
ASCE 2009;135(12):1472–5. https://doi.org/10.1061/(asce)em.1943-
• The multiple mode control mechanism of TMDI is mainly governed 7889.0000057.
by two contributions, namely the damping effects related to the [6] Chapain S, Aly AM. Vibration attenuation in high-rise buildings to achieve system-
inertance, the auxiliary mass and damping coefficient (A0 and A2), level performance under multiple hazards. Eng Struct 2019;197:109352. https://
doi.org/10.1016/j.engstruct.2019.109352.
and the negative stiffness effects related to the inertance and the [7] Cao L, Laflamme S, Taylor D, et al. Simulations of a Variable Friction Device for
auxiliary mass (A1), weighted by modal coordinate difference and Multihazard Mitigation. J Struct Eng 2016;142(12):H4016001. https://doi.org/
modal coordinate of multiple modes. Within the commonly con­ 10.1061/(asce)st.1943-541x.0001580.
[8] Zuo H, Bi K, Hao H. Using multiple tuned mass dampers to control offshore wind
cerned frequency interval, the damping effects resulting from the
turbine vibrations under multiple hazards. Eng Struct 2017;141:303–15. https://
large damping coefficient of well-tuned TMDI and TID are the main doi.org/10.1016/j.engstruct.2017.03.006.
factor contributing to the multiple-mode control effects. [9] Kawamata, S., Development of a Vibration Control System of Structures by Means of
• Benefiting from the multiple-mode control effect, a single TMDI with Mass Pumps, in Institute of Industrial Science. 1973, University of Tokyo: Tokyo,
Japan.
multistory topology can suppress excessive vibrations triggered by [10] Smith MC. Synthesis of mechanical networks: the inerter. IEEE Trans Autom
both wind loads and earthquakes. TMDI outperform MTMDI Control 2002;47(10):1648–62. https://doi.org/10.1109/TAC.2002.803532.

16
H. Qiao et al. Engineering Structures 266 (2022) 114666

[11] Papageorgiou C, Houghton NE, Smith MC. Experimental Testing and Analysis of [25] Giaralis A, Petrini F. Wind-induced vibration mitigation in tall buildings using the
Inerter Devices. Journal of Dynamic Systems Measurement and Control- tuned mass-damper-inerter. J Struct Eng 2017;143(9):04017127. https://doi.org/
Transactions of the Asme 2009;131(1):11. https://doi.org/10.1115/1.3023120. 10.1061/(asce)st.1943-541x.0001863.
[12] Swift SJ, Smith MC, Glover AR, Papageorgiou C, Gartner B, Houghton NE. Design [26] Xu TC, Li YC, Lai T, et al. A simplified design method of tuned inerter damper for
and modelling of a fluid inerter. Int J Control 2013;86(11):2035–51. damped civil structures: theory, validation, and application. Struct Control Health
[13] De Domenico D, Ricciardi G, Zhang RF. Optimal design and seismic performance of Monit 2021;28(9):e2798. https://doi.org/10.1002/stc.2798.
tuned fluid inerter applied to structures with friction pendulum isolators. Soil Dyn [27] Wang Z, Giaralis A. Top-storey softening for enhanced mitigation of vortex
Earthquake Eng 2020;132:106099. https://doi.org/10.1016/j. shedding induced vibrations in wind-excited optimal tuned mass damper inerter
soildyn.2020.106099. (TMDI)-equipped tall buildings. J Struct Eng 2020;147(1). https://doi.org/
[14] Zhu H, Li Y, Shen W, Zhu S. Mechanical and energy-harvesting model for 10.1061/(ASCE)ST.1943-541X.0002838.
electromagnetic inertial mass dampers. Mech Syst Sig Process 2019;120:203–20. [28] Taylor DP. Mega brace seismic dampers for the TORRE MAYOR project at Mexico
[15] Lazar IF, Neild SA, Wagg DJ. Using an inerter-based device for structural vibration city. Taylor Devices, Inc.
suppression. Earthquake Eng Struct Dyn 2014;43(8):1129–47. https://doi.org/ [29] Newmark NM. A method of computation for structural dynamics. J Eng Mech Div,
10.1002/eqe.2390. ASCE 1959;85(3):67–94. https://doi.org/10.1061/JMCEA3.0000098.
[16] Marian L, Giaralis A. Optimal design of a novel tuned mass-damper-inerter (TMDI) [30] Zhang R, Zhang L, Pan C, et al. Targeted modal response control of structures using
passive vibration control configuration for stochastically support-excited structural inerter-based systems based on master oscillator principle. Int J Mech Sci 2021;
systems. Probab Eng Mech 2014;38:156–64. https://doi.org/10.1016/j. 206:106636. https://doi.org/10.1016/j.ijmecsci.2021.106636.
probengmech.2014.03.007. [31] Xu TC, Yang G, Li YC, et al. Influence of inerter-based damper installations on
[17] De Domenico D, Ricciardi G. An enhanced base isolation system equipped with control efficiency of building structures. Struct Control Health Monit 2022;29(5):
optimal tuned mass damper inerter (TMDI). Earthquake Eng Struct Dyn 2018;47 26. https://doi.org/10.1002/stc.2929.
(5):1169–92. https://doi.org/10.1002/eqe.3011. [32] Wang YC, Chen QJ, Zhao ZP, et al. Input energy spectra and energy characteristics
[18] De Domenico D, Qiao H, Wang Q, et al. Optimal design and seismic performance of of the hysteretic nonlinear structure with an inerter system. Struct Eng Mech 2020;
Multi-Tuned Mass Damper Inerter (MTMDI) applied to adjacent high-rise 76(6):709–24. https://doi.org/10.12989/sem.2020.76.6.709.
buildings. Struct Des Tall Special Build 2020;29(14):e1781. https://doi.org/ [33] Patsialis D, Taflanidis AA, Giaralis A. Tuned-mass-damper-inerter optimal design
10.1002/tal.1781. and performance assessment for multi-storey hysteretic buildings under seismic
[19] Zhao Z, Zhang R, Jiang Y, et al. A tuned liquid inerter system for vibration control. excitation. Bull Earthq Eng 2021. https://doi.org/10.1007/s10518-021-01236-4.
Int J Mech Sci 164;2019:105171. https://doi.org/10.1016/j. [34] Talley PC, Javidialesaadi A, Wierschem NE et al. Evaluation of steel building
ijmecsci.2019.105171. structures with inerter-based dampers under seismic loading. Eng Struct 2021;242:
[20] Wang Q, Tiwari ND, Qiao H et al. Inerter-based tuned liquid column damper for 112488. https://doi.org/10.1016/j.engstruct.2021.112488.
seismic vibration control of a single-degree-of-freedom structure. Int J Mech Sci [35] Ministry of Housing and Urban-Rural Development of the People’s Republic of
2020;184:105840. https://doi.org/10.1016/j.ijmecsci.2020.105840. China, GB50009-2012, Load code for the design of building structures. 2012, China
[21] Wang Q, Qiao H, De Domenico D, et al. Seismic performance of optimal Multi- Architecture & Building Press: Beijing, China.
Tuned Liquid Column Damper-Inerter (MTLCDI) applied to adjacent high-rise [36] Applied Technology Council and Federal Emergency Management Agency.
buildings. Soil Dyn Earthq Eng 2021;143:106653. https://doi.org/10.1016/j. Quantification of building seismic performance factor; 2009: Washington, D.C.
soildyn.2021.106653. [37] Djerouni S, Abdeddaim M, Elias S, et al. Optimum double mass tuned damper
[22] Giaralis A, Taflanidis AA. Optimal tuned mass-damper-inerter (TMDI) design for inerter for control of structure subjected to ground motions. J Build Eng, 2021;44:
seismically excited MDOF structures with model uncertainties based on reliability 103259. https://doi.org/10.1016/j.jobe.2021.103259.
criteria. Struct Control Health Monit 2018;25(2):e2082. https://doi.org/10.1002/ [38] Ministry of Housing and Urban-Rural Development of the People’s Republic of
stc.2082. China. GB 50011-2010 Code for seismic design of buildings. China Architecture &
[23] Dai J, Xu ZD, Gai PP. Tuned mass-damper-inerter control of wind-induced Building Press; 2010.
vibration of flexible structures based on inerter location. Eng Struct 2019;199: [39] Giaralis A, Petrini F. Optimum design of the tuned mass-damper-inerter for
109585. https://doi.org/10.1016/j.engstruct.2019.109585. serviceability limit state performance in wind-excited tall buildings. Procedia Eng
[24] Shen WA, Niyitangamahoro A, Feng ZQ, et al. Tuned inerter dampers for civil 2017;199:1773–8. https://doi.org/10.1016/j.proeng.2017.09.453.
structures subjected to earthquake ground motions: optimum design and seismic [40] Deb K, Pratap A, Agarwal S, et al. A fast and elitist multiobjective genetic
performance. Eng Struct 2019;198:109470. https://doi.org/10.1016/j. algorithm: NSGA-II. IEEE Trans Evol Comput 2002;6(2):182–97. https://doi.org/
engstruct.2019.109470. 10.1109/4235.996017.

17

You might also like