Вы находитесь на странице: 1из 20

A FLUID-STRUCTURE INTERACTION MODEL FOR EVALUATION OF

FLOW INDUCED VIBRATIONS IN ACTIVELY CONTROLLED


STRUCTURES

Sangalli, Lúcia Armiliato


lu.sangalli@gmail.com
Applied and Computational Mechanics Center, Federal University of Rio Grande do Sul
Av. Osvaldo Aranha, 99, 90035-190 Porto Alegre, RS, Brazil.
Aguirre, Miguel Angel
miguel.aguirre@ufrgs.br
Applied and Computational Mechanics Center, Federal University of Rio Grande do Sul
Av. Osvaldo Aranha, 99, 90035-190 Porto Alegre, RS, Brazil
Funez, Marcelo Luiz
funezmarcelo@gmail.com
Applied and Computational Mechanics Center, Federal University of Rio Grande do Sul
Av. Osvaldo Aranha, 99, 90035-190 Porto Alegre, RS, Brazil
Braun, Alexandre Luis
alexandre.braun@ufrgs.br
Applied and Computational Mechanics Center, Federal University of Rio Grande do Sul
Av. Osvaldo Aranha, 99, 90035-190 Porto Alegre, RS, Brazil.

Abstract. In the present work, a finite element formulation for fluid structure interaction is proposed
and applied to actively controlled structures subject to wind action, especially large-span bridges. In this
context, aerodynamic appendices are attached to the structure in order to reduce or even eliminate
dynamic instabilities of aeroelastic origin, which are free to move as function of flow incidence or
actively controlled using principles of linear control theory. A numerical formulation based on the
explicit two-step Taylor-Galerkin scheme is adopted using trilinear finite elements with reduced
integration. The system of flow fundamental equations is constituted by the Navier-Stokes equations
and the mass balance equation, where the pseudo-compressibility hypothesis and arbitrary Lagrangian-
Eulerian (ALE) formulation are considered. Turbulence is analyzed using Large Scale Simulation (LES
- Large Eddy Simulation) in conjunction with the Smagorinsky's sub-grid scale model. For fluid-
structure interactions, a partitioned coupling model is utilized, where the rigid body approach is
considered for kinematic description of the structure. For the solution of the matrix Riccati’s equation
to obtain the gain matrix, an iterative scheme based on the Newton-Raphson method is proposed.
Verification and validation of the proposed two-dimensional numerical model are performed by
analyzing torsional divergence in a rectangular prism and flutter analysis in a sectional model the Great
Belt East bridge. Appendices with different cross-sections and frequencies are employed in order to
evaluate its influence on displacement reduction and suppression of aeroelastic instability phenomena.

Keywords: Fluid-Structure Interaction, Finite Element Method, Long-span Suspended Bridges, Flutter
Control, Aerodynamic Appendages.

1 Introduction

Modern long-span bridges are characterized as slender structures with relatively low stiffness and
damping properties owing to the structural systems adopted, which may use suspended or cable-stayed

CILAMCE 2019
Proceedings of the XL Ibero-Latin-American Congress on Computational Methods in Engineering, ABMEC,
Natal/RN, Brazil, November 11-14, 2019.
A Fluid-Structure Interaction Model for Evaluation of Flow Induced Vibrations in Actively Controlled Structures

decks. It is well known today that long-span bridges are more susceptible to aeroelastic phenomena
induced by the wind action, such as flutter and buffeting (see Simiu and Scanlan [1]; Blevins [2]). In
order to attenuate or even suppress these dynamic instabilities, some aerodynamic devices have been
proposed, where appendices are added to the deck structure, which may be passive or mechanically
controlled. Although wind tunnel tests are usually carried out to evaluate the performance of vibration
control systems associated with bridges and buildings, numerical simulation can be also adopted in the
field of wind engineering by using specialized techniques of fluid-structure interaction (FSI) and
computational fluid dynamics (CFD).
Bridge and tall buildings have been traditionally designed as passive structures with respect to
forces produced by the wind action. However, with important developments observed in the field of
control theory applied to civil engineering in the last decades (see Soong [3]; Meirovitch [4]), much
more efficient structures can be obtained in terms of performance and economy, where structural and
geometric adaptability is a key factor for maintaining vibration levels within a desired range. In this
sense, devices are added to the structure to attenuate the structural response, which may be classified as
aerodynamic or mechanical devices and passive or active devices. Passive systems include tuned mass
dampers and many other mechanical devices incorporated to the structure in order to dissipate the energy
provided by external forces. Mechanical devices can be also controlled using an active control system
that restrains the structural response through some external energy source, such as the active tuned mass
dampers.
In the field of bridge aeroelasticity, Ostenfeld and Larsen [5] are one of the first authors to propose
flaps and winglets for long-span bridges to suppress flutter instability by using movable and fixed
appendices located below and on the deck sides. The performance of aerodynamic appendices has been
investigated using flaps attached along the deck sides and winglets localized above or below the deck
structure. These appendices may be considered as active when they are controlled by a control system
that mechanically and automatically modifies its position according to the present flow and structural
conditions (see for instance Kobayashi and Nagoka [6], Wilde and Fujino [7], Kwon e Chang [8],
Hansen e Christensen [9], Wilde et al. [10], Nissen et al. [11], Gouder et al. [12], Bakis et al. [13] e Li
et al. [14, 15, 16]), or passive when its motion is governed by the deck vibrations (see for instance Wilde
et al. [17], Omenzetter et al. [18, 19], Phan e Nguyen [20] e Bakis et al. [21]).
In this work, a two-dimensional fluid-structure interaction model is proposed for numerical
investigations on actively controlled systems applied to suppression of aeroelastic phenomena,
especially for long-span bridge decks. The flow fundamental equations are written using an Arbitrary
Lagrangian-Eulerian (ALE) kinematical description and turbulence is approximated employing Large
Eddy Simulation (LES). An explicit two-step Taylor-Galerkin scheme is adopted for discretization of
the flow field, where eight-node hexahedral elements with one-point quadrature and hourglass
stabilization are utilized. The structure is considered using a rigid body approach for large rotations and
a partitioned strong coupling model is utilized for fluid-structure interaction (FSI). A control algorithm
based on the control theory is presented using algebraic and instantaneous optimal control applied in the
context of the fluid-structure interaction. The proposed model is validated considering unstable
prismatic geometries and a typical bridge deck cross-section, where the performance of different control
techniques is investigated for different winglet configurations.

2 Mathematical formulation

2.1 Flow field

The fundamental equations for numerical simulation of wind flows are the Navier-Stokes equations
and the mass conservation equation, where three-dimensional incompressible turbulent flows in
isothermal process are considered. In addition, a Newtonian viscous fluid model is adopted for
constitutive modeling of the air.
Taking into account the arbitrary Lagrangian-Eulerian (ALE) formulation for flow kinematical
description and Large Eddy Simulation (LES) for turbulence analysis in an orthogonal Cartesian

CILAMCE 2019
Proceedings of the XL Ibero-Latin-American Congress on Computational Methods in Engineering, ABMEC,
Natal/RN, Brazil, November 11-14, 2019.
SANGALLI, L.A., AGUIRRE, M.A., FUNEZ, M.L., BRAUN, A.L.

coordinates system, the system of fundamental equations may be expressed as Balance of mass (Eq. 1),
Balance of momentum (Eq. 2) and Constitutive relations (Eq. 3):
∂p ∂v
+ ρc2 i = 0 (i = 1, 2,3), (1)
∂t ∂xi
where ρ is the fluid specific mass, c is the artificial compressibility parameter (or sound speed), p is the
thermodynamic pressure, vi are the Cartesian components of the flow velocity vector v referring to the
coordinate axes xi and t denotes time. Notice that artificial compressibility is adopted in this work in
order to obtain the pressure field explicitly. Additional information on the artificial compressibility
hypothesis may be found in Chorin [22] and Nithiarasu [23].
∂vi ∂v 1 ∂σ ij
+ (v j − wj ) i = (i , j = 1, 2,3), (2)
∂t ∂x j ρ ∂x j

where wj are the components of the mesh velocity vector w and σij are the components of the fluid stress
tensor σ. Body forces are neglected in the present work.
σ ij = − pδ ij + τ ij (i, j = 1, 2,3), (3)

where the components of the viscous stress tensor τ are expressed as follows:

 ∂vi ∂v j  ∂v
τ ij = ( µ + µt )  +  + λ k δ ij (i, j , k = 1, 2,3), (4)
 ∂x
 j ∂xi  ∂xk

where µ and λ are the fluid dynamic viscosity and bulk viscosity, respectively, and δij are the components
of the Kronecker delta. It is important to notice that the equations presented above are described in terms
of filtered variables, according to the LES approach, where large flow scales are obtained directly and
small flow scales are represented by turbulence modeling. In the present work, the Smagorinsky’s sub-
scale model (Smagorinsky [24]) is utilized to obtain the eddy viscosity µt:

µ t = ρ ( C S ∆ ) ( 2 Sij S ij )
2 12
(i , j = 1, 2,3), (5)

where CS is the Smagorinsky constant and ∆ is the box filter, which may obtained locally for a finite
element formulation as follows: ∆ = ( Ω e ) , where Ωe is the volume of element e. The components of
13

1  ∂vi ∂v j 
the strain rate tensor Sij are given as: Sij =  +  . Further information on LES applied to
2  ∂x j ∂xi 
computational wind engineering (CWE) problems may be found in Murakami [25] and Blocken [26].
Appropriate initial and boundary conditions must be specified in order to obtain the solution of the
flow problem. The flow fundamental equations are valid for a determined spatial domain Ωf, where the
flow variables are functions of the Cartesian coordinates xj and time. The flow spatial domain presents
a boundary surface Γ, such that Γ = Γv + Γp + Γσ + Γfs, where Γv and Γp are the boundary subsets where
flow velocity and pressure are prescribed, respectively, Γσ is the boundary subset where the flow traction
vector t is prescribed and Γfs denotes the fluid-structure interface, where compatibility conditions are
imposed (see Fig. 1).
The mesh velocity vector is determined arbitrarily according to the mesh motion scheme adopted.
In this work, the mesh motion is governed by the following equation:
NB

∑ w (d ) j −n
k ij
j =1
wki = NB
(i = 1,..., NI ; k = 1, 2,3), (6)
∑ (d )
−n
ij
j =1

CILAMCE 2019
Proceedings of the XL Ibero-Latin-American Congress on Computational Methods in Engineering, ABMEC,
Natal/RN, Brazil, November 11-14, 2019.
A Fluid-Structure Interaction Model for Evaluation of Flow Induced Vibrations in Actively Controlled Structures

where NB and NI are the number of ALE boundary nodes and ALE internal nodes, respectively, dij is
the Euclidian distance between an ALE internal node (i) and an ALE boundary node, which may be
located on the fluid-structure interface or on the external boundary of the moving region (see Fig. 1),
and n is a user defined parameter to control mesh flexibility. In this work, the values adopted for NB
and n are 1 and 4, respectively.

Figure 1. Computational domain: boundaries and mesh regions.

2.2 Structure

In the present formulation, the structural subsystem is analyzed using the rigid body assumption.
This approach is frequently adopted in numerical simulation of bridge sectional models, where strain
magnitudes are much smaller than those observed in the displacement field. The rigid body motion is
described using the following kinematic relations:
u& SI = u& CS + ω × rCI , (7)

u && CS + α × rCI + ω × ( ω × rCI ) ,


&& SI = u (8)

where u& and u && are the linear velocity and linear acceleration vectors, ω and α are the angular velocity
and angular acceleration vectors and rij is the relative position vector considering points i and j. The
superscripts I and C indicate points belonging to the fluid-structure interface and the center of mass of
the structure, respectively. A matrix form for the equations above may be expressed as:
& I = LU
U & C, (9)
S S

U && C + L′U
&& I = LU & C, (10)
S S S

with:

 1 0 0 0 r3 − r2 
L= 0 1 0 − r3 0 r1  , (11)
 
 0 0 1 r2 − r1 0 

 0 0 0 (ω2r2 + ω3r3 ) −ω2r1 −ω3r1 


  (12)
L′ =  0 0 0 −ω1r2 (ω1r1 + ω3r3 ) −ω3r2 ,
 0 0 0 −ω1r3 −ω2r3 (ω1r1 + ω2r2 ) 
where U & and U&& are vectors containing linear as well as angular velocities and accelerations of the
structure, respectively.
CILAMCE 2019
Proceedings of the XL Ibero-Latin-American Congress on Computational Methods in Engineering, ABMEC,
Natal/RN, Brazil, November 11-14, 2019.
SANGALLI, L.A., AGUIRRE, M.A., FUNEZ, M.L., BRAUN, A.L.

Figure 2 shows a schematic drawing representing a bridge sectional model with two winglets
attached to the deck by supports, where up to 6 DOF (3 rotations + 3 translations) may be considered
for the deck and 2 DOF (1 rotation + 1 rotation) are considered for the winglets. In the present work, the
vertical and torsional DOF are taken into account in the deck motion and torsional DOF for the winglets.
Consequently, the displacement vector at the center of mass of the structure is reduced to:
U CS = ( u 2 ,θ 3 , β1 , β 2 ) . (13)

Figure 2. Bridge sectional model with winglets included.

The fundamental equations for the structural subsystem are derived from the Lagrange’s equation
of motion (Meirovitch [4]). The matrix equation of motion of the bridge structure, which includes deck
and winglets:
&& c + C U
MSU & c + K Uc = Q + F , (14)
S S S S S c

where:
 m Sθ3 S β1 Sβ2 
 

MS = 
Iθ 3 (r β1 S β1 − I β1 ) (r β2 S β2 − I β 2 ) 
, (15)
 I β1 0 
sim I β2 
 

C S = diag  2mωu2 ζ u2 2 Iθ3 ωθ3 ζ θ3 2 I β1ω β1ζ β1 2 I β 2 ωβ 2 ζ β2  , (16)

K S = diag  mωu22 Iθ3 ωθ23 I β1 ωβ21 I β 2 ωβ22  , (17)

Q = Q2c M θc3 M β1 M β2  , (18)

Fc =  0 0 M cβ1 M cβ 2  , (19)

where m is the total translational mass, ωi are the angular natural frequencies and ζi are the damping
ratios (i = u2, θ3, β1, β2), Q is the external load vector including vertical force Q2 and moment Mθ3, both
evaluated at the center of mass, and moments Mβ1 and Mβ2 evaluated at the corresponding winglet hinge.
Vector Fc contains control moments Mcβ1 and Mcβ2 applied to the winglets. The remaining physical and
geometrical parameters are the first and second mass moments of inertia, Si and Ii, respectively.

CILAMCE 2019
Proceedings of the XL Ibero-Latin-American Congress on Computational Methods in Engineering, ABMEC,
Natal/RN, Brazil, November 11-14, 2019.
A Fluid-Structure Interaction Model for Evaluation of Flow Induced Vibrations in Actively Controlled Structures

2.3 Structural control theory

Control forces are determined here using control theory applied to structural dynamics (see, for
instance, Soong [3]; Meirovitch [4]; Ogata [27]). In order to formulate the control algorithm adequately,
the system of equations of motion (Eq. 14) is rewritten in a state space representation as follows:
z& (t ) = Az (t ) + Bf (t ) z (0) = z 0 , (20)

where f is the vector of forces, z is the vector of state variables and A and B are system matrices given
as follows:
f (t ) nx1 = Q (t ) nx1 + Fc (t ) nx1 , (21)

 U c (t )   0 nxn I nxn  0 
z& (t ) 2 nx1 =  cs nx1  , A 2 nx2 n =  , B 2 nx 2 n =  nx−n1  , (22)
&&
 U s (t ) nx1 
−1
 −M S K S −M C S 
−1
S M S 
where 0 and I are the null and identity matrices and n denotes the number of degrees of freedom of the
structural system, which is set to 4 in the present work.
The control forces may be obtained considering the classical optimal control, where Fc is calculated
based on minimization of a performance index J subject to the constrain imposed by Eq. (23). The
problem is solved using the first variation of the following Lagrangian function:
tf

L= ∫ {z (t )Sz (t ) + FcT (t ) RFc (t ) + λ T (t ) [ Az (t ) + Bf (t ) − z& (t ) ]} dt ,


T
(23)
0

where λ is the vector of time-varying Lagrange multipliers, S is a 2nx2n positive semi-definite matrix
and R is a nxn positive definite matrix, both standing for weighting matrices for definition of the
magnitude of control force and the corresponding energy required. Large values of S in comparison with
R values indicate that response reduction is more important than energy consumption. Notice that the
minimization condition δL = 0 is considered with respect to the state and control variable vectors z and
Fc, which leads to the following system of equations:
λ& = − Aλ − 2Sz , λ (t f ) = 0 , z (0) = z 0 , (24)

1
Fc = − R −1B T λ . (25)
2
Assuming a closed-loop control algorithm, the Lagrange multiply vector is defined as:
λ (t ) = P (t ) z (t ) . (26)
By substituting Eq. (26) into Eqs. (20), (24) and (25), considering P(tf) = 0 and assuming that Q(t)
= 0, one obtains:
1
P& (t ) + P (t ) A − P (t )BR −1B T P (t ) + A T P (t ) + 2S = 0 P (t f ) = 0 , (27)
2
which is referred to as the matrix Riccati equation, whose solution is the Riccati matrix P. In the present
work, a Newton-Raphson scheme is adopted for the evaluation of the P and the optimal control law for
the control force vector Fc is obtained:
1
Fc (t ) = G (t ) z ( t ) = − R −1B T P (t ) z ( t ) , (28)
2
where G(t) is the control gain matrix.

CILAMCE 2019
Proceedings of the XL Ibero-Latin-American Congress on Computational Methods in Engineering, ABMEC,
Natal/RN, Brazil, November 11-14, 2019.
SANGALLI, L.A., AGUIRRE, M.A., FUNEZ, M.L., BRAUN, A.L.

3 Numerical Model

3.1 Flow model

The fluid flow analysis is performed in this work using the explicit two-step Taylor-Galerkin
scheme (see, for instance, Donea [28]; Kawahara and Hirano [29]) in the context of the finite element
method, where eight-node hexahedral elements with one-point quadrature and hourglass control are
utilized. In the present scheme, approximations for the flow variables at tn+1/2 are evaluated in the first
step using the following equations:

∆t  n ∂vi
n
1 ∂p n
vin +1 2 = vin +  ( j
− − j ) − +
n
v w
2  ∂x j ρ ∂xi
(29)
 µ + µ   ∂vin ∂v j  λ ∂vkn  ∆t n ∂ 2vin 
n

 t
 +  + δ ij  + ( j j )( k k ) ∂x ∂x  ,
v − w n
v n
− w n

∂x j  ρ   ∂x j ∂xi  ρ ∂xk  4 j 


k 

∆t  ∂v 
n
n +1 2
p = p +  − ρc2 j
n
 , (30)
2  ∂x j 
1 ∆t ∂
vin +1 2 = vin +1 2 −
ρ 4 ∂xi
( p n +1 2 − p n ) . (31)

The flow variables at tn+1 are finally obtained using:


vin +1 = vin + ∆vin +1 2 , (32)

p n +1 = p n + ∆p n +1 2 , (33)
where:
n +1 2
 ∂v 1 ∂p ∂  µ + µ   ∂vi ∂v j  λ ∂vk  
∆vin +1 2 = ∆t − ( v j − w j ) i − +  t
 + + δ ij   , (34)

∂x j ρ ∂xi ∂x j  ρ   ∂x j ∂xi  ρ ∂xk  
n +1 2
n +1 2  ∂v 
∆p = ∆t  − ρ c 2 i  . (35)
 ∂xi 
Superscripts indicate the time position where the corresponding flow variable is evaluated. The
time step ∆t is obtained locally using the stability condition ∆t = α ∆x/(c + U∞), where ∆x is the
characteristic length of the element, U∞ is the undisturbed flow speed and α is a safety coefficient (α <
1). A unique time step ∆t is utilized throughout the fluid mesh, which corresponds to the minimum value
obtained among all the elements. Finite element equations are finally obtained applying the Bubnov-
Galerkin weighted residual method on Eqs. (29)-(35). Hexahedral elements with trilinear interpolation
functions are employed for approximation of both the velocity and pressure fields. Additional details on
the finite element formulation employed in this work may be found in Braun and Awruch [30] and Braun
and Awruch [31].

3.2 Structure and fluid-structure interaction models

Assuming the rigid body approach for structural motions and a partitioned coupling model for fluid-
structure interactions, where flow and structure are solved sequentially by imposing kinematic
compatibility and equilibrium conditions on the interface, and also considering the active vibration

CILAMCE 2019
Proceedings of the XL Ibero-Latin-American Congress on Computational Methods in Engineering, ABMEC,
Natal/RN, Brazil, November 11-14, 2019.
A Fluid-Structure Interaction Model for Evaluation of Flow Induced Vibrations in Actively Controlled Structures

control, the dynamic equilibrium equation of the structure may be expressed as:
&& c + C U
MS U & c + K Uc = Q + F , (36)
S S S S S c

where M S and C S are the structural mass and damping equivalent matrices, which includes terms
referring to fluid elements on the fluid-structure interface:


( 
)
nel
M S = M S + ρ ∑ Te M eII Te  ,
T
(37)
 e =1 
 nel

CS = CS + ρ ∑ Te M IIe T′e + Te ( A IIe + DIIe ) Te   .
T T
(38)
 e =1
 

In addition, the equivalent load vector Q is written as follows:


 nel
 T 
Q S = Q − ρ ∑  Te M eIF v& eF + Te ( A eIF + DeIF ) v eF − Te G eI v eI   ,
T 1 T
(39)
 e =1  ρ 
where M, A, D and G are the mass, advection, diffusion and gradient matrix referring to a fluid element
e in contact with the fluid-structure interface, v eF and v& eF are the nodal velocity and acceleration vectors
referring to fluid element e, Q is the external load vector (except flow loads), ρ is the fluid specific mass
and nel denotes the total number of fluid elements on the interface. Subscripts I (I = 1,4) and F (F = 5,8)
indicate fluid nodes on the interface and nodes within the flow domain, respectively (see Fig. 3).
Transposed matrices TT and T’T include all the nodal translation matrices L and L’ (see Eqs. 11 and 12)
of element e:

= [ L1 L 2 K L8 ]( 6 x 24) , = [ L1′ L′2 K L′8 ]( 6 x 24 ) ,


T T
T e
T′ e
(40)

where translation matrices related to nodes within the flow domain (subscripts F) are null.

Figure 3. Fluid element on the fluid-structure interface.

Equation (42) is obtained by applying the finite element representation of the compatibility and
equilibrium equations on nodes belonging to the fluid-structure interface, which is solved in the time
domain using the implicit Newmark method. A detailed description of the coupling model adopted in
the present paper may be found in Braun and Awruch [32].

3.3 Numerical algorithm

The structural response of actively controlled structures is obtained here using the numerical
procedures described above, which may be organized according to the following algorithm:
For any instant of time tn over the time integration process:
1. Solve the flow problem (Eqs. 29-35) to obtain vn+1 and pn+1 (subject to boundary conditions and
compatibility conditions) from previous values vn and pn;
2. Obtain the mass and damping equivalent matrices (Eqs. 37 and 38);
3. Obtain the equivalent load vector (Eq. 39) considering the present state conditions to evaluate
the control force vector (Eq. 28) when control devices are activated;
4. Solve the dynamic equilibrium equation at the center of mass of the structure (Eq. 36);
5. Impose compatibility conditions (Eqs. 7, 8) on the interface considering the structural response
evaluated at the center of mass of the structure at tn+1;
CILAMCE 2019
Proceedings of the XL Ibero-Latin-American Congress on Computational Methods in Engineering, ABMEC,
Natal/RN, Brazil, November 11-14, 2019.
SANGALLI, L.A., AGUIRRE, M.A., FUNEZ, M.L., BRAUN, A.L.

6. Update the finite element mesh using the mesh motion scheme given by Eq. 6 and the continuity
conditions.
7. Return to step 1.

4 Applications

4.1 Rectangular prism

The present application shows a numerical investigation where different control approaches are
analyzed in order to obtain a stable response for a rectangular prism subject to a viscous fluid flow,
which naturally leads to large rotations. Computational domain and boundary conditions utilized here
are shown in Fig. 4, where 37,440 hexahedral elements and 75,920 nodal points are employed for a
rectangular prism with a cross-section area of 2.5m x 0.5m. The smallest grid spacing in the finite
element mesh is referred to elements located on the prism surface, where the characteristic lengths are
approximately ∆x = 0.01 H (with H = 0.5 length units) and the time step is ∆t = 1.5x10-4 s. The fluid
properties are specified considering a Reynolds number of Re = 1000 and the dimensionless structure
parameters are defined as follows: translational mass m* = 195.57, mass moment of inertia I* = 105.94,
vertical stiffness k*y = 0.7864, angular stiffness k*θ = 17.05, vertical damping cy* = 0.0325 and angular
damping cθ* = 0.0.

Figure 4. Rectangular prism: computational domain and boundary conditions.

The control algorithm is then applied to suppress dynamic instability by using different
combinations of the control matrices S and R (see Eq. 23). Figure 5 presents a partial record related to
the torsional response obtained here during the time interval t ∈ [90, 243], which is used as a reference
response for the control analysis performed in this example.
0.5
0.1 Without control Without control
Angular displacement [rad]

Moment [dimensionless]

0.25

0 0

-0.25

-0.1
-0.5
121.5 243 121.5 243
(a) Time [s] (b) Time [s]

Figure 5. Rectangular prism: (a) structural response and (b) flow induced moment.

Control efficiency is initially investigated considering different values for S while R is maintained
equal to 1 (see Fig. 6). Notice that the control force (control moment) is applied on the rotational DOF
CILAMCE 2019
Proceedings of the XL Ibero-Latin-American Congress on Computational Methods in Engineering, ABMEC,
Natal/RN, Brazil, November 11-14, 2019.
A Fluid-Structure Interaction Model for Evaluation of Flow Induced Vibrations in Actively Controlled Structures

only and the same value is attributed to all elements of matrix S. It is observed that a damped response
is obtained even for R = S = 1 with a relatively small amount of energy required, considering that the
maximum moment applied to the structure due to vibration control is approximately 8.0x10-2 moment
units. As S values are increased, amplitudes of the structural response are reduced while the control
algorithm requires higher energy to maintain the structural response restricted to small amplitudes.
0.08 0.1

Control moment [dimensionless]


Angular displacement [rad]

S:R - 1:1 S:R - 1:1


0.04

0 0

-0.04

-0.08 -0.1
121.5 243 121.5 243
Time [s] Time [s]
0.04 0.55

Control moment [dimensionless]


Angular displacement [rad]

2 2
S:R - 1:10 S:R - 1:10

-0.04 -0.22
121.5 243 121.5 243
Time [s] Time [s]

Figure 6. Rectangular prism: structural response and control moments for two different S/R ratios.

Figure 7 summarizes results obtained from the control efficiency analysis, where maximum
displacement and maximum control force are presented as functions of S values. One can see that after
these graphs are obtained for a given structure, the amount of control force to reduce the displacement
amplitudes to a specified limit can be determined.

(a) (b)
Figure 7. Rectangular prism: (a) maximum displacement and (b) maximum control moment for
different S values.

Control efficiency is now investigated considering different values for R while S is set to unity.
Figure 8 shows structural responses and maximum control moments obtained according to the S/R ratio
adopted in the present simulations, where one can see that stable responses are obtained only for R = 1
and R = 10. Results indicate that a minimum amount of energy is required in order to stabilize the
structural response, considering that higher values of R denote priority to save energy over displacement
reduction.

CILAMCE 2019
Proceedings of the XL Ibero-Latin-American Congress on Computational Methods in Engineering, ABMEC,
Natal/RN, Brazil, November 11-14, 2019.
SANGALLI, L.A., AGUIRRE, M.A., FUNEZ, M.L., BRAUN, A.L.

Displacement x R Control Moment x R

Maximum control moment


0.12
Displacement at instant 0.08
0.1 0.07

[dimensionless]
0.06
0.08 0.05
214 s

0.06 0.04
0.04 0.03
0.02
0.02 0.01
0 0
1E+00 1E+01 1E+02 1E+03 1E+04 1E+05 1E+00 1E+01 1E+02 1E+03 1E+04 1E+05
(a) R (b) R
Figure 8. Rectangular prism: (a) structural response for different S/R ratio, (b) control moment as
function of R.

The influence of aerodynamic appendices on the control efficiency is investigated considering two
winglets attached to the prism according to the geometric configuration presented in Fig. 9. Two winglet
cross-section geometries are analyzed: a thin rectangular plate and a biconvex airfoil based on the
NACA012 geometry.

Figure 9. Rectangular prism with winglets: geometrical characteristics.

Computational domain and boundary conditions utilized in the present analysis are shown in Fig.
10, where the finite element mesh is composed by 29,855 hexahedral elements and 60,524 nodes. Fluid
and structural properties related to the prism are the same as those adopted in the previous analyses.
Dimensionless structural parameters associated with the winglets are defined as follows: mass moment
of inertia – I*β1 = I*β2 = 5.14x10-3; first mass moments of inertia – S*β1 = S*β2 = 6.72x10-2 and S*θ3 =
145.70; torsional stiffness – K*β1 = K*β2 = 0.1264; torsional damping – c*β1 = c*β2 = 0.0.

Figure 10. Rectangular prism with winglets: computational domain and boundary conditions.

CILAMCE 2019
Proceedings of the XL Ibero-Latin-American Congress on Computational Methods in Engineering, ABMEC,
Natal/RN, Brazil, November 11-14, 2019.
A Fluid-Structure Interaction Model for Evaluation of Flow Induced Vibrations in Actively Controlled Structures

The structural system composed by prism and winglets is first analyzed taking into account that the
winglets are not controlled. Two conditions are investigated: (a) the winglets are free to vibrate due to
the flow action and (b) the winglets are rigidly linked to the prism. In the present investigation, only
winglets with rectangular shape are utilized.
Figure 11a shows the structural response in terms of prism angular displacements referring to the
different conditions analyzed here. One can see that the winglets alone are not able to suppress dynamic
instability induced by the fluid flow if a control algorithm is not employed. It is observed that the
condition with winglets rigidly attached to the prism leads to amplification of the structural response
and anticipation of instability effects, although a smaller growth rate can be identified when the winglets
are free to vibrate. Figure 11b presents the angular displacements at the winglets corresponding to the
condition with winglets free to vibrate, where one can see that the windward winglet shows a structural
response with significantly higher amplitudes than those observed at the leeward winglet. The winglets
frequency (fβ1 = 0.877 hz; fβ2 = 0.855 hz) is approximately 18 times higher than the prism frequency (fθ3
= 0.048 hz).
0.4
Prism with uncontrolled rigid winglets
Prism with uncontrolled hinged winglets
0.3
Angular displacement [rad]

Prism without winglets and no control


0.2

0.1

-0.1

-0.2

-0.3
90 140 190 240 290 340 390 440 490
(a) Time [s]

Winglet 1 - rectangular cross section - without control


0.6 Winglet 2 - rectangular cross section - without control
Angular displacement[rad]

0.4
0.2
0
-0.2
-0.4
-0.6
90 100 110 120 130 140
(b) Time [s]
Figure 11. Rectangular prism with uncontrolled winglets: structural response (a) at the prism and (b) at
the winglets.

The structural system is analyzed now with winglets hinged to the prism and actively controlled. A
preliminary analysis is carried out in order to determine the optimal set of control parameters (see
matrices S and R, Eq. 23). The main goal is to obtain a gain matrix G which leads to a stable response
with relatively low displacement amplitudes and relatively small control moments applied on the
winglets by using different combinations of R and S values. Structural responses are recorded during a
time period corresponding to the first 50 s of the FSI analysis.
Results obtained in this preliminary analysis shows that angular displacements obtained at the prism
are significantly reduced for Sθ3 > 106. On the other hand, angular displacements and control moments
at the winglets are significantly increased for Sθ3 values greater than 104, which may lead winglets to
present dynamic instability. These high angular displacements and control moments at the winglets are
CILAMCE 2019
Proceedings of the XL Ibero-Latin-American Congress on Computational Methods in Engineering, ABMEC,
Natal/RN, Brazil, November 11-14, 2019.
SANGALLI, L.A., AGUIRRE, M.A., FUNEZ, M.L., BRAUN, A.L.

reduced as Sβ1 and Sβ2 are increased. Taking into account the angular displacements related to the prism,
one can see that the influence of the actively controlled winglets is reduced as Sβ1 and Sβ2 are increased.
Finally, as expected, angular displacements and control moments on the winglets are reduced when Rβ1
and Rβ2 are increased owing to the importance of energy saving over displacement reduction manifested
by large values of R. On the other hand, angular displacements referring to the prism are increased
significantly as Rβ1 and Rβ2 are increased, which indicates that the influence of the winglets on the prism
structural response is reduced for large values of R.
After this preliminary analysis on the control matrices R and S, the structural response of the
structural system (prism + actively controlled winglets) is obtained considering an optimized set of
control parameters given by: R = diag[1 1000 1000] and S = diag[106 102 102 106 102 102]. Figure 12
shows the structural response obtained for the prism with controlled winglets, where one can observe
that the control algorithm was able to obtain a stable response and maintain the angular displacement
amplitudes limited to a maximum of 3°.
0.25
Prism with controlled winglets
Angular displacement [rad]

Winglet 1
0.125 Winglet 2

-0.125

-0.25
90 140 190 240 290 340 390 440 490
Time [s]
Figure 12. Rectangular prism with actively controlled winglets structural response at the prism and at
the winglets.

The structural responses referring to prism and winglets can be comparatively analyzed with Fig.
12. It is observed that prism and winglets present the same frequency, although a phase shift can be
identified between the winglets. Moreover, one can see that the leeward winglet response is
synchronized with the prism response, while the windward winglet response presents a phase shift of
½T with respect to the prism response, where T is the vibration period. Details of the finite element
mesh near the prism are shown in Fig. 13 considering the different time instants indicated in Fig. 12.

Figure 13. Rectangular prism with actively controlled winglets: finite element mesh configurations for
different time instants.

4.2 Great Belt East Bridge

In this example, a numerical simulation of the wind action on a bridge sectional model with airfoil
winglets is performed, where the well known Great Belt East Bridge’s deck geometry is utilized.
Detailed information on the Great Belt East (GBE) Bridge may be found, for instance, in Larsen [33].
The flow field is characterized here using a Reynolds number Re = 105 and the structure parameters
related to the bridge deck are the following: translational mass – m = 2.27x103 kg/m; mass moment of
inertia – Iθ3 = 2.47x106 kg.m2/m; vertical natural frequency – ω2 = 0.628 rad/s; torsional natural
frequency – ωθ3 = 1.747 rad/s; damping ratio – ζ2 = ζθ3 = 0.2%. Geometrical characteristics related to
the bridge deck and winglets utilized in the present analyses are shown in Fig. 14, where the finite
element mesh and boundary conditions employed here are also presented. The finite element mesh uses
CILAMCE 2019
Proceedings of the XL Ibero-Latin-American Congress on Computational Methods in Engineering, ABMEC,
Natal/RN, Brazil, November 11-14, 2019.
A Fluid-Structure Interaction Model for Evaluation of Flow Induced Vibrations in Actively Controlled Structures

60,400 hexahedral elements and 122,130 nodal points, considering that the smallest elements are located
on the winglet surface with characteristic lengths ∆x = 0.005H (H = 4.4 m). The winglets are considered
as symmetric and biconvex airfoils based on the NACA0012 and NACA0021 geometries.

(a)

(b)
Figure 14. Great Belt East Bridge: (a) geometrical characteristics; (b) computational domain and
boundary conditions.

A preliminary investigation is carried out in order to obtain the critical flutter speed considering the
bridge deck with no aerodynamic appendices. In addition, aerodynamic analyses are also performed to
obtain the aerodynamic coefficients, taking into account the bridge deck and the different winglet
configurations adopted in this work. The flutter analysis is accomplished here using a direct method in
which bridge structural responses are obtained for different flow speeds considering that the bridge deck
is released from an initial condition with a small angular displacement. A growth/decay rate is then
calculated for each of the structural responses obtained and plotted against flow reduced speed. The
critical flutter speed is determined at the point where the curve crosses the reduced speed axis.
Instantaneous pressure fields obtained during the aerodynamic analyses performed here are
presented in Fig. 15. One can see that the presence of winglets above the deck modified the flow field
on the bridge upper surface. In addition, it is observed that the vortex shedding phenomenon at the
winglets can be identified for both the NACA airfoils tested in this work, especially for NACA0021,
where pressure loads acting on the winglets are more significant in this case. These observations may
be better expressed considering time histories of drag, lift and moment coefficients referring to the
predictions obtained from the present study (see Fig. 16).
Notice that the inclusion of airfoils modified significantly the aerodynamic force resultants acting
on the structure, as demonstrated in Table 1. Moreover, higher frequencies are observed owing to the
presence of smaller vortices associated with the vortex shedding at the winglets. Results obtained in this
work show a good agreement with numerical and experimental predictions presented by Kuroda [34].

CILAMCE 2019
Proceedings of the XL Ibero-Latin-American Congress on Computational Methods in Engineering, ABMEC,
Natal/RN, Brazil, November 11-14, 2019.
SANGALLI, L.A., AGUIRRE, M.A., FUNEZ, M.L., BRAUN, A.L.

(a) (b)
Figure 15. Great Belt East – instantaneous pressure fields: (a) bridge deck with NACA0012 airfoil; (b)
bridge deck with NACA0021 airfoil; (c) bridge deck without winglets.

(a) (b)

(c)
Figure 16. Great Belt East – aerodynamic coefficients: (a) bridge deck with NACA0012 airfoil; (b)
bridge deck with NACA0021 airfoil; (c) bridge deck without winglets.

Simulations are performed considering now the airfoil winglets included, which may be actively
controlled or free to vibrate under the wind action. The influence of the winglet shape and frequency on
the bridge aeroelastic behavior is investigated using NACA0012 and NACA0021 airfoils with different
spring stiffnesses, which are defined according to the following frequency ratios: ωβ/ωθ3 = 5, 10, 14 and
20, where ωθ3 = 1.747 rad/s and ωβ1 = ωβ2 = ωβ. A slightly supercritical flow speed (73,6 m/s) is adopted
in order to evaluate the performance of airfoil winglets under flutter instability condition. The structure
parameters utilized in the present analysis are the following: translational mass – m = 2.395x103
(NACA0021) and m = 2.391x103 (NACA0012); mass moments of inertia – Iθ3 = 2.52x106 kg.m2/m, Iβ1
= Iβ2 = 4.815x102 kg.m2/m (NACA0021) and Iβ1 = Iβ2 = 3.787x102 kg.m2/m (NACA0012); first mass
CILAMCE 2019
Proceedings of the XL Ibero-Latin-American Congress on Computational Methods in Engineering, ABMEC,
Natal/RN, Brazil, November 11-14, 2019.
A Fluid-Structure Interaction Model for Evaluation of Flow Induced Vibrations in Actively Controlled Structures

moments of inertia – Sθ3 = 3.092x105 kg.m/m (NACA0021) and Sθ3 = 3.078x105 kg.m/m (NACA0012),
Sβ1 = Sβ2 = 3.316x102 kg.m/m (NACA0021) and Sβ1 = Sβ2 = 2.609x102 kg.m/m (NACA0012). No
structural damping is considered on the winglets.

Table 1 - Great Belt East – time-averaged aerodynamic coefficients


Bridge configuration Cd Cl Cm
Deck without winglets 0,59 0,02 -0,042
Deck with NACA 012 airfoil 0,931 -0,062 -0,066
Deck with NACA 021 airfoil 0,996 -0,077 -0,067
FX FY FY
CD = ; CL = ; CM = . (41)
1/ 2 ρU H
2
1/ 2 ρU B
2
1/ 2 ρU 2 B 2
The structural responses referring to a frequency ratio ωβ/ωθ3=14 are shown in Fig. 17. In this case,
the deck displacements are more effectively reduced when NACA0012 airfoil is utilized, although the
angular displacements at the winglets are significantly higher for NACA0012 airfoil, which show a
tendency to become unstable. On the other hand, windward winglets present a structural response with
displacement amplitudes higher than those observed in the leeward winglets. Notice that the bridge
configuration with NACA0021 airfoil leads to a deck response which may become unstable over time.
0.15
Angular displacement [rad]

Deck - NACA 012


0.1 Deck - NACA 021

0.05
0
-0.05
-0.1
-0.15 0 5 10 15 20 25 30
Time [s]
Angular displacement [rad]

0.4 Airfoil 1 - NACA 012


Airfoil 1 - NACA 021
0.2

-0.2

-0.4
0 5 10 15 20 25 30
Time [s]
Angular displacement [rad]

0.4

0.2

-0.2
Airfoil 2 - NACA 012
-0.4 Airfoil 2 - NACA 021
0 5 10 15 20 25 30
Time [s]
Figure 17. Great Belt East Bridge with NACA airfoils: structural response for ωβ/ωθ3 = 14.

Figure 18 summarizes all the results obtained in the present investigation, where the averaged
logarithmic decrement associated with the deck angular responses is presented as a function of the
different frequency ratio ωβ/ωθ3 utilized. Notice that, in general, NACA0021 shows better performance
than NACA0012 when displacement reduction of the deck response is considered. However, it is
CILAMCE 2019
Proceedings of the XL Ibero-Latin-American Congress on Computational Methods in Engineering, ABMEC,
Natal/RN, Brazil, November 11-14, 2019.
SANGALLI, L.A., AGUIRRE, M.A., FUNEZ, M.L., BRAUN, A.L.

observed that a stable response over time is not guaranteed when uncontrolled winglets are utilized.
25%
20%
15%
ζ Airfoil NACA 012
10%
Airfoil NACA 021
5%
0%
0 10 ωβ/ωα 20
Figure 18. Great Belt East Bridge with NACA airfoils: averaged logarithmic decrement as a function
of frequency ratio ωβ/ωθ3.

A final investigation is performed considering actively controlled winglets based on NACA0012


and NACA0021 airfoils. The structural system is characterized with the same structural parameters
utilized in the previous analysis, except the frequency ratio ωβ/ωθ3, which is maintained at ωβ/ωθ3 = 14.
Notice that the structure is released from the equilibrium configuration. The control scheme adopted
here is defined using a set of optimized control parameters obtained previously for the present
application: R = diag[1 10 10] and S = diag[1014 109 109 1014 109 109]. The control moments are applied
at t = 2.45 s, when the deck displacement presents its maximum value.
0.15
Angular displacement [rad]

Deck - NACA 012 - without control


0.1 Deck - NACA 021 - without control
0.05 Deck - NACA 012 - with control
Deck - NACA 021 - with control
0
-0.05
-0.1
-0.15 0 5 10 15 20 25 30
Time [s]
Angular displacement [rad]

0.25

-0.25
Airfoil 1 - NACA 012 - without control
-0.5 Airfoil 1 - NACA 021 - without control
Airfoil 1 - NACA 012 - with control
-0.75 Airfoil 1 - NACA 021 - with control
0 5 10 15 20 25 30
Time [s]
Figure 19. Great Belt East Bridge with NACA airfoils: structural responses using controlled and
uncontrolled leading winglets with frequency ratio ωβ/ωθ3 = 14.

Deck and winglet responses obtained under the present conditions are shown in Fig. 19, where
responses referring to the uncontrolled system are also presented. Notice that at t = 2.45 s, when the
control moments are applied, both winglets start oscillating with a frequency similar to that observed at
the bridge deck. Moreover, the leeward winglet starts vibrating approximately in phase with the bridge
deck, while the windward winglet shows a phase angle of π radians with respect to the leeward winglet.
Wilde and Fujino [7] pointed out that a controlled system works optimally when that behavior is
observed. That structural pattern is also observed here for different frequency ratios ωβ/ωθ3.
One can see that the overall structural response referring to the controlled system shows a
significant reduction in terms of displacement amplitude and time-averaged displacement when it is
compared with the corresponding responses obtained with uncontrolled system. A dynamically stable
CILAMCE 2019
Proceedings of the XL Ibero-Latin-American Congress on Computational Methods in Engineering, ABMEC,
Natal/RN, Brazil, November 11-14, 2019.
A Fluid-Structure Interaction Model for Evaluation of Flow Induced Vibrations in Actively Controlled Structures

behavior is clearly observed in the controlled system after the control algorithm is applied to the
structure.
Time series of control moments produced by the control algorithm are presented in Fig. 20, where
a frequency ratio ωβ/ωθ3 = 14 is adopted. Other frequency ratios are also utilized and the corresponding
maximum control moments obtained during the analysis are presented in Fig. 21a. as functions of the
frequency ratio. Notice that the maximum control moment is generally observed as soon as the control
algorithm begins to be applied and the control moments required to reduce the vibration amplitude at
the bridge deck are similar for both the airfoils tested in the present investigation. As can be seen, the
control moments at the winglets are always in opposite direction. The averaged logarithmic decrement
of the deck angular responses as a function of the ratio ωβ/ωθ3 is increased for all cases when the airfoils
are actively controlled in comparison with the uncontrolled cases (see Fig. 21b). It was not possible to
obtain results for both cross sections with ωβ/ωθ3 = 5 because the displacements imposed by the control
forces reach high values already in the first seconds of analysis, leading to the instability of the airfoils.
6.0E+04 CM1 - NACA 012
Control moment [rad]

4.5E+04 CM2 - NACA 012


3.0E+04 CM1 - NACA 021
CM2 - NACA 021
1.5E+04
0.0E+00
-1.5E+04
-3.0E+04
0 5 10 15 20 25 30
Time [s]
Figure 20. Great Belt East Bridge with controlled NACA airfoils: control moments for ωβ/ωθ3 = 14.

Airfoil NACA 0012 - CM-1 Airfoil NACA 0012


Airfoil NACA 0021 - CM-1 Airfoil NACA 0021
Airfoil NACA 0012 - CM-2 Airfoil NACA 0012 - with control
Máximum control moment

Airfoil NACA 0021 - CM-2 Airfoil NACA 0021 - with control


200000 30%

150000
[kN.m]

20%
ζ

100000
10%
50000

0 0%
0 5 10 15 20 25 0 5 10 15 20 25
(a) ωβ/ωα (b) ωβ/ωα

Figure 21. Great Belt East Bridge with controlled NACA airfoils: a) maximum control moments and
b) averaged logarithmic decrement of the deck motion, both as functions of frequency ratio ωβ/ωθ3.

5 Conclusions

A fluid-structure interaction model for numerical simulation of actively controlled structures under
the wind action was proposed in this work, where optimal control theory was utilized to investigate the
influence of actively controlled appendices on the structural response of structures subject to aeroelastic
instabilities. The finite element formulation was verified using FSI classic applications of torsional
flutter in rectangular prism, where a partitioned coupling scheme and the rigid body approach were
adopted. Predictions obtained here reproduced very well numerical results obtained by other authors.
The control algorithm was initially applied to suppress rectangular cylinder undergoing large rotations
induced by the fluid flow was also analyzed taking into account different control techniques and

CILAMCE 2019
Proceedings of the XL Ibero-Latin-American Congress on Computational Methods in Engineering, ABMEC,
Natal/RN, Brazil, November 11-14, 2019.
SANGALLI, L.A., AGUIRRE, M.A., FUNEZ, M.L., BRAUN, A.L.

structural appendices to stabilize the structural response, where uncontrolled and actively controlled
winglets were employed. The influence of the control matrices S and R on the control efficiency was
investigated using different S/R ratios. It was observed that vibration amplitudes and control forces
obtained from a given number of simulations can be plotted as functions of S and R values in order to
evaluate efficiency in terms of displacement reduction and energy required by the control system. When
structures with uncontrolled appendices were considered, stable responses were obtained sometimes for
subcritical flow speeds. However, for structures undergoing dynamic instability, only actively controlled
appendices were able to suppress instabilities. In these cases, synchronization between the main
structure and appendices was observed and the winglets vibrated in opposite direction. On the other
hand, it was concluded that the use of thick airfoils as aerodynamic appendices may lead to significant
increase of aerodynamic forces on the structure. Therefore, thin airfoils are recommended. The mesh
motion scheme utilized here worked well in general, although some difficulties were observed in the
mesh region between the main structure and appendices, where elements near the structure failed by
excessive distortion when large relative rotations were observed. For future works, some modifications
must be done in the mesh motion scheme in order to obtain a more local evaluation of mesh velocity
(Eq. 6), especially for nodes located between bodies with large relative rotations.

Acknowledgments

The authors would like to thank CNPq and CAPES for the financial support. The present research
was developed using computational resources provided by CESUP/UFRGS, NACAD/UFRJ and
CENAPAD/UNICAMP.

References

[1] E. Simiu E, R.H. Scanlan. Wind effects on structures: fundamentals and applications to design, third
ed., John Wiley and Sons, New York, 1996.
[2] R.D. Blevins RD. Flow-induced Vibration, second ed., Van Nostrand Reinhold, New York, 1990.
[3] T.T. Soong. Active Structural Control: Theory and Practice, Longman Scientific and technical,
Essex, UK, 1990.
[4] L. Meirovitch. Dynamics and Control of Structures, John Wiley and Sons, New York, 1990.
[5] K.H. Ostenfeld. A. Larsen, 1992. Bridge engineering and aerodynamics, In: Proceedings of the 1st
International Symposium on Aerodynamics of Large Bridges, pp. 3–22.
[6] H. Kobayashi, H. Nagaoka. Active control of flutter of a suspension bridge. Journal of Wind
Engineering and Industrial Aerodynamics, n. 41, pp. 143–151, 1992.
[7] K. Wilde, Y. Fujino. Aerodynamic control of bridge deck flutter by active surfaces. Journal of
Engineering Mechanics, n. 124, pp. 718–727, 1998.
[8] S.D. Kwon, S.P. Chang. Suppression of flutter and gust response of bridges using actively controlled
edge surfaces. Journal of Wind Engineering and Industrial Aerodynamics, n. 88, pp. (2000) 263–281.
[9] H.I. Hansen, P.T. Christensen. Active flap control of long suspension bridges. Journal of Structural
Control, n. 8, pp. 33-82, 2001.
[10] K. Wilde, P. Omenzetter, Y. Fujino. Suppression of bridge flutter by active deck-flaps control
system. Journal of Engineering Mechanics, n. 127, pp. 80-89, 2001.
[11] A.D. Nissen, P.H. Sorensen, O. Jannerup. Active aerodynamic stabilization of long suspension
bridges. Journal of Wind Engineering and Industrial Aerodynamics, n. 92, pp. 829-847, 2004.
[12] K. Gouder, X. Zhao, D.J.N. Limebeer, J.M.R. Graham. Experimental aerodynamic control of a
long-span suspension bridge section using leading- and trailing-edge control surfaces. Transactions on
Control Systems Technology, n. 24, pp. 1441-1453, 2016.
[13] K.N. Bakis, M. Massaro, D.J.N. Limebeer, M.S. Williams. Aeroelastic control of long-span
suspension bridges with controllable winglets. Journal of Structural Control and Health Monitoring, n.
23, pp. 1417–1441, 2016.

CILAMCE 2019
Proceedings of the XL Ibero-Latin-American Congress on Computational Methods in Engineering, ABMEC,
Natal/RN, Brazil, November 11-14, 2019.
A Fluid-Structure Interaction Model for Evaluation of Flow Induced Vibrations in Actively Controlled Structures

[14] K. Li, L. Zhao, Y.J. Ge, Z.W. Guo. Theoretical framework of feedback aerodynamic control of
flutter oscillation for long-span suspension bridges by the twin-winglet system. Journal of Wind
Engineering and Industrial Aerodynamics, n. 145, pp. 166–177, 2015.
[15] K. Li, L. Zhao, Y.J. Ge, Z.W. Guo. Flutter suppression of a suspension bridge sectional model by
the feedback controlled twin-winglet system. Journal of Wind Engineering and Industrial
Aerodynamics, n. 168, pp. 101–109, 2017.
[16] K. Li, L. Zhao, Y.J. Ge, Z.W. Guo. Numerical simulation of feedback flutter control for a single-
box-girder suspension bridge by twin-winglet system. Journal of Wind Engineering and Industrial
Aerodynamics, n. 169, pp. 77-93, 2017.
[17] K. Wilde, Y. Fujino, T. Kawakami. Analytical and experimental study on passive aerodynamic
control of flutter of a bridge deck. Journal of Wind Engineering and Industrial Aerodynamics, n. 80, pp.
105-119,1999.
[18] P. Omenzetter, K. Wilde, Y. Fujino. Study of passive deck-flaps flutter control system on full bridge
model. I: Theory. Journal of Engineering Mechanics, n. 128, pp. 264-279, 2002.
[19] P. Omenzetter, K. Wilde, Y. Fujino. Study of passive deck-flaps flutter control system on full bridge
model. II: Results. Journal of Engineering Mechanics, n. 128, pp. 280-286, 2002.
[20] D.H. Phan, N.T. Nguyen. Flutter and buffeting control of long-span suspension bridge by passive
flaps: experiment and numerical simulation. International Journal of Aeronautical and Space Sciences,
n. 14, pp. 46-57, 2013.
[21] K.N. Bakis, D.J.N. Limebeer, M.S. Williams, J.M.R. Graham. Passive aeroelastic control of a
suspension bridge during erection. Journal of Fluids and Structures, n. 66, pp. 543-570, 2016.
[22] A.J. Chorin. A numerical method for solving incompressible viscous flow problems. Journal of
Computational Physics, n. 2, pp. 12-26, 1967.
[23] P. Nithiarasu. An efficient artificial compressibility (AC) scheme based on the characteristic based
split (CBS) method for incompressible flows. International Journal for Numerical Methods in
Engineering, n. 56, pp. 1815-1845, 2003.
[24] J. Smagorinsky. General circulation experiments with the primitive equations: the basic
experiment. Monthly Weather Review, n. 91, pp. 99-135, 1963.
[25] S. Murakami. Current status and future trends in Computational Wind Engineering. Journal of
Wind Engineering and Industrial Aerodynamics, n. 67-68, pp. 3-34, 1997.
[26] B. Blocken. 50 years of Computational Wind Engineering: past, present and future. Journal of Wind
Engineering and Industrial Aerodynamics, n. 129, pp. 69-102, 2014.
[27] K. Ogata. Modern Control Engineering, fifth ed., Pearson Prentice Hall, New York, 2010.
[28] J. Donea. A Taylor-Galerkin method for convective transport problems. International Journal for
Numerical Methods in Engineering, n. 20, pp. 101-119, 1984.
[29] M. Kawahara, H. Hirano. A finite element method for high Reynolds number viscous fluid flow
using two step explicit scheme. International Journal for Numerical Methods in Fluids, n. 3, pp. 137-
163, 1983.
[30] A.L. Braun, A.M. Awruch. Finite element simulation of the wind action over bridge sectional
models: Application to the Guamá River Bridge (Pará State, Brazil). Finite Elements in Analysis and
Design, n. 44, pp. 105-122, 2008.
[31] A.L. Braun, A.M. Awruch. Aerodynamic and aeroelastic analyses on the CAARC standard tall
building model using numerical simulation. Journal of Computers and Structures, 87 ,pp.564-581, 2009.
[32] A.L. Braun, A.M. Awruch. Aerodynamic and aeroelastic analysis of bundled cables by numerical
simulation. Journal of Sound and Vibration, n. 284, pp. 51-73, 2005.
[33] A. Larsen. Aerodynamic aspects of the final design of the 1624 m suspension bridge across the
Great Belt. Journal of Wind Engineering and Industrial Aerodynamics, n. 48, pp. 261-285, 1993.
[34] S. Kuroda. Numerical simulation of flow around a box girder of a long span suspension bridge.
Journal of Wind Engineering and Industrial Aerodynamics, n. 67&68, pp. 239-252, 1997.

CILAMCE 2019
Proceedings of the XL Ibero-Latin-American Congress on Computational Methods in Engineering, ABMEC,
Natal/RN, Brazil, November 11-14, 2019.

Вам также может понравиться