Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

J. Wind Eng. Ind. Aerodyn.

100 (2012) 30–37

Contents lists available at SciVerse ScienceDirect

Journal of Wind Engineering


and Industrial Aerodynamics
journal homepage: www.elsevier.com/locate/jweia

Coupling between structural and fluid dynamic problems applied to vortex


shedding in a 90 m steel chimney
Ali Vasallo Belver a,n, Antolı́n Lorenzana Ibán b, Carlos E. Lavı́n Martı́n a
a
CARTIF Centro Tecnológico, Parque Tecnológico de Boecillo, parcela 205, 47151 Boecillo (Valladolid), Spain
b
ITAP, University of Valladolid, Paseo del cauce 59, 47011 Valladolid, Spain

a r t i c l e i n f o a b s t r a c t

Article history: A simplified fluid–structure interaction approach is used to study the dynamic behaviour of a particular
Received 2 December 2010 90 m steel chimney under vortex-induced vibrations. Navier–Stokes equations for incompressible flow
Received in revised form are solved in 2D in several transverse planes of the line-like structure. The resultant pressure field is
13 October 2011
introduced using standard FEM interpolation techniques, together with the dynamical behaviour of the
Accepted 15 October 2011
structure and its boundary conditions. A fractional step scheme is used to solve the fluid field. In each
Available online 22 November 2011
fluid plane, the displacements are taken into account considering an Arbitrary Lagrangian Eulerian
Keywords: approach. The stabilisation of incompressibility and convection is achieved through orthogonal quasi-
Steel chimney static subscales, an approach that is believed to provide a first step towards turbulence modelling. In
Fluid–structure interaction
order to solve the structural problem, a special one-dimensional element for thin walled cross-section
Vortex induced vibrations
beams is implemented. The standard second-order Bossak method is used for the time integration of
Across-wind vibration
the dynamic problem. The wind is modelled as an incompressible fluid acting on the structure in a
series of planes, transverse to the structure that are supposed to be independent among them. For each
period of time, the fluid problem is solved, the aeroelastic analysis is carried out and the geometry of
the mesh of each fluid plane is updated according to the structure displacements.
& 2011 Elsevier Ltd. All rights reserved.

1. Introduction Research carried out in latest years about analytical and semi-
empirical models to study the dynamic response in vertical
Multidisciplinary methods for fluid–structure interaction (FSI) structures due to the vortex shedding have followed two distinct
can be used to describe a wide range of interesting problems in lines. The first one was aimed at determining the maximum
engineering. Essentially, FSI phenomena are caused by the pres- effects due to along-wind response, whereas the second extends
sure forces from the flow around a structure resulting in a original methods from along-wind response to cross-wind and
deformation of the structure, which in turn alters the flow in a torsional responses.
dynamic way. In particular, vibrations generated by vortex shed- The first research line was focused on along-wind response and
ding are of practical interest to civil and industrial engineering derived from the observation that equivalent static force, as
and slender structures such as bridges, tall buildings, industrial defined by Davenport (1967), usually produces correct mean
chimneys or cables are designed according to rules for cross-wind maximum displacements but may give rise to other effects (above
vibrations. all bending moments and shear forces) marked by noteworthy
The significance of vortex-induced vibrations (VIVs) and the errors. Efforts were made by Kasperski (1992) and Holmes (1994)
complexity of the cross-wind forcing mechanisms have led to a to develop new analytical methods for defining the equivalent
considerable amount of research work, either experimental or static force in a more realistic and simplified manner. Analogous
theoretical, the latter through nonlinear-oscillator modelling methods were applied by several authors to different structural
approaches and, recently, also through Computational Fluid types (Holmes, 2002), sometimes arriving at a closed form solution
Dynamics (CFD). In the review by Williamson and Govardhan (Dyrbye and Hansen, 1997). The progress of analytical methods for
(2008) fundamental results and discoveries concerning VIV are studying the wind-induced response of structures, with special
discussed. regard to cantilever vertical structures, was presented by Solari
(2002).
The second research line was aimed at determining the three-
dimensional wind-induced response of structures. With reference to
n
Corresponding author. Tel.: þ34 983546504. slender structures and structural elements in the atmospheric
E-mail address: alivas@cartif.es (A.V. Belver). boundary layer, Piccardo and Solari (1996) schematised along-wind,

0167-6105/$ - see front matter & 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jweia.2011.10.007
A.V. Belver et al. / J. Wind Eng. Ind. Aerodyn. 100 (2012) 30–37 31

cross-wind and torsional actions, by quasi-steady theory, as a linear


combination of the longitudinal, lateral and vertical turbulence
components; cross-wind forces and torsional moments due to the
vortex shedding were superposed considering these as independent
of the turbulence actions. Along-wind, cross-wind and torsional
responses, dealt with as uncoupled and only dependent on the
related fundamental mode shapes, were determined in closed form
(Piccardo and Solari, 2000) by the generalised equivalent spectrum
technique (Piccardo and Solari, 1998).
In the recent years, the researchers were focused on refining
the analytical models to predict accurate results for vortex
shedding at high Reynolds numbers (Sampaio and Coutinho,
2000), to account for mode shape corrections (Zhou et al., 2002)
or to account for fatigue induced by vortex shedding phenomenon
(Repetto and Solari, 2004). Other authors, like Gorski (2009) or Fig. 1. Strouhal number vs. Reynolds number for circular cylinders (Techet, 2005).

Verboom and Koten (2010), improved the known model of


Vickery–Basu, Basu and Vickery (1983), by including some com- along-wind, Fx(t), and across-wind, Fx(t), directions, which can be
plements to study the along- and across-wind response of tall rendered dimensionless and expressed in terms of lift and drag
industrial chimneys due to the vortex excitation. coefficients as
On the other hand, besides these 2D analytical and semi-
F y ðtÞ
empirical models for the study of vibrations caused by vortex C L ðtÞ ¼ ð1Þ
1=2rU 2 d
shedding, there are some models based on the CFD. Unlike the
other methods, the pure numerical methodology used to solve
F x ðtÞ
aeroelastic problems in structural engineering is characterised by C D ðtÞ ¼ ð2Þ
1=2rU 2 d
not requiring experimental data, although model updating tech-
niques should be used. When a vortex is formed on one side of the structure, wind
The first numerical researches on air flow around a cylinder speed increases on the other side and, according to theory of
were carried out in 1969 by Son and Hanratty (1969). Also Braza Bernoulli, this cause a reduction in the pressure value. Thus, the
et al. (1986) used the finite volume method to analyse the vortex structure is subjected to a lateral force or vortex shedding force
shedding in cylinders, with a Reynolds number less than 1000. away from the side where a vortex is formed. As vortices are shed
Later, Dawes (1993) studied the same problem with an adaptive alternately from one to the other side, a harmonic across-wind
method based on the finite volume method, including the ability varying lateral load with the same frequency as the frequency of
of refining the mesh in function of the solution obtained, to get a the vortex shedding ns is formed:
computationally economical solution. Other researchers, like Rodi
U
(1997), studied the vortex shedding phenomenon in non-aero- ns ¼ St ð3Þ
d
dynamic bodies. Larsen and Walther (1998) used the discrete
vortex method and were able to calculate aerodynamic para- where d is any relevant transversal dimension of the structure
meters, like the lift and drag coefficients or the Strouhal number, and St is the Strouhal number, which depends on the shape of the
for different cross-sections. Other authors, like Jan and Sheu cross section, the surface roughness and the Reynolds number.
(2004) and Prasanth and Mittal (2008), have also used the finite Experimental values for cylinders are shown in Fig. 1.
element method to analyse the vortex shedding phenomenon. For lightly damped structures, which are free to oscillate, large
The present work focuses on the case of FSI problems amplitude vibrations may occur if the dominating frequency of
characterised by long and slender structures where the local vortex shedding is in resonance with one of the natural frequen-
deformation is negligible compared to the displacement. This is cies for the structure bending in a mode in the cross-wind
the case e.g. for masts and industrial chimneys, and means that direction. Vortex shedding also generates a periodic along-wind
the structure can be accurately modelled with a beam finite load with a frequency of 2ns, but being the amplitude of this force
element formulation. The air is assumed to be an incompressible much narrower than the one in the across-wind direction, its
fluid acting on the structure in a series of planes that are effects are usually neglected. Although in the numerical methods,
independent among them, and transverse to the structure. With based on CFD for steady winds, this variation on the along-wind
these assumptions, VIVs are studied by coupling a one-dimen- load appears, for atmospheric winds it is always masked by
sional structural model to the fluid flow using a simplified FSI turbulences and gusts, not considered in the model. Besides, it
algorithm and the across-wind response for a particular 90 m is known (Ruscheweyh, 2009) that standard turbulences and
steel chimney under the wind action is shown. gusts are not able to increase the across-wind response up to its
The paper is organised as follows. In Section 2 the physical maximum, so numerical results based on steady winds could
background about vortex-induced vibrations is briefly explained. be appropriated for predicting the across-wind behaviour, as
Section 3 describes the structural model, Section 4 the fluid model intended in this paper.
and Section 5 the fluid–structure coupling algorithm. Section 6
presents the example and Section 7 concludes by summarizing
the main aspects of the work. 3. Structural model

An analytical one-dimensional model was developed to


2. Physical background describe the behaviour of 3D thin-walled beams. Detailed
description of the model can be seen in the work of Mediavilla
Vortex-induced vibrations occur when vortices are shed alter- et al. (2007). 14 degrees of freedom are used for each straight
nately from opposite sides of the structure. For steady flows, this beam element (2 nodes) accounting for extension, shearing,
gives rise to a fluctuating load Fext ðz,tÞ with components in the bending, bimoment and torsion. The methodology proposed uses
32 A.V. Belver et al. / J. Wind Eng. Ind. Aerodyn. 100 (2012) 30–37

y

jy

 n
ix x s
O 
ds j

ds
dz O' dz

s i


k
s z

kz
z

Fig. 2. Coordinate systems.

a Love–Kirchhoff shell model to relate the stresses and strains in equations can be expressed as
the shell. In these relations, the generalised beam deformations @v 2
corresponding to the Navier–Bernoulli and Vlasov models are þðvUrÞvnr v þ rp ¼ b ð7Þ
@t
introduced through geometric considerations but deformations of
the cross-section as ovalization are neglected, according to beam rUv ¼ 0 ð8Þ
element modelling.
The solution procedure is based on a fractional step scheme
Fig. 2a shows the geometry and coordinate system for a thin-
(Codina, 2001), namely, a second order algorithm based on the
walled beam with an arbitrary cross-section, while Fig. 2b shows
Crank–Nicolson discretization for the viscous and convective
similar information for any shell portion composing the beam. z is
terms and a second order pressure splitting, leaving the pressure
the longitudinal coordinate for the beam, which is the line-like
gradient at a given time level in the first step and computing its
structure, and no curvature is considered in this direction.
increment in the second one (Dadvand et al., 2010).
To describe the kinematics, Uz, Ux and Uy will be considered to
The fractional step scheme is described briefly below, in its
be the displacement of the beam along Oz, Ox and Oy global axes,
basic form, without the introduction of any stabilisation for the
respectively; bx and by will be the rotations of the cross-section
pressure term. The starting point is the spatial discretization of
beam about the Ox and Oy axes, respectively, and f will be the
the Navier–Stokes Eqs. (7) and (8). The matrix form of the
twist of each section around the Oz axis. The local shell displace-
problem using the Galerkin approach is given by
ments in the O0zns system will be indicated as u, vn and vt. Merging
the Love–Kirchhoff theory for shells together with the Navier– Vn þ 1 Vn
M þ KðVn þ y ÞVn þ y þ GPn þ 1 ¼ Fn þ y ð9Þ
Bernoulli and Vlasov hypothesis for beams (Monleón, 2001; Jung dt
et al., 2002; Jung and Park, 2005 ), a matrix relation between
external forces, Fext ðz,tÞ, and generalised displacements, x, at both DVn þ 1 ¼ 0 ð10Þ
ends of the beam element can be achieved in the form: where V and P are the arrays of nodal velocities and pressures,
M x€ þ C x_ þKx ¼ Fext ðz,tÞ ð4Þ respectively. If the node indexes are denoted with superscripts a,
b, the space indexes with subscripts i, j, and the standard shape
where K is the stiffness matrix, M is the mass matrix and C is the function of node a by Na, the components of the arrays involved in
damping matrix. The generalised displacement vector is these equations are
n ot
x ¼ U z U x U y f bx by f,z ð5Þ Mab a b
ij ¼ ðN ,N Þdij ðdij is the Kronecker dÞ,
1
where f,z is the axial warping and the generalised force vector is KðVn þ y Þab a nþy
ij ¼ ðN ,v UrN b Þdij þ ðN a ,ðrUvn þ y ÞN b Þdij þ nðrN a , rN b Þdij ,
2
given by
n ot Gab a b
i ¼ ð@i N ,N Þ,
Fext ðz,tÞ ¼ N V x V y T s M x M y M o ð6Þ Fai ¼ /N a ,f i S,

where N is the axial force; Vx and Vy are the transverse shear Dab a b
j ¼ ðN ,@j N Þ:

forces; Mx and My are the bending moments about the x and the y The fractional step scheme is based on the introduction of an
directions, respectively; Ts is the St. Venant twisting moment and nþ1
auxiliary variable V~ , which allows to rewrite the time deriva-
Mo is the Vlasov bimoment. tive as
Once defined the governing equations in the local coordinate
nþ1 nþ1
system, standard assembly procedures are used and boundary Vn þ 1 Vn Vn þ 1 V~ V~ Vn
¼ þ ð11Þ
conditions imposed. The resulting system of differential equations dt dt dt
is solved using the second-order Bossak scheme. and the fractional step scheme can be presented as
nþ1
V~ Vn
M þKðVn þ y ÞVn þ y þGPn ¼ Fn þ y ð12Þ
4. Computational fluid dynamics dt
nþ1
Assuming that the air is an incompressible fluid, typically Vn þ 1 V~
M þ GðPn þ 1 Pn Þ ¼ 0 ð13Þ
accepted in the civil engineering applications, the Navier–Stokes dt
A.V. Belver et al. / J. Wind Eng. Ind. Aerodyn. 100 (2012) 30–37 33

DVn þ 1 ¼ 0 ð14Þ the friction resistance, where a time-accurate viscous flow solver is
needed. In our case, the interest focuses on the behaviour of a beam
Using the following approximation:
subjected to a flow orthogonal to the beam axis. Under this
nþy nþy assumption, the fluid domain was modelled with a number of
KðVn þ y ÞVn þ y  KðV~ ÞV~ ð15Þ
independent planes of fluid where the problem can be solved
where separately. Conceptually, the CFD solution on each fluid plane
nþy nþy provides a force density, acting on the cross-section beam, which is
V~ ¼ yV~ þ ð1yÞVn ð16Þ
obtained by integrating the pressure of the fluid over the boundary of
nþ1
and expressing V in terms of V~
nþ1
using Eq. (13) and inserting the cross-section. So, a time-varying distributed load over the
the result in Eq. (14), the set of equations to be solved is structure is obtained by interpolation between consecutive fluid
nþ1 planes. The deformation of the structure (and the necessary move-
V~ Vn nþy nþy
ment of the fluid mesh) provides a kind of correlation between the
M þ KðV~ ÞV~ þGPn ¼ Fn þ y ð17Þ
dt motion of the different sections and fluid planes.
nþ1 It is interesting to make a conceptual consideration on the
dtDM1 GðPn þ 1 Pn Þ ¼ DV~ ð18Þ nature of the coupling. The beam formulation makes use of the
nþ1
small strain hypothesis, which implies that the reference and
Vn þ 1 V~ deformed configuration are considered to coalesce in writing the
M þ GðPn þ 1 Pn Þ ¼ 0 ð19Þ
dt equilibrium. On the other hand, the motion of the cross-section is
Note that DM  1G can be replaced by the Laplacian operator L and obtained exactly once given the motion of the corresponding axis,
hence the fractional step scheme takes the form: without taking advantage of the small strain hypothesis. This
implies that the loads acting on the structure will be allowed
 Step A depending in a non-linear way on the motion of the beam. This
nþ1
feature may become important for the cases in which the model
V~ Vn nþy nþy moves at the limits of the small strain formulation, which is not
M þKðV~ ÞV~ þ GPn ¼ Fn þ y ð20Þ
dt the case for the example to be shown.
Finally, it is needed to choose a suitable fluid–structure
 Step B coupling algorithm. In this work, the aeroelastic problem is
nþ1 characterised by large Reynolds numbers and flow separation
dtLðPn þ 1 Pn Þ ¼ DV~ ð21Þ around the circular cross-section. It is well known that loose
coupling procedures (Dadvand et al., 2010) are very efficient for
 Step C such a problem. In the coupling scheme used to solve the fluid–
nþ1
structure interaction problem, the structural solution is inte-
Vn þ 1 V~ grated in the fractional step procedure and the Arbitrary Lagran-
M þ GðPn þ 1 Pn Þ ¼ 0 ð22Þ
dt gian-Eulerian (ALE) formulation is used to take into account the
movement of the fluid mesh.
The coupling scheme is based on the following steps
The fluid solver used in this work is based on the use of
Orthogonal Sub-Grid Scale (OSS) stabilisation, as described for 1. Solve the structural problem for an initial guess pressure,
example in Codina (2002). Such stabilisation falls into the Xnn þ 1s ¼ Xnn þ 1s ðPnn þ 1 Þ.
category of Variational Multi-Scale (VMS) techniques, which have 2. Move the fluid domain according to the structure motion,
been deeply investigated during the last decade. An important Xnn þ 1 ¼ Xnn þ 1 ðPnn þ 1 Þ.
nþ1
property of such techniques is that, aside of allowing a stable 3. Step A, V~ .
Finite Element solution, they tend to introduce a dissipation into 4. Step B, Pn þ 1. Impose pressure boundary conditions.
the solution that acts similarly to a turbulence model. A mathe- 5. Solve the structural problem for the calculated pressure,
matical discussion on the effectiveness of such models in model- Xn þ 1s ¼ Xn þ 1s ðPn þ 1 Þ.
ling the turbulence can be found for example in the work of 6. Step C, Vn þ 1. Impose velocity boundary conditions.
Codina et al. (2010), Principe et al. (2010) or Oñate et al. (2007),
while some empirical evidence of their success in simulating high The structural model, the CFD abilities and the former numerical
Reynolds number flows around bluff bodies can be found in the methods have been implemented in Kratos (Dadvand et al., 2010,
work of Rossi and Oñate (2010). Belver et al., 2010) (an object-oriented framework for developing
finite element codes for multi-disciplinary applications).

5. Fluid-structure coupling
6. Steel chimney under wind action
To perform the interaction between the structural and fluid
problems, it is necessary to identify the geometry of the cross- Across-wind vibrations caused by vortex shedding in a parti-
section of the beam and the distribution of pressure or force cular 90 m steel chimney are studied using the former approach.
density on its boundary. Under the usual assumptions, the motion For low damping ratios, the stack is likely to oscillate strongly
of the beam axis describes the motion of the whole cross-section, (Areemit and Warnitchai, 2001) in a plane perpendicular to the
which can be imagined as a rigid surface, which follows the mean wind direction for speeds around 9 m/s. The diameter of the
translation and the rotation of the beam axis. Any section of finite chimney varies from 2.20 m at the top to 5.50 m at the base and
dimensions in the xy plane is associated to the corresponding the thickness of the shell from 22 mm to 12 mm, as shown in
point of the one-dimensional beam element, assumed to be Fig. 3. The average mass per unit height is 1916 kg/m and the
oriented in the Oz axis. diameter dm ¼4.0 m is adopted as reference to get dimensionless
Current work focuses on bluff bodies, in particular a circular cross- values for displacements. The elastic properties are Young’s
section body, with a clear predominance of the shape resistance over modulus E¼2.1  1011 Pa, Poisson’s ratio n ¼0.3 and density
34 A.V. Belver et al. / J. Wind Eng. Ind. Aerodyn. 100 (2012) 30–37

Table 1
Inflow velocities and Reynolds numbers.
10 m 2.20 m; 12 mm
h [m] vx [m/s] Re L1 [m] L2 [m] W [m]

20 m 30 10.62 3.68Eþ 06 31.2 176.8 104


3.20 m; 15 mm 60 11.78 3.22Eþ 06 24.6 139.4 82
80 12.30 2.62Eþ 06 19.2 108.8 64
90 12.52 1.84Eþ 06 13.2 74.8 44

90 m 30 m 4.10 m; 19 mm

v 

S
u d W

30 m 5.20 m; 22 mm W/2

y
L1 L2
Fig. 3. (a) 90 m steel chimney in Rayong, (b) basic dimensions of the chimney
x
(Areemit and Warnitchai, 2001).
Fig. 5. Boundary conditions’ definition.

Fig. 4. Chimney model under wind action.

Fig. 6. Computational mesh.


r ¼7850 Kg/m3. The resultant first natural frequencies ne1, ne2
are 0.91 Hz and 3.6 Hz. Using the special beam element presen-
ted in Section 3, and considering the simplicity of the the dimensions of the meshed fluid domains, as marked in Fig. 5.
geometry, just 4 elements are enough to capture the dynamic With these dimensions, mesh convergence and time integration
response. requirements are fulfilled using 29,694 triangle fluid elements
Dynamic behaviour is very dependent on structural damping. (Fig. 6) and a time step of 0.002 s.
Rayleigh damping, which is commonly used in structural The boundary conditions for the fluid problem are the follow-
dynamics, is assumed so that C ¼ aMþ bK. By taking b ¼0, higher ing: traction-free boundary condition at the edge Gd, no-slip
modes of vibration would be less damped than the first ones, condition at the cross-section boundary S of the structure, vy ¼0
which is the standard situation in this kind of structures where at Go, and the specified (Table 1) inflow velocity (vx, 0) at edge Gu.
the first mode of vibration is predominant. As a reference for the The time step value depends on the period of the first mode of
ideal chimney with high damping ratio, a value for a ¼0.2 has vibration of the structure, the wind speed and the characteristic
been taken. Results for half and a quarter of this value are also size of the mesh, which is defined as the smallest dimension of
presented. the finite elements. Taking the appropriated time step is essential
The density of fluid (air) is ra ¼1.21 kg/m3 and its dynamic to solve correctly both the dynamic structural problem and the
viscosity m ¼1.8  10  5 Ns/m2. Influence of the number of fluid fluid field.
planes used to model the wind action was studied by Belver The across-wind behaviour of the chimney under the specified
(2009), considering several cases with different fluid planes wind action is studied for 40 s assuming the chimney is at rest at
located at different heights. For this case, convergence is achieved t¼0 s and after this period of time, the fluid is stopped at once, so
using just four planes of fluid located at 30, 60, 80 and 90 m from the free vibration of the structure can be studied and structural
the clamped edge. damping checked using free decay technique.
The wind profile is characterised by the reference wind The across-wind displacement time history of the cross-sec-
velocity, u10. The following exponential expression is assumed tion of the chimney located at the tip is plotted in Fig. 7. While
for the profile of the wind speed in the z-axis (Fig. 4): wind is acting, the chimney vibrates at different frequencies
 z e (Fig. 8a) as a consequence of the vortex shedding in different
uðzÞ ¼ u10 ð23Þ planes. Once the wind stops, the chimney only vibrates according
10
to its natural frequencies (Fig. 8b).
where e¼0.15. For the speed of interest, u10 ¼ 9 m/s, acting in the Fig. 9 shows the oscillation of the lift coefficient for the same
x-direction, the inflow velocities in each plane of fluid are shown cross-section. The corresponding FFT analysis of the lift coefficient
in Table 1 together with the corresponding Reynolds numbers and is shown in Fig. 10.
A.V. Belver et al. / J. Wind Eng. Ind. Aerodyn. 100 (2012) 30–37 35

0.16
0.12

disply / dm x 102
0.08
0.04
0
-0.04
-0.08
-0.12
-0.16
0 5 10 15 20 25 30 35 40 45 50
t [s]

Fig. 7. Across-wind vibration of the chimney.

0.16 0.004
ns1= 0.55Hz
0.14 0.0035
ns2 = 0.89Hz ne1 = 0.91Hz
0.12 0.003
ns3 = 1.13Hz
0.1 0.0025

Rm
Rm

0.08 0.002
0.06 0.0015
0.04 0.001
ns4 = 1.62Hz ne2 = 3.6Hz
0.02 0.0005
0 0
0 1 2 3 4 5 0 2 4 6 8 10
fr [Hz] fr [Hz]

Fig. 8. FFT analysis for the force vibration, 0 ot o 40 s (a) and free vibration, 40 ot o 50 s (b).

0.65
0.5
0.35
0.2
CL

0.05
-0.1
-0.25
-0.4
-0.55
0 5 10 15 20 25 30 35 40 45 50
t [s]
Fig. 9. Lift coefficient.

3500 Table 2
Vortex shedding frequencies and frequency response of the chimney.
ns =1.62Hz
3000 Diameter [m] Velocity [m/s] Theoretical freq. [Hz] Frequency resp. [Hz]

5.2 10.62 0.58 0.55


2500
4.1 11.78 0.82 0.89
3.2 12.30 1.09 1.13
2000 2.2 12.52 1.62 1.62
Rm

1500
As the frequency of the oscillation of the lift coefficient coin-
1000 cides with the vortex shedding frequency, the Strouhal number can
be evaluated considering the Strouhal relation, Eq. (3):
500 d 2:2
St ¼ ns ¼ 1:62 ¼ 0:284 ð24Þ
U 12:52
0
The Root Mean Squared (RMS) value of the lift coefficient
0 1 2 3 4 5
(Fig. 9) is 0.3. Both this value and the Strouhal number, St¼ 0.284,
fr [Hz]
agree with the experimental results (see Fig. 1) for the corre-
Fig. 10. FFT analysis for the lift coefficient. sponding Re (1.84  106)
36 A.V. Belver et al. / J. Wind Eng. Ind. Aerodyn. 100 (2012) 30–37

Taking into account the Strouhal relation and considering Acknowledgements


St¼0.284, the theoretical vortex shedding frequencies are given
in Table 2 for the different cross-sections of the chimney. It is Authors wish to acknowledge to the International Committee
observed that the theoretical frequencies are near the computed on Industrial Chimneys (CICIND), the International Center for
response frequencies of the chimney (Fig. 8). Numerical Methods in Engineering (CIMNE) and Kratos develop-
In order to compare the results obtained by means of the FSI ment team for all the technical support.
approach considered in this work, the across-wind response
amplitude will be evaluated using the so called ‘‘Correlation References
Length Model’’ (Holmes, 2001, Ruscheweyh, 2009). Although
some other methodologies have been proposed (Muñoz et al., Areemit, N., Warnitchai, P. 2001, Vibration suppression of a 90-m-tall steel stack
2009) and included in some international design codes, results using a high-damping tuned mass damper, In: Proceedings of the Eighth East
obtained are very different, even more than 5 times between Asia-Pacific Conference on Structural Engineering and Construction, Nanyang
Technological University, Singapore.
them, and therefore a reliable comparison is not possible. Apply-
Basu, R.I., Vickery, B.J., 1983. Across-wind vibrations of structures of circular cross-
ing the ‘‘Correlation Length Model’’, included in DIN 4133 and in section. Part II. Development of a mathematical model for full-scale applica-
EN 1991-1-4, the standard deviation of the across-wind vibration tion. Journal of Wind Engineering and Industrial Aerodynamics 12 (1), 75–97.
amplitude yRMS can be calculated by Belver, A.V. 2009, Analysis of aeroelastic vibrations in slender structures under
wind loads, Ph.D. Dissertation, University of Valladolid, Spain.
R Belver, A.V., Mediavilla, A.F., Iban, A.L., Rossi, R., 2010. Fluid–structure coupling
ra pC L 0h u2 ðzÞdðzÞfðzÞ dz analysis and simulation of a slender composite beam. Science and Engineering
yRMS ¼ Rh 2
ð25Þ of Composite Materials 17 (1), 47–77.
2dð2pne1 Þ2 0 mðzÞf ðzÞ dz Braza, M., Chassaing, P., Minh, H.H., 1986. Numerical Study and Physical Analysis
of the Pressure and Velocity Fields in the near Wake of a Circular Cylinder.
where m(z) is the mass distribution along the chimney axis, d is Journal of Fluid Mechanics 165, 79–130.
the logarithmic decrement of damping, f(z) is the unity normal- Codina, R., 2001. Pressure stability in fractional step finite Element methods for
incompressible flows. Journal of Computational Physics 170 (1), 112–140.
ised mode shape. Codina, R., 2002. Stabilized finite element approximation of transient incompres-
The logarithmic decrement of damping is evaluated from the sible flows using orthogonal subscales. Computer Methods in Applied
across-wind response (Fig. 7) for more than 40 s using free-decay Mechanics and Engineering 191 (39–40), 4295–4321.
Codina, R., Principe, J., Avila, M., 2010. Finite element approximation of turbulent
technique, resulting in 0.6. By applying Eq. (25) the resultant thermally coupled incompressible flows with numerical sub-grid scale mod-
across-wind response of the chimney is yRMS/dm  102 ¼0.095, elling. International Journal of Numerical Methods for Heat & Fluid Flow 20,
which is congruent with the RMS value of the across-wind 492–516.
Dadvand, P., Rossi, R., Oñate, E., 2010. An object-oriented environment for
response presented in Fig. 7, which is 0.062.
developing finite Element codes for multi-disciplinary applications. Archives
The estimated values, using Eq. (25), when damping ratios are half of Computational Methods in Engineering 17 (3), 253–297.
and a quarter of the considered one, are 0.196 and 0.412, respectively, Davenport, A.G., 1967. Gust loading factors. Journal of the Structural Division,
whereas computed values using the presented numerical technique ASCE 93, 11–34.
Dawes, W.N., 1993. Simulating Unsteady Turbomachinery Flows on Unstructured
are 0.117 and 0.260. Eq. (25) usually overestimates and the differ- Meshes which Adapt Both in Time and Space. International Gas Turbine and
ences between estimated values and computed ones are similar to Aeroengine Congress and Exposition, Cincinnati, Ohio.
the ones between estimated values and experimentally measured Dyrbye, C., Hansen, S.O., 1997. Wind Loads on Structures. Wiley, NY, USA.
Gorski, P., 2009. Some aspects of the dynamic cross-wind response of tall
values presented by Ruscheweyh (2009). Also computed values are industrial chimney. Wind and Structures 12 (3), 259–279.
congruent with the estimations presented in Areemit and Warnitchai, Holmes, J.D., 1994. Along-wind response of lattice towers: Part I-derivation of
2001 for the same chimney with different damping ratios. expressions for gust response factors. Engineering Structures 16, 287–292.
Holmes, J.D., 2001. Wind Loading of Structures. Spon Press, London.
Holmes, J.D., 2002. Effective static load distributions in wind engineering. Journal
of Wind Engineering and Industrial Aerodynamics 90, 91–109.
Jan, Y.J., Sheu, T.W.H., 2004. Finite element analysis of vortex shedding oscillations
from cylinders in the straight channel. Computational Mechanics 33 (2),
7. Conclusions 81–94.
Jung, S.N., Nagaraj, V.T., Chopra, I., 2002. Refined structural model for thin- and
A simplified numerical method for wind–structure interaction thick-walled composite rotor blades. AIAA Journal 40 (1), 105–116.
Jung, S.N., Park, I.J., 2005. Structural behavior of thin and thick-walled composite
analysis of line slender structures has been developed. Despite its
blades with multi-cell sections. AIAA Journal 43 (3), 572–581.
simplicity, the model can successfully capture the theoretical and Kasperski, M., 1992. Extreme wind load distributions for linear and nonlinear
experimental behaviour of slender structures, where the dynamic design. Engineering Structures 14, 27–34.
problem is solved not by modal superposition but using a time Larsen, A., Walther, J.H., 1998. Discrete vortex simulation of flow around five
generic bridge deck sections. Journal of Wind Engineering and Industrial
integration method. Frequencies in the across-wind direction Aerodynamics 77–78, 591–602.
match with the theoretical results, being St number and lift Mediavilla, A.F., Garcia, J.A.G., Belver, A.V., 2007. One-dimensional model for the
coefficient in good agreement with the experimental values. Also analysis of thin-walled composite beams. Revista Internacional de Métodos
Numéricos para Cálculo y Diseño en Ingenierı́a 23 (2), 225–242.
amplitudes are similar to the ones predicted by simplified Monleón, S., 2001. Análisis de Vigas, Arcos, Placas y Láminas: Una presentación
semiempirical methods. unificada, 2nd Ed. Servicio de Publicaciones Universidad Politécnica de
This work provides new opportunities for evaluation of the Valencia, ISBN: 978-84-7721-769-5.
Muñoz, C.J., Hernández, H., López, A., 2009. A comparison of cross-wind response
effects induced by vortex shedding. Forces are calculated in 2D evaluation for chimneys following different international codes.In: Proceed-
fluid planes using the theory of computational fluid dynamics ings of the Eleventh American Conference on Wind Engineering, International
coupled with a structural model. In spite of the simplifying Association for Wind Engineering, San Juan, Puerto Rico.
Oñate, E., Valls, A., Garcia, J., 2007. Computation of turbulent flows using a finite
assumption that the flow on each of the fluid planes is indepen- calculus-finite element formulation. International Journal for Numerical
dent of the flow on the neighbouring ones and no tip effects are Methods in Engineering 54, 609–637.
considered, the simulation provides an interesting alternative to Piccardo, G., Solari, G., 1996. Arefined model for calculating 3-D equivalent static
wind forces on structures. Journal of Wind Engineering and Industrial Aero-
experimental tests and clearly improves the classic 2D computa-
dynamics 65, 21–30.
tional approach. Piccardo, G., Solari, G., 1998. Generalized equivalent spectrum technique.
All these results bring new possibilities in the field of struc- Wind and Structures 1, 161–174.
tural assessment and control simulation: the effectiveness of Piccardo, G., Solari, G., 2000. 3-D wind-excited response of slender structures:
closed form solution. Journal of Structural Engineering, ASCE 126, 936–943.
passive control devices for the suppression of vibrations can Prasanth, T.K., Mittal, S., 2008. Vortex-induced vibrations of a circular cylinder at
now be estimated in early design stages. low Reynolds numbers. Journal of Fluid Mechanics 594, 463–491.
A.V. Belver et al. / J. Wind Eng. Ind. Aerodyn. 100 (2012) 30–37 37

Principe, J., Codina, R., Henke, F., 2010. The dissipative structure of variational Solari, G., 2002. The role of analytical methods for evaluating the wind-induced
multiscale methods for incompressible flows. Computer Methods in Applied response of structures. Journal of Wind Engineering and Industrial Aerody-
Mechanics and Engineering 199, 791–801. namics 90, 1453–1477.
Repetto, M.P., Solari, G., 2004. Directional wind-Induced fatigue of slender vertical. Son, J.S., Hanratty, T.J., 1969. Numerical solution for the flow around a cylinder at
structures. Journal of Structural Engineering, ASCE 130, 10032–11040. Reynolds number of 40, 200, 500. Journal of Fluid Mechanics 35, 369–386.
Rodi, W., 1997. Comparison of LES and RANS calculations of the flow around bluff Techet, A.H., 2005. 13.42 Lecture: Vortex Induced Vibrations, /http://web.mit.
bodies. Journal of Wind Engineering and Industrial Aerodynamics 69–71, eduS.
55–75. Verboom, G.K., Koten, H., 2010. Vortex excitation: three design rules tested on 13
Rossi, R., Oñate, E., 2010. Analysis of some partitioned algorithms for fluid– industrial chimneys. Journal of Wind Engineering and Industrial Aerodynamics
structure interaction. Engineering with Computers 27 (1), 20–56. 98 (3), 145–154.
Ruscheweyh, H., 2009. Experience with vortex-induced vibrations, CICIND Inter- Williamson, C.H.K., Govardhan, R., 2008. A brief review of recent results in vortex
national Committee on Industrial Chimneys. Technical Report 26 (2), 49–57. induced vibrations. Journal of Wind Engineering and Industrial Aerodynamics
Sampaio, P.A.B., Coutinho, A.L.G.A., 2000. Simulating vortex shedding at high 96 (6–7), 713–735.
Reynolds numbers. In: Proceedings of the Tenth International Offshore and Zhou, Y., Kareem, A., Gu, M., 2002. Mode shape corrections for wind load effects.
Polar Conference, Seattle, USA. Journal of Engineering Mechanics 128 (1), 15–23.

You might also like