Vous êtes sur la page 1sur 14

Engineering Structures 23 (2001) 10541067

www.elsevier.com/locate/engstruct

Control of wind-excited towers by active tuned liquid column


damper
T. Balendra
a

a,*

, C.M. Wang a, N. Yan

Department of Civil Engineering, National University of Singapore, 10 Kent Ridge Crescent, Singapore 119260
b
79 Anson Road, Unit 20-01, Singapore 079906
Received 12 May 2000; received in revised form 25 January 2001; accepted 31 January 2001

Abstract
An active control system involving a tuned liquid column damper (TLCD) is developed for the vibration control of a singledegree-of-freedom tower subjected to wind excitation. This new control system will be referred to as the active tuned liquid column
damper (ATLCD) system. It is found that the wind-induced response of the tower can be reduced substantially by the ATLCD
system. The most efficient type of sensor is found to be the acceleration sensor when compared with displacement- and velocitytype sensors. To assess the effectiveness of such a control system, a comparison study is made between the ATLCD and the active
passive tuned mass damper (APTMD). In seeking the optimal control parameters for both systems, the norm of the root-meansquare of the system output under actual wind excitation, Hrms, is adopted as the objective function instead of the commonly used
H2 (i.e., the norm of the system output under a unit white noise). The results show that the proposed ATLCD system is a good
alternative to the APTMD system. Considering the potential advantages of the liquid column damper, the ATLCD system may be
a better choice for vibration control of wind-induced towers. 2001 Elsevier Science Ltd. All rights reserved.
Keywords: Active tuned liquid column damper; Wind-excited vibration; Towers; Optimal control technique

1. Introduction
Modern tall buildings and towers are relatively light
and slender, and possess low natural damping characteristics. Thus, they are prone to excessive wind-induced
oscillations in severe wind environments. The oscillations may cause failure from the viewpoint of serviceability. It is therefore important to design an effective and economical means of suppressing structural
vibrations induced by wind loading.
To date, many auxiliary damping control systems have
been proposed to suppress excessive structural
vibrations. The most commonly used control systems are
the passive control system and the active control system.
In the passive control system, tuned mass dampers
(TMDs) have been found to be effective in reducing the
wind-induced response of tall buildings and towers (see,
e.g., [1,2]). A newly proposed passive device is the tuned
liquid damper (TLD), which relies on the motion of shal-

* Corresponding author.

low liquid inside a rigid tank for changing the dynamic


characteristics of a structure and dissipating its vibration
energy [3]. Sakai et al. [4] suggested the use of a tuned
liquid column damper (TLCD), which suppresses the
wind-induced motion by dissipating the energy through
the motion of the liquid mass in a tube-like container
fitted with orifices. Balendra et al. [5] investigated the
effectiveness of such a device for vibration control of
towers subjected to along-wind excitation. They determined the optimum parameters of the TLCD for maximum
suppression of wind-induced acceleration and deformation for a wide range of tower dimensions. Such a
TLD or TLCD has several potential advantages which
include low cost, easy installation, simple maintenance
requirements, non-restriction to unidirectional excitation
and multiple uses, e.g., containers may be used as water
tanks for drinking and fire fighting.
Although the passive damper system offers several
benefits, its effectiveness is limited. To overcome the
limitation of the passive damper system, active tuned
mass damper (ATMD) devices have been developed.
These devices require a control algorithm that analyzes

0141-0296/01/$ - see front matter 2001 Elsevier Science Ltd. All rights reserved.
PII: S 0 1 4 1 - 0 2 9 6 ( 0 1 ) 0 0 0 1 5 - 3

T. Balendra et al. / Engineering Structures 23 (2001) 10541067

Nomenclature
A
parameter affecting energy distribution on lower frequency range
frontal area of tower
A0
APTMD activepassive tuned mass damper
ATLCD active tuned liquid column damper
ATMD active tuned mass damper
B
width of tube
breadth of tower
B0
drag coefficient, which is taken to be equal to 1.2
C0
CD, CL damping of platform and liquid in the ATLCD system
CP, CA damping of passive TMD and active TMD, respectively
damping of tower
CS
D
depth of tower
f(t)
along-wind force
frequency of vortex shedding
fS
H
transfer function matrix
norm of root-mean-square of the system output under unit white noise
H2
Hrms norm of root-mean-square of the system output under actual wind load
turbulence intensity
Iu
stiffness of the platform in the ATLCD system
KD
KS, KP, KA stiffness of tower, passive TMD and active TMD, respectively
proportional constant between the control force and the movement of the hydraulic piston
Kt
ground surface drag coefficient, which is taken to be equal to 0.03
k0
wavelength, which is taken to be equal to 1200 m
lx
MD, ML mass of platform and liquid in the ATLCD system
MS, MP, MA mass of tower, passive TMD and active TMD, respectively
frequency of tower
nS
feedback gain of the transducer in the servomechanism
1/R0
collective loop gain of the electrohydraulic servomechanism
R1
S
Strouhal number
proportional constant between the sensed structure response and the output voltage from the sensor
S0
wind power spectral density matrix
SF
power spectral density of wind force
SFF
SxS, SxD, Sy power spectral density of tower, platform and liquid of the ATLCD system, respectively
SxP, SxA power spectral density of passive TMD and active TMD of the APTMD system, respectively
TLCD tuned liquid column damper
TMD tuned mass damper
displacement response of tower in frequency domain
XS
XP, XA relative displacement response of passive TMD and active TMD in frequency domain, respectively
XD, XL relative displacement response of platform and displacement of liquid in frequency domain,
respectively
displacement response of tower in time domain
xS
xP, xA relative displacement response of passive TMD and active TMD in time domain, respectively
xD, xL relative displacement response of platform and displacement of liquid in time domain, respectively
mean wind speed at the reference height, 10 m
U10
mean wind speed at height z
Uz
U(w) control force in frequency domain
u(t)
control force in time domain
z
height of tower
a
widthlength ratio in the ATLCD system
b
normalized frequency ratio
d
head loss
e
normalized loop gain

1055

1056

T. Balendra et al. / Engineering Structures 23 (2001) 10541067

zS, zP, zA damping factor of tower, passive TMD and active TMD, respectively
zD, zL damping factor of platform and liquid, respectively
mP, mA mass ratio of passive TMD and active TMD over tower, respectively
mD, mL mass ratio of platform and liquid over tower, respectively
density of air, which is taken to be equal to 1.2 kg/m3
r0
sF
standard deviation of along-wind force
standard deviation of lift coefficient
sL
sx S
standard deviation of acceleration of tower in the ATLCD and APTMD system, respectively
sxP, sxA standard deviation of displacement of the passive TMD and active TMD, respectively
standard deviation of displacement response of platform in the ATLCD system
sxD
standard deviation response of control force
su
standard deviation response of structural acceleration with damper control system normalized by that
s x rS
without any control system
standard deviation response of dampers displacements with control system normalized by the
s xrD
maximum permissible stroke
standard deviation response of liquids displacement with control system normalized by the maximum
s yr
permissible motion of liquid
standard deviation response of control force normalized by the wind loading
s ur
t
normalized feedback gain
c
damper mass ratio in the APTMD system
P, A frequency ratio of passive TMD and active TMD over tower, respectively
D, L frequency ratio of platform and liquid over tower, respectively
w
circular frequency of along-wind excitation
wS, wP, wA circular frequency of tower, passive TMD and active TMD, respectively
wD, wL circular frequency of platform and liquid, respectively

the dynamic structural response and produces a feedback


control force which drives a mass to counter the
response. Extensive investigations have shown that use
of such a control system can reduce structural vibration
further over that of the passive control system. For
example, Kwok and Samali [6] found that the effectiveness of a TMD in suppressing wind-induced tall building
motion could be enhanced by about 33% by the addition
of an active control force. Suhardjo et al. [7] proposed
a frequency-domain approach in the optimal control of
wind-excited buildings. The control parameters are
obtained by minimizing the H2 of the transfer function
from the external disturbance to the regulated output. To
illustrate the methodology and to show the reduction in
the acceleration response of the building, a 60-storey
building under an along-wind excitation was analyzed.
The results showed that the proposed approach was flexible and useful. Xu [8] carried out a parametric study
for selecting the design parameters of active mass dampers for wind-excited tall buildings. Recently, Nishimura
et al. [9] proposed a hybrid control system by attaching
an active TMD to a passive TMD. Such a composite
activepassive tuned mass damper (APTMD) system is
found to have a similar control effect to the ATMD system, but involving a smaller active mass.
Although a TLCD operates on the same basic principles as a TMD, some of the drawbacks of a TMD are

not present in a TLCD. Due to the simple physical concepts on which the restoring force is provided in a
TLCD, no activation mechanism is required. Therefore,
maintenance cost is minimized. A TLCD is a flexible
system in meeting architectural demands since it can be
fitted easily in a variety of ways. Due to simplicity of
installation, it may be used in existing structures that
often have severe space constraints, even for temporary
use. Furthermore, a TLCD provides an excellent reservoir of water high up in the building for fire suppression.
Therefore, to make use of several potential advantages
of a TLCD, in this paper, an active tuned liquid column
damper system (ATLCD) is proposed. In the ATLCD
device shown in Fig. 1(a), the TLCD is fixed on a movable platform at the top of the tower. The movement of
the platform is controlled by a spring and a dashpot
when it is driven by a control force. A servo-actuator is
used to generate the control force based on the feedback
from the sensor attached to the top of the tower, which is
modelled as a single-degree-of-freedom system. Before
assessing the effectiveness of ATLCD, the most beneficial sensor type is first selected among the available
displacement-, velocity- and acceleration-type sensors.
Because the liquid in the tube oscillates in sympathy
with the rigid-body motion of the tube under a control
force, the proposed system may be viewed as a hybrid
control system similar to the APTMD. Therefore, the

T. Balendra et al. / Engineering Structures 23 (2001) 10541067

1057

Fig. 1. Damper control system fitted on a single-degree-of-freedom tower: (a) active tuned liquid damper (ATLCD) system, (b) activepassive
tuned mass damper (APTMD) system.

optimum performance of the ATLCD system is compared with that of the APTMD system by using a frequency-based control method. The optimum control
parameters for both ATLCD and APTMD systems are
achieved by minimizing the root-mean-square (rms)
norm of the system output when the wind spectral densities corresponding to along-wind and across-wind excitation have been adopted as the input excitation.

2. Governing equations for control system


Consider a tower assumed as a single-degree-of-freedom structure subjected to wind force f(t). The structural
displacement xS(t) is to be controlled by two types of
damping device, namely an active tuned liquid column
damper (ATLCD) [Fig. 1(a)] and a hybrid activepassive
tuned mass damper (APTMD) [Fig. 1(b)].
2.1. TowerATLCD system

1
rALy rAdyy2rAgyrAB(xSxD),
2

where MS, KS, CS are the mass, stiffness and damping


of the tower, respectively; MD is the total mass of the
ATLCD system i.e., mass of the platform and mass
of the liquid damper (including the liquid mass as well
as the weight of the tank); KD, CD are respectively the
stiffness and damping of the platform; r is the density
of the liquid; A is the cross-sectional area of the tube;
B is the width of the tube; L is the length of the liquid
measured along the centreline of the tube; d is the coefficient of head loss governed by the opening ratio of the
orifice; g is gravitational acceleration; u(t) is the control
force; xD is the relative displacement responses of the
active damper; and y is the displacement of the liquid.
The superdot denotes differentiation with respect to time.
Eq. (3) is non-linear due to the liquid damping. However, the equation may be linearized using the equivalent
linearization technique [10]. Thus, Eqs. (2) and (3) can
be rewritten as
MD(xSxD)aMLyCDxDKDxDu(t)0

For the towerATLCD system shown in Fig. 1(a), the


governing equations are given by

and

MSx SCSxSKSxSCDxDKDxDu(t)f(t)

aML(xSxD)MLyCLyMLw2Ly0,

(1)

and
MD(xSxD)rAByCDxDKDxDu(t)0,

(3)

(4)
(5)

where
(2)

while the equation of motion of the liquid column is


given by [4]

MLrAL,

(6a)

B
a ,
L

(6b)

1058

T. Balendra et al. / Engineering Structures 23 (2001) 10541067

wL2g/L

(6c)

and
CL

p L s ,
2MLd

(6d)

sy being the standard deviation of the liquid velocity.


Since wind loads on a building are often modeled as
stochastic processes that are stationary in time and nonhomogeneous in space, the spectral method is used to
determine the response. In view of the towerATLCD
system being linear, Eqs. (1, 3) and (4) may be cast into
the form given by


xS(t)

xD(t)

y(t)

XS(w)

f(t)

XD(w)

Y(w)

h11 h12 h13


h21 h22 h23
h31 h32 h33

eiwt,

(7)

where XS(w), XD(w), Y(w) and F0 are the Fourier transforms of xS(t), xD(t), y(t) and f(t), respectively, and w is
the frequency of the wind excitation.
For the control force u(t), Roorda [11] has established
the following relationship between the control force
U(w) and the controlled tip displacement of the tower
XS(w) in the frequency domain
te
X (w),
U(w)Kt(ib)
e+ib S
r

(8)

where

r 1 for the velocity-type sensor

XD(w) 0 ,

Y(w)

(10)

h111b2g1(2zSg2)(ib),

(11a)

h12mD2D2mDDzD(ib),

(11b)

h130,

(11c)

h21g1b2mDg2(ib),

(11d)

h22(2Db2)mD2mDDzD(ib),

(11e)

h23ab2mL,

(11f)

h31ab2mL,

(11g)

h32ab2mL,

(11h)

h33(2Lb2)mL2mLLzL(ib)

(11i)

(9a)

MD
mD ,
MS

(12a)

ML
,
MS

(12b)

wD
,
wS

(12c)

wL
L ,
wS

(12d)

CS
,
zS
2MSwS

(12e)

CD
,
zD
2MDwD

(12f)

mL

2 for the acceleration-type sensor

n
b ,
nS

(9b)

R1
e
2pnS

(9c)

and
(2pnS)rS0
.
R0

and

0 for the displacement-type sensor


XS(w)

where

F0

servomechanism; R1 the collective loop gain of the electrohydraulic servomechanism; and S0 the proportional
constant between the sensed structural response and the
output voltage from the sensor.
In view of Eqs. (7)(9), the governing equations (1)
and (4), (5) can be expressed as

(9d)

In Eqs. (8) and (9), the parameter Kt represents a proportionality constant between the control force and the
movement of the hydraulic piston; the parameters e and
t are the normalized loop gain and feedback gain,
respectively; b is the normalized excitation frequency; n
the along-wind excitation frequency; nS the frequency of
tower; 1/R0 the feedback gain of the transducer in the

CL
zL

2MLwL

2p s
1 ad
9

(12g)

and
F0
.
F
MSw2S

(12h)

T. Balendra et al. / Engineering Structures 23 (2001) 10541067

The expressions of g1 and g2 are given in Table 1.


By solving Eq. (10), one obtains the following transfer functions
XS(w) c1
,
HS(ib)
F
c2

(13)

XD(w) 1 h11c1
HD(ib)

F
h12 h12c2
and

(14)

(15)

where

(21a)

U
r 10,
U
Uz

(21b)

UzU10

z
10

0.3

(21c)

wSlx
,
2p4U10

(21d)

F0p4 F0p42L
F

MSw2SB 2gaMS

(21e)

and

h23h32h22h33
c 1
h12

(16a)

and
c2h21h33h23h31

h11(h23h32h22h33)
.
h12

(16b)

The displacement power spectra of the tower, the platform and the liquid can be expressed, respectively, as
SxSHS(ib)2SFF(b),

(17)

SxDHD(ib)2SFF(b)

(18)

and
SyHL(ib)2SFF(b),

(19)

where SFF(b) is the non-dimensional power spectral density of the along-wind excitation or the across-wind excitation.
For the along-wind force, the two-sided Harris spectrum is adopted. This non-dimensional power spectral
density S lFF is given by [5]
S lFF(b)

2r p4yb
U
f
,
[2+(p3yb)2]5/6

h22
Y
h11h22 h21 c1
HL(ib)

,
F
h12h23 h12h23 h23 c2

1059

4k0F 2
f,
b

(20)

in which

F0r0A0C0U2z .

k0 is the ground surface drag coefficient, lx the wavelength, z the height of the tower, wS the circular frequency of the tower, U10 the mean wind speed at the
reference height of 10 m, r0 the density of air, A0 the
frontal area of the tower, C0 the drag coefficient and Uz
the mean wind speed at height z.
For the across-wind excitation, a new empirical formula for the across-wind spectra proposed by Choi and
Kanda [12] has been adopted for the present study. The
spectral form can be decomposed into two parts, namely
a narrow spectral peak due to the regular vortex shedding
and a wideband spectral distribution caused by vorticity
in the separated shear layer and the shear layertrailing
edge direct interaction. The non-dimensional spectra
density S cFF of the across-wind force can be written as
follows:

1 ln f+0.5d20

2
d0

g1

g2

Displacement
sensor

Kt te2
KS e2+b2

Kt te

KS e2+b2

Velocity sensor

Kt teb
KS e2+b2

Kt te
KS e2+b2

Acceleration sensor

Kt te2b2

KS e2+b2

Kt teb2
KS e2+b2

(22)

where

B1

Type of sensor

F 2y s2L
(1B1)
(f/k)
B1AC

exp
S cFF(b)
A 5/A

b
[1+(f/k) ]
2pd0

AC

Table 1
Expressions of g1 and g2 for different sensors

(21f)

A(5.0/A)
,
(4.0/A)(1.0/A)


Iu

k1.58

d I
0

(23a)

z
3D
1
3
5B0

z
3

d01.58

for D/B03.0,

B0
D

k1.58

6 0.8

(23c)

for D/B 1.0,


0

B I 3
D

(23b)

(23d)
(23e)

for D/B01.0,

(23f)

1060

T. Balendra et al. / Engineering Structures 23 (2001) 10541067

It is noted that the power spectra SxS, SxD and Sy are


functions of s y . Thus, Eq. (28) has to be evaluated first
using an iterative procedure. The false position method
is used for solving the transcendental equation (28) for
s y .

F0
F y
,
MSw2Sz

(23g)

1
F0 r0A0U2z ,
2

(23h)

fS n
fS

(23i)

2.2. TowerAPTMD system

(23j)

The effectiveness of the ATLCD system will be


assessed by comparing it with another hybrid APTMD
control system. The governing equations of the tower
APTMD system shown in Fig. 1(b) are given by [13]

and
SUz
fS .
B0

In Eqs. (22) and (23), sL represents the standard deviation of the lift coefficient; () the gamma function; fS
the frequency of vortex shedding; S the Strouhal number;
Iu the turbulence intensity; and B0 and D are the breadth
and depth of the tower block, respectively. The parameter A affects the energy distribution in the lower frequency range. Here, A is assumed to be 2.0 and thus the
value of AC becomes 1.26.
Similarly, the power spectrum of the control force is
described as [8]

te
S (b).
Su Kt(ib)
e+ib xS
r

(24)

The root-mean-square responses are determined from

s x S

sxSp4
Bw2S

1 4

b SxS db
p

1/2

(25)

s xD

sxDp4
B

1/2

S db
p xD

(26)

s y

syp
1

S db
B
p y

1/2

(27)

MSx SCSx SKSxSCPx PKPxPf(t),

(30)

MP(x Sx P)CPx PKPxPCAx AKxAu(t)0

(31)

and
MA(x Sx Px A)CAx AKxAu(t)0,

where MP, MA are the masses of the passive and active


tuned mass dampers, respectively; KP, KA the stiffnesses
of the passive and active tuned mass dampers, respectively; CP, CA are the damping coefficients of the passive
and active tuned mass dampers, respectively; xP, xA the
relative displacements of the passive and active mass
dampers, respectively. The control force and wind load
are the same as for the towerATLCD system.
As in the case of the towerATLCD system, Eqs.
(30)(32) in the frequency domain may be expressed as
Eq. (10) with XD(w) and Y(w) being replaced by XP(w)
and XA(w), but with
h111b22zS(ib),

(33a)

h12mP2P2mPPzP(ib),

(33b)

h130,

(33c)

h21g1b2mPg2(ib),

(33d)

h22(2Pb2)mP2mPPzP(ib),

(33e)

h23 m 2mAAzA(ib),

(33f)

h31g1b2mAg2(ib),

(33g)

h32b mA,

(33h)

h33(2Ab2)mA2mAAzA(ib),

(33i)

2
A A

and

1/2

syp4
1 2
s y

b Sy db
BwS
p

(28)

To ascertain the level of control force required to


reduce the tower motion to a particular level, when
designing an active damper system, the normalized standard deviation of the control force is defined as [8]

1/2

Su db

su

sF

SFF db

(29)

(32)

and
MP
mP ,
MS

(34a)

MA
mA ,
MS

(34b)

wP
,
wS

(34c)

wA
,
wS

(34d)

T. Balendra et al. / Engineering Structures 23 (2001) 10541067

1061

zS

CS
,
2MSwS

(34e)

zP

CP
,
2MPwP

tem when the input is defined by a wind spectral density


matrix SF(b), the output spectral density matrix Sx(b) of
the system is given by

(34f)

Sx(b)H(b)SF(b)H(b),

(34g)

where H(b) is the transfer function matrix and H(b) is


the complex conjugate transpose of H(b), which can be
written as

CA
zA
2MAwA

(41)

and
F

F0
.
MSw2S

(34h)

The expressions of g1 and g2 are the same as those given


in Table 1. The transfer functions are the same as Eqs.
(13)(15), but with HD(ib), HL(ib), XD and Y being
replaced by HP(ib), HA(ib), XP and XA, respectively.
The displacement power spectra of the tower, the
passive and active tuned mass dampers can be expressed,
respectively, as
SxSHS(ib)2SFF(b),

(35)

SxPHP(ib)2SFF(b)

(36)

and

h11 h12 h13

H(b) h21 h22 h23


h31 h32 h33

(42)

The expressions hij (i,j=1, 2, 3) for the ATLCD are given


by Eqs. (11a, 11b, 11c, 11d, 11e, 11f, 11g, 11h) and
(11i) and for the APTMD by Eqs. (33a, 33b, 33c, 33d,
33e, 33f, 33g, 33h) and (33i). The matrix SF(b) for wind
excitation is


SFF 0 0

SxAHA(ib)2SFF(b).

(37)

SF 0
0

The expression of the power spectrum of the control


force is the same as Eq. (24), but Eq. (35) should be
used to replace Eq. (17) for the term SxS.
The root-mean-square responses are then determined
from


sx Sp4

s x S

Bw

2
S

sxPp
B

1
S db
p xP

(38)

1/2

(39)

and

s xA

sxAp4
B

S db
p xA

0 0

where SFF is the power spectral density of either the


along-wind or the across-wind excitation.
In order to minimize the output, an appropriate objective function would be [14]

1
Hrms
2p

1/2

trace[H(b)SF(b)H (b)] db

s xP

(43)

1/2

1 4

b SxS db
p

0 0 ,

(40)

The expression of the normalized standard deviation


of the control force is the same as Eq. (29).
3. Hrms control technique
For a general control system, if the system is
described as an arbitrary multi-degree-of-freedom sys-

(44)

where the rms values of the diagonal terms in Eq. (41)


are added with equal weight.
For the special case where the input is a unit intensity
white-noise excitation, i.e., SF(b)=I, where I is the identity matrix, Eq. (44) becomes

1/2

1
H2
2p

1/2

trace[H(b)H (b)] db

(45)

Since H2 optimal control is known to be equivalent to


conventional linear quadratic Gaussian (LQG) optimal
control in the time domain, the H2 method was thus
adopted by many researchers (e.g., [7,15]). In order to
introduce the external disturbance, such as wind loads,
a wind fluctuation model for which the input is a vector
of unit white noise and the output is a vector process
with spectral density matrix is constructed first. However, the process of construction is quite complex

1062

T. Balendra et al. / Engineering Structures 23 (2001) 10541067

because realization of the transfer function of wind needs


to be expressed as the form in a state space representation, so as to be consistent with the form of the governing equations of the structure. Moreover, a series of
transformations and inversions are required. In view of
this, the power spectral density of wind is introduced
directly in this study. Thus, Eq. (44) instead of Eq. (45)
becomes the objective function and is minimized. The
optimal control parameters and structural parameters can
be determined using any standard optimization algorithm. The Generalized Reduced Gradient (GRG) algorithm [1618] is used for this purpose.

4. Numerical examples
We first seek the most appropriate type of sensor
among the displacement, velocity and acceleration sensors. Next, the optimum performance of the ATLCD is
compared with that of the APTMD to show the effectiveness of the former system.
4.1. Tower and damper parameters
The towers studied by Xu [8] and Holmes [19] are
considered here. Their characteristics are listed in
Table 2.
The towers considered here are limited to square section with a fixed aspect ratio D:B0=1:1. The behavior of
the towers subjected to wind loads has been studied by
the authors when they are installed: (1) without any
damper, (2) with optimum TMD, (3) with optimum
ATMD, and (4) with optimum APTMD. The results can
be found in the papers by Wang et al. [17] and Yan et
al. [13].
Based on the sensitivity study done by Yan et al. [13],
for active control, the structural responses have been
found to be rather insensitive to the total damperstructure mass ratios, the damperstructure frequency ratios
and the damping ratios of dampers for both along-wind
and across-wind excitation. In view of this, typical
properties of the dampers have been chosen as shown in
Table 3.

Table 2
Properties of towers
Height (m)

Mass (kg)

Fundamental Damping
frequency
factor
(Hz)

Reference

81
186
278

1.0106
1.52107
3.7107

0.49
0.20
0.16

Holmes [19]
Xu [8]
Holmes [19]

1%
1%
1%

4.2. Choice of sensors


Among displacement, velocity and acceleration sensors, it is desirable to know the type of sensor that will
provide the largest reduction in structural motion. It is
also important to monitor the displacements of the dampers so as to avoid excessive movements. For the platform, the maximum permissible value of the stroke is
taken as 1 m [6]. In the case of a liquid column damper,
the displacement of the liquid should be less than (L
B)/2 to prevent the liquid from entering the horizontal
part of the tube. For the ATLCD system under alongwind load, based on Balendra et al. [5], the head loss d
is set to 10 for a practical orifice opening ratio. The
widthlength ratio a is taken to be 0.9.
Fig. 2(a)(c) shows the effects of using displacement,
velocity and acceleration sensors, respectively, on the
structural acceleration normalized by the corresponding
response without any damper. It can be seen from Fig.
2(a) that the acceleration sensor gives the most reduction
when t8. Using control parameters e5 may lead to
a 10% reduction in the structural acceleration compared
with the case of e=2. From Fig. 2(b), it is found that
when t8, the structural acceleration with the velocity
sensor is less than that with the acceleration sensor.
However, this range is not stable as shown by sharp
peaks. The effect of the displacement sensor is plotted
in Fig. 2(c). In contrast to the acceleration sensor, the
displacement sensor fares the worst. Therefore, the
acceleration sensor is the most effective.
Fig. 3(a)(c) shows the effects of different types of
sensor on the displacement of liquid normalized by the
permissible movement of liquid. It is indicated from Fig.
3(a) and (b) that the stable range of t should be larger
than 8 for the acceleration sensor and 6 for the velocity
sensor. Comparing Fig. 3(a) and (b), it is found that,
in the desirable range, the acceleration sensor produces
somewhat less displacement of liquid. For instance,
when t=20, the displacement of liquid with the acceleration sensor is about 90% of that with the velocity
sensor. The displacement sensor is not as good because
it requires greater movement of liquid that may force the
liquid to enter the horizontal part of the tube when t40.
Hence, from the point of liquid displacement, it is also
found that the acceleration sensor is the best choice.
The effects of the acceleration, the velocity and the
displacement sensors on the displacement of the platform normalized by the maximum stroke are depicted in
Fig. 4(a)(c). It is seen that the acceleration sensor
causes the smallest displacement of the platform and the
largest desirable region where the platform moves within
the limitation.
As shown in Fig. 5(a)(c), the normalized control
forces are affected by the different types of sensor. It
is seen that, in the considerable region where t8, the
normalized control force with the acceleration sensor

T. Balendra et al. / Engineering Structures 23 (2001) 10541067

1063

Table 3
Properties of damper systems
Damper system

Properties of damper system

Remarks

ATLCD

mL=mD=0.02, D=1.0, zD=9.75%

APTMD

mP+mA=0.02, c=0.3, P=A=1.0, zP=zA=9.75%

The mass of the platform is ignored, i.e., the total mass of ATLCD is
equal to the mass of liquid
c is the damper mass ratio, which is mass ratio of the active mass
damper over the total mass dampers

Fig. 2. Effect of different types of sensor on structural acceleration normalized by corresponding response without any damper: (a) acceleration
sensor, (b) velocity sensor, (c) displacement sensor.

Fig. 3. Effect of different types of sensor on displacement of liquid normalized by corresponding response without any damper: (a) acceleration
sensor, (b) velocity sensor, (c) displacement sensor.

increases slower than that with the velocity sensor when


t increases. So less control force is required with the
acceleration sensor for a given e and t. For example,
when e=5 and t=40, the normalized control force with
the acceleration sensor is about 67% of that with the
velocity sensor. Unlike the acceleration sensor and the
velocity sensor, the displacement sensor appears ineffective.
From the above observations, it is seen that the acceleration sensor is the most effective for use in the ATLCD
system in controlling wind-excited towers. The same
observations were made with across-wind excitations

and for other cases of a=0.6, 0.7 and 0.8, but the results
have not been presented as they are similar in nature to
those shown in Figs. 25.
4.3. Comparison of optimum performance between
ATLCD and APTMD
It is to be noted that the liquid in the tube has the
freedom to oscillate in respect to the motion of the tower,
besides undergoing a rigid-body motion with the platform, under a control force. In this regard the ATLCD
system is a hybrid system, like the APTMD system,

1064

T. Balendra et al. / Engineering Structures 23 (2001) 10541067

Fig. 4. Effect of different types of sensor on displacement of platform normalized by corresponding response without any damper: (a) acceleration
sensor, (b) velocity sensor, (c) displacement sensor.

Fig. 5. Effect of different types of sensor on control force normalized by total wind force: (a) acceleration sensor, (b) velocity sensor, (c)
displacement sensor.

which has been shown to be better than the ATMD system [9,13]. Thus, to investigate the performance of the
ATLCD further, the ATLCD system will be compared
with the APTMD system. It is noted that the most beneficial sensor for the APTMD system was found to be
the acceleration-type sensor. Therefore, the accelerationtype sensor is used in the following study.
To investigate the effects of the control parameters e
and t on the performance of the ATLCD system, the
variations in structural acceleration, structural displacement, displacement of the platform, displacement of the
liquid, control force and Hrms with respect to the control parameters are plotted in Fig. 6(a)(f). Fig. 6(a)(f)
are associated with along-wind excitation. It is seen that
the structural acceleration and displacement responses
can be reduced to an acceptable range with the ATLCD.
In view of the results in Fig. 6(a)(f), it is also found that
the dynamic responses of the system are not sensitive to
the loop gain e but to the feedback gain t. The same
insensitivity of the APTMD system to the loop gain e
is also observed. Therefore, in the following study, the
loop gain is set to a fixed value (say, e=5) while the

feedback gains t for both ATLCD and APTMD are


determined so as to minimize the Hrms as given in Eq.
(44). In the optimization problem, the displacement and
the acceleration of the structure are bounded for safety
and serviceability requirements. Based on the British
Standard BS 6611, the constraint for the displacement
takes the form of a ratio of the top displacement of the
structure to the entire height, which is set to be less than
1/500 [20]. For the acceleration, the constraint is that the
maximum acceleration should be less than 20 cm/s2 [20]
so that the occupants do not feel uncomfortable. Further,
the stroke of the platform must be within the permissible
range of 1 m and the movement of the liquid cannot
exceed the limit value of (LB)/2 so as to prevent the
liquid from entering the horizontal part of the tube.
Based on the above constraints, the optimum feedback
gains of the ATLCD and the APTMD systems installed
on three different high towers are determined by minimizing Hrms using the GRG algorithm.
Table 4 lists the optimal feedback gains for the
ATLCD and the APTMD under the along-wind excitation with m=0.02 and e=5. The widthlength ratio a

T. Balendra et al. / Engineering Structures 23 (2001) 10541067

1065

Fig. 6. Variations of (a) structural acceleration, (b) structural displacements, (c) platform displacement, (d) liquid displacement, (e) control force
and (f) Hrms with ATLCD against control parameters under along-wind excitation.

for the ATLCD system is set to be 0.6. The values of


the corresponding responses for these two hybrid systems are also given in the same table. Table 5 gives the
corresponding results under across-wind excitation.
From Tables 4 and 5, it can be seen that the values of
the optimal feedback gains for the ATLCD are a little
larger than those for the APTMD for different heights
of the tower. Larger optimal feedback gains are required
for taller towers. The dynamic responses show that the
structural accelerations and displacements, the platform

(active part) displacement and movement of the liquid


level (passive part) with the ATLCD are less than their
corresponding parts in the APTMD for both along-wind
and across-wind excitations. Contrary to the case under
the along-wind excitation, where the control force with
the ATLCD is smaller than that with the APTMD, the
control force in the ATLCD is a little larger under the
across-wind excitation.
To compare the effectiveness of the ATLCD with that
of the APTMD, the ratios of each response for these two

1066

T. Balendra et al. / Engineering Structures 23 (2001) 10541067

Table 4
Optimal feedback gains and corresponding responses for along-wind excitation (e=5)
Height
(m)

Control
system

topt

sx S (cm/s2)

sxS (cm)

s xD or s xA (cm) s y or s xP (cm)

su (kN)

Min. Hrms

81

ATLCD
APTMD
ATLCD
APTMD
ATLCD
APTMD

8.65
8.16
9.10
8.67
9.84
9.35

2.67
2.92
2.48
2.70
2.18
2.45

2.59
2.67
2.38
2.58
2.21
2.35

17.70
19.15
17.74
19.61
10.05
10.93

43.06
43.50
43.07
43.49
42.97
43.85

4.27
4.68
4.24
4.57
2.39
3.27

186
278

27.67
39.51
27.36
36.89
15.22
21.13

Table 5
Optimal feedback gains and corresponding responses for across-wind excitation (e=5)
Height
(m)

Control
system

topt

sx S (cm/s2)

sxS (cm)

s xD or s xA (cm) s y or s xP (cm)

su (kN)

Min. Hrms

81

ATLCD
APTMD
ATLCD
APTMD
ATLCD
APTMD

9.23
8.72
9.40
8.94
10.7
9.78

2.68
2.93
2.51
2.75
2.18
2.44

2.36
2.68
2.24
2.43
2.16
2.25

19.75
20.84
19.63
21.30
10.02
10.90

43.03
42.90
43.18
43.00
42.98
42.81

4.32
4.68
5.14
5.39
2.58
3.35

186
278

systems are shown in Table 6. From Table 6 it is clear


that, for both along-wind and across-wind excitation, the
accelerations and displacements of the towers with the
ATLCD are 90% and 93% of that with the APTMD.
Also, the displacements of the platforms (active part)
and the liquids (passive part) in the ATLCD are smaller
by 8% and 28% than the corresponding values of active
and passive mass dampers in the APTMD, respectively.
The control forces in these two hybrid systems are more
or less same. Although the control force required by the
ATLCD is 0.4% more in the across-wind excitation, it is
still preferred over the APTMD, considering its potential
advantages. Therefore, it can be concluded that the
ATLCD is an effective hybrid control system in reducing
vibrations of towers. It can be treated as an alternative
to the APTMD. In view of its potential advantages, the
ATLCD is a better choice for wind-induced motions
of towers.

23.53
33.14
37.50
49.20
16.91
23.16

5. Conclusions
An active tuned liquid column damper system has
been proposed for the vibration control of towers under
wind excitation. The effectiveness of the ATLCD system
in suppressing wind-induced acceleration response has
been demonstrated for a single-degree-of-freedom tower
in the frequency domain. The results obtained from the
current research have shown that the proposed ATLCD
system can be used to significantly reduce the windinduced acceleration response of towers. The acceleration type of sensor has been found to be most effective
compared with displacement and velocity sensors. Since
the proposed ATLCD is a hybrid control system, its performance is compared with the APTMD, which has previously been shown to be better than the well-known
ATMD. Through the comparison of the optimum performance of the ATLCD and the APTMD, it is found

Table 6
Response ratios under along-wind and across-wind excitations when using optimum feedback gains
Excitation

Height (m)

sATLCD
x S

sATLCD
xS

sATLCD
xD

APTMD
x S

APTMD
xS

APTMD
xA

s
Along-wind

Across-wind

81
186
278
81
186
278

0.91
0.92
0.90
0.91
0.91
0.89

0.97
0.92
0.94
0.88
0.92
0.96

0.92
0.90
0.92
0.95
0.92
0.92

sATLCD
y
sAPTMD
xP

sATLCD
u
sAPTMD
u

0.70
0.74
0.72
0.71
0.76
0.73

0.99
0.99
0.98
1.003
1.004
1.004

T. Balendra et al. / Engineering Structures 23 (2001) 10541067

that the ATLCD system is an effective system and can


be used as an alternative to the APTMD for the control
of both along-wind and across-wind motion. In spite of
a little larger control force required for the across-wind
excitation for certain widthlength ratios, the ATLCD
system is still preferred because of potential advantages
of the liquid column damper, such as relatively low cost,
multiple uses, and less platform displacements as well
as smaller liquid displacements.
In light of the complexity of the across-wind response,
being dependent not only on the aspect ratio and structural dynamic properties but also on the cross-section
shape, further study will be carried out to investigate
the influence of these factors on the effectiveness of the
ATLCD system. Some structures, such as tall buildings,
cannot be accurately represented by single-degree of
freedom systems. For these structures, multi-degree-offreedom models are required. Thus, the current study
will be extended to investigate the effectiveness of the
ATLCD system on multi-degree-of-freedom systems.

References
[1] Kawaguchi A, Terammura A, Omote Y. Time history response
of a tall building with a tuned mass damper. J Wind Eng Ind
Aerodyn 1992;41-44:194960.
[2] Xu YL, Kwok KCS, Samali B. Control of wind-induced tall
building vibration by tuned mass dampers. J Wind Eng Ind Aerodyn 1992;40:132.
[3] Fujino Y, Pacheco BM, Chaiseri P, Sun LM. Parametric studies
on tuned liquid damper (TLD) using circular containers by freeoscillation experiment. J Struct Engrs/Earthquake Eng
1988;5(1):38191.
[4] Sakai F, Takaeda S, Tamaki T. Tuned liquid column damper
new type device for suppression of building vibrations. In: Proceedings of International Conference on Highrise Buildings,
Nanjing, China, 1989:92631.

1067

[5] Balendra T, Wang CM, Cheong HF. Effectiveness of tuned liquid


column dampers for vibration control of towers. Eng Struct
1995;17(9):66875.
[6] Kwok KCS, Samali B. Performance of tuned mass dampers under
wind loads. Eng Struct 1995;17(9):65567.
[7] Suhardjo J, Spencer BF Jr., Kareem A. Frequency domain optimal control of wind-excited buildings. J Eng Mech
1992;118(12):246381.
[8] Xu YL. Parametric study of active mass dampers for windexcited tall buildings. Eng Struct 1996;18(1):6476.
[9] Nishimura I, Kobori T, Sakamoto M, Yamada T, Koshika N,
Sasaki K et al. Active passive composite tuned mass damper.
In: Proceedings of Seminar on Seismic Isolation, Passive Energy
Dissipation and Active Control. San Francisco, Applied Technology Council, 171, (ATC-17-1), 1993:73748.
[10] Iwan WD, Yang I. Application of statistical linearization techniques to nonlinear multidegree of freedom system. J Appl Mech
1972;39:54550.
[11] Roorda J. Tendon control in tall structures. J Struct Div, ASCE
1975;101(ST3):50521.
[12] Choi H, Kanda J. Proposed formulae for the power spectral densities of fluctuating lift and torque on rectangular 3-D cylinders.
J Wind Eng Ind Aerodyn 1993;46-47:50716.
[13] Yan N, Wang CM, Balendra T. Composite mass dampers for
vibration control of wind-excited towers. J Sound Vibr
1998;213(2):30116.
[14] Boyd SP, Barratt CH. Linear controller design limits of performance. Englewood Cliffs (NJ): Prentice-Hall, 1991.
[15] Dyke SJ, Spencer BF Jr., Quast P, Sain MK, Kaspari DC Jr.,
Soong TT. Acceleration feedback control of MDOF structures. J
Eng Mech 1996;122(9):90717.
[16] Lasdon LS, Waren AD. Generalised reduced gradient software
for linearly and nonlinearly constrained problems. In: Greenberg
H, editor. Design and implementation of optimization software.
Sijthoff and Noordhoff, 1979.
[17] Wang CM, Yan N, Balendra T. Control on dynamic structural
response using activepassive composite tuned mass dampers. J
Vibr Control 1999;5(3):47589.
[18] Yan N, Wang CM, Balendra T. Optimal damper characteristics
of ATMD for building under wind loads. J Struct Eng ASCE
1999;125(12):137683.
[19] Holmes D. Listing of installations. Eng Struct 1995;17(9):6767.
[20] Balendra T. Vibration of buildings to wind and earthquake loads.
London: Springer-Verlag, 1993.

Vous aimerez peut-être aussi