Anda di halaman 1dari 12

Journal of Wind Engineering

and Industrial Aerodynamics 83 (1999) 197}208

Vibration control of various types of buildings


using TLCD
T. Balendra*, C.M. Wang, G. Rakesh
Center for Wind Resistant Structures, Department of Civil Engineering, National University of Singapore,
10 Kent Ridge Crescent, Singapore 119260, Singapore

Abstract

The e!ectiveness of the TLCD in reducing the along-wind response of tall buildings is
investigated. Variety of buildings with di!erent mass-sti!ness distributions (e.g. uniform, linear,
abrupt variation) modeled as shear wall and rigid frame systems are studied. A continuum
formulation is adopted, which provides response statistics along the entire height of the
structure. The performance of the TLCD is discussed with respect to the mode shapes of
buildings. A numerical example, illustrating that a second damper could greatly improve the
overall response of certain type of buildings is also presented. ( 1999 Elsevier Science Ltd. All
rights reserved.

Keywords: Vibration control; Tuned liquid column dampers; TLCD

1. Introduction

Tall and slender structures when subjected to strong winds can undergo signi"cant
vibrations which may become unacceptable from the viewpoint of serviceability and
safety. As such, the suppression of these vibrations has become an important design
consideration in recent years. Many active and passive control devices such as the
active mass dampers (AMD), active tendons (AT), tuned liquid dampers (TLD),
impact dampers (ID) and tuned liquid column dampers (TLCD) have been proposed
to mitigate these excessive vibrations.
The TLCD, "rst proposed by Sakai et al. [1] is studied here. The TLCD dissipates
the structural vibrational energy by the combined action involving the movement
of the liquid mass in the container, the restoring force on the liquid due to gravity and
the damping e!ect due to the ori"ces. The performance of the TLCD can generally be

* Corresponding author.

0167-6105/99/$ - see front matter ( 1999 Elsevier Science Ltd. All rights reserved.
PII: S 0 1 6 7 - 6 1 0 5 ( 9 9 ) 0 0 0 7 2 - 0
198 T. Balendra et al. / J. Wind Eng. Ind. Aerodyn. 83 (1999) 197}208

said to be dependent on the structural, damper and excitation characteristics [2].


Much of the previous research studies [3,4] were targeted at varying damper proper-
ties to optimise its e!ectiveness. There have also been a number of studies made on
di!erent kinds of excitation loads (e.g. wind, stationary and non-stationary earth-
quake loads). However, there are relatively fewer studies made on the damper
performance due to variation in structural characteristics.
Hence, the e!ectiveness of the TLCD in reducing the acceleration and displacement
in various types of buildings is investigated. In a recent study by Balendra et al. [5]
only uniform buildings were considered. Herein, in addition to uniform buildings,
buildings with a soft "rst storey, buildings with linearly varying properties along the
height and buildings with sudden changes in mass-sti!ness distribution are con-
sidered. These buildings are modelled as purely shear and #exural beams, representing
rigid frame and shear-wall structural systems, respectively. Unlike commonly used
lumped-mass studies, a continuum formulation is adopted. This allows response
quantities to be obtained along the entire height of the structure and also permits the
study on the e!ects of the damper position. Higher modes have been included in the
analysis. Also, the usefulness of a second damper in improving the overall perfor-
mance of certain types of buildings is investigated.

2. Analytical procedure

Consider a tall building as shown in Fig. 1, vibrating under an external excitation


force f (z, t) where z is the height along the structure, and t the time of response. The
structure is "tted with n liquid dampers. The force exerted by the jth damper on the
building, at the height z"c H is
j
F (t)"m vK (c H, t)#oA B uK , (1)
j j j j j j
where c is the position of the damper (varies from 0 to 1) on the structure, H the
j
height of the building, m the mass of the damper, A the cross-sectional area of the
j j
damper, B the width of the damper, o the density of the liquid in the damper, u"u(t)
j
the displacement of the liquid column and v"v(z, t) the horizontal displacement of
the building with the superdots representing di!erentiation with respect to time (refer
to Fig. 2).
Adopting a shear-#exure beam model for the building, the governing equation of
motion is given by
n
m (z)vK #c (z)v5 #[EI(z)vA]A![aH2EI(z)v@]@"f (z, t)! + F (t)d (z!c H)
4 4 j H j
j/1
(2)
where EI, c and m are the #exural rigidity, damping coe$cient and the mass per unit
4 4
length of the structure respectively, aH2"GA/EI where GA is the shear rigidity and
d the dirac delta function. The primes denote di!erentiation with respect to dis-
H
tance. The behavior of the rigid frame and shear-wall structures can be obtained as
special cases of the shear-#exure cantilever beam model. A value of aHH"0 would
T. Balendra et al. / J. Wind Eng. Ind. Aerodyn. 83 (1999) 197}208 199

Fig. 1. Building subjected to wind loading.

Fig. 2. Con"guration of rectangular TLCD.

correspond to a purely #exural building and the value of aHH of about 100 would
correspond to a purely shear building.
The equation proposed by Sakai et al. [1] is used to described the motion of the
water in the TLCD. This equation which has a non-linear damping term is linearised
using the equalivaent linearisation technique suggested by Iwan et al. [6]. The
linearised equation governing the liquid #ow within the jth damper is given as

m uK #m a vK (c H, t)#c uK #m u2 u "0 (3)


j j j j j j j j $j j
where a "B / , c "J(2/n)(d / )m p 5 j the damping coe$cient of the liquid, d the
j j j j j j j u j
coe$cient of the headloss governed by the opening ratio of the ori"ce, p 5 j the RMS
u
value of the liquid velocity and u "J2g/ the natural frequency of the liquid
$j j
column.
200 T. Balendra et al. / J. Wind Eng. Ind. Aerodyn. 83 (1999) 197}208

Substituting Eq. (1) into Eq. (2) leads to

n
m (z)vK # + [m vK (c H, t)#oA B uK ]d (z!c H)#c (z)v5
4 j j j j j H j 4
j/1
#[EI(z)vA]A![aH2EI(z)v@]@"f (z, t). (4)

2.1. Boundary conditions

The boundary conditions are taken to be

v(z, t)D "0


z/0 base of the structure, (5)
v@(z, t)D "0
z/0
EI(z)vA(z, t)D "0,
z/H top of the structure. (6)
[(EI(z)vA(z, t))@!aH2EIv@(z, t)]D "0,
z/H

2.2. Frequency response matrix

The structural displacement "eld is assumed to be of the form

m
v(z, t)" + / (z)q (t), (7)
i i
i/1
where m is the number of shape functions, / the shape functions which are the mode
i
shapes corresponding to the free vibration of a shear-#exure cantilever beam and
q the generalised coordinates. Eq. (4) describing the structural motion is then
i
multiplied with m shape functions and integrated over the entire structural height so
as to yield m independent equations. Depending on the number of dampers, n, used,
a total of m#n equations is "nally obtained. The frequency response method is then
used to obtain the system transfer matrix [7].

2.3. Force matrix

The along-wind force is modeled as a stochastic process, stationary in time but


non-homogeneous in space. The longitudinal turbulence spectrum proposed by
Harris and Deaves [8] has been used to simulate the random wind loading. The
double-sided Harris velocity spectrum is given as

4pK <2 x6
S (u)" 0 10 , (8)
v u [2#(x6 )2]5@6

where K is the drag coe$cient, x6 "ul /2p< , l the length constant, < is the
0 x 10 x 10
velocity at 10 m height and u the frequency in rad/s. The height wise correlation
T. Balendra et al. / J. Wind Eng. Ind. Aerodyn. 83 (1999) 197}208 201

function is given by

A B
!C uDz !z D
Coh(z , z , u)"exp z i j , (9)
i j p(< i #< j )
z z
where C is the exponential decay constant and < the velocity at any height z which
z z
is given by the power law model. The force power spectrum is then furnished by

PP
H H
S& (u)" / (z )/ (z )o2B2C2 < < S (u)Coh(z , z , u) dz dz , (10)
ij i 1 j 2 ! 4 $ z1 z2 v 1 2 1 2
0 0
where B is the projected width of the structure subjected to wind loading and C is
4 $
the surface drag coe$cient.
Having obtained the frequency response and force matrices, using the classical
spectral analysis technique, the mean square response of the structural and liquid
motion can be found. An interactive procedure was adopted to "rst determine the liquid
velocities since the response matrix is a function of the RMS liquid velocities [6].

3. Response measures

Spectral analysis was carried out and the liquid and structural response parameters
were determined. These would include the liquid displacement u, liquid velocity u5 ,
structural displacement v and structural acceleration vK . The performance of the TLCD
is quanti"ed using the following proposed ratios (in percentage):
p K !p K $
R " v v ]100, (11)
! pK
v
p !p $
R " v v ]100 (12)
$ p
v
where p K $ and p K are the RMS values of the structural accelerations with and without
v v
the damper, respectively. The acceleration reduction ratio R , given in Eq. (11)
!
indicates the reduction of the structural acceleration due to the TLCD while the
displacement reduction ration R , given in Eq. (12) indicates the reduction in the
$
structural displacement.

4. Input data

4.1. Damper parameters

When only a single damper is used, the following values for the damper parameters
have been adopted in the analysis:
m /2 (c H) u
k " 1 1 1 "0.02, a "0.9, X " $1 "1.0 (13)
1 M 1 1 u
11 11
202 T. Balendra et al. / J. Wind Eng. Ind. Aerodyn. 83 (1999) 197}208

where M ":H/ (z)m (z)/ (z)dz is the generalised mass corresponding to the "rst
11 0 1 4 1
mode of vibration and u is the fundamental structural frequency. The chosen mass
11
ratio, k can be considered as a practical upper limit while the a and X values can be
1 1 1
taken to provide almost optimum damper performance.

4.2. Excitation parameters

The following values for the wind force parameters have been used:

K "0.03, C "1.2, l "1200 m, C "8.0, < "20 m/s. (14)


0 $ x z 10

4.3. Structural parameters

The structural period maybe approximately related to the structural period by [9]

H
" . (15)
55

Unless otherwise stated, the typical values of the structural parameters adopted are

o "200 kg/m3, f "0.01, B "40 m, (16)


" i "
where o is the mass density of the building, B the base-width of the building and
" "
f the damping ratio of the structure for the i-th mode. It has been assumed that all the
i
modes have equal damping ratios.

5. Numerical study

5.1. Non-uniform buildings

Non-uniform buildings of height 200 m are considered. Table 1 shows the vari-
ations in the mass-sti!ness [10] considered. Model I represents a uniform building;
Model II a building with a soft "rst storey; Model III a regular building with linearly
varying mass and sti!ness and Model IV represents a building with a sudden
mass-sti!ness drop at the top one-"fth height of the building. The natural frequencies
as a ratio of the fundamental frequency are given in Table 2, when the buildings are
modelled as shear type.
The fundamental frequency was taken to be 0.275 Hz for all four models. The "rst
three mode shapes of the building are depicted in Fig. 3.
Fig. 4 shows the variation of the acceleration reduction ratio, R , along the height,
!
z when the mean velocity of the wind at 10 m height is taken to be 20 m/s for a City
Centre fetch. The maximum reduction occurs somewhere between 0.6 H to 0.85 H
where H is the height of the building. It can be seen that the building with uniform
setback, Model III, and penthouse, Model IV, have lower reduction ratios as
T. Balendra et al. / J. Wind Eng. Ind. Aerodyn. 83 (1999) 197}208 203

Table 1
Mass-sti!ness variation along the height of building

Variation Mass Sti!ness

Model I II III IV I II III IV

Top 1.00 1.00 1.00 0.20 1.00 1.00 1.00 0.20


1.00 1.00 1.25 1.00 1.00 1.25 1.25 1.25
1.00 1.00 1.50 1.00 1.00 1.50 1.50 1.50
1.00 1.00 1.75 1.00 1.00 1.75 1.75 1.75
Bottom 1.00 1.00 2.00 1.00 1.00 0.40 2.00 2.00

Table 2
Normalised frequencies for non-uniform building

Model type f f f f f
1 2 3 4 5
I 1.00 3.00 5.00 9.00 11.00
II 1.00 2.99 5.29 7.62 9.90
III 1.00 2.64 4.34 6.04 7.69
IV 1.00 2.66 3.49 5.07 6.92

Fig. 3. Mode shapes of non-uniform buildings.


204 T. Balendra et al. / J. Wind Eng. Ind. Aerodyn. 83 (1999) 197}208

Fig. 4. Acceleration reduction ratio for non-uniform buildings.

compared to Models I and II. This is due to the greater contribution of the higher
modes due to closer spacing of natural frequencies.

5.2. Tapered buildings

As a further investigation into the higher mode contribution, tapered buildings


which well represent buildings with uniform setbacks (Model III) were studied. The
degree of taper, j, ratio to top width to bottom width, has been varied from 0 to 1.
A ratio of 1 corresponds to a uniform building of square cross-section while a ratio of
0 would correspond to a pyramid-shaped building. The sti!ness and geometric
dimensions vary linearly with height, depending on the degree of taper. The buildings
are modelled as either purely #exural or purely shear beams. For the case of j"0, due
to lack of space the damper was placed at 0.75H instead of the top of the building. For
all the other cases,the damper was at the top.
The acceleration reduction ratios along the height, for tapered #exural and shear
buildings are shown in Fig. 5. The degree of taper has marginal in#uence on the
#exural buildings, whereas for the shear type of buildings, the acceleration reduction
for the pyramid building is only half of that of the uniform building. This is because
higher frequencies of tapered shear buildings are closely spaced compared to uni-
form shear buildings and thus the contribution of higher modes becomes important.
Thus, in order to have greater acceleration reduction in tapered shear buildings,
multiple dampers are required. Higher natural frequencies of tapered buildings for
both the shear and #exure type buildings are shown in Table 3 for various degrees of
taper.
T. Balendra et al. / J. Wind Eng. Ind. Aerodyn. 83 (1999) 197}208 205

Fig. 5. Acceleration reduction ratio for tapered buildings.

Table 3
Higher natural frequencies of tapered buildings

Building type Flexure Shear

Degree of f f f f f f f f f
1 2 3 1 2 3 4 5 6
taper, j

1.00 1.0 6.24 17.88 1.00 3.00 5.00 7.00 9.00 11.00
0.75 1.0 5.72 15.62 1.00 2.78 4.61 6.44 8.28 10.11
0.50 1.0 5.09 13.42 1.00 2.56 4.18 5.82 7.46 9.11
0.25 1.0 4.40 11.03 1.00 2.36 3.76 5.18 6.62 8.06
0.00 1.0 4.06 9.64 1.00 2.30 3.60 4.90 6.21 7.51

5.3. Application of two dampers

Analysis revealed that for a shear building with Model IV type of mass-sti!ness
distribution, the second mode contributed about 30% of the total acceleration
response. Hence, a second damper tuned to the second natural frequency of the
building was placed at the top of the building.
Fig. 6 shows the acceleration reduction ratio for various combinations of mass
ratios of the dampers when the total damper mass is kept constant. It is seen that the
206 T. Balendra et al. / J. Wind Eng. Ind. Aerodyn. 83 (1999) 197}208

Fig. 6. Di!erent combinations of mass ratio of dampers.

Fig. 7. Variation of reduction ratios with change in position of second damper.

mass ratio combination of 1.5% and 0.5% for the "rst and second dampers respective-
ly, produces better performance throughout the height of the building than a single
damper. The introduction of the second damper improves the performance of the "rst
damper by 40%, which amounts to an overall increase of 12% in acceleration
T. Balendra et al. / J. Wind Eng. Ind. Aerodyn. 83 (1999) 197}208 207

reduction. For a mass combination of 1.75% and 0.25%, the single damper
provides better reduction over the height range of 0.75H and 0.5H because the
contribution of the second mode is minimal over this height range. This can be
observed for Fig. 6.
To study the in#uence of the second damper's position, c, for the Model IV type
of mass-sti!ness variation, the position of the second damper with 0.5% mass ratio
was varied along the building height. As seen clearly from Fig. 7, the accele-
ration reduction ratio is seen to be highly sensitive to the position of the
second damper, with the best results obtained when both dampers are at the top.
Minimum reduction is obtained when the second damper is around. 0.6 H, which
is a nodal point for the second mode, as evident from Fig. 3. The displacement
reduction ratio on the other hand is insensitive to the variations in the
damper position since the higher mode contribution is insigni"cant for displacement
response.

6. Conclusion

The acceleration reduction capacity of the TLCD is highly dependent on the


structural system. For the shear building, when only the fundamental mode is
controlled the acceleration reduction for the tapered and penthouse buildings is
smaller than the uniform and soft-storey building. Greater reduction is observed for
the #exural type of building than the shear type of building. The reduction for tapered
shear building is about half of uniform shear building. The tapering has marginal
e!ect on the acceleration reduction of #exural buildings. When higher modes are
signi"cant, the second damper tuned to the higher natural frequency can greatly
improve the performance.

References

[1] F. Sakai, S. Takeda, T. Tamaki, Tuned liquid column damper } new device for suppression of building
vibration, Proceedings of the International Conference on Highrise Buildings, Nanjing, China, 1989,
pp. 936}931.
[2] T. Balendra, C.M. Wang, H.F. Cheong, E!ectiveness of tuned liquid column dampers for vibration
control of towers, J. Eng. Struct. 17 (9) (1994) 668}675.
[3] Y.L. Xu, B. Samali, K.C.S. Kwok, Control of along wind response of structures by mass and liquid
dampers, J. Eng. Mech. 118 (1) (1992) 20}39.
[4] P.A. Htichcock, K.C.S. Kwok, R.D. Watkins, B. Samali, Characteristics of liquid column vibration
absorbers (LCVA) } Parts 1 and 2, J. Eng. Struct. 19 (2) (1997) 126}145.
[5] T. Balendra, C.M. Wang, G. Rakesh, E!ectiveness of TLCD on various structural systems J. Eng.
Struct., in press.
[6] W.D. Iwan, I. Yang, Application of statistical linearisation techniques to non-linear multi-degree of
freedom systems, J. Appl. Mech. 39 (2) (1972) 545}550.
[7] G. Rakesh, Tuned liquid column damper for vibration control of tall structures, M. Eng. Thesis,
Department of Civil Engineering, 1997, National University of Singapore.
208 T. Balendra et al. / J. Wind Eng. Ind. Aerodyn. 83 (1999) 197}208

[8] R.I. Harris, H.M. Deaves, Wind engineering in the eighties, Proceedings of the CIRIA Conference,
London, 1980, pp. 4.31}4.57.
[9] S. Lagomarsino, Forecast models for damping and vibration periods of buildings, J. Wind Eng. Ind.
Aerodyn. 48 (1993) 221}239.
[10] Y. Ishiyama, Distribution of lateral seismic force along the height of the buildings, Research Paper
No. 120, 1986, Building Research Institute, Japan.

Anda mungkin juga menyukai