seismic vibration control of bridges with nonlinear tuned

39
Seismic vibration control of bridges with nonlinear tuned mass dampers. Final Report December 2020 Rajesh Rupakhety Said Elias

Upload: others

Post on 18-Dec-2021

9 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Seismic vibration control of bridges with nonlinear tuned

Seismic vibration control of bridges with

nonlinear tuned mass dampers.

Final Report

December 2020

Rajesh Rupakhety

Said Elias

Page 2: Seismic vibration control of bridges with nonlinear tuned

Seismic vibration control of bridges with

nonlinear tuned mass dampers.

Final report

Rajesh Rupakhety

Said Elias

Report No. 2020-002

Selfoss 2020

Page 3: Seismic vibration control of bridges with nonlinear tuned

Rajesh Rupakhety, Said Elias, Seismic vibration control of bridges with

nonlinear tuned mass dampers. Earthquake Engineering Research

Centre, University of Iceland, Report No. 2020-002 Selfoss, 2020

© Earthquake Engineering Research Centre, University of Iceland, , and

authors

All rights reserved. No part of this publication may be reproduced,

transmitted, or distributed in any form or by any means, electronic or

mechanical, including photocopying, without permission from the

Earthquake Engineering Research Centre or the authors.

Page 4: Seismic vibration control of bridges with nonlinear tuned

Abstract

This study presents the design methodology and effectiveness of a

hysteretic tuned mass dampers in controlling seismic response of

structures. The hysteretic dampers (n-TVAs) consist of a mass attached

by a hysteretic spring to the host structure. The stiffness and strength of

the spring is tuned in such a way that the hysteretic energy dissipated by

the damper is equal to the viscous energy dissipated by a linear optimal

tuned mass damper. Due to the non-linear behaviour of the device,

optimal design is excitation dependent. A general procedure to design the

optimal parameters of the device is proposed in this study. The procedure

relies on an iterative solution of the equation of motion on a nonlinear

system. Analysis using some typical ground motions show that the

iterative procedure converges rapidly. The resulting n-TVAs are found

to be effective in controlling seismic response of a reinforced concrete

bridge structure, used here as a case study example. The proposed device

is at least as effective as existing TMD solutions. Investigation of the

proposed device using recorded and simulated ground motions

corresponding to a typical near-fault scenario in South Iceland shows that

the effectiveness of the device is highly dependent on the ground motion,

and for those ground motions which exert excessive demands on the

uncontrolled structure, the proposed devices can reduce the base shear

and mid-span displacement demands of the case study bridge by as much

as 25%.

Page 5: Seismic vibration control of bridges with nonlinear tuned

Contents

1. Introduction ......................................................................................................................7

2. Design and evaluation methodology ................................................................................9

3. Numerical study ..............................................................................................................13 3.1. Force deformation curves ............................................................................................................. 14 3.2. System transfer function ................................................................................................................ 14 3.3. Energy assessment ........................................................................................................................... 16 3.4. Evaluation of the effectiveness of TVAs ................................................................................... 17

4. Evaluation of effectives is South Iceland ......................................................................22 4.1. Design of the n-TVAs ....................................................................................................................... 23 4.2. Robustness of the n-TVAs .............................................................................................................. 24

5. Conclusions ......................................................................................................................29

ACKNOWLEDGEMENTS ...............................................................................................31

DISCLAIMER ....................................................................................................................31

BIBLIOGRAPHY ...............................................................................................................32

Page 6: Seismic vibration control of bridges with nonlinear tuned

List of figures

FIGURE 1. (A) MODEL OF CONTINUOUS BRIDGE; (B) REPRESENTATION OF THE TVA IN X

AND Y DIRECTIONS; (C) FORCE DEFORMATION CURVE OF THE N-TVA. 10 FIGURE 2. FLOW CHART FOR DESIGN OF THE NON-LINEAR SPRING OF THE N-TVAS. 14 FIGURE 3. FORCE DEFORMATION CURVES OF THE NON-LINEAR SPRING OF THE N-

TVAS SUBJECTED TO DIFFERENT GROUND MOTIONS. 15 FIGURE 4. NORMALIZED TRANSFER FUNCTIONS OF THE BRIDGE MID-SPAN

ACCELERATION TO WHITE NOISE GROUND ACCELERATION IN LONGITUDINAL

AND TRANSVERSE DIRECTIONS OF THE BRIDGE. 16 FIGURE 5. INPUT, DAMPING AND STRAIN ENERGY FOR DIFFERENT TVA SCHEMES

SUBJECTED TO THREE DIFFERENT GROUND MOTIONS. 17 FIGURE 6. TIME VARIATION OF THE NORMALIZED PIER BASE SHEAR UNDER THE 1940

IMPERIAL VALLEY, EARTHQUAKE GROUND MOTION FOR NC, C-TVAS, O-TVAS,

AND N-TVAS. 18 FIGURE 7. SAME AS IN FIGURE 6, BUT FOR THE 1989 LOMA PRIETA EARTHQUAKE GROUND

MOTION. 18 FIGURE 8. SAME AS IN FIGURE 6, BUT FOR THE 1995 KOBE EARTHQUAKE GROUND MOTION.

19 FIGURE 9. TIME VARIATION OF THE MID-SPAN DISPLACEMENT UNDER THE 1940 IMPERIAL

VALLEY EARTHQUAKE GROUND MOTION. 19 FIGURE 10. SAME AS IN FIGURE 9 BUT FOR THE 1989 LOMA PRIETA EARTHQUAKE GROUND

MOTION. 20 FIGURE 11. SAME AS IN FIGURE 9 BUT FOR THE 1995 KOBE EARTHQUAKE GROUND MOTION.

20 FIGURE 12. COMPARISON OF THE PERFORMANCE OF N-TVA WITH/WITHOUT DASHPOT TO

THAT OF O-TVA. THE TOP PANEL SHOWS MID-SPAN DISPLACEMENT IN THE LONGITUDINAL AND TRANSVERSE DIRECTIONS, AND THE BOTTOM PANEL SHOWS CORRESPONDING STROKE OF THE DEVICES. 22

FIGURE 13. ELASTIC RESPONSE SPECTRA (5% DAMPED) OF THE RECORDED AND SIMULATED GROUND MOTIONS USED IN THIS STUDY. THE RIGHT AND LEFT COLUMNS REPRESENT SEISMIC ACTION IN THE TRANSVERSE AND THE LONGITUDINAL DIRECTIONS OF THE BRIDGE, RESPECTIVELY. 23

FIGURE 14. BASE SHEAR DEMAND ON THE STRUCTURE (TOP: LONGITUDINAL AND BOTTOM: TRANSVERSE) OF THE UNCONTROLLED AND CONTROLLED STRUCTURE SUBJECTED TO SIMULATE GROUND MOTION NUMBER 2. 25

FIGURE 15. SAME AS IN FIGURE 13, BUT FOR GROUND MOTION RECORDED AT HELLA STATION. 26

FIGURE 16. SAME AS IN FIGURE 14 BUT FOR KALDARHOLT STATION. 26 FIGURE 17. SAME AS IN FIGURE 14 BUT FOR SELSUND STATION. 27 FIGURE 18. SAME AS IN FIGURE 14 BUT FOR THORSARBRU STATION. 27 FIGURE 19. DISPLACEMENT RESPONSE OF THE CONTROLLED AND UNCONTROLLED

STRUCTURE WHEN SUBJECTED TO GROUND MOTION RECORDED AT THE FLAGBJARNARHOLT STATION; THE TOP AND BOTTOM ROWS CORRESPOND TO LONGITUDINAL AND TRANSVERSE DISPLACEMENTS, RESPECTIVELY, AT THE MID-SPAN OF THE BRIDGE. 28

FIGURE 20. SAME AS IN FIGURE 19 BUT FOR HELLA STATION. 28 FIGURE 21. SAME AS IN FIGURE 19 BUT FOR KALDARHOLT STATION. 29

Page 7: Seismic vibration control of bridges with nonlinear tuned

List of tables

TABLE 1. PEAK RESPONSES FOR DIFFERENT CONTROL SCHEMES. ..................................................... 21 TABLE 2. RESPONSE REDUCTION AND STOKE OF THE N-TVAS CORRESPONDING TO THE 8

DESIGN GROUND MOTIONS. ......................................................................................................................... 24 TABLE 3. BASE-SHEAR REDUCTION DUE TO N-TVAS COMPARED TO C-TVAS AND O-

TVAS ....................................................................................................................................................................... 24

Page 8: Seismic vibration control of bridges with nonlinear tuned

1. Introduction

Lifeline structures like bridges need to be especially robust against natural hazards such as

earthquakes and strong winds. One approach to performance enhancement of bridges

against dynamic loads is through vibration control using passive, active, semi-active and

hybrid systems. Passive devices have the advantages of being simple to install and cost-

effective as compared to active, semi-active and hybrid systems. Generally, these controlled

structures assumed to remain in elastic range. An alternate, but time consuming, and

perhaps, not as insightful theoretically, is the Monte Carlo approach of response simulation.

A well-controlled structure is not expected to be stressed too much beyond its elastic

capacity, and the general trend in analysis of vibration control methods has been to assume

structures responding elastically. Base isolation, a passive device, is one of the most widely

used technology to reduce seismic response of bridges (Nagarajaiah et al., 1991; Jangid,

2004; Sahasrabudhe and Nagarajaiah, 2005; Agrawal et al, 2009; Nagarajaiah et al., 2009;

Madhekar and Jangid, 2010; Attary et al., 2015; Elias and Matsagar, 2017, 2019). The other

simple passive vibration control devices are the so-called tuned vibration absorbers (TVAs).

TVAs are widely used in tall buildings and bridges around the world. For broadband

efficiency in seismic circumstances, one of the practical options is to use the multiple TVAs

(MTVAs) with distributed natural frequencies, which have been studied by Li (2000, 2002),

Li and Liu (2003). Studies have shown that both MTVAs and tuned tandem mass dampers

(TTMDI) are robust devices for vibration mitigation of structures (Li and Cao, 2019; Cao

and Li, 2019; Chang and Li, 2019). Robustness of the vibration control scheme is as off-

tuning can have adverse effects on the structure. Performance enhancement of TVAs is

gaining widespread research interest. Lin et al. (2009) reported that the TVAs are more

effective in vibration control of structures under impulse‐like ground motion with more

forward‐and‐backward cycles. Aldemir et al. (2012) compared the performance of

passive and active TVAs in seismic vibration control of structures. They found that

conventional optimally designed TVAs could reduce the displacement response of the

structure but were not effective in controlling acceleration response. Active TVAs were

more found to be more effective to suppress both displacement and acceleration responses.

Generally, TVAs are useful for i) wind response mitigation of long span bridges, ii)

mitigation of vertical vibration of bridges caused by vehicles, iii) vibration control of

pedestrian bridges, and iv) seismic response mitigation of bridges.

Effectiveness of TVAs in reducing wind-induced vibrations of long-span bridges has been

studied by Lin et al. (2000a and 2000b), Gu et al. (2001), Pourzeynali and Datta (2002),

Chen and Kareem (2003), and Kwon and Park (2004). These studies showed that TVAs are

effective in reducing buffeting response of such bridges. Chen and Wu (2008) studied the

efficiency of multiple TVAs (MTVAs) in controlling wind-induced vibration of bridges.

Casciati and Giuliano (2009) demonstrated the usefulness of MTVAs in reducing gust

response of towers in suspension bridges. Casalotti et al. (2014) estimated the parameters

of hysteretic TVAs for multi-mode flutter mitigation in long-span suspension bridges.

Bortoluzzi et al. (2015) reported significant reduction in mid-span vibration of TVA-

Page 9: Seismic vibration control of bridges with nonlinear tuned

8

controlled bridges subjected to wind forces. Ubertini (2010) and Ubertini et al. (2015)

present a probabilistic approach for arranging TVAs for flutter suppression in long-span

bridges. Verstraelen et al. (2016) found that the effectiveness of the absorbers obtained from

wind tunnel tests was higher than mathematically predicted.

Similarly, many researchers (Yang et al., 1997; Moreno and Dos-Santos, 1997; Kwon et al.,

1998; Wang et al., 2003; Chen and Cai, 2004; Chen and Chen, 2004; Chen and Huang,

2004; Yau and Yang, 2004a; Yau and Yang, 2004b; Li et al., 2005; Lin et al., 2005; Wu and

Cai, 2007; Hijmissen and van Horssen, 2007; and Hijmissen et al., 2009) have demonstrated

the usefulness of TVAs in reducing vertical displacements, absolute accelerations, end

rotations, and train accelerations during resonant speeds in bridges. Improved performance

of the MTVAs as compared to the TVAs was also reported in these studies. Liu et al. (2012)

showed the efficiency of MTVAs in mitigating vibrations of bridges caused by high speed

trains. TVAs have also been found to be useful in reducing vibrations of footbridges

(Dallard et al., 2001a, 2001b; Poovarodom et al. 2001, 2002, 2003; Yang et al., 2007; and

Hoang et al., 2008). Li et al. (2010) conclude that MTVAs designed by random optimization

were more efficient than conventionally designed ones in minimizing dynamic response

during crowd-footbridge resonance, and that appropriate frequency spacing improvement

efficiently diminishes the off-tuning effect of the MTVAs. Daniel et al. (2012) studied the

use of MTVAs for multi-modal control of pedestrian bridges. They concluded that MTVAs

are more robust than TVAs. Andersson et al. (2014) studied the use of passive and adaptive

damping systems to mitigate vibrations in a railway bridge during resonance. It was reported

that passive TVA and pendulum dampers showed a significant improvement under the same

detuned conditions. Recently, van Nimmen et al. (2016) concluded that a simplified design

procedure based solely on the contribution of the resonant mode is not appropriate for

evaluating the dynamic performance of footbridges installed with TVAs. Lievens et al.

(2016) studied robust TVAs in vibration mitigation of footbridges. Tubino et al. (2016)

studied MTVAs for vibration mitigation of footbridges subjected to several loading

conditions. They concluded that the MTVAs reduced the vibration amplitudes considerably

under almost all loading conditions considered in their study. Earthquake-induced vibration

is also a major concern in design, operation and maintenance of bridges. As reported by Lee

et al. (2013), many bridges in the United States of America (USA) were damaged or totally

collapsed by earthquakes during 1980-2012. The study showed that concrete bridges are

more vulnerable to earthquakes than steel bridges. Vestroni et al. (2014), and Carpineto et

al. (2014) reported experimental tests on a simply-supported beam with hysteretic TMDs

under harmonic and random base excitations. The hysteretic TMDs placed at the beam mid-

span was found to diminish vibrations in the first flexural mode by up to 98% for harmonic

excitations and 87% for random excitations. Pisal and Jangid (2016) conclude that, for

seismic excitation, optimized multiple tuned vibration friction absorbers (MTVFAs)

positioned at mid-span are more efficient in response reduction than optimal single TVFA

(STVFA) and MTVFAs distributed along the length of the bridge. Miguel et al. (2016a and

2016b) showed the effectiveness of robust TVAs in controlling the vibration of bridges

subjected to ground motion. They showed that classical analytical techniques could be

effectively used in robust optimal design of the STVA. Matin et al. (2014, 2017, and 2018)

Page 10: Seismic vibration control of bridges with nonlinear tuned

9

presented multi-mode control of continuous concrete bridges using MTVAs. They reported

that MTVAs can reduce seismic response considerably, and that their performance is

dependent on their spatial placement on the bridge. In a recent study, Lu et al. (2018)

reported that the TVA damping coefficient was significantly increased by the introduction

of eddy‐current damping mechanism. A detail review of nonlinear dissipative devices is

available in Lu et al. (2018).

In most studies, TVAs installed on bridges are assumed to be elastic. This assumption

implies unlimited displacement capacity of such devices. In practice, such devices are

expected to yield under certain conditions. It is therefore realistic to model their inelastic

behavior. In fact, non-linear yielding behavior can be utilized to reduce displacement

demand on the structure. The main objective of the present study is to investigate design of

non-linear TVAs (n-TVAs) and their efficiency in reducing seismic response of bridges. In

this study, the springs of the TVAs are modelled with hysteretic behavior. Design of TVAs

with non-linear springs is formulated and the performance of the device is compared to their

linear counterparts, such as conventional linear TVSs (c-TVAs) and optimum linear TVAs

(o-TVAs).

2. Design and evaluation methodology

A reinforced concrete (RC) continuous span bridge, as shown in Figure 1, is considered in

this study. The piers and rigid abutments support the conventional bridge bearings as

illustrated in Figure 1(a). In Figure 1(b), kx and cx are the stiffness and damping coefficients

of the TVA in x(longitudinal) direction. Similarly, ky and cy are the stiffness and damping

coefficients in y (transverse) direction. The mass m is attached to both springs to control the

bidirectional responses of the bridge. Figure 1 shows the details of TVA for only one of the

spans of the bridge. Similar TVAs are provided in all the spans of the bridge. The springs

are inelastic with hysteresis curves like the one shown in Figure 1(c). The equations of

motion of the bridge equipped with linear TVAs is

s s s tM Q C Q K Q F+ + = (1)

where ][ sM , ][ sC , and ][ sK are the mass, damping, and stiffness matrices of the controlled

bridge, respectively of order )22()22( nNnN ++ ; N is the number of degrees of freedom

of the bridge model and n is the number of TVAs, respectively;

T

1 2 N 1 n 1 2 N 1 n, ,{ } { , , , , , }Q U U U u u V V V v v= , }{Q , and }{Q are the displacement, velocity, and

acceleration vectors, respectively. In the equation, {Ft} is the forcing function given by

s x gx s y gyM r a M r a− − , where xr and yr are the influence coefficient vectors in the

longitudinal and transverse directions, and gxa and gya are the corresponding ground

Page 11: Seismic vibration control of bridges with nonlinear tuned

10

accelerations. Moreover, Ui and Vi are the displacements of the ith degree of freedom of the

bridge in the longitudinal and transverse directions, respectively, and iu , iv are those of the

ith TVA. When the TVAs have nonlinear springs, the equation of motion needs to be solved

numerically using iterative methods.

Figure 1. (a) Model of continuous bridge; (b) Representation of the TVA in X and Y

directions; (c) Force deformation curve of the n-TVA.

The design parameters of linear conventional TVAs (c-TVAs) are given by

2

x l xk m= (2-a)

2

y l yk m= (2-b)

where, kx and ky are the stiffness of the c-TVAs in longitudinal and transverse directions

respectively; ml is the mass of c-TVAs placed on each span; and x and

y are the

corresponding vibration frequencies, which are assumed to be the same as the fundamental

frequency of the bridge in the X and Y directions. For a given damping ratio d , the

damping coefficients (xc and

yc ) of the c-TVAs are calculated as:

2x d l xc m = (3-a)

2y d l yc m = (3-b)

In case of the optimum linear TVAs (o-TVAs), the frequency of the device in each direction

is a fixed proportion ( optf ) of the corresponding fundamental frequency of the bridge (1 1,x y

).

1 1, ,x y optx x opty yf f = (4-a)

The fixed proportion is known as the optimum tuning frequency ratio. This ratio, for lightly

damped structure, like the bridge being considered here, is given by Sadek et al. (1997)

11

1 1

x

optx optx

x x

f

= − + +

(4-b)

Page 12: Seismic vibration control of bridges with nonlinear tuned

11

11

1 1

y

opty opty

y y

f

= − + +

(4-c)

where, the optimum damping ratios are given by

1 1

x

doptx x

= + + +

(5-a)

1 1

y

dopty y

= + + +

(5-b)

In Equations (4) and (5), is the ratio of the mass of the TVA to the total mass of the

bridge; x and y are the mode shape amplitudes of the bridge at the location of TVA in

longitudinal (X) and transverse (Y) directions, respectively. Similarly, x and y are the

damping ratios of the bridge in X and Y directions, respectively. Having obtained the

optimized frequencies from equations 4 and 5, the stiffness and damping coefficients of the

o-TVA can be computed from equations 2 and 3, respectively.

In case of non-linear TVAs (n-TVAs), the hysteretic behavior of the springs is assumed to

follow the Wen (1976) model. According to this model, the restoring force of the n-TVAs

in x and y directions are given as,

(1 )nx nx xx xF ak u a F Z= + − (6-a)

(1 )ny ny yy yF ak v a F Z= + − (6-b)

where, xxF , and

yyF are the yield strengths of the spring in x and y directions, respectively;

a is the ratio of post-yielding stiffness to initial stiffness, represented by nxk and

nyk ; u and

v are the displacements of the springs. xZ , and

yZ are the hysteretic displacement

components satisfying the following 1st order differential equations.

1p p

yield x x xu Z Au u Z u Z −

= − − (7-a)

1p p

yield y y yv Z Av v Z v Z −

= − − (7-b)

where, yieldu and yieldv are the yield displacement and , , p , and A are dimensionless

parameters of the model. Here, p controls the curve smoothness during transition from

elastic to plastic state, and its value is fixed at 10 to simulate bilinear behavior. The

parameters and control the shape of hysteresis loop; and A is the restoring force

amplitude (Wen 1976). In this study, and A are taken as 0.5, and 1, respectively. The n-

TVAs are designed in such a way that their effective stiffness (secant stiffness at maximum

displacement) is equal to the stiffness of the o-TVAs. Let us consider peak strokes of xD

Page 13: Seismic vibration control of bridges with nonlinear tuned

12

and yD in X and Y directions. Then the ratio of yield displacement to peak stroke is defined

as /x yield xR u D= and /y yield yR v D= . This ratio controls the extent of plastic deformation in the

device and is reciprocal of ductility demand. A value of this ratio less than 1 implies inelastic

behavior. This can be treated as a design variable. Based on the assumption that the energy

dissipated by a o-TVA is the same as that dissipated by an elastic-plastic spring of the n-

TVA, an approximate value of R can be computed from the damping ratio of the o-TVA

by using the following equation.

, ,1

2x y doptx doptyR

= − (8)

For the structure being studied here, the value of R is approximately 0.75 and is taken to

be equal in the X and Y directions. For a bilinear model, the initial stiffness of the n-TVA

is related to its effective stiffness by the following equations.

( )1

x

nx

x x

kk

R a R=

− + (9-a)

( )1

y

ny

y y

kk

R a R=

− +

(9-b)

Because the peak stroke of the device is not known a priori but depends on the excitation,

an iterative procedure needs to be followed to estimate the parameters of the n-TVAs. In

the proposed method, an initial value of the peak stoke is assumed. This value can be taken

as the peak stroke of the o-TVA for a given ground motion. With this peak stroke and given

R , an initial value of yield displacement is computed, which is then used to estimate the

yield force by multiplying with the initial stiffness. With these parameters established, non-

linear time history analysis is carried out with different ground motions and the actual peak

stroke corresponding to each ground motion is computed. In the next iteration the computed

peak stroke is used, and the process is repeated until the assumed peak stroke is equal to the

computed peak stroke within some reasonable tolerance. The design and iteration is carried

out independently in X and Y directions. The results, in each direction and for each ground

motion, an estimate of the yield displacement and yield force. For each direction, the final

yield force is taken as the maximum of the yield forces induced by the different ground

motions (see Figure 2). This implies that the design is optimal for the strongest ground

motion used in the analysis but might not dissipate as much hysteretic energy when

subjected to less demanding ground motions. In this study, three different ground motions

were used, as will be described subsequently, and it was found that the iterations converged

very fast. In practical scenarios, design ground motion specified by relevant seismic

provision can be used in the same manner as is described here. Unlike the o-TVAs, viscous

dashpots are not used with n-TVAs, where damping is provided by hysteretic energy

dissipation.

Page 14: Seismic vibration control of bridges with nonlinear tuned

13

3. Numerical study

The responses of both controlled and uncontrolled bridges are computed for bidirectional

ground shaking. Seismic response of long-span structures is affected by wave propagation

and incoherency of ground motion (see, for example, Zerva 2016). In the example

considered here, the span length is only 30m and the fundamental frequency of the bridge

is ~2 Hz. In such cases, the lagged coherency is almost equal to 1 (see, for example,

Rupakhety and Sigbjörnsson, 2012 and AfifChaouch et al., 2016). In this study, for sake of

simplicity, wave passage and incoherency effects are not considered, and it is assumed that

the same ground motion is experienced by all the supports of the bridge. Further, standard

codes such as Eurocode 8 (2005); American Society of Civil Engineers (ASCE) standards

7-10 (2010); and Federal Emergency Management Agency-356 (FEMA-356, 2000) require

at least three earthquake ground motions for time history analysis of structures for seismic

design and evaluation. The ground motions used in this study are the horizontal components

of the 1940 Imperial Valley Earthquake (El Centro station); 1989 Loma Prieta Earthquake

(Los Gatos Presentation Centre); and the 1995 Kobe Earthquake (Japan Meteorological

Agency - JMA station). The peak ground acceleration (PGA) of the Imperial Valley, Loma

Prieta, and Kobe earthquake ground motions applied on the bridge are: 0.35g, 0.57g, and

0.86g in the longitudinal direction and 0.21g, 0.61g, and 0.82g in the transverse direction,

respectively, where g denotes the gravitational acceleration.

A three-span continuous reinforced concrete (RC) bridge is considered in this study. The

details of the RC bridge sections are given in Elias and Matsagar (2017). Cross-sectional

area of deck and piers are respectively, 3.57 m2 and 1.767 m2. The moment of inertia of the

deck in both principal directions of the cross-section are equal and considered to be 2.08

m4. Similarly, the moment of inertia of the piers is 0.902 m4 in both principal directions of

its cross-section. Young’s modulus of elasticity and mass per unit volume respectively are

3.6 × 107 kN/m2 and 23.536 kN/m3 for the RC material used in this bridge. Each span is 30

m long and pier height is 10 m. The fundamental frequency of the RC bridge is 1.86 Hz in

both longitudinal and transverse directions. Damping is assumed to be 5% of critical.

The c-TVAs, o-TVAs and n-TVAs are located at the center of each span. A mass ratio of

0.01 is assumed for all the types of TVAs. The total mass ratio considering TVAs in each

of the three spans is therefore 0.03. The stiffness, damping coefficient, and tuning frequency

ratio of the c-TVAs are 1.250 107 N/m, 2.133 105 N-sec/m, and 1, respectively. These

properties of the o-TVAs are 1.214 107 N/m), 3.132 105 N-sec/m. The properties of the n-

TVAs were designed in an iterative manner described above. The initial stiffness in the

longitudinal and transverse directions is 1.592 107 N/m and the corresponding yield

displacement are 18.4 cm and 9.6 cm. A flowchart of the design process of the n-TVA is

presented in Figure 2.

Page 15: Seismic vibration control of bridges with nonlinear tuned

14

Figure 2. Flow chart for design of the non-linear spring of the n-TVAs.

3.1. Force deformation curves

The force deformation curves of the n-TVA springs in longitudinal and transverse directions

are shown in Figure 3 for three different ground motions. The forces are normalized by the

weight of the n-TVAs. The design of the n-TVAs was controlled by the Kobe Earthquake

ground motion in the transverse direction and Loma Prieta Earthquake ground motion in the

longitudinal direction. The hysteresis curves shown in Figure 3 show largest inelasticity for

these ground motions (see the middle panel in the top row and the right panel in the bottom

row). When subjected to the Imperial Valley Earthquake ground motion, the n-TVAs remain

elastic.

3.2. System transfer function

The differences in dynamic properties of the different control schemes can be studied by

comparing the transfer functions of the controlled structure to that of the uncontrolled

structure. This section presents the transfer function, relating ground acceleration to the

acceleration response of the bridge mid-span. If ( )AxS and ( )AyS denote the auto-power

spectral density of the acceleration response in longitudinal and transverse directions

respectively, and ( )FxS and ( )FyS denote auto-power spectral density of corresponding

ground acceleration, the transfer functions are defined as

( )( )

( )

2 Ax

x

Fx

SH

S

=

(10-a)

( )( )

( )

2 Ay

y

Fy

SH

S

=

(10-b)

Figure 6 shows transfer functions normalized by the maximum ordinate of the transfer

function of the uncontrolled structure. The excitation is taken as white noise (see, for

example, Elias et al., 2016; 2018). The effect of the TVAs is to reduce the amplitude of the

Page 16: Seismic vibration control of bridges with nonlinear tuned

15

resonant peaks, making them wider, which is a consequence of additional damping provided

by the devices. The reduction in resonant peak is more prominent when using o-TVAs as

compared to the c-TVAs. The area under the transfer function, for a white noise excitation,

provides the ratio between the root mean square response and noise intensity. It is therefore

an indicator of the effectiveness of vibration transmission through a structure; lower area

implying reduction in vibration. These areas for different TVA schemes are indicated in

Figure 4, which shows that the n-TVAs are the most effective, although not very different

from o-TVAs. It should be noted that these conclusions apply only to white noise excitation

and reduction in root mean square response. The effectiveness of the devices in reducing

peak response is investigated in the following sections.

Figure 3. Force deformation curves of the non-linear spring of the n-TVAs subjected to

different ground motions.

-3

-2

-1

0

1

2

3

-0.2 -0.1 0.0 0.1 0.2

Norm

aliz

ed F

orc

e

Longitudinal

Stroke (m)

Imperial Valley, 1940

PGA = 0.35g PGA = 0.57g PGA = 0.86g

PGA = 0.82gPGA = 0.61gPGA = 0.21g

-3

-2

-1

0

1

2

3

-0.2 -0.1 0.0 0.1 0.2

Stroke (m)

Loma Prieta, 1989-3

-2

-1

0

1

2

3

-0.2 -0.1 0.0 0.1 0.2

Stroke (m)

Kobe, 1995

-2

-1

0

1

2

-0.1 0.0 0.1

Norm

aliz

ed F

orc

e

Transverse

Stroke (m)

-2

-1

0

1

2

-0.1 0.0 0.1

Stroke (m)

-2

-1

0

1

2

-0.1 0.0 0.1

Stroke (m)

Page 17: Seismic vibration control of bridges with nonlinear tuned

16

Figure 4. Normalized transfer functions of the bridge mid-span acceleration to white noise

ground acceleration in longitudinal and transverse directions of the bridge.

3.3. Energy assessment

Ground shaking imparts energy into a structure, part of which is dissipated through damping

and hysteresis in the structure and the remaining energy accounts for strain and kinetic

energy of the structure. The input energy at the ith time step normalized by the total mass of

the structure (Ei) is given by

( )

T T

i

1 1t

1 i i

i s x gx i s y gyE Q M r a Q M r aM

= − +

(13)

where iQ is the incremental displacement vector between two consecutive time steps. It

was reported by Elias et al. (2016) and Elias (2018) that the structures installed with the

TVAs were able to dissipate more energy than those without TVAs. The damping energy at

the ith time step per unit mass of the structure is

( )

T

1t

1 i

di i s i

i

E Q C QM =

= −

(14)

The strain energy per unit mass at the ith time step is

( )

T

t

1 1

2si i s iE Q K Q

M

=

(15)

In these calculations, the energy dissipated by the TVAs are not considered, i.e., the matrices

and vectors in equations 13-15 contain only the bridge degrees of freedom. Figure 6 shows

the accumulation of different types of energies defined above. The top panel shows damping

energy dissipated by the structure along with the input energy (dashed curve). The bottom

panel shows strain energy. The energy corresponding to different TVAs and ground motions

are marked as indicated in the figure. The results show that the damping energy

corresponding to the structural degrees of freedom is considerably reduced when TVAs are

used, which is a direct consequence of reduced vibration amplitudes. The reduction in

damping energy is up to about 50% for one of the ground motions considered here. The

1.6 1.8 2.0 2.20.0

0.2

0.4

0.6

0.8

1.0A

NC= 9.40

Longitudinal

No

rmal

ized

Tra

nsf

er F

un

ctio

n H

()

2 f

or

Acc

eler

atio

n a

t M

id-S

pan

(M

agn

itu

de)

Frequency (Hz)

NC

c-TVAs

o-TVAs

n-TVAs

Ac-TVAs

= 8.61

An-TVAs

= 7.95

Ao-TVAs

= 8.11

1.6 1.7 1.8 1.9 2.0 2.1 2.2 2.30.0

0.2

0.4

0.6

0.8

1.0

Transverse

Frequency (Hz)

ANC

= 9.30

Ac-TVAs

= 8.44

An-TVAs

= 7.32

Ao-TVAs

= 7.77

Page 18: Seismic vibration control of bridges with nonlinear tuned

17

performance of n-TVAs is much better than c-TVAs and comparable to that of o-TVAs.

Similarly, the strain energy of the bridge is considerably reduced by the TVAs.

Figure 5. Input, damping and strain energy for different TVA schemes subjected to three

different ground motions.

3.4. Evaluation of the effectiveness of TVAs

This section presents different response parameters of the uncontrolled structure and

different control schemes. Figure 6, Figure 7, and Figure 8 present base shear of the bridge

normalized by its weight for different ground motions and control schemes.

Displacement response at the mid-span is presented in Figure 9 through Figure 11. The results

show considerable decrease in base shear due to the TVAs. The peak base shears and

percentage in peak reductions (in bracket) are listed in Table 1 for all control schemes and

ground motions. The n-TVA induces up to 33.5% and 34% reduction in base shear in the

longitudinal and transverse directions, respectively. In some cases, the c-TVAs seem to

amplify base shear. Although the performance of o-TVAs and n-TVAs are, in general

comparable, n-TVAs are found to be more effective than o-TVAs for some ground motions.

In Figure 9 to Figure 111, it is observed that the displacement at mid-span of the RC bridge

is significantly reduced by o-TVAs and n-TVAs. It is evident from the results that, for the

ground motions considered here, a maximum response reduction of 22% is achieved by c-

0.00

0.09

0.18

0.27

0.36

0 10 20 30 40

50%

Ener

gy (

m2/s

ec2)

Damping Energy

Time (sec)

Imperial Valley, 1940

PGA = 0.35g PGA = 0.57g PGA = 0.86g

PGA = 0.82gPGA = 0.61gPGA = 0.21g

0.0

0.3

0.6

0.9

1.2

1.5

1.8

0 10 20 30 40 50

45%

Time (sec)

Loma Prieta, 19890.00

0.45

0.90

1.35

0 10 20 30 40

Time (sec)

Kobe, 1995

45%

0.00

0.03

0.06

0.09

0.12

0 5 10 15 20 25

Ener

gy (

m2/s

ec2)

Strain Energy

Time (sec)

0.00

0.09

0.18

0.27

0.36

10 15 20 25

NC

c-TVAs

o-TVAs

n-TVAs

Time (sec)

0.00

0.09

0.18

0.27

0.36

10 15 20 25

Time (sec)

Page 19: Seismic vibration control of bridges with nonlinear tuned

18

TVAs. At the same time, the response is amplified by 6% for one of the ground motions.

Optimization of the TVAs based on the formulations proposed by Sadek et al. (1997)

improved the performance.

Figure 6. Time variation of the normalized pier base shear under the 1940 Imperial Valley,

Earthquake ground motion for NC, c-TVAs, o-TVAs, and n-TVAs.

Figure 7. Same as in Figure 6, but for the 1989 Loma Prieta Earthquake ground motion.

-0.6

-0.3

0.0

0.3

0.6

0 5 10 15 20 25

PGA = 0.35g

Bas

eshea

r/W

eight

of

th B

ridge

(W)

Longitudinal

Time (sec)

-0.6

-0.3

0.0

0.3

0.6

0 5 10 15 20 25

Time (sec)

Imperial Valley, 1940

-0.6

-0.3

0.0

0.3

0.6

0 5 10 15 20 25

Time (sec)

-0.4

-0.2

0.0

0.2

0.4

0 5 10 15 20 25

PGA = 0.21g

Bas

eshea

r/W

eight

of

th B

ridge

(W)

Transverse

Time (sec)

-0.4

-0.2

0.0

0.2

0.4

0 5 10 15 20 25

Time (sec)

c-TVAs

o-TVAs

NC

c-TVAs

o-TVAs

n-TVAs-0.4

-0.2

0.0

0.2

0.4

0 5 10 15 20 25

Time (sec)

-1.6

-0.8

0.0

0.8

1.6

0 5 10 15 20 25

PGA = 0.57g

Bas

eshea

r/W

eight

of

th B

ridge

(W)

Longitudinal

Time (sec)

-1.6

-0.8

0.0

0.8

1.6

0 5 10 15 20 25

Time (sec)

Loma Prieta, 1989-1.6

-0.8

0.0

0.8

1.6

0 5 10 15 20 25

Time (sec)

-0.6

-0.3

0.0

0.3

0.6

0 5 10 15 20 25

PGA = 0.61g

Bas

eshea

r/W

eight

of

th B

ridge

(W)

Transverse

Time (sec)

-0.6

-0.3

0.0

0.3

0.6

0 5 10 15 20 25

Time (sec)

c-TVAs

o-TVAs

NC

c-TVAs

o-TVAs

n-TVAs

-0.6

-0.3

0.0

0.3

0.6

0 5 10 15 20 25

Time (sec)

Page 20: Seismic vibration control of bridges with nonlinear tuned

19

Figure 8. Same as in Figure 6, but for the 1995 Kobe Earthquake ground motion.

Figure 9. Time variation of the mid-span displacement under the 1940 Imperial Valley

Earthquake ground motion.

-1.6

-0.8

0.0

0.8

1.6

0 5 10 15 20 25

PGA = 0.86g

Bas

eshea

r/W

eight

of

th B

ridge

( W)

Longitudinal

Time (sec)

-1.6

-0.8

0.0

0.8

1.6

0 5 10 15 20 25

Time (sec)

Kobe, 1995-1.6

-0.8

0.0

0.8

1.6

0 5 10 15 20 25

Time (sec)

-0.8

-0.4

0.0

0.4

0.8

0 5 10 15 20 25

PGA = 0.82g

Bas

eshea

r/W

eight

of

th B

ridge

(W)

Transverse

Time (sec)

-0.8

-0.4

0.0

0.4

0.8

0 5 10 15 20 25

Time (sec)

c-TVAs

o-TVAs

NC

c-TVAs

o-TVAs

n-TVAs-0.8

-0.4

0.0

0.4

0.8

0 5 10 15 20 25

Time (sec)

-0.06

-0.03

0.00

0.03

0.06

0 5 10 15 20 25

PGA = 0.35g

Dis

pla

cem

ent

at M

id-S

pan

(m

)

Longitudinal

Time (sec)

-0.06

-0.03

0.00

0.03

0.06

0 5 10 15 20 25

Time (sec)

Imperial Valley, 1940

-0.06

-0.03

0.00

0.03

0.06

0 5 10 15 20 25

Time (sec)

-0.04

-0.02

0.00

0.02

0.04

0 5 10 15 20 25

c-TVAs

o-TVAs

PGA = 0.21g

Dis

pla

cem

ent

at M

id-S

pan

(m

)

Transverse

Time (sec)

-0.04

-0.02

0.00

0.02

0.04

0 5 10 15 20 25

NC

c-TVAs

Time (sec)

-0.04

-0.02

0.00

0.02

0.04

0 5 10 15 20 25

o-TVAs

n-TVAs

Time (sec)

Page 21: Seismic vibration control of bridges with nonlinear tuned

20

Figure 10. Same as in Figure 9 but for the 1989 Loma Prieta Earthquake ground motion.

Figure 11. Same as in Figure 9 but for the 1995 Kobe Earthquake ground motion.

The maximum displacement response reduction due to o-TVAs was in order of 18 % for

longitudinal direction and 33% for transverse direction. The maximum response reduction

by n-TVAs is 33% along the longitudinal and transverse directions. The n-TVAs, in general,

are at least as effective as the o-TVAs, with an exception of transverse response caused by

the Imperial Valley Earthquake ground motion. It is interesting to note that, when subjected

to the Kobe Earthquake ground motion, which produces the largest response of the three

-0.12

-0.06

0.00

0.06

0.12

0 5 10 15 20 25

PGA = 0.57g

Dis

pla

cem

ent

at M

id-S

pan

(m

)

Longitudinal

Time (sec)

-0.12

-0.06

0.00

0.06

0.12

0 5 10 15 20 25

Time (sec)

Loma Prieta, 1989-0.12

-0.06

0.00

0.06

0.12

0 5 10 15 20 25

Time (sec)

-0.06

-0.03

0.00

0.03

0.06

0 5 10 15 20 25

c-TVAs

o-TVAs

PGA = 0.61g

Dis

pla

cem

ent

at M

id-S

pan

(m

)

Transverse

Time (sec)

-0.06

-0.03

0.00

0.03

0.06

0 5 10 15 20 25

NC

c-TVAs

Time (sec)

-0.06

-0.03

0.00

0.03

0.06

0 5 10 15 20 25

o-TVAs

n-TVAs

Time (sec)

-0.12

-0.06

0.00

0.06

0.12

0 5 10 15 20 25

PGA = 0.86 g

Dis

pla

cem

ent

at M

id-S

pan

(m

)

Longitudinal

Time (sec)

-0.12

-0.06

0.00

0.06

0.12

0 5 10 15 20 25

Time (sec)

Kobe, 1995-0.12

-0.06

0.00

0.06

0.12

0 5 10 15 20 25

Time (sec)

-0.08

-0.04

0.00

0.04

0.08

0 5 10 15 20 25

c-TVAs

o-TVAs

PGA = 0.82g

Dis

pla

cem

ent

at M

id-S

pan

(m

)

Transverse

Time (sec)

-0.08

-0.04

0.00

0.04

0.08

0 5 10 15 20 25

NC

c-TVAs

Time (sec)

-0.08

-0.04

0.00

0.04

0.08

0 5 10 15 20 25

o-TVAs

n-TVAs

Time (sec)

Page 22: Seismic vibration control of bridges with nonlinear tuned

21

ground motions, the n-TVAs are considerably more efficient than the o-TVAs. It is also

noteworthy that the n-TVAs are more efficient than the o-TVAs for the ground motions for

which they were designed.

Table 1. Peak Responses for different control schemes.

Ground

Motion

Control

scheme

Base

shear/Weight of

the Bridge

Mid-Span

Displacement

(m)

Stroke (m)

X Y X Y X Y

Imper

ial

Val

ley,

1940 NC 0.848 0.386 0.068 0.046 - -

c-TVAs 0.753 0.318 0.060 0.036 0.133 0.079

o-TVAs 0.705 0.297 0.056 0.031 0.107 0.065

n-TVAs 0.641 0.298 0.051 0.032 0.118 0.076

Lom

a

Pri

eta,

1989 NC 1.722 0.741 0.136 0.074 - -

c-TVAs 1.493 0.585 0.118 0.062 0.289 0.133

o-TVAs 1.372 0.525 0.108 0.056 0.277 0.109

n-TVAs 1.156 0.490 0.091 0.049 0.245 0.117

Kobe,

1995 NC 1.534 0.900 0.122 0.083 - -

c-TVAs 1.336 0.908 0.107 0.088 0.265 0.159

o-TVAs 1.252 0.826 0.100 0.079 0.209 0.135

n-TVAs 1.019 0.788 0.081 0.077 0.238 0.128

From practical considerations, it might be desirable to control the stroke of the device. The

results indicate that the n-TVAs result in similar stroke as o-TVAs. The stroke in n-TVAs

might be large than that in o-TVAs if the excitation is not strong enough to cause inelastic

deformation because the n-TVAs are not equipped with viscous dashpots. A solution to

control the stroke of the n-TVAs is to reduce the value of R . It is to be noted that the its

value used in this study assumes that the energy dissipated by o-TVA is the same as that

done by n-TVA. If more energy dissipation is desired, a lower R value can be used. A few

values of R in the range of 0.25 to 0.75 were investigated and it was found that designing

with lower R values is very efficient in controlling the stroke. However, the efficiency of

the device in controlling structural response reduced with decreased value of R , in some

cases, its performance being worse than that of o-TVA. We then investigated other solutions

to control stroke, and the most promising one seems to be an addition of viscous dashpot to

the n-TVAs. Dashpots like the ones used with o-TVAs were added to the n-TVAs. The

results obtained with o-TVA, n-TVA, and nd-TVA denoting n-TVA with added dashpot are

shown in Figure 12

The results show that that the dashpots are effective in controlling the stroke. However, with

the inclusion of the dashpots, efficiency of the device in controlling structural response

decreases slightly, but its performance is better than that of o-TVAs.

Page 23: Seismic vibration control of bridges with nonlinear tuned

22

Figure 12. Comparison of the performance of n-TVA with/without dashpot to that of o-

TVA. The top panel shows mid-span displacement in the longitudinal and transverse

directions, and the bottom panel shows corresponding stroke of the devices.

4. Evaluation of effectives is South Iceland

As a preliminary evaluation of the effectiveness of the proposed devices in hazard scenarios

corresponding to the South Iceland lowland. For this purpose, 5 accelerograms recorded

during the 17 June 2000 earthquakes are considered. These accelerograms were recorded at

Flagbjarnaholt, Hella, Kaldarholt, Selsund, and Thorsarbru. In addition, 9 artificial

accelerograms whose 5% damped response spectra are compatible to the average response

spectra of the recorded accelerograms are simulated.

Page 24: Seismic vibration control of bridges with nonlinear tuned

23

Figure 13. Elastic response spectra (5% damped) of the recorded and simulated ground motions used

in this study. The right and left columns represent seismic action in the transverse and the

longitudinal directions of the bridge, respectively.

4.1. Design of the n-TVAs

To design the nTVS, 8 out of these 14 ground motions are considered, and the rest of the

ground motions are used to check the robustness of the design methodology. In iterative

methodology is used for each ground motion until the stroke of the TMD device converges

as explained in previous sections. The resulting strokes and response reduction factors

corresponding to the 8 ground motion used in the design process are presented in Table 2.

It can be observed that the stroke of the device caused by these ground motions are very

variable. This is due to the variability of the ground motions used here. The fundamental

period of vibration of the bridge being studied is around 0.6s, at which the spectral

displacement (see Figure 13) varies between about 2 o 10 cm. Therefore, although the

ground motion has similar spectral content on the average, the record to record variability

is quite high, as is expected close to earthquake sources. It is also apparent from Table 2

that the response reduction is higher when the stroke of the device is higher. This

corresponds to larger energy dissipation and therefore more effective vibration control.

Small values of peak stroke might indicate lack of yielding in the n-TVAs and therefore

lesser energy dissipation. In general, when the demand on the structure is high (peak stroke

0.0 0.5 1.0 1.5 2.0

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0S

pec

tral

Acc

elra

tion (

g)

Period (sec)

Ground Motion

Average

0.0 0.5 1.0 1.5 2.0

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

Spec

tral

Acc

elra

tion (

g)

Period (sec)

Ground Motion

Average

0.0 0.5 1.0 1.5 2.0

0.0

0.2

0.4

Spec

tral

Dis

pla

cem

ent

(m)

Period (sec)

Ground Motion

Average

0.0 0.5 1.0 1.5 2.0

0.0

0.2

0.4

Sp

ectr

al D

isp

lace

men

t (m

)

Period (sec)

Ground Motion

Average

Page 25: Seismic vibration control of bridges with nonlinear tuned

24

is higher), the reduction in base shear and mid-span displacement is up to 25%. However,

the reduction is not significant when the demand on the structure is low.

Table 2. Response reduction and stoke of the n-TVAs corresponding to the 8 design ground motions.

Ground

Motion

Control

scheme

Base shear

Reduction (%)

Mid-Span

Displacement (%) Stroke (m)

X Y X Y X Y

Sim

ula

ted

1

n-TVAs

25.13 12.75 24.28 23.28 0.215 0.218

2 18.31 8.78 18.41 8.66 0.184 0.199

3 10.02 12.17 10.45 12.79 0.200 0.244

South Iceland

2000 (station

Flagbjarnarholt)

6.81 15.67 5.00 16.47 0.042 0.179

4 20.57 5.48 19.23 4.66 0.160 0.118

5 10.17 15.89 2.14 9.32 0.210 0.104

6 12.55 15.45 11.84 19.09 0.210 0.102

7 20.30 19.79 20.23 10.66 0.189 0.142

4.2. Robustness of the n-TVAs

The n-TVAs and their effectiveness depend on the excitation. For practical purposes, a

suitable design needs to be selected, which needs to be robust against variability of ground

motion from the use used in designing the device. To test the robustness of vibration control

devices, the properties of the n-TVAs are taken as the average of the design properties

corresponding to the 8 design ground motions listed in Table 2. The effectiveness of this

design is then compared to other TMD schemes for the 14 ground motions. Reduction in

base shear by different control schemes when subjected to these 14 ground motions are

shown in Table 3. The results show that the n-TVAs are marginally better than the other

schemes. There is a large variability in the effectiveness against different ground motions

in all schemes. In general, the performance of the n-TVAs is better when the ground motion

is very demanding on the uncontrolled structure.

Table 3. Base-shear reduction due to n-TVAs compared to c-TVAs and o-TVAs

Ground Motion c-TVAs o-TVAs n-TVAs

X Y X Y X Y

1 26.67 11.47 24.45 10.17 25.19 12.68

2 15.50 9.28 12.50 8.28 18.19 8.89

3 11.82 12.99 10.50 11.49 10.34 12.68

South Iceland

2000 (station

Flagbjarnarholt) 2.08 11.26 1.79 9.31 5.04 10.97

4 4.60 14.84 5.47 14.29 13.02 15.01

5 1.13 16.71 1.35 15.88 1.92 16.83

6 10.97 13.53 10.89 13.19 12.66 18.87

Page 26: Seismic vibration control of bridges with nonlinear tuned

25

7 20.09 21.34 19.50 16.54 20.39 21.34

8 18.11 32.09 16.75 27.94 19.84 32.08

9 16.63 -0.25 15.69 0.87 16.21 3.76

South Iceland

2000 (station

Hella)

-0.05 23.31 0.32 21.80 0.21 22.62

South Iceland

2000 (station

Kaldarholt)

-1.04 22.62 0.11 20.68 0.87 26.14

South Iceland

2000 (station

Selsund)

16.88 3.56 16.17 2.40 14.19 7.26

South Iceland

2000 (station

Thjorsarbru)

26.76 26.26 25.37 22.00 30.31 30.06

Some examples of time history response of the bridge with different control schemes are

presented in Error! Reference source not found..

Figure 14. Base shear demand on the structure (top: longitudinal and bottom: transverse) of the

uncontrolled and controlled structure subjected to simulate ground motion number 2.

Page 27: Seismic vibration control of bridges with nonlinear tuned

26

Figure 15. Same as in Figure 13, but for ground motion recorded at Hella station.

Figure 16. Same as in Figure 14 but for Kaldarholt station.

Page 28: Seismic vibration control of bridges with nonlinear tuned

27

Figure 17. Same as in Figure 14 but for Selsund station.

Figure 18. Same as in Figure 14 but for Thorsarbru station.

Page 29: Seismic vibration control of bridges with nonlinear tuned

28

Figure 19. Displacement response of the controlled and uncontrolled structure when subjected to

ground motion recorded at the Flagbjarnarholt station; the top and bottom rows correspond to

longitudinal and transverse displacements, respectively, at the mid-span of the bridge.

Figure 20. Same as in Figure 19 but for Hella station.

Page 30: Seismic vibration control of bridges with nonlinear tuned

29

Figure 21. Same as in Figure 19 but for Kaldarholt station.

5. Conclusions

The study presents the effectiveness of using inelastic tuned vibration absorbers (n-TVAs)

in seismic response mitigation of reinforced concrete (RC) bridges. We present a

methodology to design the parameters of nonlinear springs used in n-TVAs. The method

assumes that the n-TVAs without additional viscous damping dissipate the same energy as

the energy dissipated by viscous damping in o-TVAs. The design is an iterative process

starting with an initial guess of peak displacement, which is revised in consecutive steps

until convergence is achieved. In this study, the initial peak displacement was based on the

peak displacement of the o-TVAs. In practical applications, the initial peak displacement

can be estimated from elastic analysis using design ground motion.

Optimization of n-TVAs is not straightforward because the energy dissipated by hysteresis

of the device depends on excitation. This implies that the device cannot be optimized for all

potential ground motions. A device that is designed based on a given ground motion might

not yield when subjected to a weaker ground motion, and therefore hysteretic energy

dissipation is not achieved. In such situations, the performance of the n-TVAs might be

lower than that of o-TVAs, which dissipate energy due to viscous damping. This might

seem, at first look, as a limitation of n-TVAs. However, the efficiency of the device in

reducing structural response is more crucial at stronger ground motions which have the

potential to damage the structure or cause functional loss. The results of the case study

presented here show that the proposed n-TVAs are more efficient in controlling structural

displacements and base shear when subjected to design ground motion. In summary, the n-

Page 31: Seismic vibration control of bridges with nonlinear tuned

30

TVAs are at least as effective as the o-TVAs, and are preferable because they do not require

expensive viscous dashpots like the o-TVAs.

In practical applications, unlimited stroke of TVAs cannot be accommodated by the host

structure. Therefore, reducing device stroke can become a design problem. It was observed,

from the results of simulations presented here, that n-TVAs might increase device stroke

slightly, especially when the excitation is not strong enough to cause hysteretic energy

dissipation in the device. Bagheri and Rahmani-Dabbagh (2018) present the performance

of inelastic TMD and conclude that they perform better than optimal TMD. They also point

out that the performance of inelastic TMDs may be further improved by introducing

additional viscous damping into the device. Our results indicate that addition of viscous

dashpot to the n-TVA can result in reduced performance in controlling structural response.

However, this additional damping was found to be effective in controlling device stroke. It

seems then that there is a trade-off between controlling structural response and device

stroke, which need to be simultaneously optimized in practical design applications based on

allowable limits of these parameters. It is also to be noted that the n-TVAs used here were

optimized assuming that there is no viscous damping in the device. Optimization of n-TVAs

with additional viscous damping is an important issue to be further investigated.

The foregoing arguments support the use of n-TVAs for loading scenarios controlling the

design of the structure, for example, life safety and ultimate limit states when it comes to

seismic design. The modern design philosophy however leans towards performance-based

design, where different performance levels corresponding to different levels of excitation

are defined. In such a philosophy, allowable limits on response parameters are defined for

different intensities of ground motion. It appears then that a n-TVA designed for load

corresponding to ultimate limit state might not be effective in controlling the response at

lower performance levels, such as serviceability and immediate occupancy. This is because

at weaker excitations, n-TVAs behave as o-TVAs but without viscous dashpots. To account

for these scenarios, and if there is a need to control structural response at lower performance

levels, it is advisable to provide additional dashpots with the n-TVAs. In this way, the

devices can be at least as effective as o-TVAs for low intensity ground shaking. This can

result in slight performance reduction during strong shaking, but also helps to control device

stroke during very strong ground shaking. Future studies need to focus on optimization of

n-TVA parameters with additional dashpot, and its robustness against de-tuning effects

caused by shifts in natural frequencies of the structure.

Examination of the effectiveness of the devices in a seismic hazard scenario corresponding

to the South Iceland lowland shows that the effectiveness of the control devices depends, to

a large degree, on the ground motion being used. This applies not only to the n-TVAs, which

are inherently dependent on excitation, but also to other control schemes, which although

are independent of excitation, their effectiveness depends on excitation. This variability

seems to be very large and is a direct consequence of the variability of ground motion. Such

variability, as has been observed in recent earthquakes, is very high close to the earthquake

faults. Nevertheless, n-TVAs designed in the average sense were found to be robust against

Page 32: Seismic vibration control of bridges with nonlinear tuned

31

different ground motions, and although their performance was variable, they did not result

in unwanted effects such as response amplification. It is theoretically possible to increase

the effectiveness of the n-TVAs by allowing for more inelastic deformation, i.e., by

increasing the ductility capacity of the device. This would imply higher energy dissipation

even during relatively milder shaking. This approach might be problematic in a highly active

seismic zone because repeated yielding of the device during frequent small earthquakes

might warrant for frequent maintenance and/or replacement of the devices. Then there are

other considerations such as stroke of the device which can be practically accommodated

by the host structure. In this sense, future studies in vibration control of structures with n-

TVAs should focus on a performance-based approach where different design constraints

and hazard scenarios are pre-defined and practical constraints are properly addressed.

ACKNOWLEDGEMENTS

We acknowledge financial support from the Vegagerðin research grant which was used to

finance part of the involvement of the second author in this study

DISCLAIMER

The authors of the present report are responsible for its contents. The report and its findings

should not be regarded as to reflect the Icelandic Road Authority’s guidelines or policy, nor

that of the respective author’s institutions.

Page 33: Seismic vibration control of bridges with nonlinear tuned

BIBLIOGRAPHY

AfifChaouch K, Tiliouine B, Hammoutene M, Sigbjörnsson R, Rupakhety R (2016).

Estimating ground motion incoherence through finite source simulation: a case study of the

1980 El-Asnam Earthquake. Bulletin of Earthquake Engineering, 14(4), 1195-1217.

Agrawal A, Tan P, Nagarajaiah S, Zhang J (2009) Benchmark structural control problem

for a seismically excited highway bridge Part I: Phase I problem definition. Structural

Control and Health Monitoring, 16(5), 509-529.

Aldemir U, Yanik A, Bakioglu M (2012) Control of structural response under earthquake

excitation. Computer‐Aided Civil and Infrastructure Engineering, 27(8), 620-638.

Andersson A, O'Connor A, Karoumi R (2015) Passive and adaptive damping systems

for vibration mitigation and increased fatigue service life of a tied arch railway bridge.

Computer‐Aided Civil and Infrastructure Engineering, 30(9), 748-757.

Attary N, Symans M, Nagarajaiah S, Reinhorn AM, Constantinou MC, Sarlis AA, Taylor

DP (2015) Experimental shake table testing of an adaptive passive negative stiffness device

within a highway bridge model. Earthquake Spectra, 31(4), 2163-2194.

Bagheri S, Rahmani-Dabbagh V (2018) Seismic response control with inelastic tuned

mass dampers. Engineering Structures, 172, 712-722.

Bortoluzzi D, Casciati S, Elia L, Faravelli L (2015) Design of a TMD solution to mitigate

wind-induced local vibrations in an existing timber footbridge. Smart Structures and

Systems, 16(3), 459-478.

Cao L, Li C (2019) Tuned tandem mass dampers‐inerters with broadband high

effectiveness for structures under white noise base excitations. Structural Control and

Health Monitoring, 26(4), e2319.

Carpineto N, Lacarbonara W, Vestroni F (2014) Hysteretic tuned mass dampers for

structural vibration mitigation. Journal of Sound and Vibration, 333(5), 1302-1318.

Chang K, Li C (2019) Performance of tuned tandem mass dampers based on shape

memory alloy. In IOP Conference Series: Earth and Environmental Science (Vol. 304, No.

5, p. 052014). IOP Publishing.

Casalotti A, Arena A, and Lacarbonara W (2014) Mitigation of post-flutter oscillations

in suspension bridges by hysteretic tuned mass dampers. Engineering Structures, 69, 62-71.

Casciati F, Giuliano F (2009) Performance of multi-TMD in the towers of suspension

bridges. Journal of Vibration and Control, 15(6), 821-847.

Page 34: Seismic vibration control of bridges with nonlinear tuned

33

Chen SR, Cai CS (2004) Coupled vibration control with tuned mass damper for long-

span bridges. Journal of Sound and Vibration, 278(1-2), 449-459.

Chen SR, Wu J (2008) Performance enhancement of bridge infrastructure systems: long-

span bridge, moving trucks and wind with tuned mass dampers. Engineering Structures,

30(11), 3316-3324.

Chen X, Kareem A (2003) Efficacy of tuned mass dampers for bridge flutter control.

Journal of Structural Engineering, American Society of Civil Engineers (ASCE), 129(10),

1290-1300.

Chen YH, Chen DS (2004) Timoshenko beam with tuned mass dampers to moving

Loads. Journal of Bridge Engineering, American Society of Civil Engineers (ASCE), 9(2),

167-177.

Chen Y-H, Huang Y-H (2004) Timoshenko beam with tuned mass dampers and its

design curves. Journal of Sound and Vibration, 278(4-5), 873-888.

Dallard P, Fitzpatrick AJ, Flint A, Le Bourva S, Low A, Ridsdill Smith RM, Willford,

M (2001b) The London Millennium footbridge. The Structural Engineer, 79(22), 17-33.

Dallard P, Fitzpatrick T, Flint A, Low A, Smith R, Willford M, Roche M (2001a).

London Millennium Bridge: pedestrian-induced lateral vibration. Journal of Bridge

Engineering, American Society of Civil Engineers (ASCE), 6(412), 412-417.

Poovarodom N, Kanchanosot S, Warnitchai P (2001) Control of man-induced vibrations

on a pedestrian bridge by nonlinear multiple tuned mass dampers. The Eighth East Asia-

Pacific Conference on Structural Engineering and Construction, Paper No.1344, Singapore,

December, 5-7.

Daniel Y, Lavan O, Levy R (2012) Multiple tuned mass dampers for multimodal control

of pedestrian bridges. Journal of Structural Engineering, American Society of Civil

Engineers (ASCE), 138(9), 1173-1178.

Elias S (2018) Seismic energy assessment of buildings with tuned vibration absorbers.

Shock and Vibration, 2018.

Elias S, Matsagar V (2017) Effectiveness of tuned mass dampers in seismic response

control of isolated bridges including soil-structure interaction. Latin American Journal of

Solids and Structures, 14(13), 2324-2341.

Elias S, Matsagar V (2019) Passive-hybrid system of base-isolated bridge with tuned

mass absorbers. In International Conference on Earthquake Engineering and Structural

Dynamics, 95-109.

Page 35: Seismic vibration control of bridges with nonlinear tuned

34

Elias S, Matsagar V, Datta TK (2016) Effectiveness of distributed tuned mass dampers

for multi-mode control of chimney under earthquakes. Engineering Structures, 124, 1-16.

Elias S, Matsagar V, Datta TK (2018) Along‐wind response control of chimneys with

distributed multiple tuned mass dampers. Structural Control and Health Monitoring, e2275.

Gu M, Chen SR, Chang CC (2001) Parametric study on multiple tuned mass dampers for

buffeting control of Yangpu Bridge. Journal of Wind Engineering and Industrial

Aerodynamics, 89(11-12), 987-1000.

Hijmissen JW, van Horssen WT (2007) On aspect of damping for a vertical beam with a

tuned mass damper at the top. Nonlinear Dynamics, 50(1), 169-190.

Hijmissen JW, van den Heuvel NW, van Horssen WT (2009) On the effect of the bending

stiffness on the damping properties of a tensioned cable with an attached tuned-mass-

damper. Engineering Structures, 31(5), 1276-1285.

Hoang N, Fujino Y, Warnitchai P (2008) Optimal tuned mass damper for seismic

applications and practical design formulas. Engineering Structures, 30(3) 707-715.

Jangid RS (2004) Seismic response of isolated bridges. Journal of Bridge Engineering,

American Society of Civil Engineers (ASCE), 9(2), 156-166.

Kwon HC, Kim MC, Lee IW (1998) Vibration control of bridges under moving loads.

Computers and Structures, 66(4), 473-480.

Kwon SD, Park KS (2004) Suppression of bridge flutter using tuned mass dampers based

on robust performance design. Journal of Wind Engineering and Industrial Aerodynamics,

92(11), 919-934.

Li C (2000) Performance of multiple tuned mass dampers for attenuating undesirable

oscillations of structures under the ground acceleration. Earthquake Engineering and

Structural Dynamics, 29(9), 1405‐1421.

Li C (2002) Optimum multiple tuned mass dampers for structures under the ground

acceleration based on DDMF and ADMF. Earthquake Engineering and Structural

Dynamics, 31(4), 897‐919.

Li C, Liu Y (2003) Optimum multiple tuned mass dampers for structures under ground

acceleration based on the uniform distribution of system parameters. Earthquake

Engineering and Structural Dynamics, 32(5), 671‐690.

Li C, Cao L (2019) Active tuned tandem mass dampers for seismic structures.

Earthquakes and Structures, 17(2), 143-162.

Page 36: Seismic vibration control of bridges with nonlinear tuned

35

Li JZ, Su MB, Fan LC (2005) Vibration control of railway bridges under high speed

trains using multiple tuned mass dampers. Journal of Bridge Engineering, American Society

of Civil Engineers (ASCE),10(3), 312-320.

Li Q, Fan J, Nie J, Li Q, Chen Y (2010) Crowd-induced random vibration of footbridge

and vibration control using multiple tuned mass dampers. Journal of Sound and Vibration,

329(19), 4068-4092.

Lievens K, Lombaert G, De-Roeck G, Van-den-Broeck P (2016) Robust design of a

TMD for the vibration serviceability of a footbridge”. Engineering Structures, 123, 408-

418.

Lin CC, Chen CL, Wang JF (2010) Vibration control of structures with initially

accelerated passive tuned mass dampers under near‐fault earthquake excitation. Computer‐

Aided Civil and Infrastructure Engineering, 25(1), 69-75.

Lin CC, Wang JF, Chen BL (2005) Train induced vibration control of high speed railway

bridges equipped with multiple tuned mass dampers. Journal of Bridge Engineering,

American Society of Civil Engineers (ASCE), 10(4), 398-414.

Lin Y-Y, Cheng C-M, Lee C-H (2000a) A tuned mass damper for suppressing the

coupled flexural and torsional buffeting response of long-span bridges. Engineering

Structures, 22(9), 1195-1204.

Lin Y-Y, Cheng C-Y, Sun D (2000b) Wind-induced vibration control of long-span

bridges by multiple tuned mass dampers. Tamkang Journal of Science and Engineering,

3(1), 1-13.

Lu X, Zhang Q, Wu W, Shan J (2018) Data‐driven two‐level performance evaluation of

eddy‐current tuned mass damper for building structures using shaking table and field

testing. Computer‐Aided Civil and Infrastructure Engineering.

Lu Z, Wang Z, Zhou Y, Lu X (2018) Nonlinear dissipative devices in structural vibration

control: A review. Journal of Sound and Vibration, 423, 18-49.

Lu M, Zabel V, and Könke C (2012) An optimization method of multi-resonant response

of high-speed train bridges using TMDs. Finite Elements in Analysis and Design, 53, 13-

23.

Madhekar SN, Jangid RS (2010) Seismic performance of benchmark highway bridge

with variable friction pendulum system. Advances in Structural Engineering, 13(4), 561-

589.

Page 37: Seismic vibration control of bridges with nonlinear tuned

36

Matin A, Elias S, Matsagar V (2014) Seismic control of continuous span concrete bridges

with multiple tuned mass dampers. Proceeding of the Second European Conference on

Earthquake Engineering and Seismology (2ECEES), Istanbul, Turkey.

Matin A, Elias S, Matsagar V (2017) Seismic response control of reinforced concrete

bridges with soil-structure interaction. The Bridge and Structural Engineer, ING/IABSE,

47(1), 46-53.

Miguel, LFF, Lopez RH, Torii AJ, Miguel LFF, and Beck AT (2016b) Robust design

optimization of TMDs in vehicle-bridge coupled vibration problems. Engineering

Structures, 126, 703-711.

Moreno DR, Dos-Santos RC (1997) Modeling of railway bridge-vehicle interaction on

high-speed tracks. Computer and Structures, 63(3), 511-521.

Nagarajaiah S, Narasimhan S, Agrawal A, Tan P (2009) Benchmark structural control

problem for a seismically excited highway bridge Part III: Phase II sample controller for the

fully base-isolated case. Structural Control and Health Monitoring, 16(5), 549-563.

Nagarajaiah S, Reinhorn AM, Constantinou MC (1991) Nonlinear dynamic analysis of

3D-base isolated structures. Journal of Structural, Engineering American Society of Civil

Engineers (ASCE), 117(7), 2035-2054.

Pisal AY, Jangid RS (2016) Vibration control of bridge subjected to multi-axle vehicle

using multiple tuned mass friction dampers. International Journal of Advanced Structural

Engineering, 1-15.

Poovarodom N, Kanchanosot S, Warnitchai P (2003) Application of nonlinear multiple

tuned mass dampers to suppress man-induced vibrations of a pedestrian bridge. Earthquake

Engineering and Structural Dynamics, 32(7), 1117-1131.

Poovarodom N, Mekanannapha C, Nawakijphaitoon S (2002) Vibration problem

identification of steel pedestrian bridges and control measures. The Third World Conference

on Structural Control, Como, Italy, April 7-12.

Pourzeynali S, Datta TK (2002) Control of flutter of suspension bridge deck using TMD.

Wind and Structures, 5(5), 407-422.

Rupakhety R, Sigbjörnsson R (2012) Spatial variability of strong ground motion: novel

system-based technique applying parametric time series modeling. Bulletin of Earthquake

Engineering, 10(4), 1193-1204.

Sadek F, Mohraz B, Taylor AW, Chung RM (1997) A Method of estimating the

parameters of tuned mass dampers for seismic applications. Earthquake Engineering and

Structural Dynamics, 26(6), 617-635.

Page 38: Seismic vibration control of bridges with nonlinear tuned

37

Sahasrabudhe SS, Nagarajaiah S (2005) Semi-active control of sliding isolated bridges

using MR dampers: An experimental and numerical study. Earthquake Engineering and

Structural Dynamics, 34(8), 965-983.

Tubino F, Carassale L, Piccardo G (2016) Human-induced vibrations on two lively

footbridges in Milan. Journal of Bridge Engineering, American Society of Civil Engineers

(ASCE), 21(8), Article Number C4015002.

Ubertini F (2010) Prevention of suspension bridge flutter using multiple tuned mass

dampers. Wind and Structures, 13(3), 235-256.

Ubertini F, Comanducci G, Laflamme S (2015) A parametric study on reliability-based

tuned-mass damper design against bridge flutter. Journal of Vibration and Control,

1077546315595304.

van Nimmen K, Verbeke P, Lombaert G, De Roeck G, Peter van den Broeck PVD (2016)

Numerical and experimental evaluation of the dynamic performance of a footbridge with

tuned mass dampers. Journal of Bridge Engineering, American Society of Civil Engineers

(ASCE), C4016001-1.

Verstraelen E, Habib G, Kerschen G, Dimitriadis G. (2016) Experimental passive flutter

mitigation using a linear tuned vibrations absorber. The 34th IMAC, A Conference and

Exposition on Structural Dynamics, Volume 1: Nonlinear Dynamics, 389-403.

Vestroni F, Lacarbonara W, Carpineto N (2014) Hysteretic tuned mass damper device

(TMD) for passive control of mechanical vibrations. IT Patent No. RM2011A000434

(8/10/2011) and PCT28718 (8/9/2012), Sapienza University of Rome, Italy.

Wang JF, Lin CC, Chen BL (2003) Vibration suppression for high-speed railway bridges

using tuned mass dampers. International Journal of Solids and Structures, 40(2), 465-491.

Wen YK (1976) Method for random vibration of hysteretic system. Journal of

Engineering Mechanics Division, 102, 249-263.

Wu WJ, Cai CS (2007) Theoretical exploration of a taut cable and a TMD system.

Engineering Structures, 29(6), 962-972.

Yang F, Sedaghati R, Esmailzadeh E (2007) Passive damping control of Timoshenko

beams using optimal tuned mass dampers. 21st Canadian Congress of Applied Mechanics,

590-601.

Yang YB, Yau JD, Hsu LC (1997) Vibration of simple beams due to trains moving at

high speeds. Engineering Structures, 19(11), 936-944.

Page 39: Seismic vibration control of bridges with nonlinear tuned

38

Yau JD, Yang YB (2004a) A wideband MTMD system for reducing the dynamic

response of continuous truss bridges to moving train loads. Engineering Structures, 26(12),

1795-1807.

Yau JD, Yang YB (2004b) Vibration reduction for cable-stayed bridges traveled by high-

speed trains. Finite Elements in Analysis and Design, 40(3), 341-359.

Zerva A (2016). Spatial variation of seismic ground motions: modeling and engineering

applications. Crc Press.