You are on page 1of 18

STRUCTURAL CONTROL AND HEALTH MONITORING

Struct. Control Health Monit. 2015; 22:13251342


Published online 26 March 2015 in Wiley Online Library (wileyonlinelibrary.com). DOI: 10.1002/stc.1740

Tuned liquid column ball damper for seismic vibration control


Sourav Gur, Koushik Roy and Sudib Kumar Mishra*,
Department of Civil Engineering, Indian Institute of Technology Kanpur, Kanpur, UP 208016, India

SUMMARY
Tuned liquid column damper (TLCD) is a passive control device for vibration control of structures. Recently, the
tuned liquid column ball damper (TLCBD) has been introduced for performance enhancement over the original
TLCD. In this paper, the efciency of the TLCBD is assessed under random earthquake events. The response
of a single-degree-of-freedom system with TLCBD is evaluated by stochastic analysis and compared with the
TLCD. The study reveals signicantly improved performance of TLCBD over TLCD in terms of reducing response of the structure as well as the liquid column. The performance largely hinges on the optimum tuning ratio
and the ball-to-tube diameter ratio values, which are proposed herein. The parametric studies establish the performance robustness of the TLCBD under loading and parametric variations. Furthermore, the accuracy of the optimal stochastic responses is veried through deterministic response analysis under recorded seismic motions.
Copyright 2015 John Wiley & Sons, Ltd.
Received 22 April 2014; Revised 2 December 2014; Accepted 27 January 2015
KEY WORDS:

tuned liquid column ball damper; vibration control; earthquakes; optimization; stochastic

1. INTRODUCTION
Space constraint in the process of rapid urbanization has triggered the construction of tall, light, and
exible structures. Because of their slender and lightly damped nature, these structures are quite sensitive to dynamic loading from strong wind and earthquake events. The vibration control technologies
are widely adopted in order to reduce the dynamic load effects on such slender structures, primarily by
adopting various types of damping systems installed on them. The tuned liquid column damper
(TLCD) was proposed and tested in the past for vibration mitigating under both wind and
earthquake-induced effects on structures. The TLCD was introduced by Sakai et al. [1] and subsequently attracted attention of the vibration control community because of its appealing features, such
as ease of installation and consistent performance under a wide range of exciting frequencies.
The TLCD consists of a U-shaped tube lled with liquid (preferably water) with an orice at the
middle of the horizontal portion. During the course of motion, the liquid in the tube ows from one
arm of the tube to another to trigger actuation of the control force. The control efciency of TLCD
is adjusted by tuning the frequency of the liquid column with respect to the frequency of the main
structure [2]. In addition, the ow through orice also helps in dissipating part of input energy of vibration. Research works [3,4] have been conducted to assess the effectiveness of the TLCD and to obtain the optimal parameters in order to ensure best performance. A number of modications have been
proposed on the original TLCD conguration consequently in order to improve its performance, which
includes introduction of a spring and a viscous damper among the structure and the TLCD [5], variable
orice in the TLCD in order to adjust the damping [6] adaptively, semi-active actuator to change the
*Correspondence to: Sudib Kumar Mishra, Department of Civil Engineering, Indian Institute of Technology Kanpur, Kanpur,
UP 208016, India.

E-mail: smishra@iitk.ac.in

Copyright 2015 John Wiley & Sons, Ltd.

1326

S. GUR, K. ROY AND S. K. MISHRA

cross section of the TLCD [7] and usage of magneto-rheological uid [8]. A version of the TLCD with
varying cross section of the tube was proposed, popularly referred as the liquid column vibration absorber [9]. It was revealed [10] that increased cross section of the horizontal tube further helps in reducing the controlled response of the structure. The potential of TLCD in controlling the lateral
vibration of structure was demonstrated by Xu et al. [11] and Haroun et al. [12]. Wu et al. [13] has
applied a special conguration of TLCD in order to control the torsional vibration in the torsionally
coupled building. The usage of multiple TLCDs was tested by Gao and coworkers [14] for vibration
control of multi degree of freedom system. Recently, Shum [15] has proposed closed-form expression
for the optimal parameters for a TLCD subjected to harmonic excitations. The performance of the liquid column vibration absorber was extensively studied by Hitchcock et al.[16],Watkins and Hitchcock
[17], and Chang and Hus [18].
Very recently, a modied version of the TLCD has been introduced by Saif et al. [19] by replacing
the xed orice with a rolling steel ball in the horizontal portion of the TLCD. This is conveniently
referred as the tuned liquid column ball damper (TLCBD). The ball essentially acts as a movable orice modulating the ow in the liquid column to improve the response reduction capability of the
damper. The optimal performance of the proposed system is ensured by adopting parameters following
the method in another study conducted by Wu et al. [20] and Shum et al. [15], although with the proposed modication of the original TLCD, such choices are ad hoc. The type of excitations adopted for
demonstration is also conned to harmonic excitations with the view that the TLCD systems are primarily suitable for narrow-band excitation typically encountered in wind loading.
In view of the aforementioned studies, the present study assesses the performance of the TLCBD
under wide-band random seismic excitations. The optimal choices of parameters for the system are
attained through a systematic design optimization based on numerical search technique in the feasible
design space of the system variables. The optimal performance is also compared with the conventional
TLCD in order to establish a relative estimate of the improvement attained in TLCBD over the TLCD.
The robustness of the aforementioned features is established based on extensive numerical simulation
concerning the expected variations in the system parameters as well as scenarios of earthquake loading.
The response analysis for the process of optimization is based on random vibration analysis. The legitimacy of the optimal performance, so obtained, is further veried through deterministic time-history
response analysis under set of recorded ground motions pertaining to real earthquakes.

2. RESPONSE ANALYSIS OF STRUCTURE EQUIPPED WITH TLCBD


The structure adopted for the proposed control application is idealized as a single-degree-of-freedom
(SDOF) system, shown in Figure 1. It is demonstrated [21,22] that a damper-assisted MDOF structural
system can be accurately idealized as an SDOF system if the structure possesses well-separated modes
of vibrations. Moreover, the TLCD or TLCBD is typically designed to suppress a specic mode of

Figure 1. Simple single-degree-of-freedom idealization of structure equipped with tuned liquid column ball damper.
Copyright 2015 John Wiley & Sons, Ltd.

Struct. Control Health Monit. 2015; 22: 13251342


DOI: 10.1002/stc

TUNED LIQUID COLUMN BALL DAMPER

1327

vibration (preferably the fundamental mode) that allows the controlled structure to be adequately
modeled as SDOF system.
The SDOF system equipped with the TLCBD is shown in Figure 1 and adopted for the present analysis. The TLCBD is a U-shaped tube lled with liquid and a ball (spherical and made of steel) placed at
the middle of the horizontal portion of the tube. The ball moves in the liquid during motion. As the
control strategies substantially reduce the response of the structure, the behavior of the controlled structure is reasonably assumed to be linear.The governing equation of motion for the structure-TLCBD is
derived by using the Lagrangian formulation. The equation of motion for the ball is written as [19]



 



Jb
Jb 
2ml gR2bt
x s xg d eq x_ l
xl
mb 2 x b deq x_ b
Le
Rb
R2b

(1:1)

in which mb is the mass of the ball, ml is the mass of the liquid in the tube, Jb is the mass moment of
inertia, and Rb is the radius of the ball. The equivalent viscous damping for the ball in the liquid is denoted by deq, and the parameter Rbt is the ratio of the diameter of the ball to the diameter of the tube.
The effective length of the liquid in the column is Le, which is sum of the lengths of the horizontal (Bh)
and vertical (h) liquid column (Le = Bh + 2h). The acceleration, velocity, and displacement of the ball
are denoted by xb ; x_ b ; and xb , respectively. The quantities xg and xs denote the acceleration of the
ground and the structure. The parameter g is the gravitational acceleration. The displacement and velocity of the liquid column are given by xl and x_ l , respectively. The equation of motion of the liquid
mass in the column is written as [19]




2ml g
xl pml x s xg
(1:2)
ml xl 2ml l l x_ l
Le
where l is the head loss coefcient of the liquid in the TLCBD. This is a nonlinear function of ball-to-tube
diameter ratio (Rbt). The numerical values of l are obtained from this ratio in a table provided in the previous study [19]. The frequency of the vibration of liquid column is denoted by l. The parameter p is the
length ratio which expresses the ratio of the length of the horizontal portion to the total length of the liquid
column, that is, (p = Bh/Le). The equation of motion for the SDOF structure can be written as [19]




 
Jb
Jb
Jb
x b
ms ml 2 x s cs x_ s ks xs  ms ml 2 x g  pml xl
Rb
Rb
R2b

(1:3)



where ms is the mass, and k s ms 2s and cs(=2ms ss) are the stiffness and viscous damping of the
structure. The parameter s is the viscous damping ratio of the structure, Jb is the mass moment of inertia
of the ball, and deq is the viscous damping of the motion of the ball by the liquid in the tube. These quantities are given by [19]
J b 2mb R2b =5

(2:1)

d eq 6Rb

(2:2)

where is the kinematic viscosity of the liquid. The Equations (1.1), (1.2), (2.1), and (2.2) together are
expressed in matrix form as
Mfu g Cfu_ g Kfug Mfrgx g

(3)

in which [M], [C], and [K] are the mass, damping, and stiffness matrix for the TLCBD-structure system.
The inuence coefcient vector {r} is expressed as f 0; 0; 1 gT . The displacement vector is dened as
fug f xb ; xl ; xs gT , and the vectors of the velocities and accelerations are also dened accordingly.
An important difference among the equations of motion for the liquid in TLCD and the TLCBD is
in the damping term. The TLCBD is modeled with the assumption of linear damping, whereas the
damping in TLCD is nonlinear [19]. The linearity of damping was experimentally validated through
Copyright 2015 John Wiley & Sons, Ltd.

Struct. Control Health Monit. 2015; 22: 13251342


DOI: 10.1002/stc

1328

S. GUR, K. ROY AND S. K. MISHRA

a free vibration test conducted by Al-Saif et al. [19] under an initial pulse using electro-dynamic
shaker. In their study, the response of the oscillating water level in the tuned was monitored through
a HeNe laser. The displacement time history of the U-tube liquid level shows exponential decay in
amplitude with time, as commonly observed in case of damped free vibration response of an SDOF
system. Thus, the damping in the liquid column vibration was taken as linearly viscous, and the respective damping coefcient is determined from the logarithmic decrement technique.
The linearity of the damping is due to the resistance to the uid ow by the ball moving in the liquid. This resistance is governed by the viscosity of the uid and the inertial forces on the ball,
expressed in terms of the Reynolds number. The Reynolds number for ow in TLCDs typically ranges
from 0.7 105 to 1.05 105, which is mostly in turbulent regime [20]. The Reynolds number in
TLCBD will be in comparable range, although it is expected to be lesser than the respective TLCD,
because of relatively smoother characteristics of the ow geometry.
Further, the Reynolds number of the ow in TLCD (and so in TLCBD) does not change signicantly with the variations in the excitation frequency [20], which is typically encountered in multiple
frequency excitations (wind/earthquakes). This allows the usage of the same damping coefcient
(which is based on tests performed on a single representative excitation frequency) for multiple frequencies. The same justication can also be extended in case of TLCBD. Therefore, the results from
Al-Saif et al. [19] can be relied on for broadband seismic excitations as well.
The properties of the TLCBD-SDOF structure are conveniently normalized with respect to the system parameters as
ml =ms ; 2mb =5ms ; l =s

(4:1)

in which ; ; and are the mass ratio of liquid, mass ratio of ball, and the tuning ratio of the liquid,
respectively. The frequency of liquid mass is expressed as
l

p
2g=Le

(4:2)

These are important parameters controlling the efciency and other important response behavior of
the structure and the TLCBD itself.
The dynamic response analysis under seismic excitation requires a number of ground motions as input.
These motions could be recorded on an accelerogram or simulated ground motions compatible to design
spectrum. The wide variability in the ground motion characteristics (frequency content, amplitudes, etc.)
is taken care by selecting a suite of ground motions with widely varying characteristics. Unlike such deterministic analysis, the stochastic analysis employs a stochastic ground motion model, which is dened in
terms of its power spectral density function (PSDF). The most widely employed stochastic ground motion
model is described by the KanaiTajimi power spectral density, dened as [23,24]

h
2 

2 i  2
2
4
2
(5)
f 
2 f f
S S0 f 2 f f

in which S() is the power spectral density of the seismic acceleration at frequency , f and f are the
characteristic frequency and damping of the soil strata, and S0 represents the intensity of the rock-bed
white-noise excitations. This model is a stationary model and is obtained by ltering a white noise
process acting at the rock bed through a linear lter representing the ground. The underlying lter equations are written as
w
xf 2 f f x_ f 2f xf 

(6:1)

xg 2 f f x_ f  2f xf

(6:2)

is the white noise intensity at the rock bed with power spectral density S0, xf , x_ f , and xf are the
where w
acceleration, velocity, and displacement of the ground. The intensity of the rock-bed white-noise excitastatistically
corroborates to the root mean square (r.m.s) of the representative acceleration time
tion S0 

history 2u g as [25]

Copyright 2015 John Wiley & Sons, Ltd.

Struct. Control Health Monit. 2015; 22: 13251342


DOI: 10.1002/stc

TUNED LIQUID COLUMN BALL DAMPER



S0 2 f 2u g = 1 4 2f f

1329
(7)

This model of earthquake excitations is introduced in the governing equation of motion of the
structure-TLCBD system (Equation (3)) by substituting expression of x g from Equation (7) as


fu g  M 1 Cfu_ g  M 1 Kfug frg 2 f f x_ f 2f xf

(8)

Similarly, Equation (6.1) for the ground motion is also expressed as

x f 2 f f x_ f  2f xf  w

(9)

With the state vector {Y} is dened as fY g f fug xf fu_ g x_ f gT , the equation of motion is written in
statespace form as
d
fY g AfY g fwg
dt

(10)

where [A] is the augmented system matrix (detailed in the Appendix) and fwg f f0g 0 f0g 
w gT is a
vector of the rock bed motion. As the responses of the system is assumed to be Markovian, the evolution
equation for the covariance matrix [CYY] of the responses {Y} is governed by [26]
d
C YY  AC YY T C YY AT Sww 
dt

(11)



where the elements of the response covariance matrix [CYY] is C Y i Y j E Y i Y j . The matrix [Sww] contains
the terms of the intensity of the rock-bed white-noise excitations. The matrix [Sww] has all terms zero except
the last diagonal as 2S0. The matrix equation in Equation (11) is solved to obtain the response covariance.
The response statistics of the acceleration x s ; x l ; x b are obtained as
C Y_ Y_  AC YY AT Sww 

(12)

It is often convenient to express the r.m.s response


 of the controlled structure xs with respect to the

respective response of the uncontrolled structure uc


as
xs

xs xs = uc
xs

(13)

in which the bar () above represents the normalized response (i.e., the ratio of responses) and the superscript uc refers to the responses of the uncontrolled structure.
The optimal responses of the TLCBD-controlled structure are compared with the response of the
structure controlled with the TLCD. This requires evaluation of response of the TLCD-structure system under seismic excitations. This closely follows the procedure presented earlier for TLCBD, but
not presented herein. More details of the stochastic response analysis of structure controlled by
TLCD under random ground motion can be found elsewhere [27,28]. The present analysis assumes
that the earthquake excitation is a Gaussian stationary stochastic process and so are the responses.
However, in practice, the ground motions contain strong nonstationary characteristics in its amplitude
as well as frequency contents. The nonstationary amplitude does not affect the optimal responses and
parameters too much [29]; however, the nonstationarity in the frequency contents may have substantial effect on the response and the optimal variables, which is not discussed herein and can be extended in future work.
Copyright 2015 John Wiley & Sons, Ltd.

Struct. Control Health Monit. 2015; 22: 13251342


DOI: 10.1002/stc

1330

S. GUR, K. ROY AND S. K. MISHRA

3. OPTIMIZATION OF PERFORMANCE OF THE TLCBD


It is well-established that in a conventional TLCD with given mass ratio (), the tuning ratio () and
head loss coefcient of the orice ( l) are the two most important parameters to control the efciency
of vibration reduction [30,31]. The vibration-reduction efciency is expressed in terms of the normalized response. These parameters are chosen to maximize the vibration reduction capability by minimizing the response of the controlled structure. The relevant problem of optimization is stated as
Findf l gT to minimize xs

(14)

In case of the TLCBD, these design parameters are replaced by the tuning ratio () and the ball-totube diameter ratio (Rbt) leading to an equivalent problem of optimization stated as
Findf

Rbt gT to minimize xs

(15)

The parameter Rbt is employed to characterize the equivalent head loss coefcient ( l) in TLCBD.
The relationship among the Rbt and l is nonlinear. The value of l is obtained from a previous study for
a given Rbt [19].
It may be mentioned here that along with the control efciency, the displacement of the liquid column is also an important quantity, and its peak value must be within some permissible limit. This is
because excessive liquid displacement can cause the liquid to spill out of the tube and may impair
the effectiveness of the TLCD/TLCBD. Keeping this in view, the respective optimization problem
can be redened by incorporating the displacement of the liquid column as a constraint. The relevant
constrained optimization problem is redened as
Findf l =Rbt g to minimize xs
so that xl all
xl

(16)

where all
xl is the allowable displacement of the liquid column. Such study on TLCD has been reported
by Chakraborty et al. [30]. The present study, however, is restricted only to unconstrained optimization, and the effects of constraint can be studied in future.
The optimization referred herein ensure best performance of the controlled structure under random
ground motions, which is obtained by minimizing the stochastic response of the structure, referred as
stochastic structural optimization (SSO). It is obvious that this results in an unconstrained optimization
problem, which is solved to obtain the suitable set of design variables for the TLCBD ensuring optimal
performance. This optimization problem posed herein is solved using the routines available in the
MATLAB Optimization Toolbox (Mathworks Inc., Massachusetts, USA).

4. VERIFICATION OF OPTIMAL PERFORMANCE UNDER RECORDED GROUND MOTIONS


The consistency of the optimal response behavior of the controlled structure, as obtained through SSO
is subjected to verication by evaluating the response of the TLCBD-controlled structure with optimal
parameters as obtained through SSO. The structure so obtained is then subjected deterministic ground
accelerations recorded from real earthquakes. This is in order to assess the feasibility of the optimal behavior under real earthquakes. A set of recorded ground accelerations are selected for this purpose. The
structural responses are evaluated through nonlinear time-history analysis by integrating the equations
of motions (Equation (3)) for the structure-TLCBD system. The Newmark-beta step-by-step integration scheme with average acceleration is employed for the purpose. The adopted time steps are kept
sufciently low (t = 0.0001 s) to ensure stability and accuracy of the numerical integration
scheme. Time histories of the displacement of the structure, displacement of the liquid, and displacement of the ball (for TLCBD only) are obtained as the important response quantities of interests as
studied subsequently.
Copyright 2015 John Wiley & Sons, Ltd.

Struct. Control Health Monit. 2015; 22: 13251342


DOI: 10.1002/stc

1331

TUNED LIQUID COLUMN BALL DAMPER

5. NUMERICAL ILLUSTRATION
5.1. Stochastic response analysis and optimization
The performance of the structure equipped with TLCBD is demonstrated through numerical simulation. The optimal performance of this system is compared with the performance of structure controlled
by TLCD. Typical values of the properties of the structure, the TLCD/TLCBD, and the stochastic
ground motion model are illustrated in Table I. Whereas these are adopted as the default values, realistic ranges of variations are also considered in order to study the effects of parameter variations. The
normalized displacement of the structure, displacement of the liquid, and the displacement of the ball in
TLCBD are presented as the response quantities of interest.
The variations of responses with respect to varying tuning/frequency ratios of TLCD are shown in
Figure 2a and b and are compared with the TLCBD in Figure 2ce. Following the behavior of TLCD,
there also exist certain value of the tuning ratio minimizing the response of the structure and thus maximizing the control efciency of TLCBD as shown in Figure 2ae. Comparing the optimal responses, it
is observed that the TLCBD offers much higher control efciency than the conventional TLCD. The
increase in efciency is as high as 50%. Further, comparison of the liquid displacements (Figure 2b
and d) reveals that the liquid displacement in the TLCBD is reduced signicantly, which is around

Table I. Parameters for TLCD, TLCBD, structure and the ground motion model.
Properties of the TLCD and TLCBD
Properties of the structure
Time period = 1.5 s;
Damping ratio = 2%

Mass ratio (%)


Tuning ratio
Ball-to-tube diameter ratio
Head loss coefcient
Density of liquid (kg/m3)
Kinematic viscosity (Nm/s)
Density of ball (kg/m3)

TLCD

TLCBD

3.000
0.950

3.000
0.900
0.400

0.700
1000

Stochastic ground
motion parameters
f = 9 rad/s;
f = 0.600;
S0 = 0.050 m2/s3

1000
0.001
7500

TLCD, tuned liquid column damper; TLCBD, tuned liquid column ball damper.

Figure 2. (a) Displacement ratio of structure and (b) liquid displacement of tuned liquid column damper (TLCD)
for different tuning ratios. The variations are shown for (c) displacement ratio of structure, (d) liquid displacement,
and (e) displacement of the ball in tuned liquid column ball damper (TLCBD).
Copyright 2015 John Wiley & Sons, Ltd.

Struct. Control Health Monit. 2015; 22: 13251342


DOI: 10.1002/stc

1332

S. GUR, K. ROY AND S. K. MISHRA

60% of the respective TLCD. It is also noted that the optimal tuning ratios correspond to the maximum
displacement of the liquid. This implies that better control efciency can only be attained at the expense of larger displacement of the liquid column. Therefore, the displacement of the structure and
the resulting displacement of the liquid column result in two mutually conicting objectives in optimizing the design of TLCD, although the second criterion is often overruled. The TLCBD not only improves the control efciency but also reduces the liquid displacement substantially from that in
TLCD. However, the displacement of the ball in TLCBD becomes very high, especially for lower
values of the tuning ratio. This, however, reduces signicantly around the optimal value of the tuning
ratio.
The plots in Figure 2ae also show that the mass ratio of the liquid is an important parameter to control the efciency of both the TLCD and TLCBD systems. The control efciency generally increases
with increasing mass ratios. However, increasing liquid mass ratio also leads to enhanced displacement
of the ball in TLCBD. The existence of optimal tuning ratio can be explained by studying the variations
in the respective PSDFs of the responses under varying tuning ratios, shown in Figure 3a and b for the
TLCD and in Figure 3ce for the TLCBD. It is observed that the PSDF ordinate of the controlled structural displacement (Figure 3a and c) attains minimum for particular tuning ratio which causes certain
tuning ratio minimizing the controlled response. The same trend is also followed by the PSDFs of
the other response quantities (displacement of liquid column and the ball) as well. These are shown
in Figure 3b, d, and e.
Similar to the effect of tuning ratio, the effect of head loss coefcient of TLCD and ball-to-tubediameter ratio of TLCBD on the displacement of the controlled structure and on the displacement of
the liquid and ball in TLCBD is assessed. It is mentioned that the ball-to-tube diameter ratio of TLCBD
affects the damping provided by the liquid motion, and therefore, its role is analogous to the head loss
coefcient in the TLCD. Hence, the effects of these two quantities are expected to be similar. The response behavior of the structure, TLCD and TLCBD are shown in Figure 4ae for varying head loss
coefcient. Similar to the optimal tuning ratio, there also exists an optimal head loss coefcient value
for the TLCD to maximize the control efciency. The role of optimum head loss coefcient in TLCD is
switched to the optimal ball-to-tube diameter ratio in TLCBD (for maximizing the control efciency)
as shown in Figure 4c. Comparing the optimal responses from both the systems, it can be pointed out
that the TLCBD offers more control efciency along with simultaneous reduction of the displacement

Figure 3. Power spectral density function (PSDF) of the (a) normalized structural displacement and (b) liquid displacement for the tuned liquid column damper (TLCD) for varying tuning ratios. Similar variations are for (c)
normalized structural displacement, (d) liquid displacement, and the (e) displacement of the ball for tuned liquid
column ball damper (TLCBD).
Copyright 2015 John Wiley & Sons, Ltd.

Struct. Control Health Monit. 2015; 22: 13251342


DOI: 10.1002/stc

TUNED LIQUID COLUMN BALL DAMPER

1333

Figure 4. (a) Displacement ratio of structure and (b) liquid displacement of the tuned liquid column damper
(TLCD) under different coefcients of head loss. Similar variations in the (c) displacement ratio, (d) liquid displacement, and (e) ball displacement in tuned liquid column ball damper (TLCBD) under varying ball-to-tube
diameter ratio.

of the liquid column while compared with the conventional TLCD. The existence of the optimal head
loss coefcient (for TLCD) or ball-to-tube diameter ratio (for TLCBD) can similarly be explained by
studying the variations of the respective PSDFs, not shown herein. It is observed that the mass ratio of
the liquid signicantly affects the control efciency, and increasing mass ratio generally increases the
control efciency.
It is established from the preceding discussion that, whereas the optimal parameters for the TLCD
are the tuning ratio and head loss coefcient, the efciency of TLCBD is governed by the choice of the
tuning ratio and ball-to-tube diameter ratio. An important factor to affect these values is the mass ratio.
With the earlier discussion in view, in next, the choice of the optimal parameters and the respective
optimal responses are obtained based on systematic optimization studies. However, in order to provide
an intuitive picture of the various suboptimal combinations of the parameters and the respective responses, surface plots are employed. The variation of displacement ratio and the displacement of liquid
column are demonstrated for the TLCD under varying tuning ratios and the head loss coefcients,
shown in Figure 5a and b. For the TLCBD, similar variations, along with the displacement of the ball,
are presented in the surface plots of Figure 5ce. These plots clearly indicate to an optimal combination
of two parameters (tuning ratio and head loss coefcient/ball-to-tube diameter ratio) in order to ensure
best performance (Figure 5a and c). The optimal combination of these parameters corresponds to the
lowest point in the plot in Figure 5a and c, indicating to minimum displacement ratio. However, no optimal combination exists for minimizing the displacement of the liquid column (Figure 5b and d).
The optimal values of the design variables are presented with respect to the varying mass ratios of
the liquid in Figure 6ac. Respective responses of the optimal TLCD and TLCBD system are also
shown in Figure 6df. Two important features observed from these plots are the signicant variations
of the optimal tuning ratio of the TLCBD, which is lacking in TLCD (Figure 6a), and the insensitivity
of the optimal tuning ratio of TLCD but signicant variations of the same in TLCBD under varying
mass ratio.
The observed trends in these gures are explained. The ball acts as a moving orice to allow for
compliant tuning of the uid ow opposing the structural motion, which results in enhanced damping
in TLCBD system. Large damping shifts the tuning frequency toward the lower side, as seen in
Figure 6a. Increasing mass ratio increase the optimal head loss coefcient of liquid almost proportionately (Figure 6b) which is in line with the earlier observations [28]. However, this trend gets reversed in
Copyright 2015 John Wiley & Sons, Ltd.

Struct. Control Health Monit. 2015; 22: 13251342


DOI: 10.1002/stc

1334

S. GUR, K. ROY AND S. K. MISHRA

Figure 5. Variations of the (a) displacement ratio and (b) displacement of the liquid column for the tuned liquid
column damper (TLCD) under varying combinations of the tuning ratios and head loss coefcients. Variations
in the (c) displacement ratio, (d) liquid displacement, and (e) displacement of the ball in the tuned liquid column
ball damper (TLCBD) for varying combinations of the tuning ratio and the ratio of the ball-to-tube diameter.

Figure 6. Optimal (a) tuning ratio, (b) head loss coefcient, and (c) ball-to-tube diameter ratio and respective (d)
displacement ratio, (e) liquid displacement, and (f) ball displacement of the controlled structure for different values
of the mass ratio of the liquid. TLCD, tuned liquid column damper; TLCBD, tuned liquid column ball damper.

optimal ball-to-tube diameter ratio in TLCBD (Figure 6c), which rather decreases with increasing mass
ratio. This might be to facilitate the motion of larger amount of liquid in TLCBD.
Comparing the optimal responses, the superior control efciency of the TLCBD over the TLCD can
be easily gured out (Figure 6d) along with the signicant reduction in the liquid displacement
(Figure 6e). Expectedly, the displacement of the ball in TLCBD increases with increasing mass ratio
(Figure 6f). It is also observed in Figure 6d that the variations of the optimal response for different mass
ratios in the TLCBD is largely reduced (almost insensitive) compared with the TLCD, which implies
enhanced robustness of the TLCBD. In this context, it is important to mention that to achieve identical
Copyright 2015 John Wiley & Sons, Ltd.

Struct. Control Health Monit. 2015; 22: 13251342


DOI: 10.1002/stc

TUNED LIQUID COLUMN BALL DAMPER

1335

control efciency the TLCBD requires much lower value of mass ratio than the TLCD, which is a signicant improvement made in the TLCBD system.
In next, the performance improvement is assessed under varying system parameters as well as scenarios of earthquake loading. The optimal characteristics and performances are compared under varying exibilities of the superstructure as shown in Figure 7af. Unlike the insensitivity of the optimal
tuning ratio of TLCD to the variations of structural exibility, the optimal choice of the tuning ratio
of TLCBD strongly depends on the exibility of the structure to be controlled. It is important to note
that the TLCBD becomes increasingly efcient than the TLCD for structure with increasing exibilities. With improved efciency and reduced liquid displacement of TLCBD, a potential problem for
the TLCBD could be the large displacement of the ball, especially for controlling of highly exible
structure as depicted in the trend of Figure 7f. This might limit the applicability of such system and
must be adequately taken care while designing such system. The optimization for such system could
impose a constraint on the displacement of the ball, not discussed herein.
The effect of the length ratio is shown on the optimal parameters of TLCD and TLCBD in
Figure 8ac, along with the optimal responses in Figure 8df. The improved performance of TLCBD
over TLCD is indicated in these plots. For identical length ratio, the displacement ratio in TLCBDassisted structure is much lesser (around 33%) than the TLCD-assisted. The displacement of the liquid
column also reduces substantially. Therefore, the TLCBD will require a much lesser length of the liquid column than the TLCD. An important feature to observe in TLCBD is that the control efciencies
remain irrespective of the length ratio, whereas in TLCD, the efciencies increase with increasing
length ratio. This implies to the enhanced performance robustness of the TLCBD under varying length
ratios.
Variation of the optimal parameters for both the system are shown with respect to the varied level of
damping in the controlled structure in Figure 9ac. Respective responses are demonstrated in
Figure 9df. It is observed that the displacement of the ball reduces for lower value of structural
damping. This is an appealing feature of TLCBD in view of its potential application in lightly damped
structures.
Further, the comparative performances of the two systems are presented under varying intensities of
earthquakes, characterized by the intensities of the white noise excitations at the rock bed. The optimal
parameters are shown in Figure 10ac, and the optimal responses are in Figure 10df. It is observed
that the optimal parameters for both TLCD and TLCBD are irrespective of the seismic intensity

Figure 7. Optimal (a) tuning ratio (b) head loss coefcient (c) ball-to-tube diameter ratio and respective (d) displacement ratio, (e) liquid displacement, and (f) ball displacement of the controlled structure for different values of the
primary structure exibility. TLCD, tuned liquid column damper; TLCBD, tuned liquid column ball damper.
Copyright 2015 John Wiley & Sons, Ltd.

Struct. Control Health Monit. 2015; 22: 13251342


DOI: 10.1002/stc

1336

S. GUR, K. ROY AND S. K. MISHRA

Figure 8. Optimal (a) tuning ratio, (b) head loss coefcient, and (c) ball-to-tube diameter ratio and respective (d)
displacement ratio, (e) liquid displacement, and (f) ball displacement of the controlled structure for different values
of length ratio. TLCD, tuned liquid column damper; TLCBD, tuned liquid column ball damper.

Figure 9. Optimal (a) tuning ratio, (b) head loss coefcient, and (c) ball-to-tube diameter ratio and respective (d)
displacement ratio, (e) liquid displacement, and (f) ball displacement of the controlled structure for varied level of
damping in the primary structure. TLCD, tuned liquid column damper; TLCBD, tuned liquid column ball damper.

(Figure 10ac). This trend matches with the observed behavior for TLCD in several studies [27,28]
presented earlier. The dependency of the liquid displacement on the seismic intensity is a direct consequence of the weak nonlinearity involved in the mechanism of orice damping in the TLCD and
TLCBD systems. It is also important to note that the reduction in the liquid displacement increases under earthquakes of larger intensities, implying that the improved performance of TLCBD is increasingly availed under earthquakes of larger intensities.
Copyright 2015 John Wiley & Sons, Ltd.

Struct. Control Health Monit. 2015; 22: 13251342


DOI: 10.1002/stc

TUNED LIQUID COLUMN BALL DAMPER

1337

Figure 10. Optimal (a) tuning ratio, (b) head loss coefcient, and (c) ball-to-tube diameter ratio and respective (d)
displacement ratio, (e) liquid displacement, and (f) ball displacement of the controlled structure for different values
of the seismic intensity. TLCD, tuned liquid column damper; TLCBD, tuned liquid column ball damper.

Figure 11. Optimal (a) tuning ratio, (b) head loss coefcient, and (c) ball-to-tube diameter ratio and respective (d)
displacement ratio, (e) liquid displacement, and (f) ball displacement of the controlled structure for different values
of the dominant frequency content of the random earthquake. TLCD, tuned liquid column damper; TLCBD, tuned
liquid column ball damper.

The optimal performance and robustness of TLCBD are also veried under varying dominant
frequency content of the earthquakes. The optimal tuning ratios, respective head loss coefcient
(for TLCD) and ball-to-tube diameter ratio (in TLCBD) are shown in Figure 11ac. The respective optimal responses of the structure and control systems are shown in Figure 11df. In these plots, the dominant frequency of the motion is conveniently normalized with respect to the frequency of the structure.
Except for low frequencies (close to the resonating frequency of the SDOF structure), the optimal parameters for both TLCD and TLCBD are observed to be insensitive to the frequency content of the
Copyright 2015 John Wiley & Sons, Ltd.

Struct. Control Health Monit. 2015; 22: 13251342


DOI: 10.1002/stc

1338

S. GUR, K. ROY AND S. K. MISHRA

ground motions and so are the responses. Nonetheless, the TLCBD substantially reduces the response
along the considered frequency range than the TLCD-controlled structure.
5.2. Verication of optimal performance under recorded earthquakes
The performance of the SDOF system equipped with a TLCBD with optimal parameters obtained from
the stochastic optimization is veried by subjecting the structure-TLCBD system under a set of recorded earthquake ground motions. The set of recorded motions are listed in Table II. These are
employed to evaluate the transient response of such system. The table shows the salient characteristics,
such as the peak ground acceleration, and dominant periods (frequency content) that primarily governs
the dynamic response behavior. The variations of these quantities are kept in mind while choosing this
set of motions. These motions also belong to quite different geological and fault conditions. However,
this set is not comprehensive by any means, but they can be employed to check for certain degree of
robustness of the performance of the structure-TLCBD system. For more exhaustive study, design
ground motions compatible to specied spectrum may be used.
Typical time history responses of the SDOF structure controlled by optimal TLCBD are shown under the 1979 Imperial Valley earthquake (GM-3) motion. The time history of displacement of the structure (Figure 12a), the liquid displacements for both the TLCD and TLCBD (Figure 12b), and the
displacement of the ball in TLCBD (Figure 12c) are shown. Figure 12ac shows the largely reduced
Table II. Set of ground motions used for response evaluation.
Serials

Earthquake

Year

Station

PGA (g)

Dominant period (s)

GM-1
GM-2
GM-3
GM-4
GM-5

Superstition Hill
Kobe
Imperial Valley
Landers
San Fernando

11/24/1987
01/16/1995
10/15/1979
06/28/1992
02/09/1971

5210 Wildlife Liquef. Array


Kakogawa
6605 Delta
23 Coolwater
128 Lake Hugas # 12

0.408
0.345
0.351
0.417
0.366

0.10
0.17
0.21
0.23
0.18

PGA, peak ground acceleration;

Figure 12. Time histories of displacements of the (a) structure, (b) liquid in tuned liquid column ball damper
(TLCBD) and tuned liquid column damper (TLCD)-controlled systems, and (c) ball in TLCBD.
Copyright 2015 John Wiley & Sons, Ltd.

Struct. Control Health Monit. 2015; 22: 13251342


DOI: 10.1002/stc

TUNED LIQUID COLUMN BALL DAMPER

1339

displacement of the structure assisted by the TLCBD than the TLCD. The liquid displacement in
TLCBD is also shown to be largely reduced than the TLCD. These trends are in parity with the observed behavior of stochastic responses and therefore justify the viability of the optimal performance
of TLCBD obtained through SSO.
The variations of optimal responses under varying system parameters are subsequently presented for
comparison with the observed trends from stochastic analysis. The ground motion time histories, listed
in Table II, are employed to carry out dynamic time history response analysis for each, and the average
trend of the responses from all of the motions is obtained. The average trends are shown in thick lines,
and the responses from individual ground motions are indicated by faint ones. It is to be mentioned that
the numerical values of the responses are not being emphasized here; rather, the trend of variations of
responses are identied and compared with their stochastic counterpart presented earlier.
The variations of peak transient responses (peak values) with different mass (Figure 13ac) and
length ratio (Figure 13df) under recorded motions are found to be similar to that obtained from stochastic analysis (Figure 6af for mass ratio and Figure 8af for length ratio). However, the only exception is in the displacement of the ball with the length ratios. The stochastic analysis shows decreasing
values of the displacement of ball with increasing length ratio, but beyond certain threshold value, the
ball displacement increases. Thus, a particular length ratio minimizes the displacement of the ball. In
contrary, the transient response behavior shows monotonically increasing ball displacement with increasing length ratio of TLCBD. Such discrepancies may be attributed to the signicant disparities
among the widely varying frequency contents of the selected ground motions from the smooth
KanaiTajimi spectra pertaining to the stochastic earthquake model.
Further, the variations of the peak transient responses are also veried with respect to the stochastic
response under varying dominant frequency content and intensities of recorded motions. The intensity
of individual ground motions are calculated using Equation (7) from their respective r.m.s values.
These intensities are then multiplied by suitable factor in order to vary their magnitudes. The dominant
frequency content can be identied from the Fourier spectra of the ground motions. A shift in the dominant frequency (and for all frequencies) of a particular motion is achieved by suitably adjusting its
sampling interval. This is because the sampling interval dictates the Nyquist frequency of a particular
motion and hence causes a shift (on lower/upper side depending on the increased/reduced sampling interval) in its frequency content.

Figure 13. Peak (a) displacement of ratio, (b) displacement of liquid, and (c) displacement of the ball of the structure for different mass ratios of the optimal tuned liquid column damper and tuned liquid column ball damper.
Displacement variations for (d) ratio, (e) liquid, and (f) ball of structure for different length ratios.
Copyright 2015 John Wiley & Sons, Ltd.

Struct. Control Health Monit. 2015; 22: 13251342


DOI: 10.1002/stc

1340

S. GUR, K. ROY AND S. K. MISHRA

The variations in the peak transient responses for different seismic intensities (Figure 14ac) show
sufcient degree of match with the trend of stochastic response variations, shown earlier in
Figure 10df. Similar agreements are also observed for the response variations under varying dominant
frequency content of ground motions (Figure 14df with Figure 11df). It is also noted that no discrepancies are observed among the displacement of the ball which are otherwise present for varying response under different mass and length ratios.
Taken together, the aforementioned comparison establishes the legitimacy of the stochastic optimal
response behavior of the TLCBD under real earthquakes. Signicant improvement in the control efciency along with reduced displacement of the liquid column is shown to be achieved in TLCBD while
compared with the conventional TLCD.
A step by step procedure can be synthesized to augment the optimal design of the TLCBD system
for seismic vibration control of structure as follows.
1.
2.
3.

4.

5.

6.

For the given frequency and damping of the structure, typical value of mass ratio of liquid for
the TLCBD is adopted. The mass ratio () should preferably be within 0.5% to 5%.
For the given time period, damping of the structure and mass ratio, the optimal value of the
tuning ratio (), and the ball-to-tube diameter ratio (Rbt) can be obtained.
From the optimal tuning ratio () and structural frequency (s), the frequency of vibration of liquid column (l) can be obtained. Further, the
length of the liquid column (Le) can be obtained
p
from the frequency of liquid column l 2g=L.
With varying length ratio, the optimal tuning does not change signicantly for TLCBD.
However, the length ratio (p) corresponds to the minimum displacement of the ball (0.6) can
be adopted, which is little different than the commonly adopted value of 0.9 in TLCD [11].
From the optimal length ratio and the total length of the liquid column (Le), the horizontal
(Be) and vertical length of the liquid column (h) can be obtained.
From the adopted mass ratio () and the mass of the structure (ms), the required mass of the
liquid (ml) can be obtained. This will provide an estimate of the required diameter of the liquid
tube. The required diameter of the ball can then be obtained from the optimal ball-to-tube diameter ratio (Rbt).
With the adopted set of parameters, the maximum displacement of the liquid column and the
ball can be obtained under the design ground motions. Based on the maximum displacement

Figure 14. Displacements of (a) ratio, (b) liquid, and (c) ball of the controlled structure of the optimal system for
different seismic intensities. Similar response variations for (d) displacement ratio, (e) liquid displacement, and (f)
ball displacement of the controlled structure for different dominant frequency contents of the ground motions.
Copyright 2015 John Wiley & Sons, Ltd.

Struct. Control Health Monit. 2015; 22: 13251342


DOI: 10.1002/stc

1341

TUNED LIQUID COLUMN BALL DAMPER

of the liquid, free height of the vertical column should be provided. Further, the maximum displacement of the ball must be less than half of the length of the horizontal portion of the liquid tube.
If any of the two criteria is violated, the design is reiterated by adopting a different mass ratio.

6. CONCLUSION
The signicantly improved performance of TLCBD over the conventional TLCD is established for
controlling stochastic vibration of structures subjected to random earthquakes. The important parameters affecting the performance of the structure-TLCBD system are identied, and their optimal values
are obtained through optimization. The response evaluation for the process of optimization is based on
random vibration analysis. Extensive numerical simulations have been conducted to establish the robustness of the improved performance of TLCBD over the TLCD. Further, the viability of optimal performances of TLCBD are veried by subjecting the optimal TLCBD-assisted structure to a set of real
ground motions. The trend of the optimal stochastic response behavior matches agreeably well with the
responses under real ground motions to justify the viability of the optimal response behavior.

APPENDIX A
Details of the augmented system matrix used in Equation (12) is given as
2

0n;n

6
6 f0g1;n
6
A 6
6 M 1
K n;n
n;n
4
f0g1;n

f0gn;1

I n;n

f0g1;n

2f fr gn;1

M 1
C n;n
n;n

2f

f0g1;n

f0gn;1

7
7
7
7
2 f f frgn;1 7
5
2 f f
1

where n is the degree of freedom of the structure with the control system (n = 3 for the TLCBDstructure system).
The system property matrices for the same are given as
2

6
M  6
4 0

p

 p

6
7
7; C  6 0
4
5

c

2 l s

2 s s

0 k
6
7
6
2g
7; K  6 0
5
6
Le
4
0

7
7
0 7
7
5
2s

in which the terms used in the matrices are dened as


p 2=7; 2mb =5ms 8b Rb R2bt =15l Le ; c 45=14R2b b ; k 15l g=14b Rb
The power spectral density matrix for the rock bed seismic motion is expressed as
2

0n;n

6 f0g
6
1;n
Sww  6
6 0
4 n;n
f0g1;n
Copyright 2015 John Wiley & Sons, Ltd.

f0gn;1

0n;n

f0gn;1

f0g1;n

f0gn;1

0n;n

f0g1;n

7
7
7
f0gn;1 7
5
2S0

Struct. Control Health Monit. 2015; 22: 13251342


DOI: 10.1002/stc

1342

S. GUR, K. ROY AND S. K. MISHRA


REFERENCES

1. Sakai F, Takeda S, Tamaki T. Tuned liquid column damper-new type device for suppression of building vibration, In:
Proceedings of the International Conference on High Rise Buildings, Nanjing, China 1989; 926931.
2. Soong TT, Dargush GF. Passive Energy Dissipation Systems in Structural Engineering. John Wiley & Sons: New York,
1997.
3. Felixa Jorge LP, Balthazar Jose M, Brasilc Reyolando MLRF. On tuned liquid column dampers mounted on a structural
frame under a non-ideal excitation. Journal of Sound and Vibration 2000; 282:12851292.
4. Yalla SK, Kareem A. Optimum absorbers parameters for tuned liquid column dampers. Journal of Structural Engineering,
ASCE 2000; 125:906915.
5. Ghosh A, Basu B. Seismic vibration control of short period structures using the liquid column damper. Engineering
Structures 2004; 26:19051913.
6. Haroun MA, Pires JA. Active dampers, Proc orice control in hybrid liquid column. In 1st World Conf. on Structural
Control Vol, Los Angeles I, Lewis FL, Syrmos VL (eds.). Optimal Control. Wiley: New York, 1995.
7. Yalla S, Kareem A, Kantor JC. Semi-active tuned liquid column dampers for vibration control of structures. Engineering
Structures 2001; 23:14691479.
8. Wang JY, Ni YQ, Ko JM, Spencer JF. Magneto-rheological tuned liquid column dampers (MR-TLCDs) for vibration mitigation of tall buildings: modeling and analysis of open-loop control. Computer and Structures 2005; 83:20232034.
9. Watkins RD. Tests on a liquid column vibration absorber for tall structures. Steel Structures: Recent Research and Developments, Singapore, 1991; 22(24).
10. Gao H, Kwok KCS, Samali B. Optimization of tuned liquid column dampers. Engineering Structures 1997; 19:476486.
11. Xu YL, Samali B, Kwo KCS. Control of along-wind response of structures by mass and liquid dampers. Journal of Engineering Mechanics, ASCE 1992; 118:2039.
12. Haroun MA, Pires JA, Won AYJ. Suppression of environmental induced vibrations in tall buildings by hybrid liquid column
dampers. The Structural Design of Tall Buildings 1996; 5(1):4554.
13. Wu JC, Wang YP, Lee CL, Liao H, Chen YH. Wind-induced interaction of a non-uniform tuned liquid column damper and a
structure in pitching motion. Engineering Structures 2008; 30:35553565.
14. Gao H, Kwok KSC, Samali B. Characteristics of multiple tuned liquid column dampers in suppressing structural vibration.
Engineering Structures 1999; 21:316331.
15. Shum KM. Closed form optimal solution of a tuned liquid column damper for suppressing harmonic vibration of structures.
Engineering Structures 2009; 31:8492.
16. Hitchcock PA, Kwok KCS, Watkins RD, Samali B. Characteristics of liquid column vibration absorbers (LCVA)-I. Engineering Structures 1997; 19:12634.
th
17. Watkins RD, Hitchcock PA. Tests on various liquid column vibration absorbers. In: Proceedings of the 4 International
Conference on Motion and Vibration Control, Yokohama; 1992; 1130.
18. Chang CC, Hsu CT. Control performance of liquid column vibration absorbers. Engineering Structures 1998; 20:580586.
19. Al-Saif KA, Aldakkan KA, Foda MA. Modied liquid column damper for vibration control of structures. International
Journal of Mechanical Sciences 2011; 53:505512.
20. Wu JC, Shih MH, Lin YY, Shen YC. Design guidelines for tuned liquid column damper for structures responding to wind.
Engineering Structures 2005; 27:1893905.
21. Warburton GB, Ayorinde EO. Optimum absorber parameters for simple systems. Earthquake Engineering and Structural
Dynamics 1980; 8:197217.
22. Ayorinde EO, Warburton GB. Minimizing structural vibrations with absorbers. Earthquake Engineering and Structural
Dynamics 1980; 8:219236.
23. Kanai K. Semi-empirical formula for the seismic characteristics of the ground. Bulletin of Earthquake Research Institute,
University of Tokyo 1957; 35:309325.
24. Tajimi HA. Statistical method of determining the maximum response of a building structure during an earthquake. Proceedings of the 2nd World Conference on Earthquake Engineering 1960; 11:781798.
25. Crandall SH, Mark WD. Random Vibration in Mechanical Systems. Academic Press, New York, 1963.
26. Lutes DL, Sarkani S. Random Vibrations, Analysis of Structural and Mechanical Systems. Elsevier: Burlington, USA, 2004.
27. Taanidis AA, Beck JL, Angelides DC. Robust reliability-based design of liquid column mass dampers under earthquake
excitation using an analytical reliability approximation. Engineering Structures 2007; 29:35253537.
28. Taanidis AA, Scruggs JT. Performance measures and optimal design of linear structural systems under stochastic stationary
excitation. Structural Safety 2010; 32(5):305315.
29. Jangid RS. Stochastic response of building frames isolated by lead-rubber bearings. Journal of Structural Control and
Health Monitoring 2010; 17:122.
30. Chakraborty S, Debbarma R, Marano GC. Performance of tuned liquid dampers considering maximum motion in seismic
vibration control of structures. Journal of Sound and Vibration 2012; 331:15191531.
31. Chakraborty S, Debbarma R. Stochastic earthquake response control of structures by liquid column vibration absorber with
uncertain bounded system parameters. Structural Safety 2011; 33(2):136144.

Copyright 2015 John Wiley & Sons, Ltd.

Struct. Control Health Monit. 2015; 22: 13251342


DOI: 10.1002/stc

You might also like