You are on page 1of 19

Received: 18 April 2020 Revised: 10 August 2020 Accepted: 22 September 2020

DOI: 10.1002/stc.2644

RESEARCH ARTICLE

Input energy reduction principle of structures with generic


tuned mass damper inerter

Zhipeng Zhao1,2 | Ruifu Zhang1,2 | Chao Pan3 | Qingjun Chen1,2 |


Yiyao Jiang2

1
State Key Laboratory of Disaster
Reduction in Civil Engineering, Tongji
Summary
University, Shanghai, China The incorporation of the inerter with tuned mass damper yields the inerter-
2
Department of Disaster Mitigation for based tuned mass damper that has been verified as effective lightweight or
Structures, Tongji University, Shanghai,
performance-enhanced control devices. This study theoretically proved an
China
3 intrinsic effect, namely, the input energy reduction, that the inerter reduces
College of Civil Engineering, Yantai
University, Yantai, China the input energy transmitted into the controlled structures from ground
motion. A generic tuned mass damper inerter (GTMDI) including two separat-
Correspondence
Ruifu Zhang, State Key Laboratory of
ing inerters (the total inertance keeps constant) is analyzed to facilitate a uni-
Disaster Reduction in Civil Engineering, versal analysis of typical inerter-based tuned mass dampers. Closed-form
Tongji University, Shanghai 200092, displacement and power equations are derived for the GTMDI to reveal and
China.
Email: zhangruifu@tongji.edu.cn clarify the working mechanisms, especially the functionality of inerters. Corre-
spondingly, an energy-based design framework is established for the design of
Funding information
a new GTMDI and the retrofitting of an existing tuned mass, in which the
National Natural Science Foundation of
China, Grant/Award Numbers: 51778489, inertances distributed on the two separating inerters are optimized to make
51978525; Basic Research Project of State GTMDI outperform the conventional TMDI. Finally, a series of examples veri-
Key Laboratory of Ministry of Science and
Technology, Grant/Award Number:
fied the discovered effect and derived power equation of GTMDI. In this study,
SLDRCE19A-02; Foundation of Shanghai two derived equations confirm that the GTMDI possesses dual benefits, refer-
Science and Technology Commission : ring to the input energy reduction and the enhanced dissipation power effects.
19DZ1202500
The required physical mass of the GTMDI is relieved due to the reduced input
energy, which essentially lays the foundation for lightweight control. The
developed energy-based design framework is suitable for utilizing the dual
benefits of GTMDI with a clear physical basis, where the introduced two sepa-
rating inerters are designed with optimized inertances to guarantee the dis-
placement control demand with a minimized total energy cost.

KEYWORDS
energy dissipation, inerter, input energy, tuned mass damper, vibration control

1 | INTRODUCTION

The inerter-based system is widely accepted and effective for vibration control1,2 and has attracted many researchers
working on the relevant invention of devices3–7 and optimal designs.8–11 In recent literature, the performance evalua-
tion and benefits of inerter-based systems for the protection of building structures,12–18 wind turbine towers,19 storage
tanks,20 and semi-submersible platforms,21 the vibration suppression of cables22 and machines,23 and the wind-induced

Struct Control Health Monit. 2020;e2644. wileyonlinelibrary.com/journal/stc © 2020 John Wiley & Sons, Ltd. 1 of 19
https://doi.org/10.1002/stc.2644
2 of 19 ZHAO ET AL.

vibration mitigation of tall buildings24,25 have been examined. The ideal linear inerter is a massless two-terminal inertia
element that can produce an inertia force FI in the form of F I = minÂur ,26 whereÂur represents the relative acceleration
across the two terminals of the inerter, and min is called the inertance. In virtue of some mechanical
implementations,2,26,27 the inerter is featured by a mass enhancement effect, namely, the fact that the inertance can
exceed its gravitational mass by two orders of magnitude.2,28
In the previous century, when the term inerter had not yet been coined, Kawamata27 developed a liquid mass pump
by utilizing the inertial resistance of the flowing liquid. Arakaki et al.29,30 proposed a preliminary energy dissipation
device that employed the ball screw mechanism to amplify the damping force. Subsequently, Saito et al.31 and Ikago
et al.2 proposed a two-terminal inertial device (the tuned viscous mass damper), which explicitly used the inertial mass
enhancement effect. They also determined another intrinsic advantageous feature, which is the amplification effect of
dashpot deformation. In virtue of the potential benefits of the inerter and its flexible installation approach, inerter-
based control technology has been widely accepted to facilitate performance-improved control,32,33 enhanced isolation
systems,34,35 and lightweight tuned-type control systems.19,36–39
A classic example of the tuned-type control system is the vibration absorber that consists of a mass-spring system
without a dashpot. Patented by the Frahm,40 the undamped vibration absorber produces a spring force equal and oppo-
site to the disturbing force whereby the primary structure does not vibrate at all. Due to the absence of damping in this
vibration absorber, the full compensation is only achieved when the primary structure is excited at the frequency being
equal to the natural frequency of the absorber.41,42 As an effective solution to this problem, the tuned mass damper
(TMD)43 was proposed by employing a subconfiguration of the spring and dashpot to connect the prespecified second-
ary mass to the primary structure. The additional mass is tuned to counteract the structural motion,44 while a large
amount of mass is required for enough energy absorption to achieve the desired high control performance. Stimulated
by the phenomenon where the mass amplification effect of the inerter makes itself a potential solution to the
lightweight-based control, some inerter-based tuned mass damper have been proposed.36,37,45,46 Specifically, acting as a
massless mass amplifier in a tuned mass damper inerter (TMDI),36 the inerter is proposed to link the tuned mass with
the floor (or the ground) that the TMD connects to. The reason for the high efficiency of the TMDI is that the inerter is
grounded to ground whereby the inertance directly augments the tuned mass of the TMD. Due to the increased TMD
mass, including the physical mass and virtual mass due to the grounded inerter, the compensation of the excitation
force is enhanced. The optimal design methods are developed in terms of displacement- and energy-based philosophies,
of which the solutions are in analytical and numerical forms, respectively. While in the tuned inerter damper (TID),37
the tuned mass is completely surrogated by the inerter, which is basically an interstory-drift-type inerter system. The
inerter element is also designed to be set in parallel with the spring and dashpot to form a tuned inerter mass system
(TIMS) as a solution to vibration control under human-induced excitation45 or seismic excitation.47 Other complex vari-
ants have been exploited for robust control in the frequency domain48 and performance improvement of wind turbine
towers.19
As summarized above, there has been increasing attention to develop the inerter-based tuned mass damper by par-
tially relaxing or completely eliminating the additional tuned mass. For the optimal design to facilitate its application,
the inerter-based tuned mass damper is mainly designed by following a similar approach for a conventional TMD36 or
the phenomenon-inspired numerical solution.34,49 An intuition is generally accepted as the underlying benefit of the
inerter-based tuned mass dampers that the inerter introduces or increases the inertia of the absorber for energy storage.
However, the inerters in the existing inerter-based tuned mass dampers are artificially designated by a fixed position,
such as the location of the inerter in the TMDI, TIMS, and TID. The separating pattern of inerter and the inertance of
theses dampers considered so far have not been analyzed and optimized.
Dealing with the uniqueness that differentiates the inerter-based tuned mass damper from others, this study discov-
ered an intrinsic effect, namely, the input energy reduction, that the inerter reduces the input energy or power and fur-
ther relieves the energy dissipation burden of the entire controlled system. Stochastic response and energy balance
analysis were performed for a generic tuned mass damper inerter (GTMDI) to establish the theoretical basis of the
reduction effect. The GTMDI features two separating inerter (without increasing inertance) and facilitates an overall
analysis of typical inerter-based tuned dampers including the TMDI, TID, and TIMS. Inspired by the theoretical essence
of the intrinsic energy-based benefit, an energy-based optimal design framework is established, in which the inertance
distributed on the two separating inerters is optimized to make GTMDI outperform the conventional inerter-based
tuned mass dampers. Finally, numerical examples are presented to validate the theoretical results and design criterion
considering a structure-GTMDI system subject to both artificial and natural excitations.
ZHAO ET AL. 3 of 19

2 | THEORETICAL STUDY OF GENERIC TUNED MASS DAMPER INERTER

2.1 | Mechanical models

Consider a viscously damped structure tuned by a GTMDI (Figure 1) that can be represented by a 2-degree of freedom
(DOF) system (primary structure and tuned mass). The primary structure is defined by mass m, stiffness k, and
damping c, whereas the symbols of the GTMDI are elaborated as the tuned mass mt, the dashpot cd, the spring kd, and
the inerter min with a distribution factor δ. The mechanical system shown in Figure 1 is described in a nondimensional
form, whose parameters used in this study are presented in Table 1.
The GTMDI is born from a traditional tuned mass damper, which consists of a tuned mass connecting to the mass
of the primary structure through a dashpot and a spring. The existing inerter-based tuned mass dampers are
implemented by inserting an inerter element between the tuned mass and the mass of the primary structure or the gro-
und, which are widely known as the TMDI and the TIMS (Figure 1). Dealing with the TMDI ideally grounded to the
ground, it is difficult for high-rise buildings by using the current technologies. The TMDI exhibits more effective vibra-
tion control performance by spanning more than one floor with the inerter connections.50 To pursuit the intrinsic bene-
fit that reduces the input power and energy with enhanced efficiency, this study considered a GTMDI (Figure 1) by
introducing the distribution factor δ(0 ≤ δ ≤ 1). The inerter in the GTMDI is separately divided into two portions, of
which δmin is set to link the ground (or floor) and the tuned mass, and (1 − δ)min is set to link the tuned mass and the
primary structure. The sum of the two portions is equal to the entire inertance employed in the TMDI or TIMS models,
which potentially yields a feasible and fair approach to distribute the inerter and optimize its position without increas-
ing the expense of inertance in comparison with the TMDI or TIMS.
It should be noted that the GTMDI extends the mechanical topological form of the inerter-based tuned mass
dampers and essentially aims to seek the optimal topological form for the reduction in input energy and energy dissipa-
tion burden. After introducing δ, TMDI and TIMS are two special cases of the GTMDI when the distribution factor δ is
1 and 0, respectively. With additional consideration of changing the values of β and μ, the other tuned-type control sys-
tems, which are TMD and TID (Figure 1), have also been involved in the mechanical topological form of the GTMDI.
The considered GTMDI employs a generic and classic form to accommodate the control features of the existing inerter-
based tuned mass dampers.

F I G U R E 1 Mechanical
model of a single degree of
freedom (SDOF) structure with
a GTMDI

T A B L E 1 Mechanical parameters Parameters Equations Definitions


of the structure-GTMDI system pffiffiffiffiffiffiffiffiffi
Structural parameters ω0 = k=m Structural circular frequency
ζ = c/2mω0 Inherent damping ratio
GTMDI Mechanical elements μ = min/m Inertance-mass ratio
β = mt/m Tuned mass ratio
κ = kd/k Frequency ratio
pffiffiffiffiffiffiffi
ξ = cd =2 km Nominal damping ratio
Position factor δ Distribution factor
4 of 19 ZHAO ET AL.

2.2 | Stochastic response analysis and displacement distribution equation

The governing equations of motion of the structure-GTMDI system are established as follows:
(
€ + β ðu
u €+u €d Þ + δμðu
€+u €d Þ + 2ζω0 u_ + ω20 u = − ð1 + βÞag
, ð1Þ
€+u
β ðu €d Þ + δμðu€+u €d Þ + ð1 −δÞμðu€+u €d Þ + 2ξω0 u_ d + κω20 ud = − βag

where the over-dot signifies a derivative with respect to time. u is the displacement of the primary structure relative to
the ground (Figure 1) and ud is the displacement of the tuned mass relative to the primary structure. ag denotes the
acceleration of the ground motion. By means of a Laplace transformation,51 the differential equations in Equation 1
can be converted into the algebraic form as
(
s2 U + s2 βðU + U d Þ + s2 δμðU + U d Þ + 2sζω0 U + ω20 U = − ð1 + βÞAg
: ð2Þ
s2 βðU + U d Þ + s2 δμðU + U d Þ + s2 ð1 − δÞμðU + U d Þ + 2sξω0 U d + κω20 U d = −βAg

Here, s = iω, where i is the imaginary unit; ω is the circular frequency of the excitation; and U, Ud, and Ag are the
Laplace transformations of u, ud, and ag, respectively. By solving Equation 2, the displacement responses with respect
to U and Ud can be obtained, and the corresponding transfer functions are given as
  
U ðsÞ  s2 ðμ + βð1 + μ −δμÞÞ + 2sð1 + βÞξω0 + ð1 + βÞκω20
H U ðiωÞ = =−     !,
Ag ðsÞs = iω s βð1 + μ −δμÞ + μ 1 + δμ −δ2 μ + 2s3 ðζμ + ξ + δμξ + βðζ + ξÞÞω0
4

+ s2 ðβ + κ + βκ + μ + δκμ + 4ζξÞω20 + 2sðζκ + ξÞω30 + κω40


 2  ð3Þ
U d ðsÞ s δμ −2sβζω0 − βω20
H U d ðiωÞ = =     !:
Ag ðsÞ s = iω s4 βð1 + μ− δμÞ + μ 1 + δμ −δ2 μ + 2s3 ðζμ + ξ + δμξ + βðζ + ξÞÞω0
+ s2 ðβ + κ + βκ + μ + δκμ + 4ζξÞω20 + 2sðζκ + ξÞω30 + κω40

In the hypothetical condition of the white noise excitation, the mean-square of the displacement response σ 2U and
the relative displacement of the tuned mass σ 2U d are obtained with an input power spectra amplitude S0:
ð∞ ð∞
σ 2U = jH U ðiωÞj2 S0 dω,σ 2U d = jH U d ðiωÞj2 S0 dω: ð4Þ
−∞ −∞

In virtue of integrating Equation 4,513 the mean-square responses are simplified into the following closed-form

expressions. Multiplying σ 2U and σ 2U d by πS00 , the mean-square responses can be made independent to ω0 and S0 and sim-
plified in nondimensional forms σ~2U and σ~2U d . Multiplying σ~2U and σ~2U d by ζ and ξ, respectively, the sum of σ~2U ζ and σ~2U d ξ
reads
   
κ + κβ2 + β2 D1 + κ + κβ2 + β2 D2 β2
σ~2U ζ + σ~2U d ξ = = ð1 + β Þ2 + , ð5Þ
κD1 + κD2 κ

where D1 and D2 are defined here as

 
D1 = 4ζð1 + β + δμÞξ3 + β2 + 4βζ2 ð1 + κÞ + 2βδμ + δ2 μ2 + 4ζ 2 ðκ + μ + δκμÞ ξ2 + ζ2 κ 2 ðβ + δμÞ2
  !
β2 ð1 + κ2 Þ + μ2 + 2κμ −1 + 2ζ 2 + ð −1 + δÞδμ + ðκ + δκμÞ2 ð6Þ
D2 = ζ    ξ:
+ 2β μ + κ −1 + 2ζ2 + κ + ð −1 + δ + δκ Þμ
ZHAO ET AL. 5 of 19

Repeating the same derivation procedure for an uncontrolled structure, the normalized mean-square displacement
response under white-noise excitation can be expressed as σ~2U,0 = 1=ζ. Substituting the expression of σ~2U,0 into Equation 5,
it can be rewritten as

σ~2U σ~2U d ξ 2 β2
+ 2  = ð1 + β Þ + : ð7Þ
σ~U,0 σ~U,0 ζ
2
κ

The structural displacement response ratio γ and the displacement response ratio of the tuned mass γ d are defined

γ ðβ, μ, κ, ξ, ζ, δÞ = σ~U =~
σ U,0 , γ d ðβ, μ, κ, ξ, ζ, δÞ = σ~U d =~
σ U,0 : ð8Þ

In this set of expressions, both γ and γ d are defined such that they are normalized with respect to the displacement
of σ~U,0 . Substituting Equation 8 into Equation 7, a closed-form displacement distribution equation of the structure-
GTMDI system is finally obtained:

ξ
γ 2 + γ 2d = θ ð9Þ
ζ

where θ = ð1 + βÞ2 + βκ is defined as the displacement response factor of the entire structure-GTMDI system. Note that,
2

if the tuned mass ratio β is assumed zero, the form in the derived displacement distribution Equation 9 is in consistent
with that for the TID in Zhang et al.52

2.3 | Input energy reduction

Apart from the closed-form displacement response of the structure-GTMDI system that is widely considered in existing
works, this section aims to elaborate the discovered reduction effect of input energy and its transmission flow. The
necessity of the introduced distribution factor δ is clarified as a progress of the existing inerter-based tuned mass dampers.

2.3.1 | Closed-form power equation

An energy analysis is a commonly accepted measurement to evaluate the structural performance and the control
expense.53,54 To characterize the dynamic behavior of a structure-GTMDI system in this way, the energy analysis model
is established by premultiplying Equation 1 by fu, _ u_ d gT and then integrating over the time domain. The input energy
transferred into the entire system can finally be distributed into the following contributions,

E k,s ðt Þ + E d,s ðt Þ + E e,s ðt Þ + E e,GTMDI ðt Þ + E k,GTMDI ðt Þ + E d,GTMDI ðt Þ = E input ðtÞ, ð10Þ


Ðt Ðt Ðt 2 Ðt
where E k,s ðt Þ = u€_ udt, _
E d,s ðt Þ = u2ζω _
0 udt, _ 0 udt,
E e,s ðt Þ = uω and E input ðt Þ = u€_ ug dt are the integrals of
0 0 0 0
the kinetic energy, the viscous damping energy, the elastic strain energy of the primary structure, and
Ðt _ _ Ðt _
the input energy,  respectively. E d,GTMDIðt Þ = u2ξω0 udt, E e,GTMDI ðt Þ = uκω20 ud dt, and
Ðt _ _ _ 0 d d 0 d
E k,GTMDI ðt Þ = €+u
u + u ðδμ + βÞðu €d Þ + uð1 −δÞμ€ ud dt are the viscous damping energy, the elastic strain energy,
0 d d
and the kinetic energy of the GTMDI, respectively. The energy balance for unit time can be established as

ek,s + ed,s + ee,s + ee,GTMDI + ek,GTMDI + ed,GTMDI = einput : ð11Þ

where the generic symbol ex represents the rate of energy Ex at the time instance t. The five items on the left-hand side
are summed in Equation 11 to determine the value of the total input energy that is globally transferred to the structure-
6 of 19 ZHAO ET AL.

€g , ex can be evaluated stochasti-


GTMDI system, which is the right-hand side. Dealing with the stochastic excitation u
cally by applying the expectation operator E[] to the terms of the integral in Equation 11. The expected values in the
ergodic and stationary condition are summarized consistently with the conservation of mechanical energy53 as

h _ i 1 dσ 2
E½ek,s  = E u €u = V
= 0,
2 dt
  
_ _ _ ðδμ + βÞ dσ 2V + V d ð1 −δÞμ dσ 2V d
E½ek,GTMDI  = ðδμ + βÞE u + u ðu €+u€d Þ + ð1 −δÞμE uu €d = + = 0, ð12Þ
d d 2 dt 2 dt
h _ i ω2 dσ 2 
_ κω2 dσ 2
E½ee,s  = ω20 E u u = 0 U = 0, E½ee,GTMDI  = κω20 E ud u = 0 U d = 0,
2 dt d 2 dt

and

h_ _i 
_ _
E½ed,s  = 2ζω0 E uu = 2ζω0 σ V ,E½ed,GTMDI  = 2ξω0 E uu = 2ζω0 σ 2V d
2
ð13Þ
dd

where the non-zero items E[ed,s] and E[ed,GTMDI] represent the portions of energy dissipated by the structural inherent
damping and the dashpot of the GTMDI. By doing this operation, the expected value of the rate of the total input
energy to the structure-GTMDI system in the ergodic and stationary condition is finally obtained,54



E einput = E½ed,s  + E½ed,GTMDI  = 2ζω0 σ 2V + 2ξω0 σ 2V d : ð14Þ

>Following the same procedure in Section 2.2, the mean-square velocity responses, the structural velocity σ 2U , and
the relative velocity of the tuned mass σ 2V d are also derived in a closed-form. In virtue of the similar derivation proce-
dure in Section 2.2, σ 2V and σ 2V d are multiplied by 2 ζω0 and 2 ξω0, respectively, then the sum of 2 ζω0 σ 2V to 2 ξω0 σ 2V d is
obtained:

     

δ + ð1 + βÞ2 ð1 −δÞ μðD1 + D2 Þ + β + β2 D1 + β + β2 D2
E einput = 2ζω0 σ 2V + 2ξω0 σ 2V d = πS0
βð1 + ð1 −δÞμÞ + μð1 + ð1 −δÞδμÞðD1 + D2 Þ
 2  ð15Þ
δ + ð1 + βÞ ð1 −δÞ μ + βð1 + βÞ
= πS0 :
βð1 + ð1 −δÞμÞ + μð1 + ð1 −δÞδμÞ

In terms of the uncontrolled structure, the normalized mean-square velocity response under white-noise excitation
can be expressed as σ 2V ,0 = πS0 =ð2ζω0 Þ. After being equipped with the GTMDI, the structural velocity response ratio α
and the velocity response ratio of the tuned mass αd are defined as

αðβ, μ,κ, ξ,ζ, δÞ = σ V =σ V ,0 , αd ðβ, μ,κ, ξ, ζ, δÞ = σ V d =σ V ,0 : ð16Þ

Here, an Input Energy Index is proposed to deal with the global protection of the structure-GTMDI system. In virtue
of the stochastic energy balance, η is defined as the ratio, in terms of the expected

values
between the rate of the energy
input into the structure-GTMDI system and that of an uncontrolled structure E einput,0 = 2ζω0 σ 2V ,0 = πS0 :
ZHAO ET AL. 7 of 19




E einput E einput
η=
= : ð17Þ
E einput,0 πS0

Substituting the defined velocity response ratios in Equation 16 and deriving Equation 15 by πS0, the closed-form
power equation can finally be obtained:

ξ
α2 + α2d = η, ð18Þ
ζ

where
 
βð1 + βÞ + δ + ð1 + βÞ2 ð1 −δÞ μ
η= : ð19Þ
βð1 + ð1 −δÞμÞ + μð1 + ð1 −δÞδμÞ

Normalized by the input power of the uncontrolled structure, the dissipation powers of primary structure and
GTMDI can be obtained as α2 and α2d ξ=ζ, respectively. The named input energy index η yields a closed-form approach
to evaluate the global energy performance, that is, the input power, of the structure-GTMDI system. The derivate of η
with respect to the inertance-mass ratio μ is obtained as
 
∂η δ ð1 −δÞμ2 + 2βð1 −δÞμð1 + ð1 −δÞμÞ + β2 ð1 + ð1 −δÞμÞ2
=−   2 ≤ 0, ð20Þ
∂μ βð1 + μ −δμÞ + μ 1 + δμ− δ2 μ

where ∂η/∂μ = 0 can be reached when δ= 0 or 1. Considering the distributed inerter 0 < δ < 1, the negative value of this
derivate ∂η/∂μ < 0 implies the fact that an increase of the distributed inertance μ results in a reduction in the input
energy index η. The discovered reduction effect of input power and energy can be conceptually understood in Figure 2,
where the input energy in vertical coordinate is plotted versus the time of excitation in horizontal coordinate. As
assumed in the stochastic energy analysis in Section 2.3.1, white noise is considered as the excitation with constant
input power that is represented by the slope of the lines in Figure 2. Owing to the distributed and grounded inerter μ of
GTMDI, the input energy power η (corresponding to the slope of the line) of structure-GTMDI is reduced and lower
than that of the original structure. Correspondingly, the reduced input power results in a preferable reduction of input
energy as marked by the red area.
The input energy reduction effect results from the fundamentals of the excitation force compensation provided by
the GTMDI. Due to the increased TMD-mass enhanced by the distributed and grounded inerter μ in GTMDI, the excita-
tion force can be compensated more significantly, which yields a smaller net force acting on the primary structure. As a
consequence, the input energy of the GTMDI-structure system is smaller.

FIGURE 2 Conceptual diagram of input energy reduction


8 of 19 ZHAO ET AL.

2.3.2 | Discussion

Referring to the derivation before, the closed-form displacement distribution equation (Equation 9) and the power
equation (Equation 18) are established for a structure with a GTMDI, which yield a concise way to quantify the
responses of the structure-GTMDI system. ξ/ζ in Equations 9 and 18 is the equivalent damping ratio of the GTMDI cal-
culated by comparing the damping ratio of the dashpot in the GTMDI with that of the primary structure.
Inspecting Equation 9, the derived displacement distribution equation reveals the essential relationship between the
displacement responses of the primary structure γ 2 and the tuned mass γ 2d , which simultaneously reflects the essence of
the control mechanism of the GTMDI for displacement reduction. At the expense of the GTMDI with a large damping
ratio ξ, a larger product of ξ/ζ and γ 2d is companied by a reduction in structural displacement response γ 2. Besides, it
can be inferred that the introduction of the tuned mass with mass ratio β increases the displacement response of the
entire system θ = ð1 + βÞ2 + βκ , while the implement of a stiff spring (corresponding to a large κ) can reduce the entire
2

displacement responses θ by limiting the stroke of the tuned mass.


As for the power equation in Equation 18, it establishes the energy balance of the structure-GTMDI system in a
closed-form manner, so that the energy responses associated with the input and the dissipated portions can be obtained
analytically. As given in the right-hand side of Equation 18, the expression of η explicitly reports the fact that the
GTMDI is effective to adjust the input energy transferred into the entire structure-GTMDI system by designing the
physical mass of the tuned mass β, the added inertance μ, and the separating pattern δ of the inerter. The input energy
reduction effect, namely, the reduction in η, counts on the force compensation mechanism of the GTMDI and can be
realized by the grounded inerter with optimized distribution. A greater overall understanding of this equation will be
discussed in the following section.

2.4 | Benefits of the GTMDI with two separating inerters

On the basis of the displacement distribution and power equations, apart from the vibration control effect like other
inerter-based tuned mass dampers, the variably separated inerter makes the GTMDI effective to reduce the input
energy. Dealing with the contribution of the inerter to the GTMDI, a parametric analysis is conducted in this section to
expectedly seek the rational design criteria for the determination of the GTMDI parameters.
In particular, the mechanical parameters, that is, β, μ, κ, ξ, and the distribution pattern factor δ, are considered for
investigation. The performance factors primarily evaluated are summarized in Table 2. For this investigation, unless it
is explicitly stated otherwise, the inherent damping ratio ζ = 0.02 is adopted as a typical example.
The contribution of the inerter element to the structural vibration control lies in the specific values of the inertance-
mass ratio μ and the distribution pattern factor δ. The variations in the five selected performance indices are calculated
according to the closed-form expressions in Section 2.3 to quantify the displacement and energy responses of the
structure-GTMDI system. Given an arbitrary parameter set of κ = 0.10 and ξ = 0.10, the variation results in γ, β, α2,
α2d ξ=ζ, and η are shown in Figures 3 and 4, which represent cases of a GTMDI with β = 0 and 0.10, respectively. Note
that the obtained variations and concluded results also hold for cases with other parameters. In these figures, the con-
cerned performance indices are reported by the ordinate.
Regarding the configuration of the GTMDI without a tuned mass β = 0, the GTMDI is retrieved into a generic TID
(GTID), whose inerter can be distributed arbitrarily. Once δ = 1, the generic TID is correspondingly converted into a

Evaluation indices Description


T A B L E 2 Evaluation indices for a
structure-GTMDI system
Displacement γ Structural displacement response ratio
γd Displacement response ratio of the
tuned mass (dashpot)
Energy related α2 Ratio of energy dissipated by the
primary structure
α2d ξ=ζ Ratio of energy dissipated by the
dashpot in the GTMDI
η Input energy index
ZHAO ET AL. 9 of 19

F I G U R E 3 Displacement and energy response results of the structure-GTMDI system for inherent the damping ratio ζ = 0.02, the
tuned mass ratio β = 0, the inertance-mass ratio μ  [0.05,1.0], and the distribution pattern factor δ  (0,1]: (a) structural displacement
response ratio γ, (b) displacement response ratio of the dashpot γ d, (c) ratio of the dissipation power of the primary structure α2, (d) ratio of
the dissipation power of the GTMDI α2d ξ=ζ, and (e) the input energy index η

conventional TID with the inerter entirely located on the ground. By inspecting the performance indices in Figure 3, as
anticipated, the displacement and energy performances of the structure improves (γ and α2 decrease correspondingly in
(a) and (c)) as μ increases from 0.005 to 1.00, which is consistent with the parametric analysis results that a large
inertance is preferred for effective vibration control. It is interesting to note that in a comparison with μ, the distribution
pattern factor δ plays a dominant role in adjusting the vibration control effect of the GTMDI. As described in each slice
with a specific μ, the increase of δ from zero to the unit (corresponding to the fact that more inertance is grounded) is
accompanied by a lower γ and α2. At the same time, the enhanced connection between the tuned mass and the ground
correspondingly leads to larger responses of the tuned mass in terms of γ d and α2d ξ=ζ in (b) and (d), respectively. This
inverse impact of the GTMDI imposed on the tuned mass and the primary structure requires an overall measurement
to evaluate the performance of the entire structure-GTMDI system. Dealing with this, the variation pattern of the input
energy index η is plotted in (e), where the adjustment of the distribution pattern factor δ yields an effective approach
for reducing the energy input into the entire structure-GTMDI system, compared with the uncontrolled structure. Tak-
ing the slight influence provided by μ, the minimum η is realized when δ ≈ 0.5. In this situation, the employed inerter
is divided into two portions with almost the same inertances-mass ratio (0.5μ) and is distributed evenly as a connection
of the ground-(tuned-mass)-structure.
As for the GTMDI with a nonzero tuned mass (β > 0), the variation patterns of the concerned displacement and
energy responses in Figure 4 are similar to those of the structure-GTMDI (β = 0) reported in Figure 3. The working
mechanism of the two separating inerters in terms of the GTMDI with the existence of a tuned mass is similar to that
of the generic TID. Note that the introduction of the distribution pattern factor δ provides a possibility to simulta-
neously compare the TIMS (δ = 0), TMDI (δ = 1), and generic TID (1 > δ > 0). As the GTMDI transforms from the
TIMS to a generic TID, and finally to the TMDI, the primary structure exhibits lower displacement and energy
responses at the cost of an increased displacement and energy response of the tuned mass. When the inerter is invalid
(μ = 0), the GTMDI is restricted as a special case, that is, the TMD. Benefiting from the implemented inerter with a dif-
ferent distribution pattern, the GTMDI with μ > 0 and δ > 0.5 features a more significant displacement control effect
and less dissipated energy. In particular, the GTMDI with an almost evenly distributed inerter (δ ≈ 0.6) is characterized
10 of 19 ZHAO ET AL.

F I G U R E 4 Displacement and energy response results of the structure-GTMDI system for inherent the damping ratio ζ = 0.02, the
tuned mass ratio β = 0.10, the inertance-mass ratio μ  [0.0,1.0], and the distribution pattern factor δ  [0,1]: (a) structural displacement
response ratio γ, (b) displacement response ratio of the dashpot γ d, (c) ratio of the dissipation power of the primary structure α2, (d) ratio of
the dissipation power of the GTMDI α2d ξ=ζ, and (e) the input energy index η

by the input energy transferred to the entire structure-GTMDI system being much lower in comparison with the
structure-TMD system. In the situation of this section, approximately 25% of the input energy is reduced, releasing the
burden of the energy dissipation imposed on the structure-GTMDI system.

3 | ENERGY-BASED DESIGN FRAMEWORK

3.1 | Introduction of the proposed framework

As revealed in the closed-form displacement distribution and power equations, Equations 9 and 18, the implemented
inerter with an adjusted distribution pattern guarantees the GTMDI a twofold benefit, making it feasible to simulta-
neously reduce the displacement and the input energy (equal to the energy dissipation demand). Inspired by the
corresponding parametric analysis results in Section 2.4, the differences between the variation patterns of the displace-
ment and energy indices stimulate a design framework to pursue the tradeoff between the displacement control effect
and the energy-based control cost.
Dealing with this and to adequately utilize the functionality of the inerter and its distribution pattern, a design
framework is proposed in this study by synthetically optimizing the displacement and energy performance of the
structure-GTMDI system.

3.2 | Design of a new GTMDI

The design procedure of a GTMDI includes five key parameters, namely, μ, κ, ξ, β, and δ. As stated in the conceptual
design, the essential measurements of the structure-GTMDI system, which are the displacement response ratio γ and
the input energy index η, are the desired performances to be optimized. Under the constraint condition of γ ≤ γ t (γ t rep-
resents the target performance demand), the minimization of η can be achieved by the collaboration of the tuned mass,
ZHAO ET AL. 11 of 19

inerter, and its distribution pattern. It is viewed as the lowest cost of the energy dissipation burden imposed on the
entire structure-GTMDI system, which is designated as the energy-based twofold design strategy. This η minimization
condition design for the GTMDI can be expressed mathematically as

minimize input energy ratio η
: ð21Þ
subject to γ ðμ,κ, ξ, β, δÞ ≤ γ t

Dealing with the minimization of η, it substantially searches for the optimal parameters of β, μ, and δ. The derivate
of η with respect to β is obtained by referring to Equation 19 as

∂η ðβ + ð1 + βÞð1 −δÞμÞðβð1 + ð1 −δÞμÞ + μð1 + δ + 2ð1 −δÞδμÞÞ


= >0 ð22Þ
∂β ðβð1 + ð1 −δÞμÞ + μð1 + ð1 −δÞδμÞÞ2

Hence, a low value of η can be reached when β = 0,

1
ηjβ = 0 = : ð23Þ
1 + ð1 −δÞδμ

In the view of energy-based performance, this obtained result implies that the tuned mass of the GTMDI (for exam-
ple, the TMDI) can be totally replaced by the inerter (such as the TID). In this condition, the inerter of the GTMDI is
distributed as δ = 0.50 to pursue the minimum η as ηmin = 1 + 0:25μ
1
. In virtue of the derivation above, β = 0 and δ = 0.5
are suggested for a new GTMDI to relieve the burden of the entire structure-GTMDI system for energy dissipation. The
optimal design problem is simplified with three undetermined parameters, that is, μ, κ, and ξ. Considering the control
effectiveness of the GTID, a rational γ t (treated as the constraint in Equation 21) can be satisfied at the optimal values
of β = 0 and δ = 0.5. It should be remarked here that, if the target performance demand γ t is excessively low, no optimal
solution can be obtained to produce a satisfactory level of displacement response; we recommend to reset a relatively
larger γ t.

3.2.1 | Minimization η

Dealing with the optimization of the three, one optional solution is to substitute ηmin into Equation 21 and interpret the
design formula as
8
> 1
>
> minimize
< For a new GTMDI 1 + 0:25μ
(
> γ ðμ, κ,ξÞ ≤ γ t : ð24Þ
>
>
: subject to
β = 0, δ = 0:5

Note that μ is optimized as its upper bound μmax; correspondingly, κ and ξ are optimized under the constraint condi-
tions of μmax and γ t, respectively.

3.2.2 | Minimization η with enhanced dissipation power

With the given β = 0 and δ = 0.5, the GTMDI is retrieved as a generic TID with a balanced inerter, of which the con-
cerned displacement performances include the structural displacement (γ) and the dashpot deformation (γ d). In terms
of the dashpot deformation and velocity for energy dissipation, we consider a typical case for μ = 0.1,2,55 with continu-
ously changing κ and ξ. Figure 5 reports the results of γ d/γ and α2d =α2 , simultaneously providing a unit reference plane
with γ d/γ = 1 and α2d =α2 = 1 marked in a transparent form. For the GTMDI with a small nominal damping ratio
12 of 19 ZHAO ET AL.

F I G U R E 5 Displacement and
velocity results of GTMDI-structure
with the inherent damping ratio
ζ = 0.02, inertance-mass ratio μ = 0.1,
stiffness ratio κ  [0.01,1.0], and
nominal damping ratio ξ  [0.001,1.0].
(a) γ d/γ and (b) α2d =α2

(ξ < 0.05), the deformation of the dashpot in GTMDI is amplified and is larger than those of the entire GTMDI, which
is a similar phenomenon discovered for the TID.52 It is proved in this study that the amplification of the dashpot yields
the enhanced dissipation power of the GTMDI in Figure 5b, which refers to another benefit of the GTMDI for dissipat-
ing energy in an efficient way. Comparing the values of γ d/γ and α2d =α2 in (a) and (b), a small inertance is preferential
to pursue the maximum enhanced dissipation power for the dashpot. Note that once the GTMDI is designed as β = 0,
δ = 0.5, and μ > 0, it is feasible to reduce the input energy and energy dissipation demand, even if only a small
inertance is employed.
Hence, in order to further utilize the enhanced dissipation power, another optional solution for the optimization of
the input energy is maximizing the value of γ d/γ of α2d =α2 . The corresponding mathematical form is
8
> maximize γ d =γ or α2d =α2
>
< For a new GTMDI
(
γ ðμ, κ, ξÞ ≤ γ t : ð25Þ
>
>
: subject to
β = 0,δ = 0:5

Solving this parameter design constraint, the maximum value of γ d/γ or α2d =α2 can be finally obtained, and the opti-
mal parameters β = 0, μ, κ, and ξ and the distribution pattern of the inerter δ = 0.5 are simultaneously obtained. This
proposed optimal design basically developed the damping enhancement design in Zhang et al.52 by additionally consid-
ering the dissipation power α2d =α2 . More important, the balanced inerter of the GTMDI makes itself effective to reduce
the input power in comparison with the TID.
As a summary, as depicted in the closed-form displacement and power equations (Equations 9 and 18, respectively),
the role of the tuned mass can be replaced by the inerter, ideally at almost no cost of the physical mass. Following the
developed energy-based design strategies, it is a rational choice to omit the tuned mass in order to guarantee the dis-
placement control effect with a lower energy expense.

3.3 | Design of retrofitting existing tuned mass

As for a GTMDI with an existing tuned mass, β0, the design formula in Equation 21 holds available when substituting
β = β0 into Equation 19. In this condition, the minimum η can be reached when the partial derivate of η with respect to
δ is set zero,

∂η μðβ0 + ð1 + β0 ð1 −δÞ −2δÞμÞðμ + β0 ð1 + ð1 −δÞμÞÞ


=− = 0: ð26Þ
∂δ ðβ0 ð1 + ð1 −δÞμÞ + μð1 + ð1 −δÞδμÞÞ2

By solving this,
ZHAO ET AL. 13 of 19

β0 + ð1 + β0 Þμ ð 2 + β 0 Þ2
δ= and ηmin = : ð27Þ
ð2 + β0 Þμ 4 + β0 + μ

The input energy of the entire structure-GTMDI system with an existing tuned mass can be minimized by solving
the following problem in the premise of Equation 27,
8
>
> ð2 + β 0 Þ 2
>
> minimize
>
> GTMDI with β = β0 4 + β 0 + μ
>
> 8
< γ ðμ,κ, ξÞ ≤ γ t
>
>
>
< : ð28Þ
>
> β0 + ð1 + β0 Þμ
>
> subject to δ =
>
> > ð2 + β0 Þμ
>
> >
>
: :
β = β0

As discussed above, a rational approach of the energy-based control strategy is to satisfy the target displacement
demand with a minimized input energy correspondingly representing the total energy dissipation burden. This way is
preferentially suggested for a new GTMDI in order to meet the tradeoff between the control performance and cost,
although it is also available for the retrofitting of a given tuned mass. Considering that a large oscillating response of a
tuned mass is not preferred, it is not suggested for a GTMDI with a given tuned mass to pursue the enhanced dissipa-
tion power.

3.4 | Procedure of the energy-based design for GTMDI

In virtue of the energy-based design principle, the solution to the parameter determination of the GTMDI is summa-
rized in Figure 6, and the optimal design equations are discussed in detail in the following section. It is noted here that
this provided solution involves the design of a new GTMDI and the upgrading design of an existing tuned mass. The
two design conditions are both solved by means of the inerter with an adjustable distribution pattern utilized for dis-
placement control and input energy reduction.

FIGURE 6 Energy-based design procedure for the parameter determination of the GTMDI
14 of 19 ZHAO ET AL.

4 | D E S I G N C AS E S

Following the proposed energy-based design framework, strategies I to III in Figure 6 for optimal GTMDI parameters
are implemented in this section in a series of examples of the structure-GTMDI system. The new GTMDI and retro-
fitting design of an existing tuned mass have been considered, while the difference in the results of the different design
strategies is emphasized. In the comparison analysis, the functionality of the two separating inerters is clarified.

4.1 | Performance quantification and design details

An SDOF structure with ζ = 0.02 and T = 2π/ω0 = 0.54 s56 is adopted as the original uncontrolled structure, of which
the structural dynamic performance is arranged to be improved by the proposed GTMDI. For design strategies I to III,
we assume the same target γ t = 0.40 (corresponding 60% reductions in the structural displacement compared to the
uncontrolled structure) for a fair comparison. In virtue of the design strategies I and II for a new GTMDI, the design
results are designated as Cases I and II. Similarly, following design strategy III for a given tuned mass β0, Cases III is
defined. Given the specified values of γ t in Cases I to III, the GTMDI parameters can be obtained numerically by solving
Equations 24, 25, and 28, respectively. The upper bound of the inertance-mass ratio is theoretically assumed to be 1.0 as
an example. The detailed information and design results of the three cases are summarized in Table 3.

4.2 | Results and discussion

Under the same displacement demand, that is, γ t in Cases I to III, the three optimal design strategies are all effective in
providing the desired displacement performance to the controlled structure and ensuring that the actual γ is equal to
the prespecified γ t. To achieve the minimization of input energy η, the parameter sets of large values of κ and ξ are
required in Cases I and III. The inertance-mass ratio μ is optimized into its upper bound of the feasible domain (1.0 as
an example in this section) to obtain the utmost use of the inerter for input energy reduction. If the enhanced dissipa-
tion power the GTMDI is simultaneously utilized, the combination of small μ, κ, and ξ is preferred. In Case II, the target
displacement demand is satisfied by the dual benefits of the input energy reduction (η = 0.93 < 1.0) and the amplifica-
tion effect of the dashpot deformation (γ d/γ = 1.59 > 1) and velocity (αd/α = 1.58 > 1). On the precondition of an exis-
ting tuned mass, the optimal μ is non-zero, and the optimal δ is neither zero nor the unit, which results from the fact
that the proposed GTMDI outperforms the conventional TMD, TMDI, and TIMS by optimizing the distribution pattern
of the inerter.
Adopting the GTMDI parameters in Table 3 designed by using stochastic design indices, this section first conducted
a series of time history analyses for verification. Thirty randomly generated stationary white-noise waves were used as
external excitations, of which the average root mean square values of the structural response ratios were calculated and

TABLE 3 Designed parameters of the GTMDI

Prespecified Structural performance parameters Designed GTMDI parameters

Case ID γt ηmin α 2
γ γd/γ α2d =α2 β κ ξ μ δ
Case I 0.50 0.80 0.17 0.50 0.31 0.34 0 0.608 0.639 1.000 0.50
Case II 0.50 0.93 0.23 0.50 1.59 1.58 0 0.259 0.024 0.313 0.50
Case III 0.50 0.86 0.15 0.50 0.30 0.34 0.1 0.602 0.815 1.000 0.571

T A B L E 4 Response results of the


Displacement performance Energy performance
structure-GTMDI system under white
Case γ γ d/γ η α2 noise excitations
a
Case I 0.5042 (0.50) 0.3066 (0.31) 0.8103 (0.80) 0.1730 (0.17)
Case II 0.5029 (0.50) 1.5855 (1.59) 0.9429 (0.93) 0.2354 (0.23)
Case III 0.5038 (0.50) 0.3095 (0.30) 0.8720 (0.86) 0.1545 (0.15)
a
The values in the round bracket are the stochastic results in Table 3.
ZHAO ET AL. 15 of 19

are reported in Table 4. By inspecting the results, the average values in terms of the displacement (γ and γ d/γ) and the
energy responses (η and α2) obtained from the time history analysis match closely to those obtained from the stochastic
dynamic analysis shown in the round bracket in Table 4. By doing this, it illustrates the reliability of the different pro-
posed design strategies and formulae summarized in Figure 6 in terms of optimizing the GTMDI parameters or retro-
fitting an existing tuned mass under a specific control demand and philosophy. More important, the verified results of
γ d/γ and η represent the dual benefits of the novel GTMDI with a variably distributed inerter that amplify the dashpot
deformation (corresponding to enhanced dissipation power) and reduce the input energy.
The frequency-domain transfer function curves of the displacements and the velocity of the structures with and
without the GTMDI are also provided for Cases I to III in Figure 7. It can be observed that the resonant vibration can
be suppressed significantly in the three cases. For Case II, the amplification of the dashpot deformation in the GTMDI
essentially adjusts the stiffness ratio κ as a tuning effect (shown in the second peaks of HU(iω) and HV(iω)). Compared
with the other two cases, the amplification effect of the dashpot represents an increased dissipation power.
To evaluate the dynamic performance of the structure-GTMDI systems subject to some typical seismic excitations,
time history analyses were repeated for the Kobe (RSN 1120), Coalinga (RSN 338), El Centro 1940 N–S, and Chi-Chi
(RSN 1540) earthquakes.57 In the premise of the assumed linear behavior of the structure-GTMDI system, the intensity
measure of the excitations (for instance, the peak ground acceleration) had no effect on the values of the dimensionless
performance ratios. Taking Case II as an example, the structural displacement u and energy responses are shown in
Figures 8 to 9, in which the corresponding γ, γ d/γ, η, and α2 were also calculated and given in the figures. In the case of
various seismic excitations, the displacement and energy performance levels observed, γ, γ d/γ, η, and α2, are close to or
lower than the specified values obtained from the design procedure shown in the title of each figure. In virtue of the
designed GTMDI, the structural displacement can be reduced significantly; simultaneously, less input energy is trans-
ferred into the entire system and much less energy is dissipated by the primary structure. Although only a limited num-
ber of ground motions have been conducted, the analysis result suggests that the two proposed closed form equations

F I G U R E 7 Frequency-domain
transfer function curves of the structural
displacement and velocity responses in
Cases I to III

F I G U R E 8 Displacement responses
of structures with and without a GTMDI
in Case II (γ t = 0.50 and γ d/γ = 1.59)
16 of 19 ZHAO ET AL.

F I G U R E 9 Energy responses of the


structures with and without a GTMDI in
Case II (α2 = 0.23, and η = 0.93)

(Equations 9 and 18) hold true for a structure-GTMDI subjected to actual seismic excitations; and the proposed design
formulae based on the stochastic performance indices can guarantee a satisfactory performance for structures.

5 | C ON C L US I ON

For the inerter-based tuned mass dampers, this study discovered an intrinsic effect that the inerter reduces the input
energy transmitted into the entire controlled structures. Closed-form dynamic responses and power equation are
derived for a generic tuned mass damper inerter (GTMDI) to establish the theoretical basis of displacement control
effect and reduction in the input energy. Inspired by the theoretical essence of the intrinsic energy-based benefit, an
energy-based optimal design framework is established. The primary conclusions of this study can be summarized as
follows:

1. It is an intrinsic benefit of the GTMDI that the inerter is capable of reducing the input energy transmitted into
the controlled structure. The derived energy dissipation equation theoretically explains this benefit, where an
analytical factor, that is, the input energy index to evaluate the input power and energy in a closed form,
counts on the total gravitational masses, inerter, and its distribution pattern. The introduction of the variable
distribution of the inerter in GTMDI facilitates the possibility to pursue a higher efficiency of the input energy
reduction.
2. The derived displacement distribution and power equations explicitly report the basic relationship between the
displacements and the energies of the structure-GTMDI systems, and the key parameters of the GTMDI in a
concise manner. Combining the two equations, the GTMDI is characterized with dual benefits, referring to the
enhanced dissipation power and the reduction effect of the input energy transferred into the entire structure-
GTMDI system.
3. Following the design philosophy of saving total energy costs (equal to input energy), the developed energy-based
design strategies can explicitly quantify the distribution pattern of the inerter by the power equation. After the acqui-
sition of the distribution pattern of the inerter, both the numerical and analytical solutions to the design formulae
are effective in providing the primary structure with a satisfactory displacement performance, while simultaneously
ensuring low energy responses. From the perspective of input energy reduction, the tuned mass of the GTMDI is
optimized zero and totally replaced by the separating inerters, which is basically consistent with the philosophy of
the lightweight control.
4. The proposed design equations and strategies are suitable for an SDOF structure with a GTMDI considering
the white noise and historic ground motions. In general, this study provides insight into the essential advanta-
geous feature of the GTMDI with two separating inerters can reduce the input energy, which also contributes
to the displacement distribution and power equations in the closed-form to reveal the working mechanism of
the GTMDI, and especially the quantitative functionality of the inerter. Further research into the actual
experimental behavior of the GTMDI device as well as the implementation of the proposed GTMDI
scheme through laboratory tests are definitely required to validate the theoretical results performed in
this work.
ZHAO ET AL. 17 of 19

ACK NO WLE DGE MEN TS


This study was supported by the National Natural Science Foundation of China (grant no. 51978525 and 51778489); the
Basic Research Project of State Key Laboratory of Ministry of Science and Technology (grant no. SLDRCE19A-02); and
the Foundation of Shanghai Science and Technology Commission (19DZ1202500).

ORCID
Zhipeng Zhao https://orcid.org/0000-0002-6324-5895
Ruifu Zhang https://orcid.org/0000-0003-2988-4412
Chao Pan https://orcid.org/0000-0003-2001-6766

R EF E RE N C E S
1. Takewaki I, Murakami S, Yoshitomi S, Tsuji M. Fundamental mechanism of earthquake response reduction in building structures with
inertial dampers. Struct Control Health Monit. 2012;19:590-608. https://doi.org/10.1002/stc.457
2. Ikago K, Saito K, Inoue N. Seismic control of single-degree-of-freedom structure using tuned viscous mass damper. Earthq Eng Struct
Dyn. 2012;41:453-474. https://doi.org/10.1002/eqe.1138
3. Papageorgiou C, Houghton NE, Smith MC. Experimental testing and analysis of inerter devices. J Dyn Syst Meas Control. 2008;131(1):
011001. https://doi.org/10.1115/1.3023120
4. Málaga-Chuquitaype C, Menendez-Vicente C, Thiers-Moggia R. Experimental and numerical assessment of the seismic response of steel
structures with clutched inerters. Soil Dyn Earthq Eng. 2019;121:200-211. https://doi.org/10.1016/j.soildyn.2019.03.016
5. Zhang Z, Lu ZQ, Ding H, Chen LQ. An inertial nonlinear energy sink. J Sound Vib. 2019;450:199-213. https://doi.org/10.1016/j.jsv.2019.
03.014
6. Garrido H, Curadelli O, Ambrosini D. Resettable-inertance inerter: a semiactive control device for energy absorption. Struct Control
Health Monit. 2019;26:e2415. https://doi.org/10.1002/stc.2415
7. Kras A, Gardonio P. Active vibration control unit with a flywheel inertial actuator. J Sound Vib. 2020;464:114987. https://doi.org/10.
1016/j.jsv.2019.114987
8. Masri SF, Caffrey JP. Transient response of a SDOF system with an inerter to nonstationary stochastic excitation. J Appl Mech. 2017;84:
041005. https://doi.org/10.1115/1.4035930
9. Chen QJ, Zhao ZP, Zhang RF, Pan C. Impact of soil–structure interaction on structures with inerter system. J Sound Vib. 2018;433:1-15.
https://doi.org/10.1016/j.jsv.2018.07.008
10. Zhao ZP, Chen QJ, Zhang RF, Pan C, Jiang YY. Energy dissipation mechanism of inerter systems. Int J Mech Sci. 2020;184:105845.
https://doi.org/10.1016/j.ijmecsci.2020.105845
11. Javidialesaadi A, Wierschem NE. Optimal design of rotational inertial double tuned mass dampers under random excitation. Eng Struct.
2018;165:412-421. https://doi.org/10.1016/j.engstruct.2018.03.033
12. Asai T, Watanabe Y. Outrigger tuned inertial mass electromagnetic transducers for high-rise buildings subject to long period earth-
quakes. Eng Struct. 2017;153:404-410. https://doi.org/10.1016/j.engstruct.2017.10.040
13. Basili M, De Angelis M, Pietrosanti D. Modal analysis and dynamic response of two adjacent single-degree-of-freedom systems linked by
spring-dashpot-inerter elements. Eng Struct. 2018;174:736-752. https://doi.org/10.1016/j.engstruct.2018.07.048
14. Krenk S. Resonant inerter based vibration absorbers on flexible structures. J Franklin Inst. 2019;356:7704-7730. https://doi.org/10.1016/j.
jfranklin.2018.11.038
15. Taflanidis AA, Giaralis A, Patsialis D. Multi-objective optimal design of inerter-based vibration absorbers for earthquake protection of
multi-storey building structures. J Franklin Inst. 2019;356:7754-7784. https://doi.org/10.1016/j.jfranklin.2019.02.022
16. Huang Z, Hua X, Chen Z, Niu H. Optimal design of TVMD with linear and nonlinear viscous damping for SDOF systems subjected to
harmonic excitation. Struct Control Health Monit. 2019;26:e2413. https://doi.org/10.1002/stc.2413
17. Palacios-Quiñonero F, Rubió-Massegú J, Rossell JM, Karimi HR. Design of inerter-based multi-actuator systems for vibration control of
adjacent structures. J Franklin Inst. 2019;356:7785-7809. https://doi.org/10.1016/j.jfranklin.2019.03.010
18. Zhao ZP, Chen QJ, Zhang RF, Jiang YY, Xia YY. Interaction of two adjacent structures coupled by inerter-based system considering soil
conditions. J Earthq Eng. 2020;21:1-21. https://doi.org/10.1080/13632469.2020.1778585
19. Zhang RF, Zhao ZP, Dai KS. Seismic response mitigation of a wind turbine tower using a tuned parallel inerter mass system. Eng Struct.
2019;180:29-39. http://doi.org/10.1016/j.engstruct.2018.11.020
20. Jiang YY, Zhao ZP, Zhang RF, De Domenico D, Pan C. Optimal design based on analytical solution for storage tank with inerter isola-
tion system. Soil Dyn Earthq Eng. 2019;129:105924. https://doi.org/10.1016/j.soildyn.2019.105924
21. Ma R, Bi K, Hao H. Mitigation of heave response of semi-submersible platform (SSP) using tuned heave plate inerter (THPI). Eng Struct.
2018;177:357-373. https://doi.org/10.1016/j.engstruct.2018.09.085
22. Lazar IF, Neild SA, Wagg DJ. Vibration suppression of cables using tuned inerter dampers. Eng Struct. 2016;122:62-71. https://doi.org/
10.1016/j.engstruct.2016.04.017
23. Wang FC, Lee CH, Zheng RQ. Benefits of the inerter in vibration suppression of a milling machine. J Franklin Inst. 2019;356:7689-7703.
https://doi.org/10.1016/j.jfranklin.2019.02.002
18 of 19 ZHAO ET AL.

24. Giaralis A, Petrini F. Wind-induced vibration mitigation in tall buildings using the tuned mass-damper-inerter. J Struct Eng. 2017;143
(9):04017127. https://doi.org/10.1061/(asce)st.1943-541x.0001863
25. Dai J, Xu ZD, Gai PP. Tuned mass-damper-inerter control of wind-induced vibration of flexible structures based on inerter location. Eng
Struct. 2019;199:109585. https://doi.org/10.1016/j.engstruct.2019.109585
26. Pradono MH, Iemura H, Igarashi A, Kalantari A. Application of angular-mass dampers to base-isolated benchmark building. Struct Con-
trol Health Monit. 2008;15(5):737-745. https://doi.org/10.1002/stc.270
27. Kawamata S. Development of a vibration control system of structures by means of mass pumps. Tokyo, Japan: Institute of Industrial Sci-
ence, University of Tokyo; 1973.
28. Kida H, Watanabe Y, Nakaminami S, et al. Full-scale dynamic tests of tuned viscous mass damper with force restriction mechanism and
its analytical verification. J Struct Constr Eng. 2011;76:1271-1280. https://doi.org/10.3130/aijs.76.1271
29. Arakaki T, Kuroda H, Arima F, Inoue Y, Baba K. Development of seismic devices applied to ball screw: part 1 Basic performance test of
RD-series. J Arch Build Sci. 1999;8:239-244. https://doi.org/10.3130/aijt.5.239_1
30. Arakaki T, Kuroda H, Arima F, Inoue Y, Baba K. Development of seismic devices applied to ball screw: part 2 Performance test and eval-
uation of RD-series. J Arch Build Sci. 1999;9:365-370. https://doi.org/10.3130/aijt.5.265
31. Saito K, Inoue N. A study on optimum response control of passive control systems using viscous damper with inertial mass: substituting
equivalent nonlinear viscous elements for linear viscous elements in optimum control systems. AIJ J Technol Design. 2007;13:457-462.
32. De Domenico D, Ricciardi G. Improving the dynamic performance of base-isolated structures via tuned mass damper and inerter
devices: a comparative study. Struct Control Health Monit. 2018;25:e2234. https://doi.org/10.1002/stc.2234
33. Zhao ZP, Zhang RF, Jiang YY, Pan C. Seismic response mitigation of structures with a friction pendulum inerter system. Eng Struct.
2019;193:110-120. https://doi.org/10.1016/j.engstruct.2019.05.024
34. De Angelis M, Giaralis A, Petrini F, Pietrosanti D. Optimal tuning and assessment of inertial dampers with grounded inerter for vibra-
tion control of seismically excited base-isolated systems. Eng Struct. 2019;196:109250. https://doi.org/10.1016/j.engstruct.2019.05.091
35. Čakmak D, Tomičevic Z, Wolf H, Božic Ž, Semenski D, Trapic I. Vibration fatigue study of the helical spring in the base-excited inerter-
based isolation system. Eng Failure Anal. 2019;103:44-56. https://doi.org/10.1016/j.engfailanal.2019.04.064
36. Marian L, Giaralis A. Optimal design of a novel tuned mass-damper–inerter (TMDI) passive vibration control configuration for stochasti-
cally support-excited structural systems. Probabilist Eng Mech. 2014;38:156-164. https://doi.org/10.1016/j.probengmech.2014.03.007
37. Lazar IF, Neild SA, Wagg DJ. Using an inerter-based device for structural vibration suppression. Earthq Eng Struct Dyn. 2014;43:
1129-1147. https://doi.org/10.1002/eqe.2390
38. Zhao ZP, Zhang RF, Jiang YY, Pan C. A tuned liquid inerter system for vibration control. Int J Mech Sci. 2019;164:105171. https://doi.
org/10.1016/j.ijmecsci.2019.105171
39. Zhao ZP, Zhang RF, Lu Z. A particle inerter system for structural seismic response mitigation. J Franklin Inst. 2019;356:7669-7688.
https://doi.org/10.1016/j.jfranklin.2019.02.001
40. Frahm H. Devices for damping vibrations of bodies, in, US, 1911. 989, 958.
41. Kim SM, Wang S, Brennan MJ. Optimal and robust modal control of a flexible structure using an active dynamic vibration absorber.
Smart Mater Struct. 2011;20(4):045003. https://doi.org/10.1088/0964-1726/20/4/045003
42. Weber F. Optimal semi-active vibration absorber for harmonic excitation based on controlled semi-active damper. Smart Mater Struct.
2014;23(9):095033. https://doi.org/10.1088/0964-1726/23/9/095033
43. Den Hartog JP. Mechanical vibrations. 4th ed. New York: McGraw-Hill, Dover; 1956.
44. Warburton GB. Optimum absorber parameters for various combinations of response and excitation parameters. Earthq Eng Struct Dyn.
1982;10(3):381-401. https://doi.org/10.1002/eqe.4290100304
45. Chen QJ, Zhao ZP, Xia YY, Pan C, Luo H, Zhang RF. Comfort based floor design employing tuned inerter mass system. J Sound Vib.
2019;458:143-157. https://doi.org/10.1016/j.jsv.2019.06.019
46. Garrido H, Curadelli O, Ambrosini D. Improvement of tuned mass damper by using rotational inertia through tuned viscous mass
damper. Eng Struct. 2013;56:2149-2153. https://doi.org/10.1016/j.engstruct.2013.08.044
47. Hu Y, Wang J, Chen MZQ, Li Z, Sun Y. Load mitigation for a barge-type floating offshore wind turbine via inerter-based passive struc-
tural control. Eng Struct. 2018;177:198-209. https://doi.org/10.1016/j.engstruct.2018.09.063
48. Cao L, Li C, Chen X. Performance of multiple tuned mass dampers-inerters for structures under harmonic ground acceleration. Smart
Struct Syst. 2020;26:49-61. https://doi.org/10.12989/sss.2020.26.1.049
49. Pietrosanti D, De Angelis M, Basili M. Optimal design and performance evaluation of systems with tuned mass damper inerter (TMDI).
Earthq Eng Struct Dyn. 2017;46:1367-1388. https://doi.org/10.1002/eqe.2861
50. Petrini F, Giaralis A, Wang ZX. Optimal tuned mass-damper-inerter (TMDI) design in wind-excited tall buildings for occupants' comfort
serviceability performance and energy harvesting. Eng Struct. 2020;204:16. https://doi.org/10.1016/j.engstruct.2019.109904
51. Crandall SH, Mark WD. Random vibration in mechanical system. Academic Press; 2014.
52. Zhang RF, Zhao ZP, Pan C, Ikago K, Xue ST. Damping enhancement principle of inerter system. Struct Control Health Monit. 2020;27
(5):e2523. https://doi.org/10.1002/stc.2523
53. Reggio A, Angelis MD. Optimal energy-based seismic design of non-conventional tuned mass damper (TMD) implemented via inter-
story isolation. Earthq Eng Struct Dyn. 2015;44:1623-1642. https://doi.org/10.1002/eqe.2548
54. Uang C, Bertero V. Use of energy as a design criterion in earthquake-resistant design, in: Technical Report UCB/EERC, Earthquake
Engineering Research Center, University of California at Berkeley, 1988.
ZHAO ET AL. 19 of 19

55. Sugimura Y, Goto W, Tanizawa H, Saito K, Nimomiya T. Response control effect of steel building structure using tuned viscous mass
damper, in: 15th world conference on earthquake engineering, Lisbon, 2012.
56. Pan C, Zhang RF. Design of structure with inerter system based on stochastic response mitigation ratio. Struct Control Health Monit.
2018;25(6):e2169. https://doi.org/10.1002/stc.2169
57. PEER. PEER Ground Motion Database, in, Pacific Earthquake Engineering Research Center, 2020.

How to cite this article: Zhao Z, Zhang R, Pan C, Chen Q, Jiang Y. Input energy reduction principle of
structures with generic tuned mass damper inerter. Struct Control Health Monit. 2020;e2644. https://doi.org/10.
1002/stc.2644

You might also like