Sie sind auf Seite 1von 9

Probabilistic Engineering Mechanics 19 (2004) 269277

www.elsevier.com/locate/probengmech

Stochastic optimal control of wind-excited tall buildings


using semi-active MR-TLCDs
Y.Q. Nia,*, Z.G. Yingb, J.Y. Wanga, J.M. Koa, B.F. Spencer Jr.c
a

Department of Civil and Structural Engineering, The Hong Kong Polytechnic University, Hung Hom, Kowloon, Hong Kong, China
b
Department of Mechanics, Zhejiang University, Hangzhou 310027, China
c
Department of Civil and Environmental Engineering, University of Illinois at Urbana-Champaign, Urbana, IL 61801, USA
Received 15 November 2003; revised 15 December 2003; accepted 2 February 2004

Abstract
A new semi-active control device, magneto-rheological tuned liquid column damper (MR-TLCD), has been devised recently by the
authors for mitigation of wind-induced vibration response of tall building structures. The developed device combines the benefits of
magneto-rheological smart materials and tuned liquid column dampers. In this paper, real-time semi-active vibration control of tall building
structures incorporating nonlinear MR-TLCDs under random wind excitation is studied by means of the statistical linearization method and
the optimal linear quadratic (LQ) control strategy. The equations of motion of a tall building structure subjected to random wind loading and
controlled by using MR-TLCDs at the top floor are first derived and represented in modal coordinate. After linearizing the uncontrollable part
of MR-TLCD damping force and incorporating it with structural components, the classical linear quadratic (LQ) control strategy is applied to
the linearized structural system to determine optimal control force of the MR-TLCDs. Clipping treatment is performed to ensure the
commanded control force implementable by the MR-TLCDs. Wind-excited response of the semi-actively controlled structural system is
evaluated by using the frequency-response function and then compared with that of the passively controlled structure to determine the control
efficacy. A case study of a 50-story building structure is conducted to illustrate excellent control efficacy of the proposed semi-active MRTLCD control system.
q 2004 Elsevier Ltd. All rights reserved.
Keywords: Tall building structure; Random wind excitation; Semi-active magneto-rheological tuned liquid column damper; Stochastic optimal control

1. Introduction
Mitigating wind-induced vibration response of building
structures by using passive tuned liquid column dampers
(TLCDs) has been studied extensively [1 8]. The TLCD
consists of a U-tube container with an orifice in the middle.
It dissipates the energy of structural vibration by a combined
action of inertia force induced by the movement of the
liquid, the restoring force due to gravity on the liquid, and
the damping effect caused by an orifice. The TLCD has
attracted interest for engineers due to its cost-effectiveness,
simplicity in installation, and low maintenance costs. A
recent application of the TLCD is its implementation to the
46-story One Wall Centre in Vancouver [9]. Two TLCDs,
each consisting of a four-story high, 50,000-gal water tank,
* Corresponding author. Tel.: 852-2766-6004; fax: 852-2334-6389.
E-mail address: ceyqni@polyu.edu.hk (Y.Q. Ni).
0266-8920/$ - see front matter q 2004 Elsevier Ltd. All rights reserved.
doi:10.1016/j.probengmech.2004.02.010

were placed at the top of this combined hotel and residential


building for mitigating wind-excited vibration. Each TLCD
contains two water columns connected by a sluice gate
(orifice) to regulate water flow. The damping system is
tuned to the natural frequency of the building by regulating
water flow through the gates and also monitoring the water
levels in the tanks.
For a TLCD system, energy dissipation in the water
column is due to the passage of the liquid through an orifice
with inherent head-loss characteristics. However, the
damping force induced by the orifice is nonlinear. The
equivalent linearized damping coefficient is responsedependent, so that optimum damping condition cannot be
maintained for a wide range of disturbances. Thus, it is
highly desirable to develop damping-variable or parameteradjustable TLCDs to achieve optimal control performance
for a wide range of loading conditions and to be tolerant of
probable structural uncertainty. Some research efforts have

270

Y.Q. Ni et al. / Probabilistic Engineering Mechanics 19 (2004) 269277

been made to this end. Haroun et al. [10] proposed a concept


of hybrid liquid column damper by actively controlling the
orifice to produce variable orifice opening ratio; Yalla et al.
[11] proposed to convert passive TLCDs into semi-active
TLCDs by introducing a controllable valve to adjust the
orifice opening. However, no physical devices of TLCDs
with variable or controllable damping have been developed.
The authors have recently developed a practical semiactive TLCD by using the smart magneto-rheological (MR)
fluids [12]. An essential characteristic of MR fluids is their
ability to reversibly change from a free flowing, linear
viscous liquid to a semi-solid having a controllable yield
strength in milliseconds when exposed to a magnetic field
[13]. They are thus used as damping fluids to devise semiactive magneto-rheological tuned liquid column dampers
(MR-TLCDs) with alterable fluid viscosity. The sharply
alterable fluid viscosity results in adjustable and controllable damping force in the MR-TLCD for structural
vibration control under a wide range of loading conditions.
As a dissipative damper, the MR-TLCD generates the
controllable damping force by utilizing the relative motion
between the liquid and container and thus, does not have the
potential to destabilize the system. Two MR-TLCD
prototypes using different types of MR fluids have been
fabricated [14,15] and laboratory experiments showed that
the devised MR-TLCDs offered much better damping
performance than passive TLCDs even in open-loop control
mode. In the present paper, after briefing the MR-TLCD
design, a semi-active control strategy for tall building
structures incorporating MR-TLCDs is developed and its
effectiveness for mitigating wind-induced vibration
response of a real 50-story building is examined.

2. Design and modeling of MR-TLCDs


As shown in Figs. 1 and 2, an MR-TLCD consists of a Utube container filled with the MR fluid and having an orifice
opening in the middle of the bottom tube. Magnetic field
offered to the fluid is generated by an electromagnet of

Fig. 2. Schematic of MR-TLCD.

length Lp around the bottom tube (usually Lp p B where B


is the horizontal length of the liquid column). The MRTLCD device is rigidly connected to the primary structure
and capable of dissipating energy through oscillation of the
liquid column. With properly designed parameters of the
damper, the out-of-phase response of liquid mass exerts an
inertia force to the primary structure to counteract external
excitations. The damping force results from viscous
interaction between the liquid in motion and the rigid
container as well as from the hydrodynamic head-loss.
Because applying a magnetic field can alter the yield stress
of the MR fluid and thus the head-loss, the damping force of
an MR-TLCD is adjustable and controllable with the
applied magnetic field in a semi-active manner.
By modeling the motion of the MR fluid between fixed
poles using the parallel-plate theory [16], the governing
equation of motion for an MR-TLCD can be represented
as [12]

rAD LD y

2rAD BD y 0

where y is the displacement of the liquid relative to the


container; y0 is the horizontal displacement of the
container; AD ; LD and BD denote the cross-sectional area,
the total central length, and the horizontal length of the
liquid column, respectively; Lp and h are the length and
depth of the flow between the fixed poles; r is the liquid
mass density; g is the gravitational acceleration; c is a
coefficient relying on the flow velocity and has a value
ranging from 2.07 to 3.07 [16,17]; ty is the yield stress
developed by the applied magnetic field. The overall
head-loss coefficient d is determined by

Fig. 1. Prototype of MR-TLCD.

cty AD Lp y_
1
2rAD gy
rA dl_yl_y
2 D
h
l_yl

48
LD X

zj

2
H
H
j
Re 1
w

where w is the width of the flow between the fixed poles;


H is the height of the bottom tube; zj is the coefficient of
minor head-losses, including those of elbows and orifices.

Y.Q. Ni et al. / Probabilistic Engineering Mechanics 19 (2004) 269277

The Reynold number Re is defined by


Re

2rVw h
hwh

represented as
3

where V is the fluid velocity; h is the field-independent


viscosity.
The damping terms in Eq. (1) are nonlinear due to the
presence of orifice and applied magnetic field. Re-writing
Eq. (1) as
mD y u_y kD y 2lmD y 0

where mD rAD LD ; kD 2rgAD ; l BD =LD : The damping force u_y can be separated into a passive component
and a semi-active component and is represented by
u_y up _y us _y

with
up _y

1
rdAD y_ 2 sgn_y;
2

us _y ty cAD Lp =hsgn_y

6a
6b

where up _y is the uncontrollable part of the damping force,


which is regarded as a passive damping force and will be
incorporated into the system equation of motion; us _y is
the controllable part of the damping force, which is semiactive control force and can be determined according to an
optimal control strategy.
By adjusting the applied external voltage, the yield stress
ty of the MR fluid can alter accordingly and thus the MRTLCD produces a controllable damping force us _y in an
active manner. As seen in Eq. (6b), the controllable damping
force us _y is always in the opposite direction of the relative
motion, implying that MR-TLCDs can only exert dissipative control force. When the required optimal control force
does not meet such a condition, the clipping treatment [18]
needs to be performed to ensure the commanded damping
force implementable by MR-TLCDs.

3. Description of controlled system


Consider a high-rise building structure with n stories and
a semi-active MR-TLCD installed at the top floor under
random wind loading. For a linear elastic shear-type
structure, the equation of motion of the controlled system
can be expressed as
M X CX_ KX FW t FD

271

where X denotes the n-dimensional horizontal displacement


vector of the structure; M; C and K are the n-dimensional
mass, damping and stiffness matrices of the structure,
respectively; FW t is the n-dimensional wind loading
vector. FD denotes the n-dimensional control force vector
produced by inertial motion of the semi-active MR-TLCD
driven under an optimal control strategy and can be

FD E1 fD ;
fD

8a

1
1 2 l2 mD y u_y kD y
l

8b

where E1 {1; 0; ; 0}T is an n-dimensional constant


vector, and fD is the interactive force between the MRTLCD and the top floor of the structure. The variable y
denotes the relative displacement of the liquid to the
container and is governed by Eq. (4).
For the random wind excitation, the cross power spectral
density of along-wind force is represented by the Davenport
spectrum [19] as


hi hj a
S v
2
SWij v r2a CD2 V10
9
Ai Aj
cohhi ; hj ; v 0
2p
100
where SWij denotes the cross spectrum of the wind forces at
the ith and jth floors; S0 is the power spectral density of the
wind velocity; ra is air mass density; CD is the drag
coefficient; V10 is the mean wind velocity at 10 m height; Ai
and Aj are the equivalent projection areas about the ith and
jth floors, respectively; hi and hj are heights of the ith and jth
floors; and a is a constant exponent. The coherence function
cohhi ; hj ; v is described by
(
)
Ch lvllhi 2 hj l
10
cohhi ; hj ; v exp 2
2pV10
where Ch is an exponential decay constant. The power
spectral density of the wind velocity S0 is of the form
2
S0 v 8pKD V10

h0

h20
;
lvl1 h20 4=3

600v
pV10

11a
11b

where KD is the ground coarse coefficient.


Suppose that the structural response be expressed by first
m# n modes of the corresponding undamped system in
terms of an n m-dimensional mode matrix F normalized
with respect to the structural mass matrix M: Then the
structural displacement response can be represented by X
FY where Y {y1 ; y2 ; ; ym }T is the m-dimensional modal
coordinate vector. The governing equation of the system
then becomes
Y JY VY FT FW t FT E1 fD

12

where the m-dimensional diagonal matrices V FT K F


v2i  and J FT C F 2zi vi  in which vi and zi are the
structural natural frequency and damping ratio of the ith
mode, respectively. For the semi-active MR-TLCD installed
at the top floor, the motion of the container is same as the top
floor, i.e. x1 E1T X; and Eq. (4) can be rewritten as
y

1
k
u_y D y 2lE1T FY
mD
mD

13

272

Y.Q. Ni et al. / Probabilistic Engineering Mechanics 19 (2004) 269277

By combining Eqs. (12) and (13), the following augmented


matrix equation for the structure incorporating semi-active
MR-TLCD is obtained as
 FAW t FAs
MA Y CA Y_ KA Y FAp Y

14

in which the m 1-dimensional generalized displacement


 the m 1-dimensional generalized mass
vector Y;
matrix MA ; damping matrix CA and stiffness matrix KA ;
the m 1-dimensional generalized external force vector
FAW ; the passive control force vector FAp ; and the semiactive control force vector FAs are given by
( )
Y
Y
;
15a
y
2
3
1 2 l2
T
I
2
m
F
E
D
17
6 m
l
15b
MA 4
5;
"

lE1T F

1
#

"
AL
(
FN

Im1

2MA21 KA

2MA21 CA

0m1

2MA21 Bp
(
)
0m1
(

U

0m1

MA21 FAW
0m1
2MA21 Bp

#
;

up _y;
;

us _y

Z_ AZ Ft U

15f

FAs 2Bp us _y;


(
)
2FT E1 =l
Bp
1=mD

15g

U 2Bus _y

KA

FAW

0m

0m

15c

V 2kD FT E1 =l

0
kD =mD
( T
)
F FW
;

#
;

15d

15e

15h

where Im is the m-dimensional identity matrix. Since the


determinant of the generalized mass matrix
lMA l 1 1 2 l2 mD

m
X

f2i1 . 0

16

18d

18e

Eq. (17) is a set of nonlinear equations and can be linearized


to cater for the application of LQ control strategy. Applying
the stochastic linearization procedure [20,21] to Eq. (17)
yields

FAp Bp up _y;

"

0m

18c

where the coefficient matrix A AL AN and


2
3
0
0m1 0m1;m
5;
AN 4
0m1 0m1;m 2MA21 Bp ceq
s
2E_y2 
ceq rdAD
;
p
(
)
0m1
;
B
MA21 Bp

CA

18b

19

20a

20b
20c
20d

in which E denotes the expectation operator; ceq is the


equivalent damping efficient of the passive damping force
component up and related to the mean square velocity of the
liquid. The linearized state-space equation (19) will be used
for optimal control design.

4. Optimal control law

i1

where fi1 is the first element of the ith mode vector in mode
matrix F; then the generalized mass matrix MA is
nonsingular and its inverse matrix MA21 exists. Premultiplying Eq. (14) by MA21 and rewriting it in the state
space yield

Z_ AL Z FN Z Ft U

17

in which the 2m 2-dimensional generalized state vector


Z; the 2m 2-dimensional coefficient matrix AL ; the
2m 2-dimensional nonlinear force vector FN ; the
external force vector F; and the control force vector U are
as follows
( )
Y
Z
;
18a
Y_

Perfect and complete observation is assumed in this


study and then the optimal control design of system (19) can
be performed directly based on the dynamical programming
principle. Optimal control law commanding the semi-active
damping force component of the MR-TLCD is determined
by minimization of the following performance index J in an
infinite time interval [22]
1 T
J lim
LZt; us tdt
21
T!1 T
0
where L is a Lagrangian of continuous differential convex
functional. The corresponding dynamical programming
equation is obtained as [22]


V
min L AZ UT
0
22
us
Z

Y.Q. Ni et al. / Probabilistic Engineering Mechanics 19 (2004) 269277

where V is called the value function. Both the Lagrangian


functional L in the performance index (21) and the control
force U defined by Eq. (20d) are related to the semi-active
damping force us to be determined.
When the optimal linear quadratic (LQ) control strategy
is implemented, the Lagrangian L and the value function V
are represented by
L Z T SZ Ru2s ;

23a

V Z T PZ

23b

and the LQ control damping force is then obtained from


Eq. (22) as
us R21 BT PZ

24

where S is a positive semi-definite symmetric constant


matrix; R is a positive constant; and P is a symmetric
constant matrix which is determined from the following
algebraic Riccati equation
S AT P PT A 2 R21 PT BBT P 0

25

In general, the LQ control damping force given by Eq. (24)


is related to the state vector Z; and does not necessarily meet
Eq. (6b). The clipping treatment is therefore performed to
define the commanded control damping force ups which is
implementable by the MR-TLCD:
( p
F sgn_y F p $ 0
p
us
26
0
Fp # 0
with
F p R21 BT PZ sgn_y

27

which satisfies the dynamical programming equation (22)


under the dissipative damping force constraint (6b) according to the variational principle [23]. The commended
optimal damping force defined in Eq. (26) is implementable
by the MR-TLCD.
Note that the dissipative control force usually mitigates
structural vibration response effectively through the dissipation of structural energy. The optimal control force
defined by Eq. (26) can be further represented in the form of
dissipative force. By partitioning the matrices
"
#
S1 S2
S
;
28a
ST2 S3
"
#
P1 P2
P
;
28b
PT2 P3
"
#
A1 A2
A
28c
A3 A4
and letting S1 0; the optimal control force can be
represented as
F p R21 BTp QT3 Y_ sgn_y

29

273

where Q3 P3 MA21 which can be determined directly from


the following reduced-order Riccati equation
S3 2 C TA QT3 2 Q3 C A 2 R21 Q3 Bp BTp QT3 0
C0A

30

C 0A

and
0m1;m Bp ceq : Making
in which C A CA
use of Eq. (26) with Eq. (27) or Eq. (29), voltage (or current)
input to the MR fluid and therefore the fluid yield stress ty
can be adjusted optimally according to the clipped LQ
control strategy for the linearized system.

5. Response evaluation
To evaluate the random response of the controlled
structural system with semi-active MR-TLCD under wind
loading, the clipped optimal control damping force given in
Eq. (26) is first linearized statistically as [20,21]
T _
ups Cseq
Y

31

where the equivalent coefficient vector is





1
T T _
Q B E
lBp Q3 Ylsgn_y
Cseq
2R 3 p
Y_

32

Then the augmented matrix equation for the controlled


structural system with semi-active MR-TLCD becomes
MA Y C~ A Y_ KA Y FAW t
C 00A

C 00A

33
T
Bp Cseq
:

where C~ A C A
and

Eq. (33) expresses a linearized vibration system subjected to


external excitation. The random response of a linearized
system can be calculated by using the frequency-response
function [21]. For the system (33), the frequency-response
function matrix is
Hjv KA 2 v2 MA jvC~ A 21
34
p
where j 21: With Eq. (34), the cross power spectrum
matrix of the random response is obtained as
SY Y v HjvSF vH T 2jv

35

in which the power spectrum matrix of the random


excitation due to wind loading is expressed as
" T
#
F SW vF 0m1
SF v
36
01m
0
where SW v is given in Eq. (9). With Eq. (35), the mean
square (MS) response of the controlled system (33) can be
evaluated by
1
EY 2i 
SY i Y i vdv;
37a
21

EY_ 2i 

1
21

v2 SY i Y i vdv

37b

For the semi-actively controlled structural system, the cross


power spectrum matrix of the displacement response is then

274

Y.Q. Ni et al. / Probabilistic Engineering Mechanics 19 (2004) 269277

obtained by using the transformation X FY as


SXX v FSYY vFT

38

where the cross power spectrum matrix SYY v of the modal


displacement response is a sub-matrix of SY Y v given in
Eq. (35). Then the MS displacement and acceleration
responses of the semi-actively controlled structural system
can be evaluated by
1
SXi Xi vdv;
39a
EXi2 
21

EX 2i 

1
21

v4 SXi Xi vdv

39b

The MS optimal control damping force is


T
_ _ T
Eup2
s  Cseq EYY Cseq

40

by which the MS yield stress and the MS external voltage


(or current) input can be obtained. The MS responses of the
corresponding passively controlled structure using MRTLCD as a passive damper can be evaluated similarly by
eliminating terms involving the optimal control damping
force.
The control efficacy of the proposed optimal semi-active
control strategy can be evaluated in terms of performance
criteria [23]. The performance criteria used here are the
percentage reduction of root-mean-square (RMS) response
Kresponse and percentage RMS optimal damping force Kus
relative to the structural total weight:
Kresponse

RMSresponsep 2 RMSresponses
100%
RMSresponsep

p
Eup2
s 
Kus
100%
trMg

41
42

where RMS represents the root-mean-square value; subscripts p and s denote the passively and semi-actively
controlled responses of the structure, respectively; and tr

Fig. 3. Mass-normalized mode shapes of building.

Fig. 4. Wind power spectral density (PSD) at different heights.

is the trace operator of square matrix. Higher values of the


percentage response reduction Kresponse and the percentage
relative optimal damping force Kus indicate more efficient
control capability.

6. Case study
A 50-story residential building is now under construction
in Hong Kong. The initial design of this building was found
to not satisfy the wind-resistant requirement prescribed in
the Hong Kong design code, and therefore use of various
supplemental damping devices in this building has been
studied. The height of the building is 161.65 m and the total
mass trM 2:774 107 kg. Modal properties of the
building have been obtained from a precise three-dimensional finite element model. The natural frequencies of
the first five modes in along-wind direction are 0.216, 0.940,
2.278, 3.941 and 5.932 Hz, respectively. Fig. 3 illustrates
the mass-normalized mode shapes. With the obtained
natural frequencies and mode shapes, a lumped-mass
model of 51 DOFs (representing 51 floors) was formulated.
The modal damping ratio is assumed to be 3.0% for all

Fig. 5. Wind cross power spectral density (CPSD) between two heights.

Y.Q. Ni et al. / Probabilistic Engineering Mechanics 19 (2004) 269277

275

Fig. 6. RMS displacement response versus building height.

Fig. 8. Percentage reduction K (%) in RMS responses.

the modes. Since the first mode is dominant in the building


vibration, the MR-TLCD system is tuned to the first natural
frequency.
The parameters of the wind loading spectrum are taken
as [19]: ra 1:28 kg/m3, CD 1:2; V10 45:3 m/s, a
0:19; KD 0:02; Ch 10 and Ai 1 (for unit equivalent
projection area of the wind loading). The parameters of
the MR-TLCD are taken as mD 2:774 105 kg,
p
vD kD =mD 1:2195 rad/s (near the first natural frequency of the structure), l 0:8 and d 30 unless
otherwise mentioned. Figs. 4 and 5 show the wind
power spectral density at different heights and the wind
cross power spectral density between two heights,
respectively. The weight parameters in the performance index are taken as R 106 =trM2 ; S3
Diag0:7; 0:8; 1:2; 1:5; 1:5; 0:3:
Fig. 6 shows the RMS displacement response of the
semi-actively controlled and passively controlled structure
using the MR-TLCD and Fig. 7 shows the RMS acceleration
response of the semi-actively controlled and passively
controlled structure. It is seen that the structural response
reduction by means of the semi-active real-time control is
much more than that using the same MR-TLCD as a passive

damper. In particular, the semi-active MR-TLCD in


conjunction with the proposed control strategy reduces
significantly the acceleration response at all stories of the
building.
Fig. 8 illustrates the RMS displacement percentage
reduction Kd and the RMS acceleration percentage
reduction Ka by using the proposed control strategy with
the RMS optimal damping force being 0.0088% of the total
structural weight. About 46% displacement reduction and
72% acceleration reduction are achieved at the top floor. It is
found that the percentage reduction in RMS displacement
varies slightly with the building height in comparison with
the percentage reduction in RMS acceleration. The
percentage reduction of RMS acceleration at the higher
stories is much larger than that at the lower stories. The
RMS interstory drift percentage reduction Kid is also
provided in Fig. 8. It is almost the same as the RMS
displacement percentage reduction.
The effect of the wind spectrum parameters on the semiactive control efficacy is then studied. Figs. 9 and 10 show
the RMS displacement percentage reduction Kd and the
RMS acceleration percentage reduction Ka under different
mean wind velocity V10 but identical weight parameters in

Fig. 7. RMS acceleration response versus building height.

Fig. 9. RMS displacement percentage reduction Kd (%) under different V10 .

276

Y.Q. Ni et al. / Probabilistic Engineering Mechanics 19 (2004) 269277

and acceleration responses of the semi-actively controlled


structure using MR-TLCD.

7. Conclusions

Fig. 10. RMS acceleration percentage reduction Ka (%) under different V10 :

the performance index. It is observed that the RMS


displacement percentage reduction is enhanced with the
increase of the mean wind velocity while the RMS
acceleration percentage reduction is slightly affected by
the mean wind velocity. Fig. 11 illustrates the percentage
reduction of both RMS displacement and acceleration
responses with different RMS optimal damping force of
the MR-TLCD (response-1: percentage RMS optimal
damping force Kus 0:0088% under weight parameter
S3 Diag0:7; 0:8; 1:2; 1:5; 1:5; 0:3; response-2: percentage RMS optimal damping force Kus 0:015% under
weight parameter S3 Diag1:7; 2:5; 3:5; 5:0; 5:0; 0:5). It is
seen that both the displacement and acceleration response
reduction capabilities can be enhanced by increasing the
percentage RMS optimal damping force Kus ; but the
enhancement is not linearly proportional to the increase of
the percentage RMS optimal damping force. For allowed
maximum RMS optimal damping force, optimal values
of the weight parameters in the performance index can
be determined by minimizing the RMS displacement

An optimal control method for tall building structures


using semi-active MR-TLCDs has been developed based on
the dynamical programming principle and the statistical
linearization method. The MR-TLCDs combine the benefits
of controllable smart materials and tuned liquid column
dampers. The dissipative optimal damping force produced
by MR-TLCDs through relative motion between the liquid
and the container does not have the potential to destabilize
the structural system. The proposed real-time control
strategy is independent of the external excitation and
applicable to tall buildings with arbitrary stories. The
cases study of a 50-story tall building demonstrates that a
semi-active MR-TLCD installed at the top floor driven by
the proposed optimal control strategy can achieve significant response reduction in terms of displacement, interstory
drift and acceleration, in comparison with that obtained by
using a passive TLCD. It is also shown that the response
reduction capability by using the semi-active MR-TLCD in
conjunction with the proposed control method increases
with the increase of the wind loading intensity.

Acknowledgements
The work presented in this paper was supported by a
grant from The Hong Kong Polytechnic University through
the Area of Strategic Development Programme (Research
Centre for Urban Hazards Mitigation) and by a grant from
the Zhejiang Provincial Natural Science Foundation, China
(Grant No. 101046). These supports are gratefully
acknowledged.

References

Fig. 11. Response percentage reduction K (%) with different optimal


damping force. (response-1: Kus 0:0088% and S3 Diag0:7; 0:8;
1:2; 1:5; 1:5; 0:3; response-2: Kus 0:015% and S3 Diag1:7; 2:5; 3:5;
5:0; 5:0; 0:5).

[1] Sakai F, Takaeda S, Tamaki T. Tuned liquid column dampernew


type device for suppression of building vibrations. In: Proceedings of
the International Conference on Highrise Buildings, Nanjing, China;
1989. p. 926 31.
[2] Xu YL, Samali B, Kwok KCS. Control of along-wind response of
structures by mass and liquid dampers. ASCE J Engng Mech 1992;
118:2039.
[3] Balendra T, Wang CM, Cheong HF. Effectiveness of tuned liquid
column damper for vibration control of towers. Engng Struct 1995;17:
668 75.
[4] Gao H, Kwok KCS, Samali B. Optimization of tuned liquid column
dampers. Engng Struct 1997;19:47686.
[5] Balendra T, Wang CM, Rakesh G. Effectiveness of TLCD on various
structural systems. Engng Struct 1999;21:291 305.
[6] Hruska A, Dorfmann A. Elastic shear frames with tuned liquid column
dampers: a controlled study of small-scale models. In: Casciati F,

Y.Q. Ni et al. / Probabilistic Engineering Mechanics 19 (2004) 269277

[7]
[8]

[9]
[10]

[11]

[12]

[13]

[14]

Magonette G, editors. Structural control for civil and infrastructure


engineering. Singapore: World Scientific; 2000. p. 28192.
Yalla SK, Kareem A. Optimum absorber parameters for tuned liquid
column dampers. ASCE J Struct Engng 2000;126:906 15.
Liu MY, Chiang WL, Chu CR, Lin SS. Analytical and experimental
research on wind-induced vibration in high-rise buildings with tuned
liquid column dampers. Wind Struct 2003;6:7190.
Fortner B. Water tanks damp motion in Vancouver high-rise. ASCE
Civil Engng 2001;71(6):18 18.
Haroun MA, Pires JA, Won AYJ. Effectiveness of hybrid liquid
column dampers for suppressing structural vibrations. In: Proceedings
of the 13th International Modal Analysis Conference, Nashville,
Tennessee 1995;1:52531.
Yalla SK, Kareem A, Kantor JC. Semi-active tuned liquid column
dampers for vibration control of structures. Engng Struct 2001;23:
146979.
Wang JY, Ni YQ, Ko JM, Spencer Jr. BF. Semi-active TLCDs using
magneto-rheological fluids for vibration mitigation of tall buildings.
In: Anson M, Ko JM, Lam ESS, eds. Advances in Building
Technology, Oxford: Elsevier; 2002. p. 53744.
Jolly MR, Bender JW, Carlson JD. Properties and applications of
commercial magnetorheological fluids. J Intell Mater Syst Struct
1999;10:513.
Zhan S, Wang JY. Magneto-rheological tuned liquid column dampers
(MR-TLCDs) for vibration control: experimental investigation.

[15]

[16]

[17]

[18]

[19]
[20]
[21]
[22]
[23]

277

Research Report, Department of Civil and Structural Engineering,


The Hong Kong Polytechnic University, Hong Kong; 2003.
Ko JM, Zhan S, Ni YQ, Duan YF. Smart TLCD using synthetic
hydrocarbon based MR fluid: an experimental study. In: Liu SC,
editor. Sensors and smart structures technologies for civil, mechanical
and aerospace systems. Proceedings vol. 5391; 2004.
Yang G, Spencer Jr. BF, Carlson JD, Sain MK. Large-scale MR fluid
dampers: modeling and dynamic performance considerations. Engng
Struct 2002;24:309 23.
Spencer Jr.BF, Yang G, Carlson JD, Sain MK. Smart dampers for
seismic protection of structures: a full-scale study. In: Proceedings of
the Second World Conference on Structural Control, vol. 1.
Chichester: Wiley; 1999;1:41726.
Dyke SJ, Spencer Jr. BF, Sain MK, Carlson JD. An experimental
study of MR dampers for seismic protection. Smart Mater Struct
1998;7:693 703.
Davenport AG. The spectrum of horizontal gustiness near the ground
in high winds. J Royal Meteorol Soc 1961;87:194211.
Wen YK. Equivalent linearization for hysteretic systems under
random excitation. ASME J Appl Mech 1980;47:1504.
Roberts JB, Spanos PD. Random vibration and statistical linearization. Chichester: Wiley; 1990.
Stengel RF. Stochastic optimal control. New York: Wiley; 1986.
Ying ZG, Zhu WQ, Soong TT. A stochastic optimal semi-active
control strategy for ER/MR dampers. J Sound Vib 2003;259:45 62.

Das könnte Ihnen auch gefallen