You are on page 1of 21

332 Int. J. Materials and Structural Integrity, Vol. 3, No.

4, 2009

Passive control of seismic response of soil-structure


system by the compliant liquid column damper

Ratan Kumar Ghosh and


Aparna (Dey) Ghosh*
Department of Civil Engineering,
Bengal Engineering and Science University,
Shibpur, P.O. Botanic Gardens,
Howrah, 711 103, India
E-mail: ratan_ju@rediffmail.com
E-mail: aparna@civil.becs.ac.in
*Corresponding author

Abstract: This paper studies the seismic vibration control of flexible-base short
period structures by the compliant liquid column damper (CLCD). The
input-output relation of the CLCD-structure-soil system is formulated in the
frequency-domain. A numerical stochastic study with varying degrees of
soil-structure interaction (SSI) and for different tuning conditions indicates that
tuning to the flexible-base structural frequency is essential for effective damper
performance. The effects of SSI on the sensitivity of damper performance to
the tuning ratio and on the optimal value of the orifice damping coefficient
have also been examined. Finally, the practical design of the CLCD for an
example five-storied building, founded on soft clay and subjected to a recorded
accelerogram, has been carried out. It has been found that it is possible to
obtain a feasible configuration of the damper that will cater for the proper
tuning and orifice damping parameters to provide substantial response
reduction.

Keywords: compliant liquid column damper; CLCD; soil-structure interaction;


SSI; complex valued impedance functions; short period structures;
multi-degree-of-freedom system; MDOF system; passive control; seismic
response.

Reference to this paper should be made as follows: Ghosh, R.K. and Ghosh, A.
(2009) ‘Passive control of seismic response of soil-structure system by the
compliant liquid column damper’, Int. J. Materials and Structural Integrity,
Vol. 3, No. 4, pp.332–352.

Biographical notes: Ratan Kumar Ghosh is a former post-graduate student in


the structural engineering specialisation of the Department of Civil
Engineering, Bengal Engineering and Science University, Shibpur, India. He is
presently an Assistant Manager (Civil) of the West Bengal Power Development
Corporation, Ltd., working in the Sagardighi Thermal Power Project.

Aparna (Dey) Ghosh is an Assistant Professor at the Department of Civil


Engineering, Bengal Engineering and Science University, Shibpur, India. Her
areas of interest include vibration control of structures, primary-secondary
systems and soil-structure interaction.

Copyright © 2009 Inderscience Enterprises Ltd.


Passive control of seismic response of soil-structure system 333

1 Introduction

Passive control devices are popular in the field of structural protection from
environmental forces such as those due to wind, waves and earthquakes, chiefly due to
the non-requirement of an external supply of power. Among the passive control devices,
the liquid dampers, in particular, provide some unique benefits, such as low cost, easy
implementation, especially in existing structures, and effectiveness even for low
vibrational amplitudes, etc. The liquid column damper (LCD) is a type of liquid damper
in which the liquid oscillates in a tube-like container through orifice(s). It offers some
further advantages as it has a high volumetric efficiency for a given volume of liquid,
consistent behaviour over a wide range of excitation levels and a very definite damping
mechanism.
The LCD, as in the case of some other passive dampers like the TMD, operates on the
principle of transference of the vibrational energy of the structure to the mass of the
damper when the natural frequency of the structure and damper are near equal (i.e.,
tuned). The conventional LCD is rigidly attached to the structure and is a long period
system. Thus, from the requirement of tuning, the LCD is chiefly applicable to flexible
structures such as tall buildings, cable-stayed bridges and towers. Further, most of the
studies on the LCD relate to wind induced excitations (e.g., Sakai et al., 1991; Xu et al.,
1992; Balendra et al., 1995, 1999; Shum and Xu, 2004; Wu et al., 2005; Min et al., 2005),
while a few studies focus on control of earthquake induced ground motions (such as
Haroun et al., 1994; Sun, 1994; Won et al., 1996; Reiterer and Ziegler, 2005). Recently, it
has also been studied for the reduction of wave-induced vibrations of the floating
offshore platform by Lee et al. (2006).
The natural frequency of the LCD is the frequency at which the liquid column in the
damper oscillates and this is inversely proportional to the square root of the length of the
liquid column. This leads to constraints in applying the LCD for short period structures as
tuning the LCD to a short period structure presents difficulties in obtaining a feasible
length of the liquid column. As the frequency bandwidth of an earthquake ground motion
is generally broad and rich in high frequency content, the short and stiff structures are
vulnerable to damage from earthquakes. The non-applicability of the otherwise very
promising LCD as a seismic vibration control device to comparatively stiff structures, led
Ghosh and Basu (2004) to propose a variation of the conventional LCD that is essentially
an LCD connected by a spring to the primary structure. This modified damper, termed as
a compliant LCD (CLCD), is thus a two-DOF system, in which there is motion of the
damper relative to the structure as well as oscillation of the liquid column relative to the
container, when the structure vibrates. It can be shown (Ghosh and Basu, 2008) that the
two combined frequencies of the CLCD lie very close to two independent frequencies,
one being the natural frequency of the oscillating liquid column in the conventional LCD
without a spring and the other being the natural frequency of horizontal motion of the
damper when the liquid has no motion relative to the container (i.e., when it acts as a
mass damper). Designing the CLCD so that the latter frequency is tuned to the frequency
of the structural mode to be controlled has been shown to be effective and thus the CLCD
eliminates the requirement of tuning to the oscillating liquid column as in the case of the
conventional rigidly attached LCD. Though at this structural frequency, the frequency of
the oscillating liquid column is detuned, for the earthquake type of excitations, some
oscillating motion of the liquid column does take place even at this non-resonant
334 R.K. Ghosh and A. Ghosh

frequency of the liquid and that provides the necessary damping (orifice damping). Thus,
no additional dashpot is required in the connection between the damper and the structure
which is an advantage as compared to the use of a TMD. Furthermore, the CLCD allows
the consideration of the interaction between the liquid and the remaining system as a
function of the ratio of the orifice damping coefficient to the length of the liquid column.
This offers a wider choice of a combination of the length of the liquid column and the
coefficient of head loss to provide the same amount of liquid damping from a practical
consideration. Moreover, there is a wide range of the ratio of the orifice damping
coefficient to the length of the liquid column, where the sensitivity of the performance of
the CLCD to this ratio is low. Also, the restriction in providing the optimal damping
(through coefficient of head loss) by limiting the liquid motions (Won et al., 1996) is not
of direct concern in the CLCD, which is a significant practical advantage.
The design of any damper system for its proper functioning requires reasonably
accurate information of the properties of the structure to which it is to be attached.
Generally, the properties of the structure used in the design are those evaluated
considering the structure to be of a fixed-base type. However, if the structure is founded
on compliant soil, the soil-structure interaction (SSI) effects may cause a reduction of the
fundamental frequency, an increase in the overall damping of the system and result in a
modification of the actual foundation motion from the free-field ground motion. A
significant change in the structural properties will cause a considerable effect on the
performance of the damper which needs careful attention. Xu and Kwok (1992)
examined the response of soil-structure-TMD system in the frequency domain for the
case of wind excitations. They reported that the TMD should be tuned to the fundamental
frequency of the soil-structure system. This was found to be effective for moderate to
stiff soils but for soft soil the performance of the properly tuned, mass damper remained
poor. Takewaki (2000) developed a method for the response reduction of structures by a
combination of viscous damper and TMD considering SSI effects. Wang and Lin (2005)
investigated the vibration control of irregular buildings subjected to base motions by
multiple tuned mass dampers (MTMD), incorporating SSI effects. Ghosh and Basu
(2005) studied the performance of the conventional LCD model for seismic applications
considering SSI effects. The study indicated that for medium to soft soil, the effect of SSI
on the damper performance is not significant. However, for very soft soils, the
effectiveness of the conventionally tuned LCD is greatly diminished. For such cases, it is
essential to tune the damper to the natural frequency of the structure-foundation system.
Ghosh and Ghosh (2008) examined the effects of varying degrees of SSI on the
performance of the CLCD for a base-excited structure modelled by a single
degree-of-freedom (SDOF) system and reported that even in case of medium soft soil,
tuning the CLCD to the fundamental frequency of the structure-foundation system would
considerably improve the damper performance.
In this work, a detailed study on the CLCD-structure-soil system leading to the design
of a CLCD for the seismic vibration control of an example five-storied building founded
on soft clay has been carried out. The system has been analysed in the frequency domain
for which the transfer function of the tip displacement response of the
flexible-base structure modelled as a multi-degree-of-freedom (MDOF) system, with
CLCD, to the input base acceleration, has been formulated. A linear soil model has been
adopted and the foundation stiffness and damping, which are frequency-dependent, have
been expressed by complex-valued impedance functions (Veletsos and Wei, 1971; Wong
and Luco, 1978; Aspel and Luco, 1987). The rocking component of the free-field ground
Passive control of seismic response of soil-structure system 335

motion has been neglected. However, both translational as well as rocking motions of the
foundation have been considered. The effect of the foundation motions has been
considered through suitable modification of the actual input to the superstructure via the
sub-structure approach (e.g., Dey and Gupta, 1999).

Figure 1 Model of CLCD-MDOF structure-foundation system


336 R.K. Ghosh and A. Ghosh

2 Modelling of the flexible-base structure-CLCD system

The flexible-base MDOF structure-CLCD model investigated is shown in Figure 1. The


tube-like container of the LCD has cross-sectional area, A, and horizontal dimension, B.
It contains liquid of mass density, ρ and column length, L. The head loss coefficient,
controlled by the opening ratio of the orifices installed in the damper tube, is denoted by
ξ . The mass, stiffness and damping of the superstructure, modelled as a n-DOF lumped
mass system, are represented by [ M ] , [ K ] , [C ] , respectively and [ h ] denotes the height
from ground to where the lumped masses are assumed to be concentrated. The stiffness
and damping of the member connecting the LCD to the structure are denoted by KD and
CD respectively. The mass of the container of the LCD, exclusive of the liquid mass, is
denoted by MC. Thus the total mass of the damper system is expressed as ( M c + ρAL) . It
is assumed that the mass of the liquid in the horizontal portion of the damper tube, ρAB ,
is concentrated at a height hn from the foundation. The performance index of the CLCD
is considered to be the tip displacement of the structure. The foundation of the structure is
considered to be a rigid slab of mass M 0 , anchored to the surface of a homogeneous,
viscoelastic half-space through linear springs and viscous damping elements. The
stiffness and damping of the underlying soil is represented by complex valued impedance
functions.

3 Formulation of transfer function

By neglecting the rocking component of the free-field ground excitation and the effects
of kinematic interaction, the foundation input motion is assumed to be same as the free
field ground translation, represented by z (t ). Let the foundation undergo translation
z0 (t ) and the rotation θ0 (t ) relative to the soil medium. Let {x(t )} denote the horizontal
displacement vector of the MDOF structure relative to the foundation, thus xn (t ) denotes
the horizontal displacement of the nth DOF of the structure, relative to the foundation.
Consider u (t ) to be the change in elevation of the liquid column in the damper. Further,
let y (t ) represent the horizontal motion of the damper container relative to xn (t ) .
Assuming small displacements, the total horizontal displacement of the structure at the
nth-DOF is given by { xn (t ) + z (t ) + z0 (t ) + hnθ 0 (t )} . The interaction forces between the
foundation and the underlying soil interface are represented by Vs (t ) , the base shear,
and M s (t ) , the base moment. The equation of motion of the liquid column from
equilibrium may be written as (Saoka et al., 1988).

1
ρ ALu(t ) + ρ Aξ u (t ) u (t ) + 2 ρ Agu (t ) = − ρ AB { y (t ) + 
xn (t ) +  z0 (t ) + hnθ0 (t )} (1)
z (t ) + 
2

The non-linear system in equation (1) may be expressed by the following equivalent
linear equation

ρ ALu(t ) + 2 ρ ACP u (t ) + 2 ρ Agu (t ) = − ρ AB { y (t ) + 


xn (t ) +  z0 (t ) + hnθ0 (t )} (2)
z (t ) + 
Passive control of seismic response of soil-structure system 337

where C P represents the equivalent linear damping co-efficient and is expressed by


Xu et al. (1992).
σ u
CP = ξ (3)

C P may be obtained by minimising the mean square value of the error between equations
(1) and (2). In equation (3), σ u is the standard deviation of the liquid velocity, u (t ) ,
which is assumed to be a Gaussian process. As is evident from equation (3), the
dependence of C P on the response, σ u , of the liquid column calls for an iterative
solution procedure for C P . Normalising equation (2) with respect to the mass of liquid in
the container, ρAL, leads to

CP
u(t ) + 2
L
u (t ) + ωl 2u (t ) = −α {
y (t ) + 
xn (t ) +  z0 (t ) + hnθ0 (t )
z (t ) +  } (4)

[ ]
where ω l = 2 g / L and is defined as the frequency of the liquid oscillating in the LCD
and α [= ( B / L)] is the ratio of the horizontal portion of the liquid column to its total
length.
The dynamic equilibrium of the whole damper system, comprising of the liquid
column mass and the container mass, leads to the following equation:
( ρ AL + M C ){
y (t ) + 
xn (t ) +  z0 (t ) + hnθ0 (t )} + ρ ABu(t ) + CD y (t ) + K D y (t ) = 0
z (t ) +  (5)

On normalising equation (5) with respect to ( ρAL + M C ) , the following is obtained:

{ y (t ) + xn (t ) + z (t ) + z0 (t ) + hnθ0 (t )} + (1α+ τ ) u(t ) + 2ζ Dω D y (t ) + ω D 2 y (t ) = 0 (6)

In equation (6), ω D = {k D / ( ρ AL + M C )}
1/2
and ζ D = ⎡⎣CD / 2ω D ( ρ AL + M C )⎤⎦ denote the
natural frequency and the damping ratio of the whole damper system, considering the
liquid in the container to be relatively still. Also in equation (6), the term τ = {M C / ρAL}
represents the ratio of container mass to the liquid mass.
The equation of motion for the mass, [ M ] , of the MDOF system, is given by

[ M ]{x(t )} + [C ]{ x(t )} + [ K ]{ x(t )} = − [ M ]{1} ( z(t ) + z0 (t ) ) − [ M ]{h}θ0 (t ) + { f (t )} (7)

⎧ 0 ⎫
⎪ : ⎪
⎪⎪ ⎪⎪
Here, { f (t )} = ⎨ : ⎬ denotes the interactive force between the structure and the damper
⎪ 0 ⎪
⎪ ⎪
⎪⎩v(t ) ⎭⎪
with v(t ) = CD y (t ) + K D y (t ).
Assuming the structure to be classically damped, it is possible to expand its response
in terms of the orthogonal modes shapes in the following way:
338 R.K. Ghosh and A. Ghosh

{ x(t )} = [ϕ ]{q(t )} (8)

In equation (8), [ϕ ] is the real valued modal matrix of the structure and {q (t )} is the
vector of normal coordinates. The decoupled equation for the j th mode of the structure,
considering mass normalised mode shapes, is

q j (t ) + 2ζ j ω j q j (t ) + ω2j q j (t )
(9)
{ } [ M ]{h} θ0 (t ) + {φ( j ) } { f (t )}
T T
= −α j (  z0 (t ) ) − φ( j )
z (t ) +  j = 1, 2,..., n

{ } { }
T
where φ( j ) is the transpose of the j th mode shape ϕ ( j ) , and α j is the modal

participation factor of the j th mode and is given by {ϕ } [ M ]{1} . Furthermore, in


( j) T

equation (9), ω j and ζ j denote the natural frequency and damping ratio in the j th
mode, respectively.
The equation of equilibrium for the complete structure-foundation system in
translation can be written as
n n
VS (t ) + ∑∑ M ϕ
j =1 r =1
j
(r )
j q r (t ) + M T ( z (t ) + z0 (t ) ) + M HT θ0 (t ) = ( CD y (t ) + K D y (t ) ) (10)

⎡ n ⎤
where M T ⎢ = M 0 + ∑ Mj ⎥⎥ is the total mass of the structure-foundation system.
⎣⎢ ⎦ j=1

The equation of equilibrium for the complete structure-foundation system in rotation,


ignoring the contributions of the gravitational forces, can be written as
n n
M S (t ) + ∑∑ M
j =1 r =1
(r )
j H jϕ j q r (t ) + M HT (  z0 (t ) ) + IT θ0 (t )
z (t ) + 
(11)
= ( CD y (t ) + K D y (t ) ) hn

⎡ n ⎤
where M HT ⎢=
⎢⎣

j =1
M j h j ⎥ is the moment of the structure-foundation system about the
⎥⎦
⎡ ⎤
∑( )
n
ground level, and I T ⎢= I 0 + I j + M j h 2j ⎥ is the moment of inertia of the primary
⎢⎣ j =1 ⎥⎦
structure-foundation system about a horizontal axis at the ground level. I 0 represents the
mass moment of inertia of the foundation and I j represents the mass moment of inertia
of the jth floor mass, M j , about a horizontal axis through its mass centre.
The interaction forces between the foundation and underlying soil are related to the
foundation displacements in the frequency domain by the complex valued impedance
functions KVV (ω ) , KVM (ω ) ( = K MV (ω ) by reciprocity theorems) and K MM (ω ) . The
relationship between the base shear and overturning moment acting at the foundation soil
Passive control of seismic response of soil-structure system 339

interface and the corresponding foundation translation and rotation are expressed as
(Chopra and Gutierrez, 1974; Gupta and Trifunac, 1991; Wu and Smith, 1995; Dey and
Gupta, 1999)

⎧ VS (ω ) ⎫ ⎡ KVV (ω ) KVM (ω ) ⎤ ⎧ Z 0 (ω ) ⎫
⎨ ⎬=⎢ ⎥⎨ ⎬ (12)
⎩ M S (ω ) ⎭ ⎣ K MV (ω ) K MM (ω ) ⎦ ⎩θ 0 (ω ) ⎭

Here, VS (ω ) and M S (ω ) denote the Fourier transforms of VS (t ) and M S (t ) ,


respectively. KVV (ω ) and K MM (ω ) represent the translational and rocking impedance
functions, while KVM (ω ) (= K MV (ω ) ) is the coupling impedance function of the
foundation.
Due to the frequency dependence of the foundation impedance functions, it is
necessary to transform the time domain equations into the frequency domain. The Fourier
transformation of equation (9) leads to

⎣ { }
Q j (ω) = H j (ω) ⎡ -α j ( Z(ω)+Z 0 (ω))-γ j θ0 (ω)+ nn(j)ν(ω) ⎤ for j = 1, 2,...,n

(13)

where
1
H ( ) (ω ) =
j
2 2
for j = 1, 2,..., n. (14)
ω j − ω + 2iζ jω jω

H ( ) (ω ) represents the modal transfer function relating the displacement response of the
j

SDOF system in the j th structural mode to the input ground acceleration. In

{ }
equation (13), γ j ⎛⎜ = ϕ ( j ) [ M ]{h} ⎞⎟
T
is the modal participation factor in the j th mode
⎝ ⎠
for the base rocking. Furthermore, in equation (13), Q( ) (ω ) , Z(ω ) , Z0 (ω ) , θ0 (ω ) and
j

v(ω ) represent the Fourier transforms of the corresponding time dependent variables,
q (t ) , z(t ) , z (t ) θ (t ) and v(t ) respectively. The expression for v(ω ) is obtained by
j 0 0

the Fourier transformation of v (t ) and is given by

ν (ω ) = Y (ω )(iωC D + K D ) (15)

with Y (ω ) denoting the Fourier transform of y (t ) [see equation (6)] and is expressed as

⎡ ⎧⎛ α ⎞ ⎫ ⎤
Y (ω ) = H D (ω )⎢ω 2 ⎨⎜ ⎟ U(ω ) + X n (ω ) + Z 0 (ω ) + hnθ 0 (ω )⎬ − Z(ω )⎥ (16)
⎣ ⎩⎝ 1 + τ ⎠ ⎭ ⎦

In equation (16), U (ω ) and X n (ω ) represent the Fourier transforms of u (t ) and xn (t ) ,


respectively while H D (ω ) represents the transfer function relating the horizontal
displacement of the CLCD to the free field ground acceleration. The expression for
H D (ω ) is given by
340 R.K. Ghosh and A. Ghosh

1
H D (ω ) = (17)
ω D2 2
− ω + 2iζ Dω Dω

On taking the Fourier transformation of equation (4), U (ω ) is obtained as

U (ω ) = α H L (ω ) ⎡⎣ω 2 { Y (ω ) + X n (ω ) + Z 0 (ω ) + hnθ 0 (ω )} − Z (ω ) ⎤⎦ (18)

where

1
H L (ω ) = (19)
Cp
ω L2 2
− ω + 2i ω
L
denotes the transfer function relating the displacement of a spring-dashpot oscillator that
represents the liquid column in the LCD, to the free field ground acceleration.
On taking the Fourier transformation of equations (10) and (11), the following
equations are obtained.
n n
VS (ω ) + ∑∑ M ϕ
j =1 r =1
j
(r )
j q r (ω ) + M T ( Z(ω ) + Z0 (ω ) ) + M HT θ0 (ω ) = ( iω CD + K D ) Y (ω ) (20)

n n
M S (ω ) + ∑∑ M
j =1 r =1
(r )
j H jϕ j q r (ω ) + M HT (  z0 (ω ) ) + IT θ0 (ω ) = ( iω CD + K D ) hnY (ω )
z (ω ) +  (21)

The terms, VS (ω ) and M S (ω ) , are the Fourier transforms of VS (t ) and M S (t ) ,


respectively. Through appropriate substitutions from equations (12)–(19) in equations
(20) and (21) and the solution of the simultaneous equations in Z 0 (ω ) and θ0 (ω) lead to
(1)
Z 0 (ω) = χ ZZ (ω) Z(ω) – χ ZZ
(2)
(ω) X n (ω) (22)

θ 0 (ω ) = χθ(1)Z (ω ) Z(ω ) − χθ(2)


Z (ω ) X n (ω ) (23)
(1)
Here, χ ZZ (ω ) and χθ(1Z) (ω ) respectively, represent the transfer functions relating the
interaction displacement and rotation, Z 0 (ω ) and θ0 (ω) , to the input ground
acceleration, Z(ω ) . Similarly, χ ( 2) (ω ) and χ ( 2) (ω ) denote the transfer functions
ZZ θZ
relating the interaction displacement and rotation to the displacement of the nth-DOF of
the superstructure. The expressions for these transfer functions are given in the appendix.
It may also be observed that when the effects of SSI are negligibly small, the values of
(1)
χ ZZ (ω ) , χθ(1Z) (ω ) , χ ZZ
( 2)
(ω ) and χθ(Z2) (ω ) tend to become zero.
The substitution of equations (22) and (23) in equation (13) leads to

Q( j ) (ω ) = S Z( j ) (ω ) Z(ω ) − Sθ( j ) (ω ) X n (ω ) (24)

with,

SZ( j) (ω) = H j (ω) ⎡⎣(1−ω2χZZ


(1)
(ω))α j −γ jω2χθ(1)Z (ω) +φn( j) R(ω) −ω2φn( j) R(ω)χZZ
(1)
(ω) − hnω2χθ(1)Z (ω)⎤⎦ (25)
Passive control of seismic response of soil-structure system 341

and

Sθ( j ) (ω ) = H j (ω ) ⎡⎣ω 2 χ ZZ
(2)
(ω )α j − γ jω 2 χθ(2) ( j) 2 (2) 2 (2) ⎤ (26)
Z (ω ) + φn R(ω )ω (1 + χ ZZ (ω )) − hnω χθ Z (ω ) ⎦

Thus, the following linear, simultaneous equations are obtained, which describe the
displacements of the degrees-of-freedom of the structure in the frequency domain, in
relation to the input base acceleration.

n
X K (ω) = ∑ φ {S
K=1
( j)
K
( j)  ( j)
Z (ω) Z (ω) − Sθ (ω) X n (ω) }, k=1, 2,..., n (27)

If the ground acceleration is characterised by a white noise power spectral density


function (PSDF) of intensity S 0 , then the PSDF of the displacement response along the
kth DOF of the structure, denoted by S X k (ω ) , is expressed by Newland (1993)

2
S X k (ω ) = H X k (ω ) S0 (28)

where H X k (ω ) denotes the transfer function relating the displacement, relative to the
foundation, along the k th DOF of the structure with CLCD, to the free field ground
acceleration. The root mean square (RMS) value of this displacement response may be
numerically evaluated by computing the square root of the area under the PSDF curve as
given by equation (28).

4 Numerical study

An illustration of the transfer function of the CLCD-structure-soil system formulated in


the foregoing section is presented. An example three-DOF superstructure is considered.
The masses assumed to be concentrated at the bottom, middle and top levels are equal to
4 × 105 kg, 4 × 105 kg and 3 × 105 kg, respectively and the heights of lumping of the
masses are equal to 3.0 m, 6.0 m and 9.0 m above the base, correspondingly. The
connecting elements between the masses have equal stiffness values of 1.0 × 109 N/m.
The modal damping ratio is assumed to be 1% for all the three modes of vibration. The
values of mass ratio i.e., mass of the damper to the total mass of the structure, ( μ ) , length
of the liquid column in the damper ( L ) , horizontal length to total length of the liquid
column, (α ) and the mass of the liquid to the mass of the of the container, (τ ) , are
assumed as 1.0%, 2.0 m, 0.9 and 1.0, respectively. The values of the impedance functions
are taken from the results by Wong and Luco (1978) for the following soil parameters:
mass density ( ρ ) = 1500 kg/m3; Poisson’s ratio (λ ) = 0.3; hysteretic damping ratio
(ς ) = 0.02; and characteristic length of the rigid square foundation = 3.0 m. As in the
work of Ghosh and Basu (2004), ζ 2 = 0 and τ = 1 are considered. The values of M 0 and
I 0 are assumed to be negligible. The base input is assumed to be characterised by a white
noise PSDF with intensity, S0 = 100 cm2 s3 .
342 R.K. Ghosh and A. Ghosh

Figure 2 Displacement transfer functions of the top mass of an example three-DOF


superstructure without damper, for fixed base and flexible base with different soil
conditions (medium soft soil: υ s = 300 m s ; very soft soil: υ s = 100 m s )

Figure 3 Variation of response reduction with tuning ratio for flexible-base case with
υ s = 100 m s

The effect of different soil conditions is demonstrated by choosing different shear wave
velocities, υ s [ = G ρ ] where G is the shear modulus of soil. The varying effects of
SSI on the displacement transfer function of the top mass of the structure, without
damper, are shown in Figure 2. Medium soft soil is represented by υ s = 300 m s while
the condition of very soft soil is characterised by υ s = 100 m s. Figure 2 indicates that as
Passive control of seismic response of soil-structure system 343

the soil becomes softer, there is a significant shift in the natural frequencies of the system
from those of the fixed-base case as well as a considerable reduction in the amplitude of
the transfer function peaks, indicating the increased damping in the system due to SSI.
The peak corresponding to the fundamental frequency of 23.905 rad s is found to shift to
13.90 rad s and to 5.5 rad s for υ s = 300 m s and υ s = 100 m s , respectively. Thus, to
ensure effective performance by the damper, it is essential to tune it to the natural
frequency of the structure-foundation system. Figure 3 exhibits the sensitivity of the
damper performance to the tuning ratio when the damper is tuned to the flexible-base
frequency of the structure founded on very soft soil ( υ s = 100 m s ). Here, and in the
subsequent figures of this paper, the damper performance in terms of response reduction
is measured by the reduction in the RMS value of the tip displacement of the structure. It
is observed that the optimum tuning ratio is very close to unity (about 0.98). Hence,
ν = 1.0 may be adopted. Next, a study on the optimum value ξ / L is made in Figure 4
which compares the variation in the response reduction with ξ L for the
fixed-base case with that for the two typical flexible-base cases ( υ s = 100 m s and
υ s = 300 m s ). It is observed that the effects of SSI on the optimal value of ξ L are
negligible for medium soft soil. Hence in the design of the CLCD for the structure
founded on medium soft soil, the values of the optimum ξ L as evaluated for the
fixed-base case may be used. But for very soft soil, the optimum ξ L should be
evaluated by tuning the CLCD to the fundamental frequency of the structure-foundation
system. Otherwise, the performance efficiency of the CLCD will be up to 1/3 lower. This
is significant as in a past study considering higher mass ratio (Ghosh and Ghosh, 2008),
the effects of SSI on ( ξ L )opt were found to be negligible even for very soft soil.

Figure 4 Variation in response reduction with ξ L tuning ratio for fixed base and flexible base
conditions (medium soft soil: υ s = 300 m s ; very soft soil: υ s = 100 m s )
344 R.K. Ghosh and A. Ghosh

Figure 5 Displacement transfer functions of the top mass of an example three-DOF


superstructure founded on medium soft soil ( υ s = 300 m s ), without damper and with
damper for different tuning conditions

Considering optimum value of ξ L and the tuning ratio to be unity, the displacement
transfer function of the top mass of the example system is presented in Figure 5 and
Figure 6 for the two typical soil conditions, respectively. In each figure, the transfer
functions for the case of no damper, case of damper tuned to fundamental frequency of
fixed-base structure and case of damper tuned to fundamental frequency of
structure-foundation system are compared. It is observed from Figures 5 and 6 that the
transfer functions for the case without damper and with damper tuned to the fundamental
frequency of the fixed-base structure are almost identical, thereby indicating that tuning
the CLCD to the fixed-base structural frequency may make the CLCD practically
ineffective. By tuning the CLCD to the fundamental frequency of the
structure-foundation system, response reductions of 32.04% and 25.91% are obtained for
the medium soft soil and very soft soil conditions respectively. In order to compare these
values with the response reduction that the CLCD would achieve for the fixed-base
example system, the RMS value of the displacement of the top mass of the fixed-base
example system, with and without CLCD, was evaluated for the same base input as for
Figures 5 and 6. It was found that the damper achieves a response reduction of 48.56%
when incorporated into the fixed-base example system. This may be explained by the
reduced response of the flexible-base system as compared to the fixed-base system
(see Figure 2), due to increased system damping resulting from SSI effects. Ghosh and
Basu (2004) have observed that the effectiveness of the damper as a vibration absorber
decreases with increase in the system damping.
Passive control of seismic response of soil-structure system 345

Figure 6 Displacement transfer functions of the top mass of an example three-DOF


superstructure founded on very soft soil ( υ s = 100 m s ), without damper and with
damper for different tuning conditions

5 Design of the CLCD for an example five-storied building on soft clay

An example five-storied building is considered (see Figure 7) and the relevant data that
define the structure are given below:
beam size (250 × 400) mm
column size (400 × 500) mm
slab thickness 125 mm
wall thickness:
outer wall 250 mm
inner wall 125 mm
superimposed load:
1 for all the floor levels (except roof) = 3.0 kN/m2
2 for the roof = 2.0 kN/m2.
On modelling the structure as a shear building with five degrees-of-freedom, the natural
frequencies of this system are obtained as ω 2 = 40.31rad s [0.45s], ω 2 = 40.31rad s
[0.16s], ω 2 = 62.79 rad s [0.10s], ω 2 = 79.17 rad s [0.08s] and ω 2 = 88.05 rad s
[0.07s]. The damper would be tuned to the fundamental as it is the predominant mode for
horizontal vibration. A modal damping ratio of 2% is considered. The building is
assumed to be founded on soft clayey soil having shear wave velocity υ s = 110 m s. The
other soil and foundation parameters are taken to be the same as those considered in the
346 R.K. Ghosh and A. Ghosh

foregoing section. The seismic input is assumed to be the recorded accelerogram of the
S00E component of the 1940 Imperial Valley earthquake at the El Centro site.

Figure 7 Plan and elevation of an example five-storied building

Since the force-displacement relationships of the soil are denoted by the complex valued
impedance functions which are frequency dependent, an indirect method is employed to
obtain the response of the CLCD-structure-soil system to the input accelerogram. First,
the RMS value of the tip displacement response of the fixed-base five-DOF example
system to the input accelerogram is evaluated by a time history analysis using the 4th
order Runge-Kutta method. Next, it is assumed that the El Centro excitation, which is
broad-banded in its frequency content, may be characterised by an equivalent white noise
PSDF. In order to determine the spectral intensity of the white noise characterising the El
Centro excitation, the RMS value of the tip displacement response of the fixed-base five-
DOF example system to white noise input is evaluated in frequency domain for different
values of S0 , till the RMS value of the tip displacement response matches with that
obtained from the time history analysis. In the present case, S0 is obtained as 170cm2 s3 .
The CLCD with water as the damper liquid is assumed to be placed on the roof of the
building. Though higher mass ratios lead to greater response reduction, it may not be
practicable to go for a very high value of the mass ratio, μ . Here, μ = 1% is adopted.
Passive control of seismic response of soil-structure system 347

Now, total mass of the structure = 1.4496 × 106 kg.


∴ The total mass of the CLCD = 14,496 kg.
ρ AL
Ratio of the mass of the water to mass of the container τ = = 1.0 is considered.
MC
Mass of the water required = 7248 kg.
It is recommended that for better performance of the damper α [ = B / L ] should be as
high as possible, but due to geometric constraints this is not always possible. In the
present case the total plan area of the structure is divided into nine equal panels
(see Figure 7) and to accommodate the calculated water mass within a panel area
(4.0 m × 4.0 m), the value of α is considered as 0.75 and the value of the horizontal
length, B , is taken as 3.0 m. Thus, the length of water column of the CLCD, L , is
obtained as 4.0 m. Considering a rectangular cross-section of the damper tube of width
equal to 0.5 m, the clear dimensions of the damper container in plan are obtained as
3.5 m × 3.6 m (see Figure 8). The mass of the container, consisting of 16 mm plate,
stiffners, connections, etc. is also equal to 7248 kg.

Figure 8 Dimensions of the designed CLCD

3.6 m

1.016 m

0.1 m
0.25 m
0.5 m

2.5 m
3.5 m

Note: The values indicated are in the inner dimensions of the container.
348 R.K. Ghosh and A. Ghosh

The transfer function of the tip displacement of the flexible-base structure, without
damper, is shown in Figure 9, from which the fundamental frequency of the
structure-foundation system is obtained as 3.7 rad s. The CLCD is tuned to this value
and for tuning ratio (ν ) equal to unity, the stiffness of the connecting element of the
damper is obtained as 198.45 kN/m.
The dimension of the orifice is evaluated on the basis of the optimum value of ξ L
for the minimum RMS value of the tip structural displacement. This yields a value of
ξopt equal to 56.0. Wu et al. (2005) have provided an empirical formula relating the head
loss and blocking ratio from their experiments. The expression is given by
ξ = (−0.6ψ + 2.1ψ 0.1 )1.6 (1 −ψ )−2 (29)

where, ξ is the head loss coefficient and ψ is the blocking ratio. Here, ψ is obtained as
0.807 and the dimension of the orifice is evaluated as about 0.1 m. The response transfer
function with the designed CLCD incorporated into the structure is shown in Figure 9.
The response reduction achieved by the CLCD is obtained as 33.78%.

Figure 9 Transfer functions of the tip displacement of the five-storied building on soft clay,
without damper and with the designed damper

6 Conclusions

The displacement transfer function of a structure, modelled as a MDOF system and


founded on compliant soil, with an attached CLCD has been formulated. Using this
formulation, a numerical study has indicated that to ensure effective performance by the
damper, it is essential to tune it to the natural frequency of the structure-foundation
Passive control of seismic response of soil-structure system 349

system. The optimum value of ν is close to unity and the sensitivity of the damper
performance to the tuning ratio is similar to that for the fixed-base case. The effects of
SSI on the optimal value of ξ L are negligible for medium soft soil. Hence, in the design
of the CLCD for the structure founded on medium soft soil, the value of (ξ L )opt as
evaluated for the fixed-base case may be used. But for very soft soil, the optimum ξ L
should be evaluated by tuning the CLCD to the fundamental frequency of the
structure-foundation system. A practical design of the CLCD, for an example,
five-storied building on soft clay has been carried out. The procedure to obtain a feasible
configuration of the damper including the opening ratio of the orifice has been outlined.
The results indicate effective vibration control by the designed CLCD.

Acknowledgements

The authors would like to gratefully acknowledge the financial support provided by the
Department of Science and Technology, Government of India, under grant number
SR/FTP/ETA-12/2005, in carrying out the research work.

References
Aspel, R.J. and Luco, J.E. (1987) ‘Impedance functions for foundations embedded in a layered
medium: an integral equation approach’, Earthquake Engg. Struct. Dyn., Vol. 15, pp.213–231.
Balendra, T., Wang, C.M. and Cheong, H.F. (1995) ‘Effectiveness of tuned liquid column dampers
for vibration control of towers’, Engg. Strut., Vol. 17, No. 9, pp.668–675.
Balendra, T., Wang, C.M. and Rakesh, G. (1999) ‘Vibration control of various types of buildings
using TLCD’, Journal of Wind Engineering and Industrial Aerodynamics, Vol. 83, Nos. 1–3,
pp.197–208.
Chopra, A.K. and Gutierrez, J.A. (1974) ‘Earthquake response analysis of multi-storey buildings
including foundation interaction’, Earthquake Engg. Struct. Dyn., Vol. 3, pp.65–77.
Dey, A. and Gupta, V.K. (1999) ‘Stochastic seismic response of multiply-supported secondary
system in flexible-base structures’, Earthquake Engg. Struct. Dyn., Vol. 26, pp.351–369.
Ghosh, A. and Basu, B. (2004) ‘Seismic vibration control of short period structures using the liquid
column damper’, Engg. Struct., Vol. 26, pp.1905–1913.
Ghosh, A. and Basu, B. (2005) ‘Effects of soil interaction on the performance of the liquid column
damper for seismic applications’, Earthquake Engg. Struct. Dyn., Vol. 34, pp.1375–1389.
Ghosh, A. and Basu, B. (2008) ‘Seismic vibration control of nonlinear structures using the liquid
column damper’, J. Struct. Eng., Vol. 134, No. 1, pp.146–153.
Ghosh, R.K. and Ghosh, A.D. (2008) ‘Soil interaction effects on the performance of compliant
liquid column damper for seismic vibration control of short period structures’, Structural
Engineering and Mechanics, Vol. 28, No. 1, pp.89–106.
Gupta, V.K. and Trifunac, M.D. (1991) ‘Seismic response of multi-storey building: the effects of
the soil-structure-interaction’, Soil Dyn. and Earthquake Engg., Vol. 10, No. 8, pp.414–422.
Haroun, M.A., Pires, J.A. and Won, A.Y.J. (1994) ‘Hybrid liquid column dampers for suppression
of environmentally-induced vibrations in tall buildings’, Proc., 3rd Conf. Tall Buildings in
Seismic Regions, Los Angeles, California.
Lee, H.H., Wong, S-H. and Lee, R-S. (2006) ‘Response mitigation on the offshore floating
platform system with tuned liquid column damper’, Ocean Engineering, Vol. 33,
pp.1118–1142.
350 R.K. Ghosh and A. Ghosh

Min, K.W., Kim, H.S., Lee, S.H., Kim, H. and Ahn, S.K. (2005) ‘Performance evaluation of tuned
liquid column dampers for response control of a 76-storey benchmark building’, Engineering
Structures, Vol. 27, pp.1101–1112.
Newland, D.E. (1993) ‘An introduction to random vibrations, spectral and wavelet analysis’, p.73,
Longman.
Reiterer, M. and Ziegler, F. (2005) ‘Bi-axial seismic activation of civil engineering structures
equipped with tuned liquid column dampers’, JSEE, Vol. 7, No. 1, pp.45–60.
Sakai, F., Takaeda, S. and Tamaki, T. (1991) ‘Tuned liquid column dampers (TLCD) for
cable-stayed bridges’, Proc. Specialty Conf. Innovation in Cable-Stayed Bridges, JSCE,
Fukuoka, Japan, pp.197–205.
Saoka, Y., Sakai, F., Takaeda, S. and Tamaki, T. (1988) ‘On the suppression of vibrations by tuned
liquid column dampers’, Annual Meeting of JSCE, IJSCE, Tokyo.
Shum, K.M. and Xu, Y.L. (2004) ‘Multiple tuned liquid column dampers for reducing coupled
lateral and torsional vibration of structures’, Engineering Structures, Vol. 26, pp.745–758.
Sun, K. (1994) ‘Earthquake responses of buildings with liquid column dampers’, Proc., 5th US
National Conf. on Earthq. Eng., Chicago, Vol. 2, pp.411–420.
Takewaki, I. (2000) ‘Soil-structure random response reduction via TMD-VD simultaneous use’,
Computer Methods in Applied Mechanics and Engineering, Vol. 190, pp.677–690.
Veletsos, A.S. and Wei, Y.T. (1971) ‘Lateral and rocking vibration of footings’, Journal of Soil
Mechanics (ASCE), Vol. 97, pp.1227–1248.
Wang, J-F. and Lin, C-C. (2005) ‘Seismic performance of multiple tuned mass dampers for
soil-irregular building interaction systems’, International Journal of Solid and Structure,
Vol. 42 No. 20, pp.5536–5554.
Won, A.Y.J., Pires, J.A. and Haroun, M.A. (1996) ‘Stochastic seismic evaluation of tuned liquid
column dampers’, Earthquake Engg. and Struct. Dyn., Vol. 25, pp.1259–1274.
Wong, H.L. and Luco, J.E. (1978) ‘Tables of impedance functions and input motions for
rectangular foundations’, Ref. CE 78-15, University of Southern California, Los Angles, CA.
Wu, J.C., Shih, M.H., Lin, Y.Y. and Shen, Y.C. (2005) ‘Design guidelines for tuned liquid column
damper for structures responding to wind’, Engineering Structures, Vol. 27, No. 13,
pp.1893–1905.
Wu, W-H. and Smith, H.A. (1995) ‘Efficient model analysis for structures with soil-structure
interaction’, Earthquake Engineering Engg. DynamicsDyn., Vol. 24, pp.283–299.
Xu, Y.L. and Kwok, K.C.S. (1992) ‘Wind-induced response of soil-structure-damper system’, 8th
International Conference on Wind Engineering, Canada, pp.2057–2068.
Xu, Y.L., Samali, B. and Kwok, K.C.S. (1992) ‘Control of along-wind response of the structures by
mass and liquid dampers’, J. of Engg. Mechanics, ASCE, Vol. 118, No. 1, pp.20–39.

Appendix

Transfer functions, χ ' s


The transfer functions for the interaction displacement and rotation as in
equations (22) and (23) may be expressed as
1
(1)
χ ZZ (ω ) = ⎡ DHTH (ω )( KVM (ω ) − ω 2 DHT (ω )) − DT (ω )( K MM (ω ) − ω 2 ST (ω )) ⎤ (30)
Δ (ω ) ⎣ ⎦

1
(2)
χ ZZ (ω ) = ⎡ PK (ω )( KVM (ω ) − ω 2 DHT (ω )) − GK (ω )( K MM (ω ) − ω 2 ST (ω )) ⎤ (31)
Δ (ω ) ⎣ ⎦
Passive control of seismic response of soil-structure system 351

1 ⎡
χθ(1)Z (ω ) = DT (ω )( KVM (ω ) − ω 2 DHTH (ω )) − DHTH (ω )( KVV (ω ) − ω 2 DT (ω )) ⎤⎦ (32)
Δ (ω ) ⎣

(1) 1
⎡GK (ω )( K MV (ω ) − ω 2 DHTH (ω )) − PK (ω )( KVV (ω ) − ω 2 DT (ω )) ⎤ (33)
χ ZZ (ω ) =
Δ (ω ) ⎣ ⎦

The notations used in equations (30)–(33) are given as

Δ (ω ) = ⎡⎣ KVV (ω ) − ω 2 DT (ω ) ⎤⎦ ⎡⎣ K MM (ω ) − ω 2 ST (ω ) ⎤⎦
(34)
− ⎡⎣ K MV (ω ) − ω 2 DHTH (ω ) ⎤⎦ ⎡⎣ KVM (ω ) − ω 2 DHT (ω ) ⎤⎦

n
DHTH (ω ) = M HT + ω 2 ∑Z
r =1
'
Hr (ω ) H r (ω ) + hn R (ω ) (35)

n
DHT (ω ) = M HT + ω 2 ∑Zr =1
'
r (ω ) H r (ω ) + hn R(ω ) (36)

n
DT (ω ) = M T + ω 2 ∑M r =1
'
r (ω ) H r (ω ) + R(ω ) (37)

n
ST (ω ) = IT + ω 2 ∑I
r =1
'
r (ω ) H r (ω ) + hn2 R(ω ) (38)

⎧⎪ n ⎫⎪

⎩⎪ r =1

PK (ω ) = ⎨hn + ω 2 γ r H r (ω )ϕ n( r ) ⎬ ω 2 R(ω )
⎭⎪
(39)

⎧⎪ n ⎫⎪

⎩⎪ r =1

GK (ω ) = ⎨1 + ω 2 α r H r (ω )ϕ n( r ) ⎬ ω 2 R (ω )
⎭⎪
(40)

where
n
Z r' (ω ) = ∑M ϕ
j =1
j
(r )
j (γ r + hnϕ n( r ) R (ω )) (41)

n
'
Z Hr (ω ) = ∑M ϕ
j =1
j
(r )
j h j (α r + ϕ n( r ) R (ω )) (42)

H D (ω )(1 + β (ω ))
R (ω ) = ( K D + iω CD ) (43)
1 − ω 2 β (ω ) H 2 (ω )

with
352 R.K. Ghosh and A. Ghosh

α 2ω 2
β (ω ) = H L (ω ) (44)
1+τ
n
M r' (ω ) = ∑M ϕ
j =1
j
(r )
j (α r + ϕ n( r ) R (ω )) (45)

n
I r' (ω ) = ∑M ϕ
j =1
j
(r )
j h j (γ r + hnϕ n( r ) R (ω )) (46)

You might also like