You are on page 1of 12

Engineering Structures 30 (2008) 2644–2655

Contents lists available at ScienceDirect

Engineering Structures
journal homepage: www.elsevier.com/locate/engstruct

Modelling and preliminary design of a structure-TLD system


M.J. Tait ∗
Department of Civil Engineering, McMaster University, 1280 Main Street West, Hamilton, Ontario, Canada L8S 4L7

article info a b s t r a c t

Article history: Tuned liquid dampers are cost effective dynamic vibration absorbers that are increasingly being
Received 1 May 2007 employed to mitigate the dynamic response of tall buildings. A significant reduction in the dynamic
Received in revised form response of a tall building can be achieved if the TLD is properly tuned and has sufficient inherent
22 January 2008
damping. The required level of TLD inherent damping can be obtained by installing damping screens
Accepted 27 February 2008
Available online 18 April 2008
inside the TLD tank. This paper develops an equivalent linear mechanical model that accounts for the
energy dissipated by the damping screens. Equivalent linear damping ratio expressions are developed
Keywords:
for both sinusoidal and random excitation. An equivalent linear mechanical model is subsequently
Tuned Liquid Damper (TLD) developed using an equivalent displacement variable and the linear damping ratio corresponding to the
Vibration control type of excitation being applied to the structure. Experimental tests are conducted on a scaled model
Damping structure-TLD system subjected to both sinusoidal and random excitation to validate the proposed model.
Damping screens Comparisons are made between predicted and measured structural response motion and TLD free surface
Liquid sloshing response motion over a range of structural response amplitudes. A preliminary design procedure for
Potential flow theory initial TLD sizing and initial damping screen design for a TLD equipped with damping screens is outlined.
© 2008 Elsevier Ltd. All rights reserved.

1. Introduction determined experimentally by conducting shake table tests on the


TLD.
The ability of a Tuned Liquid Damper (TLD) to perform as an In this paper generalized properties of a TLD are first
effective passive dynamic vibration absorber (DVA) to mitigate determined, including the equivalent damping provided by
building accelerations to acceptable levels under wind loading has screens. Warnitchai and Pinkaew [7] developed an analytical
been well established [1]. To maximize the effectiveness of a DVA, method to determine the damping provided by a device located
including a TLD, it is essential to determine its optimal properties. at the centre of a tank for the case of sinusoidal excitation.
This work is expanded upon to include the equivalent damping
Optimal tuning ratio and damping ratio values, expressed in
developed for the case of random excitation. The equivalent
terms of the mass ratio, have been determined for a linear
mechanical model presented in this paper is intended to provide
TMD attached to an undamped primary structure subjected to
an initial estimate of the TLD damping ratio for a particular fluid
sinusoidal excitation [2] and white-noise excitation [3]. response amplitude. The velocity potential, expressed in terms of
Attainment of the optimal tuning ratio for a TLD requires the a set of generalized coordinates is used to determine the fluid
liquid sloshing frequency to be properly tuned to the natural velocity at the screen location [7,10,11]. The flow-induced screen
frequency of the structure’s vibration mode to be suppressed. For forces are subsequently obtained by the application of Morison’s
small liquid response amplitudes linear potential flow theory can formula [11]. Finally, using the concept of virtual work, [12] an
be employed to predict the liquid sloshing frequency [4]. Thus for a amplitude-dependent damping coefficient is calculated for the
particular tank geometry, the designer can determine the required screens.
fluid depth in order to achieve the optimal tuning ratio. The equations of motion describing the response of a structure-
The value of the TLD damping ratio relating to the energy TLD system are manipulated allowing the structure-TLD system to
dissipated in the boundary layer is often significantly lower than be expressed as an equivalent structure-TMD system. This permits
the value required for the TLD to operate optimally. An increase in optimization equations, developed for a linear TMD, to be applied
the TLD damping ratio value can be achieved by inserting energy to a linearized structure-TLD system. For verification purposes
dissipating devices, such as screens, inside the tank [5–9]. The experimental results from scaled model structure-TLD systems
additional inherent damping provided by these devices is often subjected to sinusoidal and random excitation are compared with
values calculated using the equivalent linearized structure-TLD
system developed in this paper.
The properties of an equivalent amplitude-dependent TMD,
∗ Tel.: +1 1 905 525 91; fax: +1 1 905 529 9688. having equal energy dissipation as a TLD equipped with damping
E-mail address: taitm@mcmaster.ca. screens, have been evaluated experimentally from shake table
0141-0296/$ – see front matter © 2008 Elsevier Ltd. All rights reserved.
doi:10.1016/j.engstruct.2008.02.017
M.J. Tait / Engineering Structures 30 (2008) 2644–2655 2645

The linearized boundary condition at the free surface, η, is given as

∂Φ ∂η

= . (4)
∂z
z=0 ∂t
The velocity potential which satisfies the above boundary condi-
tions can be expressed in the general form as a sum of infinite
sloshing modes

 cosh nπ(z+h)
h i

nπx

L
Φ (x, z, t) = q̇n (t) cos .
X
(5)

sinh nπL h
  
n=1
L
L

The free surface can be expressed as


Fig. 1. Definition sketch for liquid sloshing in a rectangular Tuned Liquid Damper
(TLD) with damping screens. ∞
nπx
 
η(x, t) = qn (t) cos .
X
(6)
n=1
L
tests [13] and the performance of this semi-empirical amplitude-
dependent model has been verified [14]. However, the analytical The kinematic of liquid sloshing can be completely described by the
equivalent mechanical model presented in this paper can be set of qn (t) where n = 1, 2, 3, . . . , generalized co-ordinates [7].
employed to evaluate the response of a structure-TLD system prior The energy of a system of standing waves of the simple
to conducting any experimental testing. harmonic type can be expressed in terms of the gravitational
The linearized equivalent mechanical model can be used to potential
determine a suitable screen design. For a given target structural Z L
response amplitude optimal parameters have been determined for 1
V = ρbg η2 (x, t)dx (7)
a Tuned Liquid Column Damper (TLCD) including the quadratic 2 0
damping provided by an orifice [15–17], which is analogous to and the kinematic potential
the damping provided by damping screens in a TLD. Employing a
∂Φ 2 ∂Φ
Z 0 Z L " 2 #
preliminary TLCD design procedure proposed by Vickery et al. [18] 1
 
and utilizing the proposed equivalent linearized structure-TLD T= ρb Ẋ + + dx dz (8)
2 −h 0 ∂x ∂z
system a method to estimate the optimal loss coefficient for a
particular screen configuration is outlined. where Ẋ is the horizontal velocity of the tank [4].

2. Mathematical description of a TLD 2.2. Influence of damping screens

A rectangular tuned liquid damper with damping screens Consider a number of screens placed at discrete locations, xj ,
is shown in Fig. 1 and a structure-TLD system is shown in inside the tank as shown in Fig. 1. In the derivation presented
Fig. 2(a). The structure is modelled as a generalized single- below, constant Cm , Cd and Cl values denoting inertia, drag, and
degree-of-freedom system representing the mode of vibration loss coefficients, respectively, are assumed. For screens having a
being suppressed. As the TLD tank moves with the structure the
solidity value S, defined as
contained fluid will exhibit a sloshing response motion. For the TLD
attached to the structure, shown in Fig. 2(a), the horizontal motion, A
X (t), that the tank experiences (Fig. 1) is equivalent to the motion
S= (9)
bh
of the structure, at the TLD location, in the vibration mode to be
where A is the area of the screen normal to the flow, the
suppressed, Xs (t).
relationship between the drag and loss coefficient is given as [6,
9,19]
2.1. Fluid response equations
Cl = SCd . (10)
The liquid sloshing motion of a tuned liquid damper equipped
with damping screens is formulated using potential flow theory. The force exerted on a screen can be expressed as [11]
Consider the rigid rectangular tank shown in Fig. 1 having a length f (x, z, t) = fi (x, z, t) + fd (x, z, t). (11)
L, a quiescent fluid depth h, and a tank width b. The assumptions
of inviscid, incompressible, irrotational flow and negligible surface Eq. (11) indicates that the screen force is comprised of two
tension are made. It is also assumed that the fluid response components. For screens located at or near the centre of the tank
amplitude is small compared to the fluid depth, i.e. η  h and only the horizontal component of the flow is considered in the
the screens, which are located at or near the centre of the tank, following derivation. However, this derivation could be expanded
do not significantly alter the overall flow of the sloshing liquid. For to include the contribution of the vertical velocity component to
the stated assumptions the velocity of a liquid particle relative to the total screen force.
the tank can be expressed as a gradient of the velocity potential,
Φ (x, z, t). The kinematic continuity of incompressible flow requires 2.2.1. Inertia component of screen force
The first term in Eq. (11) represents a virtual mass inertia
∂Φ ∂Φ
2 2
+ 2 =0 (1) force proportional to the horizontal component of the acceleration
∂x 2 ∂z exerted on the mass of the liquid displaced by the screen.
and the kinematic boundary conditions are given as This conservative force dissipates no energy; however, it does
∂Φ
contribute to the overall kinetic energy of the sloshing liquid [7].
u(x, z, t)|x=0,x=L = =0 (2) The inertia component of the screen force is given by
∂x
x=0,x=L

∂u xj , z, t
" #
∂Φ

w(x, z, t) |z=−h = = 0. (3) fi (xj , z, t) = ρAcs Cm (12)
∂z z=−h ∂t
2646 M.J. Tait / Engineering Structures 30 (2008) 2644–2655

Fig. 2. (a) Structure-TLD system and (b) equivalent mechanical linearized structure-TLD system.

i 3
nπ(z+h)
 h
where Acs denotes the cross-sectional area of the screen, expressed 1 ns 
nπxj
3 Z 0 cosh L
ρbCl
X
as Qn = − sin   
nπh

2 j=1
L −h sinh L
Acs = htb S (13)
× dz |q̇n | q̇n . (21)
and tb is the dimension of the screen in the direction of the flow.
The resulting kinetic energy can be expressed as The non-conservative nonlinear damping forces can be expressed
in the following compact form
!2
1 ns Z 0 ∂Φ ρbL
ρAcs Cm
X
T= dz (14)
2 j=1 −h ∂x xj Qn = −
2nπ
Cl ∆n Ξn |q̇n | q̇n (22)
i 2
nπ(z+h)
h
where

nπxj cosh
ns 2 Z 0
1

L
ρAcs Cm   dzq̇2n .
X
T= sin   
nπh

2 j=1
L −h sinh 1 1
L ∆n =  +  (23)
nπh

(15) 3 sinh
2
L

nπxj
ns  3
2.2.2. Drag component of screen force .
X
Ξn = sin (24)
The second term in Eq. (11) represents a drag force proportional j=1
L
to the square of the fluid velocity. The drag force resulting in Fig. 4 shows the relationship between ∆1 and h/L and Ξ1 and
additional velocity squared damping is given by x/L, respectively. The parameter ∆1 is found to approach a value
 h
nπ(z+h)
i 2 of 1/3 as h/L becomes large, with values of ∆1 ranging from
1 nπxj 2 cosh approximately 10 to 1 for h/L values between 0.1 and 0.3. Fig. 4(a)
 
L
fdn (xj , z, t) = ρbCl sin   
sinh nπL h

2 L shows that the damping due to the screen-induced losses related
to the horizontal component of flow in the fundamental sloshing
× |q̇n (t)| q̇n (t). (16) mode is dependent on the water depth to tank length ratio h/L.
A set of virtual horizontal displacements can be expressed as The term Ξ1 , which relates the damping to the screen position in
the tank for the fundamental sloshing mode, is found to reduce
cosh nπ(zL+h)
h i
rapidly as the screen location is positioned away from the centre
nπxj
 
δqn (xj , z, t) =  sin δqn (t). (17) of the tank. From Fig. 4 it can be concluded that the amount of
sinh nπL h

L damping introduced by the insertion of screens into the tank, due
to the fundamental sloshing mode, is a function of both h/L and the
Eq. (18) describes the virtual work done by the non-conservative location of the screen, xj , inside the tank.
drag forces as they act through the virtual displacements caused
by an arbitrary set of variations in the generalized coordinates
2.3. Generalized TLD properties
ns Z 0
δWnc = − f d n δ q n dz .
X
(18) The corresponding Lagrange’s equations have the form
j=1 −h

d ∂T ∂T ∂V
 
Substituting Eqs. (16) and (17) into Eq. (18) the work done can be − + = Qn , n = 1, 2, 3, . . . (25)
expressed as dt ∂q̇n ∂q n ∂q n
1 ns 
nπxj 3
 where T and V are the total kinetic energy and the total potential
δWnc = − ρbCl
X
sin energy of the system and Qn are the non-conservative forces. The
2 j=1
L
resulting equations of motion can be expressed as
i 3
nπ(z+h) m∗n q̈n (t) + cn∗ q̇n (t) + m∗n ω2n qn (t) = γn∗ Ẍ (t), n = 1, 2, 3, . . .
 h
Z 0 cosh L
×   dz |q̇n | q̇n δqn . (19) (26)

nπh

−h sinh L
where the generalized mass is
Alternatively, the work done can be expressed in terms of Qn non-
conservative forces 1 ρbL2
m∗n = (27)
nπh
 
ns 2 nπ tanh
δWnc = Q n δq n .
L
X
(20)
j=1 the natural frequency for the nth sloshing mode is
The resulting non-conservative nonlinear damping forces in the nπg nπh
 
direction of qn are given as ω2n = tanh (28)
L L
M.J. Tait / Engineering Structures 30 (2008) 2644–2655 2647

the generalized stiffness is given as distribution [20]. The linearized generalized damping coefficient
is found to be
ρbLg
k∗n = (29)
s
2 2 ρbL

ceq = Cl Ξ∆σq ω (37)
the equation for the excitation factor is expressed as π π
where σq is the RMS (root mean square) motion of the fluid
(1 − cos(nπ)) response in the fundamental sloshing mode. The corresponding
γn∗ = ρbL2 (30)
(nπ)2 generalized damping ratio can be expressed as
and the modal participation factor is
s
2 πh σq
 
ζeq = Cl

tanh ∆Ξ . (38)
γ∗ 2 nπh π
 
L L
Γn = n∗ = (1 − cos (nπ)) tanh . (31)
mn nπ L For the case of sinusoidal excitation the linearized generalized
damping coefficient value is given by
The additional virtual mass added to the nth generalized sloshing
mode is found to be 4ρbL

ceq = Cl ∆Ξ qω (39)

mn-virtual 3π2
  and the corresponding generalized damping ratio can be expressed
2nπh nπh
  
L
ns 
nπxj
2 sinh + as
4 L 2
= Cm ρAcs .
X
sin (32) πh

4 q
 2  
L nπh
ζeq ∆Ξ .

nπ sinh

j=1 = Cl tanh (40)
L
3π L L
The ratio of the virtual added mass for multiple screens is given as For the special case of a single vertical flow damping device located
at the centre of the tank, the damping ratio calculated using
2nπh
   
m∗n-virtual Acs ns
nπxj 2 Eq. (40) agrees with that obtained by Warnitchai and Pinkaew [7]
 
L X
= C m
1 +  sin (33) and the solution derived by Isaacson and Premasiri [21] for a
sinh 2nLπh
 
m∗n bL j=1
L
vertical baffle.
which is in agreement with Warnitchai and Pinkaew [7] for the The above equivalent damping accounts for the losses due to
the addition of damping screens. The energy dissipation associated
special case of a single damping device located at the centre of
with the liquid viscosity can be estimated as
a tank. For thin slat screens the term Acs /bL  1, and therefore,
the additional virtual mass can often be ignored. Additionally, for ν
 s
1 2h
 
small h/L values, the influence of the virtual mass on the natural ζw = 1+ + SC (41)
2h 2 ωn b
sloshing frequency of the fluid is typically less than the influence
of the nonlinear hardening behaviour that occurs in TLDs having where ν is the viscosity of water and SC is a surface contamination
h/L values less than approximately 0.3. factor often taken as unity [22]. For practical applications the
additional damping due to the liquid viscosity is often significantly
less than the damping resulting from the screens and as result can
2.4. Linearization of damping term for the fundamental sloshing mode usually be neglected in initial design calculations [23].
(n = 1)
3. Equivalent two-degree-of-freedom system
As TLDs are typically designed to operate as a dynamic vibration
absorber in their fundamental sloshing mode the following
3.1. Structure-TMD/ TLD equations of motion
derivation considers this mode exclusively. However, the damping
coefficient related to higher sloshing modes could be determined The equations of motion for a linear structure-TMD system as
using the method presented in this section. The nonlinear damping shown in Fig. 3(a) are given by
force expressed in Eq. (22) can be replaced by a linearized
(Ms + m) m 0 Ẋs 0 Xs
        
generalized damping term by minimizing the error, ε, between Ẍs C K
+ s + s
the actual damping force and the linearized generalized damping m m ẍr 0 c ẋr 0 k xr
force [20]. The error is expressed as 
F (t)

= (42)
ρbL 0
ε = Cl ∗
∆Ξ1 |q̇| q̇ − ceq q̇. (34)
2π where Ms , Cs , and Ks are the mass, damping and stiffness of the
structure, respectively.
Minimizing the error, the following condition should be satisfied
Application of Langrange’s equations to a structure equipped
∂E ε2 with a continuous vibration absorber [24], analogous to the

=0 (35) structure-TLD system shown in Fig. 2(a), results in the following
∂ceq

set of equations
where E() denotes the expected value. The linearized generalized (Ms + ρbhL) γ ∗
  " # 
0

Ẍs Cs Ẋs
damping force is obtained from Eq. (35) as +
γ∗ m∗ q̈

0 ceq q̇
ρbL 2

E |q̇| q̇ 
0 Xs
  
F (t)


ceq = Cl Ξ∆  . (36) + s
K
= . (43)
2π E q̇2 0 k∗ q 0
∗ Eq. (43) does not permit the direct application of linear
As the value of ceq is excitation dependent, the linearized
generalized damping coefficient is determined for both random TMD design procedures for a specified target structural response
and sinusoidal excitation. For the case of random excitation, amplitude. However, a TLD can be expressed as an equivalent TMD
it is assumed that the individual cycles are sinusoidal in form by introducing a displacement variable, which leads to equivalent
with the amplitude varying slowly in time and the distribution (or effective) TLD properties, and equations of motion analogous to
of the amplitude can be sufficiently represented by a Rayleigh those given by Eq. (42).
2648 M.J. Tait / Engineering Structures 30 (2008) 2644–2655

Fig. 3. (a) Linear structure-TMD system and (b) equivalent SDOF system with effective damping.

Fig. 4. Variation of (a) ∆1 with h/L and (b) Ξ1 with x/L.

3.2. Equivalent displacement variable and equivalent TMD properties 3.2.1. Equivalent TLD properties
The equivalent mass, damping, and stiffness, corresponding
An equivalent displacement variable, xr , is introduced as
to the fundamental sloshing mode, are used to represent an
q = Γ xr (44) equivalent mechanical system, which develops dynamic forces
where q is the generalized coordinate relating to the free surface equal to the forces exerted by the sloshing fluid. The equivalent
motion and is also a measure of relative motion with respect to the mass and stiffness representing the sloshing fluid in a rectangular
horizontal motion of the structure. Substituting xr into Eq. (43) the tank are given by [17,26]
following set of equations are obtained
(Ms + mw ) meq
  " 8ρbL2 πh
#   
0

Ẍs Cs Ẋs
+ meq = tanh (48)
γ∗ γ∗ ẍr

0 ceq Γ ẋr π 3 L
8ρbLg π h
 
0 F (t) 2
.
    
K Xs keq = tanh (49)
+ s = . (45) π2 L
0 k∗ Γ x r 0
Multiplying the second row in the above set of equations of motion Fig. 5(a) shows normalized values of the equivalent mass
by Γ and modifying the mass of the structure to account for the and stiffness as a function of h/L. The normalized equivalent
non-participating component of the liquid associated with the mass indicates the amount of fluid that participates in the
fundamental sloshing mode as follows [25] sloshing motion. TLDs are often designed with low h/L values in
Ms0 = Ms + ρbhL − meq

(46) order to maintain a high normalized equivalent mass value. As
permits Eq. (45) to be expressed in the following form h/L is increased, the normalized stiffness value approaches the
 0    " #  normalized mass value indicating the natural sloshing frequency
Ms + meq meq Ẍs C 0 Ẋs
+ s becomes independent of the water depth, making TLDs with
meq meq ẍr 0 c eq ẋr
large h/L values more challenging to tune in situ. Fig. 5(b) shows
0 F (t) the modal participation factor, Γ1 , given by Eq. (31) for the
    
K Xs
+ s = (47) fundamental sloshing model as a function of h/L. The value of
0 keq xr 0
which describes the motion of the equivalent structure-TMD Γ is also used in this paper to relate the relative motion of the
system shown in Fig. 2(b). The variable, xr , is the motion associated equivalent mechanical system shown in Fig. 2(b) to the actual free
with the effective properties of the TLD. This enables the direct surface response amplitude of the fundamental sloshing mode.
application of optimized structure-TMD design parameters to this From Fig. 5(b) it can be seen that q ≈ xr when h/L is approximately
equivalent structure-TLD system for a specified structural response 1/3. For h/L values less than 1/3 the free surface response is less
amplitude. than the value of xr .
M.J. Tait / Engineering Structures 30 (2008) 2644–2655 2649

Fig. 5. (a) Normalized meq (solid) and keq (dashed) for the fundamental sloshing mode and (b) modal participation factor Γ for the fundamental sloshing mode.

Fig. 6. Normalized linearized (a) equivalent damping ratio and (b) generalized damping ratio corresponding to sinusoidal excitation.

3.2.2. Equivalent TLD damping Figs. 6 and 7 show the normalized generalized and equivalent
The amplitude-dependent equivalent damping coefficient and damping ratio values corresponding to sinusoidal and random
equivalent damping ratio are determined for random or sinusoidal excitation, respectively. As shown in Fig. 6(a), for a given relative
excitation for a given value of σr or xr , respectively. Eqs. (50) response amplitude, xr /L, the normalized damping ratio is found to
and (51) are expressions for the amplitude-dependent equivalent decrease as h/L is increased. This trend is more pronounced when
damping coefficient and damping ratio, respectively, for random plotted as a function of the free surface response amplitude, q, as
excitation. seen in Fig. 6(b). The same behaviour is observed in Figs. 7(a) and
s 7(b) for the normalized damping ratio corresponding to the free
16ρbL 32 πh surface RMS relative motion resulting from random excitation.
 
∆Ξ ωσr
3
ceq = Cl tanh (50)
3 π 3 π L The equivalent system described by Eq. (47) allows direct
s application of TLD design tools. Therefore, for a given target
32 πh σr
 
response amplitude and corresponding relative response motion,
ζeq = Cl tanh
2
∆Ξ . (51)
π3 L L the required screen properties and placement can be estimated in
order to achieve the optimal TLD damping ratio.
For the case of sinusoidal excitation Eqs. (52) and (53) can be
used to determine the amplitude-dependent equivalent damping
coefficient and damping ratio. 4. Model verification

256ρbL πh
 
ceq = Cl tanh
3
∆Ξ ωxr (52) 4.1. Experimental set-up
3π 5 L
16 πh xr Experimental tests were performed to validate the proposed
 
ζeq = Cl 2 tanh2 ∆Ξ . (53) model. Fig. 8 shows the experimental set-up used in this testing
3π L L
program. The 1:10 scaled model structure had a mass Ms =
2650 M.J. Tait / Engineering Structures 30 (2008) 2644–2655

Fig. 7. Normalized linearized (a) equivalent damping ratio and (b) generalized damping ratio corresponding to random excitation.

the value of xr , as its solution contains the amplitude-dependent


equivalent damping coefficient given by Eq. (52). The solution
was found to converge rapidly, with an error between consecutive
solutions less than 1−10 set as the required convergence limit.
Figs. 9 and 10 compare experimentally measured results to
predicted values obtained using the effective damping coefficient
given by Eq. (52) for two excitation amplitudes. Fig. 9(a) and (b)
show the normalized structural response amplitude for a range
of frequency ratio values for Fo = 11.7 N and Fo = 22.9 N,
respectively. The response accelerations of the structure-TLD
system to these particular excitation amplitudes were found to be
representative of expected wind-induced serviceability response
accelerations. In general, good agreement is found between the
predicted and measured response displacement amplitudes over
the range of frequency ratio values tested. The maximum predicted
structural response amplitude was found to be within 4% of the
Fig. 8. Structure-TLD experimental test set-up.
measured value for the test results shown.
The free surface amplitude, η, can be determined at any
location inside the tank for the fundamental sloshing mode
4040 kg, stiffness Ks = 49 656 N/m and coefficient of damping
using the calculated value of xr , Eqs. (6) and (44) with n =
Cs = 14 N s/m. A rectangular tank with length L = 0.966 m
1. Figs. 10(a) and 10(b) compare the predicted and measured
and width b = 0.874 m constructed of 0.019 m thick acrylic was
normalized free surface response motion as a function of the
used. The TLD was equipped with two damping screens located at
forcing frequency ratio. In addition, the free surface response
0.4L and 0.6L, respectively, having a loss coefficient, Cl = 2.16.
data was digitally filtered to remove the presence of higher
To properly tune the TLD to the structure, the tank was partially
harmonics in order to directly compare the measured fundamental
filled with water to a depth of h = 0.119 m. The excitation
sloshing motion to predicted values. The maximum predicted free
force F (t) was applied to the structure through a driving spring
surface response amplitude values were found to be within 5% of
attached to a hydraulic actuator. Measured response parameters
the measured values corresponding to the fundamental sloshing
included structural acceleration, structural displacement and the
mode. However, the model is unable to account for the free surface
free surface response motion of the TLD. The structure-TLD
motions associated with higher sloshing harmonics and as a result
system was tested under both sinusoidal and random excitation
underestimates the maximum free surface response amplitude.
to investigate its response behaviour under different types of
The model was found to underestimate the maximum free surface
excitation. A more detailed discussion of this experimental test
response amplitude by 36% and 44%, respectively. The largest
apparatus can be found elsewhere [23].
deviation between the predicted and measured results is found
to occur at the lower frequency ratio peak, which is strongly
4.2. Structure-TLD system response — sinusoidal excitation dependent on the dynamic response of the fluid sloshing motion.

The magnitude of xr , can be calculated using for Eq. (47) for a 4.3. Structure-TLD system response — random excitation
given excitation amplitude Fo , and forcing frequency ratio β

f The structure-TLD system was subjected to band-limit white-


β= (54) noise excitation. RMS excitation amplitudes, σF , of 23.8 N and
fs
47.6 N, respectively, resulted in peak average hourly acceleration
where f is the forcing frequency and fs is the natural frequency of values representative of target serviceability response acceleration
the structure. An iterative procedure was employed to determine values. Using the closed form solution for a structure-TMD
M.J. Tait / Engineering Structures 30 (2008) 2644–2655 2651

Fig. 9. Comparison of measured and predicted frequency response of normalized structural displacement for sinusoidal excitation of (a) Fo = 11.7 N and (b) Fo = 22.9 N.

Fig. 10. Comparison of measured and predicted frequency response of normalized free surface amplitude for sinusoidal excitation of (a) Fo = 11.7 N and (b) Fo = 22.9 N.

system [27] and Eq. (50) the value of σr was determined for a primary structure subjected to random white-noise excitation the
given σF value using the iterative procedure and convergence optimal tuning ratio is expressed as [3]
conditions described in the previous section. Subsequently, the
frequency response functions for the structure, |H(f )| and TLD free 1 + µ2
q

surface |Hη (f )| were determined using the equivalent damping Ω opt


= (55)
ratio corresponding to the calculated σr value and the equations 1+µ
developed for a structure-TMD system [27]. and the optimal absorber damping ratio as
Figs. 11 and 12 compare the predicted and measured frequency v
response functions for the structure and TLD free surface,
µ + 3µ4
u 2
u
respectively. Fig. 11(a) and (b) compare the predicted and ζa =
opt
. (56)
t
measured frequency response of the structure for two different 4 + 6µ + 2µ2
excitation amplitudes. Predicted values of σs were found to be in
Using Eq. (47) to model a structure-TLD system, the mass ratio can
excellent agreement with the measured values for a broad range
of excitation amplitudes. Fig. 12(a) and (b) compare the predicted be expressed as
and measured frequency response functions, |Hη (f )| of the free meq
surface. The model is able to predict both the structural response µ= (57)
Ms0
displacement and the free surface response displacement over a
range of excitation amplitudes. the tuning ratio as
ωa
4.4. Structure-TLD effective damping and relative motion — random Ω = (58)
excitation ω0s
and absorber damping ratio as
Utilizing linear TMD theory, for a given mass ratio, µ, optimal
absorber parameters exist [3]. For the special case of an undamped ζa = ζeq + ζw . (59)
2652 M.J. Tait / Engineering Structures 30 (2008) 2644–2655

Fig. 11. Comparison of measured and predicted frequency response function of structural displacement for random excitation of (a) σF = 23.8 N and (b) σF = 47.6 N.

Fig. 12. Comparison of measured and predicted frequency response function of free surface amplitude for random excitation of (a) σF = 23.8 N and (b) σF = 47.6 N.

The performance of a vibration absorber, TMD or TLD, is commonly The ratio of the relative motion between the absorber, which is
measured by the amount of effective damping, ζeff that it adds represented by the equivalent linear mechanical model expressed
to the primary structure. Fig. 3(a) shows a linear structure- by Eq. (47) and the primary structure to the motion of the structure,
TMD system and Fig. 3(b) shows an equivalent single-degree- R is defined as
of-freedom primary structure with effective damping. Effective σr
damping is defined here as the amount of additional damping, that R= . (62)
σs
when added to the existing structural damping, shown in Fig. 3(b),
will result in an equivalent response of this primary system to the For the case of random excitation, the corresponding relative
actual response of the structure-TMD system shown in Fig. 3(a). motion ratio for an optimally designed absorber is found to be
For the case of white-noise excitation, the effective damping is
1+µ
expressed as [28] Ropt = q . (63)
3µ 2
2µ + 2
πfs
Z ∞ −1
ζeff = H(f )2 df

− ζs (60) Substituting Eq. (44) into Eq. (62), the relative motion ratio
4 0
between the TLD free surface and the primary structure is found
where fs and |H(f )| are the natural frequency and frequency to be
response function of the structure, respectively. If optimal tuning
ση
ratio and absorber damping ratio parameters are assumed for the Rη = Γ R = . (64)
absorber, the resulting effective damping value can be calculated σs
as Fig. 13(a) and (b) compare measured and predicted values
v of effective damping and relative TLD motion, respectively over
1u µ + µ2
u
ζeff = t
opt
. (61) a range of normalized RMS structural response amplitudes. The
4 1 + 34µ value of σs-target is taken as the structural response amplitude which
M.J. Tait / Engineering Structures 30 (2008) 2644–2655 2653

Fig. 13. Comparison of measured and predicted (a) effective damping, ζeff , and (b) relative TLD free surface motion ratio, Rη .

results in the highest level of effective damping for this particular relative response of the equivalent linear mechanical model, σr ,
structure-TLD system. The measured effective damping was for the specified target structural response amplitude, σs , can be
calculated by substituting |H(f )|, determined from experimental calculated. The number of screens, their location inside the tank
testing, into Eq. (60) and the relative TLD free surface motion ratio and the screen loss coefficient can be selected using Eq. (51)
was calculated by substituting measured values of ση and σs into based on the calculated value of σr corresponding to an optimally
Eq. (64). Fig. 13(a) indicates that the proposed model can accurately designed system.
predict the effective damping provided by the TLD to the structure Additionally, ση can be estimated using Eq. (64) to ensure
over a large range of structural response amplitudes. The measured that sufficient freeboard is available. Furthermore, TLD base shear
average peak hourly accelerations for the tests shown in Fig. 13(a) forces and damping screen forces, which develop from the fluid
and (b) ranged between 5 mg and 35 mg, covering the range of sloshing motion, can be estimated from ση .
typical serviceability limit response accelerations. Fig. 14 shows a TLD design procedure summary for a structure-
Due to the nonlinear free surface response behaviour of the TLD system subjected to random white-noise excitation. The
TLD, the equivalent linear model underestimates the free surface adaptability range of various TLD design parameters, including
response motion ratio as seen in Fig. 13(b). As expected, deviation µ and h/L, has been presented by Tait and Isyumov [29] and
between the predicted and measured response values was found the sensitivity of TLD performance to changes in various design
to increase with increased fluid response amplitude. However, parameters, including Ω and ζa , has also been investigated. It is
over the range of tests considered, the largest difference between important to highlight that the rapid design procedure described
predicted and measured free surface response motion ratio was above is limited to preliminary TLD sizing and damping screen
found to be approximately 13%. The measured response was selection, as the design procedure employs an equivalent linear
digitally filtered in order to compare the free surface response model, and as such, does not take into account the nonlinear
associated with the fundamental sloshing mode. It is evident from behaviour of the TLD. A more detailed structure-TLD design should
Fig. 13(b) that the linear model can accurately predict the TLD free be completed through nonlinear numerical modelling and model
surface response motion ratio corresponding to the fundamental scale experimental testing.
sloshing mode.
6. Conclusions
5. Preliminary design procedure
This paper presents an equivalent linear mechanical mathemat-
Initial tank sizing, determination of required water depth, ical model of a TLD equipped with multiple damping screens. Ex-
screen location and screen loss coefficient for a TLD can be rapidly pressions for an equivalent linearized damping ratio to model the
completed using the proposed equivalent linear model presented energy dissipated by the damping screens were developed for both
in this paper. The following design procedure demonstrates how sinusoidal and random excitation. The equivalent linearized damp-
the proposed model can be employed to estimate preliminary ing ratio was found to depend on the response amplitude of the
TLD parameters. First, the mass ratio, µ, is selected in order to TLD, the water depth, tank length, damping screen location and
achieve a specified level of effective damping, given by Eq. (61) damping screen loss coefficient.
that is required to reach a target structural response amplitude, The equations of motion for a structure-TLD system were
σs . Optimal absorber parameters, Ω opt and ζaopt can be estimated for expressed in the same form as that of a structure-TMD system by
the selected mass ratio using Eqs. (55) and (56), respectively. Using means of an equivalent displacement variable, permitting direct
the computed value of Ω opt the parameter h/L is selected using Eqs. application of TMD design theory to the preliminary design of a TLD
(58) and (28) in order to properly tune the TLD to the structure. It for a specified target structural response amplitude. Experimental
should be noted that floor space restrictions often limit the tank studies have been carried out to validate the theoretical model
length. and investigate the effectiveness of a TLD equipped with damping
Setting Eq. (59) equal to the calculated value of ζaopt , the required screens over a range of excitation amplitudes for both sinusoidal
equivalent damping, obtained by placing damping screens in and random excitation. Predicted structural response amplitudes
the TLD, ζeq , can be calculated. Using Eqs. (63) and (62) the and TLD free surface response amplitudes corresponding to the
2654 M.J. Tait / Engineering Structures 30 (2008) 2644–2655

Fig. 14. TLD design procedure summary for a structure-TLD system subjected to random white-noise excitation.

fundamental sloshing mode were found to be in good agreement Research and Development Challenge Fund (ORCDF), and the
with experimentally measured values. However, the linear model Natural Science and Engineering Research Council (NSERC). The
was unable to simulate the nonlinear response behaviour of the author would also like to acknowledge the Boundary Layer Wind
TLD and as a result was unable to predict the maximum free surface Tunnel Laboratory at the University of Western Ontario.
response amplitude.
A rapid preliminary design procedure was outlined for a
specified target structural response amplitude corresponding References
to wind-induced serviceability level accelerations. This method
permits initial TLD sizing and damping screen selection. The simple [1] Tamura Y, Fujii K, Ohtsuki T, Wakahara T, Kohsaka R. Effectiveness of
tuned liquid damper under wind excitation. Engineering Structures 1995;17:
design procedure allows a designer to apply well-known TLD 609–21.
design theory to a TLD equipped with damping screens. [2] Den Hartog JP. Mechanical vibrations. New York (NY, USA): McGraw-Hill Book
Company; 1956.
Acknowledgements [3] Warburton GB. Optimum absorber parameters for various combinations of
response and excitation parameters. Earthquake Engineering and Structural
Dynamics 1982;10: 381–01.
This research was supported by McMaster University’s Centre [4] Lamb H. Hydrodynamics. London (England): Cambridge University Press;
for Effective Design of Structures, funded trhough the Ontario 1932.
M.J. Tait / Engineering Structures 30 (2008) 2644–2655 2655

[5] Noji T, Yoshida H, Tatsumi E, Kosaka H, Hagiuda H. Study on vibration control vibration control of tall buildings. The structural design of tall buildings
damper utilizing sloshing of water. Journal of Wind Engineering and Industrial 1998;7:147–66.
Aerodynamics 1988;37:557–66. [18] Vickery BJ, Galsworthy JK, Gerges R. The behaviour of simple non-linear tuned
[6] Fediw AA, Isyumov N, Vickery BJ. Performance of a tuned sloshing water mass dampers. In: Proceedings of the sixth world congress of the council on
damper. Journal of Wind Engineering and Industrial Aerodynamics 1995;56: tall buildings and urban habitat. 2001.
237–47. [19] Baines WD, Peterson EG. An investigation of flow through screens. Transac-
[7] Warnitchai P, Pinkaew T. Modelling of liquid sloshing in rectangular tanks with tions of the ASME 1951;73:467–79.
flow-dampening devices. Engineering Structures 1998;20:593–600. [20] Caughey TK. Equivalent linearization techniques. Journal of the Acoustical
[8] Kaneko S, Ishikawa M. Modeling of tuned liquid damper with submerged nets. Society of America 1963;35:1706–11.
Journal of Pressure Vessel Technology 1999;121:334–43. [21] Isaacson M, Premasiri S. Hydrodynamic damping due to baffles in a rectangular
[9] Tait MJ, El Damatty AA, Isyumov N, Siddique MR. Numerical flow models to tank. Canadian Journal of Civil Engineering 2001;28:608–16.
simulate tuned liquid dampers (TLD) with slat screens. Journal of Fluids and [22] Miles JW. Surface wave damping in closed basins. Proceedings of the Royal
Structures 2005;20:1007–23.
Society of London 1967;297:459–75.
[10] Yu YK, Yoon SW, Kim SD. Experimental evaluation of a tuned liquid damper
[23] Tait M. The performance of 1-D and 2-D tuned liquid dampers. Ph.D. thesis.
system. Structures and Buildings 2004;157:251–62.
London (Ontario, Canada): The University of Western Ontario; 2004.
[11] Morison JR, O’Brien MP, Johnson JW, Schaaf SA. The force exerted by surface
[24] Jacquot RG, Foster JE. Optimal cantilever dynamic vibration absorbers. Journal
waves on piles. Petroleum Transactions AIME 1950;189:149–54.
of Engineering for Industry 1977;13:138–41.
[12] Cai D, Li A, Cheng W. Parametric study of TLDs with baffles. In: Proceedings of
the second world conference on structural control. 1998. p. 121–30. [25] Vandiver JK, Mitome S. Effect of liquid storage tanks on the dynamic response
[13] Tait MJ, El Damatty AA, Isyumov N. Testing of tuned liquid damper with of offshore platforms. Applied Ocean Research 1979;1:25–32.
screens and development of equivalent TMD model. Wind and Structures, An [26] Graham EW, Rodriguez AM. The characteristics of fuel motion which affect
International Journal 2004;7:215–34. airplane dynamics. Journal of Applied Mechanics 1952;19:381–8.
[14] Tait MJ, Isyumov N, El Damatty AA. The efficiency and robustness of a uni- [27] Mcnamara RJ. Tuned mass dampers for building. Journal of the Structural
directional tuned liquid damper and modelling with an equivalent TMD. Wind Division 1977;103:1785–98.
and Structures, An International Journal 2004;7:235–50. [28] Vickery BJ, Davenport AG. An Investigation of the behaviour in wind of the
[15] Gao H, Kwok KCS. Optimization of tuned liquid column dampers. Engineering proposed centrepoint tower. Research report BLWT-1-70. Sydney (Australia):
Structures 1997;19:476–86. 1970.
[16] Yalla SK, Kareem A. Optimum absorber parameters for tuned liquid column [29] Tait MJ, Isyumov N. Effectiveness of tuned liquid dampers in reducing wind-
dampers. Journal of Structural Engineering 2000;126:906–15. induced structural response. In: Twelfth international conference on wind
[17] Chang CC, Qu WL. Unified dynamic absorber design formulas for wind-induced engineering. 2007.

You might also like