Vortex Induced Vibration of A Wavy Circular Cylinder

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

Numerical Simulation of Vortex Induced Vibrations of a Flexibly

Mounted Wavy Cylinder at subcritical Reynolds number

Kai Zhanga,b , Hiroshi Katsuchib , Dai Zhoua,c,d,∗, Hitoshi Yamadab , Tao Zhangb , Zhaolong Hana,e
a School of Naval Architecture, Ocean and Civil Engineering, Shanghai Jiao Tong University, Shanghai, 200240, China
b Department of Civil Engineering, Graduate School of Urban Innovation, Yokohama National University, Yokohama,
2408501, Japan
c Collaborative Innovation Center for Advanced Ship and Deep-Sea Exploration (CISSE), Shanghai, 200240, China
d State Key Laboratory of Ocean Engineering, Shanghai Jiao Tong University, No. 800, Dongchuan Road, Shanghai,

200240, China
e Cullen College of Engineering, University of Houston, Houston, TX, 77004, USA

Abstract
Wavy cylinders have been proven effective in controlling the flow and reducing the fluid induced forces.
However, the hydro-elastic behavior of a flexibly mounted wavy cylinder remains unclear. The current
paper endeavours to present a systematic study of the flow around a spring mounted wavy cylinder
mainly at a moderate Reynolds number of 5000, the results of which are compared with a normal
cylinder with the same fluid and structural attributions. It is discovered that the wavy cylinder,
although almost eliminates the Kármán vortices in the fixed configuration, shows only limited efficacy
on the mitigation of the flow induced vibration in the case of zero structural damping, and the typical
initial-upper-lower type response curve is manifested. The hydrodynamic forces of the wavy cylinder
also magnify significantly in the flexibly mounted cases during synchronization. Moreover, the phase
lag between the lift coefficient and the displacement displays a clear change from 0◦ at the initial branch
to 180◦ at the lower branch. The association of the 2S and 2P vortex shedding modes to the different
branches is also similar to that of the normal cylinder, in which the 2S mode corresponds to the initial
branch and the 2P to the lower branch. In general, despite the absence of the primary shedding
frequency in the fixed configuration, the flexibly mounted wavy cylinder exhibits many features that is
also found in the normal cylinder. This implies that the vortex induced vibration may not be initiated
by the Kármán vortex shedding, and thus may defy the conventional view on the mechanism of the
vortex induced vibrations. Additional simulations are performed with non-zero structural damping.
It is disclosed that with sufficiently high structural damping, the vibration of the wavy cylinder could
be reduced more efficiently than the normal cylinder.
Keywords: Wavy cylinders; vortex induced vibration; flow control; structural damping.

1. Introduction
Circular cylindrical structures are widely used in engineering applications such as deepwater risers,
free spanning pipelines, overhead transmission lines, bridge stay cables, etc. Being long and flexible,
these structures are plagued by the excessive fluid forces as well as the ensuing vortex induced vibrations
5 (VIV), which increases the construction cost and undermines the fatigue life. Aiming at alleviating
these problems, considerable efforts have been dedicated to the manipulation of the flow behind a
circular cylinder [1]. Among them, span-wise modification near the separation point, referred to as
3D forcing by Choi et al. [2], has been recognized as an effective control method for the reduction of
hydrodynamic forces on a circular cylinder. Typical examples of this category include sinus axis and
10 helical bumps [3], small size tabs [4], span-wise blowing and sucking [5], etc.
Another particular form of 3D forcing that has recently drawn the attention of seveal researchers
is the span-wise waviness, i.e., straight axis with sinusoidally varying diameter. Ahmed et al. [6, 7]

∗ Corresponding author. Email Address: zhoudai@sjtu.edu.cn

Preprint submitted to Ocean Engineering 21st September 2016


were among the first the have experimentally studied this kind of structures. Lam and his colleagues
have also been active in this field. They conducted a series of systematic investigations focusing on
15 the flow physics and the shape optimizations of the sinusoidal wavy cylinder, with Reynolds number
spanning from 100 to 50,000 [8–11]. Xu et al. [12] numerically examined the compressible flow past a
wavy cylinder. Lee and Nguyen [13] and Zhang et al. [14] both employed flow visualization techniques
to study the wavy cylinder wakes. Based on these works, some salient features of this particular shape
could be summarized here. (i). The wavy geometry enforces span-wise non-uniformity in the separation
20 angle. The free shear layers emanating from the undulated separation line are contorted in an 3
dimensional fashion; (ii). The 3D free shear layers are more resistant to rolling-up, thus the Kármán
vortices are suppressed and develop further downstream; (iii). The drag and lift forces, owing to the
elongated vortex formation length, are significantly mitigated. The stabilizing effect of the artificially
generated span-wise waviness on an otherwise two dimensional wake base flow is further justified from
25 the perspective of stability analysis by Hwang et al. [15]. Based on the above experimental, numerical
and theoretical investigations, it could be concluded that the span-wise waviness presents a promising
morphology if one is seeking for the mitigation of fluid induced forces.
The previous researchers have mainly focused on the forces and wakes of the fixed wavy circular
cylinders, coming to the conclusion that this very shape could suppress the Kármán vortex shedding
30 and reduce the drag and lift forces. An implicit conclusion has been inferred that VIV could also be
mitigated if the wavy cylinders are elastically mounted. However, based on Owen et al. [3]’s research
on flow past cylinders with sinus axis and helical bumps, in which vortex induced vibrations was not
eliminated even though vortex shedding was totally suppressed for the same fixed body, Dong et al. [16]
pointed out that some of the flow control techniques may be effective only for the stationary cylinders
35 rather than flexibly mounted ones. Another example attesting to this statement is provided by Pastò
[17]. He incorporated the effect of surface roughness and flow turbulence into an effective Reynolds
number Reef f and found out that the cylinder might experience VIV even in the critical Reef f re-
gimes in which coherent vortex shedding ceases to persist in the steady configuration. Extending our
discussion a little wider to bridge models, El-Gammal et al. [18] revealed that when attaching a span-
40 wise sinusoidal perturbation module (SSPM) to a plate girder bridge model, considerable discrepancy
could be noted in the flexibly mounted case and the fixed case: SSPM is effective in the former while
ineffective in the latter. The anticipation that the wavy cylinders are capable of mitigating VIV is
confounded by these conflictive statements. A detailed investigation on the response of the flexibly
mounted wavy cylinders is in need to demystify this issue.
45 Vortex induced vibrations of circular cylinders involve quite a number of factors such as mass
ratio, damping, stiffness, Reynolds number, incoming turbulence intensity, etc., the study of which has
fostered voluminous publications that are based on various methods. Comprehensive reviews on these
subjects could be found in Sarpkaya [19], Gabbai and Benaroya [20], Williamson and Govardhan [21]
and so on. The VIV phenomenon is generally recognized as a resonance effect with nonlinear feedback
50 [22]. It is featured by self-excited and self-limited oscillations over a range of reduced velocities in
which the vibration frequency locks itself to the natural frequency of the system. The variation of
maximum vibration amplitude with respect to reduced velocity could be fit into Feng’s two branch
(initial-lower) curve [23], typical for larger mass ratio and damping, or three branch (initial-upper-
lower) curve of Khalak and Williamson [24] for lower mass-damping. Further more, Govardhan and
55 Williamson [25, 26] identified a critical mass ratio of around 0.54, below which the vortex induced
vibration might occur for all the reduced velocities larger than 5. If the cylinder is allowed to vibrate
in both the transverse and in-line directions, the ”supper-upper” branch, characterized by massive
amplitude of 3 diameters peak-to-peak, is discovered by Jauvtis and Williamson [27]. These findings
have important indications on the practical applications, especially the oncean engineering for which
60 the mass ratio of the structure is usually small.
Owing to the advent of high performance digital computers and the development of sophisticated
numerical techniques, the recent widespread use of Computational Fluid Dynamics (CFD) has greatly
availed the research of VIV and consolidated the understanding of its mechanisms. The vast majority
of the computational work on VIV has been done in the laminar regime [28–32] as it is free of the
65 turbulence effects, and more importantly, less resource consuming. Higher Re generally encourages
larger vibration amplitude, more complicated flow patterns and more elusive forces. The numerical

2
prediction of vortex induced vibrations at moderate Reynolds numbers also presents a challenging
task because of the high computational resources needed to resolve the boundary layer as well as the
small scale flow details. A number of 2D numerical investigations in conjunction with either RANS
70 or LES turbulence model have been practiced at sufficiently high Reynolds numbers [33–37], with
the assumption of full correlation in the span-wise direction. Lately, riding on the crest of the ever-
increasing computing power, some 3D studies on VIV of circular cylinders at turbulent regimes have
emerged [38–44]. These studies have been successful in retrieving useful insights in many aspects
associated with the VIV problems.
75 Following these works, the current paper endeavors to investigate the vortex induced vibration of
a wavy cylinder, which, to the authors’ knowledge, has not been studied in literature. The aim is to
verify whether this particular geometry could reduce the vortex induced vibrations. A wavy cylinder
with mass ratio of m∗ = 2.55 is investigated mainly at Re = 5000. It is anticipated this work would
bring about deep insights into the hydro-elasticity of the wavy cylinders. The rest of the paper is
80 organized as follows. Section 2 covers the numerical aspects in the current work. In Section 3 we will
present the details of the case set-up, followed by a mesh dependency test in Section 4. In Section 5
we present and the discuss the results. The last section concludes this paper.

2. Governing equations

The Navier-Stokes equations are spatially filtered to yield the governing equations of Large Eddy
85 Simulation (LES), which, in the Arbitrary-Lagrangian-Eulerian (ALE) frame, read as follows:

∂ui
= 0, (1)
∂xi
   
∂ui ∂ui ∂p ∂ ∂ui ∂uj
+ (uj − ũj ) =− + ν + + τij , (2)
∂t ∂xj ∂xi ∂xj ∂xj ∂xi
where (x1 , x2 , x3 ) = (x, y, z) are the Cartesian coordinates, ui is the filtered velocity tensor and p is
the filtered pressure, in which the constant fluid density ρ has been incorporated. ũi is the velocity
component of mesh moving in the xi direction. τij = ui uj − ui uj is the sub-grid scale (SGS) stress
90 that requires extra modeling. The SGS stress is expressed in a wavy reminiscent of the Boussinesq
hypothesis with the introduciton of a turbulent eddy viscosity νt :
2
τij − kt δij = −2νt S ij , (3)
3
 
1 ∂ui ∂uj
in which kt = τkk /2 is the SGS turbulent kinetic energy and S ij = + is the strain
2 ∂xj ∂xi
rate tensor calculated directly from the resolved scales. In the current paper, the dynamic k-equation
model is adopted to solve for the kt and νt , i.e.,
 
∂kt ∂ ∂ ∂kt
+ (uj kt ) = P + (ν + νt ) − ,
∂t ∂xj ∂xj ∂xj
P = 2νt S ij S ij , (4)
 = C kt1.5 ∆−1 ,
95
νt = Ck ∆kt0.5 , (5)
The model coefficients C and Ck are dynamically computed as part of the solution based on the
ˆ = 2∆ by the least square minimization procedure proposed
Germano identity [45] with test filter ∆
by Lilly [46].
The above equations are solved by the finite volume method (FVM) based open source CFD
100 toolkit OpenFOAM (Open Filed Operation and manipulation) [47]. pimpleDyMFoam, a transient
solver for incompressible flow on a moving mesh, is used to tackle with the coupled velocity and
pressure equations. The peculiarity of the PIMPLE algorithm lies in the fact that more than one

3
outer loop (pimple.loop) is practiced at each time step. This enables the adoption of a larger time step
without loosing computational stability. Thus, the demanding requirement on the courant number
105 could be relaxed to larger value, in contrast to the conventional PISO algorithm where a maximum
value below one should be maintained.
The motion of the rigid cylinder body is governed by:

d2 Y dY
m +c + kY = FY , (6)
dt2 dt
where Y denotes the transverse displacement of the cylinder, m, c and k are the structural mass,
damping and stiffness, respectively. FY represents the lift force in the cross-flow direction and is
110 calculated by integrating the pressure and wall shear forces over the whole cylinder surface. The
sixDoFRigidBodyMotion, an OpenFOAM built-in solver developed in the spirit of Dullweber et al.
[48], is employed to integrate Eq. 6. The coupled fluid-structural problem are solved through a
partitioned, weakly-coupled algorithm in which the two subsystems advance in a staggered fashion, as
is depicted in Fig. 1. The structure solver is invoked at the beginning of each time step, the output
115 of which drives the motion of the mesh and serves as the boundary condition for the fluid solver.
As has been discussed by He [49–51], in contrast to the strong coupling scheme, the weak coupling
technique has the advantage of being less computational intensive. However, it does not assure the
exact satisfaction of the equilibrium on the fluid-structure interface so that the cumulated errors may
lead to a spurious solution or even a failure. This problem is frequently encountered in the case of
120 small mass ratio and strong added mass effect. However, considering that for the current problem, the
weakly-coupled algorithm has been reported to yield robust FSI results [52–54], we decided to stick
with the current scheme. In order to alleviate the numerical instability initiated from the time-lag
effect, we choose a small enough time-step based on the mesh dependency test, which will be presented
in Section 4.

%HJLQQLQJRISURJUDP (QGRISURJUDP
WUXH

IDOVH
WQ WQ ǻW W 7HQG"

)LUVWOHDSIURJXSGDWH
SRVLWLRQ WXUEXOHQFH! FRUUHFW

WUXH

&DOFXODWHIRUFHV IDOVH
WXUE&RUU

6HFRQGOHDSIURJ IDOVH
XSGDWHWKHDFFHOHUDWLRQ
6ROYHWKHSUHVVXUH WUXH SLPSOHFRUUHFW
FRUUHFWRUHTXDWLRQ
6WUXFWXUHVROYHU

0RYHPHVK
6ROYHWKHPRPHQWXP
SUHGLFWRUHTXDWLRQ

WUXH
0DNHIOX[HVUHODWLYH IDOVH
SLPSOHORRS

)OXLGVROYHU

Figure 1: Algorithm of the solution procedure

4
125 3. Case setup
A schematic of the geometry of interest with coordinate system and the corresponding mesh is
presented in Fig. 2. The diameter of the wavy cylinder varies sinusoidally along its span-wise direction
by the following equation:
Dz = Dm + 2a cos (2πz/λ) , (7)
where Dz denotes the local diameter of the wavy cylinder and Dm = (Dmin + Dmax )/2 is the averaged
130 diameter which is used in the calculation of Reynolds number as well as the averaged force coefficients.
Following the nomenclatures in Lam et al. [11], λ is the wavelength in the span-wise direction and
the amplitude a is the maximum deviation of local radius from the average radius Dm /2. The axial
locations at which the local diameter maximizes (Dmax ) or minimizes (Dmin ) are referred to as ’node’
and ’saddle’ respectively. In the current paper, we assign λ/Dm = 2 and a/Dm = 0.175.
135 In the current work, the reduced velocity, defined as Ur = U∞ / (fn Dm ), where U∞ is the freestream
velocity, fn is the natural frequency of the cylinder without considering the added mass effect, is used
as the only parameter. Unlike the scenario in the experiments, where the freestream velocity U∞
is usually varied, we modify the reduced velocity by changing the stiffness k of the spring, which
consequently sets the value of the natural frequency of the cylinder. The mass ratio of the cylinder,
140 m∗ = 4m/(ρπDm 2
) (m is the mass of the cylinders, ρ is the density of the fluid), is fixed at 2.55.
In this way, the effect of Reynolds number and mass ratio to the VIV response could be eliminated.
Unless otherwise specified, the structural damping c is set to zero to encourage the maximum possible
vibration amplitude. All simulations are started from the initial steady state. The performance of the
wavy cylinder is evaluated based on the comparison with the results of flow around a normal cylinder
145 with identical fluid and structural parameters. In this regard, the present work also contributes to the
database of numerical prediction of the normal circular cylinders’ VIV at moderate Reynolds number.

D

N
'PLQ
8’ 'PD[ F

]
Ȝ \

(a) Coordinate system (b) An overview of the mesh

Figure 2: Schematic of the coordinate system and overview of mesh

The aspect ratio, i.e., length-to-diameter ratio, L/D = 4 is chosen for the normal and wavy
cylinders. Although higher values of aspect ratio is preferred in predicting flow properties especially
the span-wise correlations [43], the resulted computational cost increases enormously. According to
150 Lei et al. [55], good agreements of force coefficients and Strouhal number with experimental data
could be achieved with span-wise length above twice the cylinder diameter. As for the wavy cylinders,
L/D = 4 accommodates two wavelengths. It was revealed that one wavelength height is sufficient
for the simulations regardless of whether the flow is laminar [11] or turbulent [10]. Based on the
above statements, the currently chosen aspect ratio should ensure fairly accurate results of the force
155 coefficients and the VIV responses.
The Cartesian coordinate is used with the x axis aligning with the incoming flow direction and
z axis pointing at the span-wise direction. The displacement of the cylinder is restricted to the y

5
direction, where a spring-damper system of stiffness k and damping c is installed. No-slip boundary
condition is imposed on the surfaces of the cylinders, i.e., the velocity of the fluid on the cylinder
160 surface is the same with the cylinder’s vibration speed. At the inlet a constant velocity of U∞ is
specified. The zero-gradient condition for velocity is used for the outlet, where the reference pressure
p = 0 is assigned. In the span-wise ends, we choose the slip boundary condition since it enables us to
assess the axial correlation with full cylinder length. In contrast, if the cyclic boundary condition were
employed, only half of the axial length could be used for the evaluation of the correlation.
165 The cylinder is placed in the center of a circular computation domain, which has a radial extension
of 20D. This domain is discretized with uniformly spaced grids in the azimuthal coordinate and
exponentially stretched ones in the radial direction. The mesh resolution and time-step are chosen after
a careful mesh dependency tst, as will be presented in Section 4. The grids in the immediate vicinity
of the cylinder wall moves rigidly with the cylinder; in this way the uniformity of the near wall mesh
170 resolution is preserved for all cases. The outermost grids are kept fixed to alleviate the computational
requirements. The region in between is where the mesh deformation takes place. This area is given
sufficient space so that fair mesh quality is ensured even in the case of large VIV amplitude.

4. Mesh dependency test

The mesh resolution should be delicately determined in order to achieve satisfactory accuracy
175 without posing excessive burden on the computational resources. Mesh dependency tests, as is detailed
in Table 1, are conducted priorly at Re = 5000 to determine the appropriate spatial grid resolution as
well as the time steps. Regarding the case naming conventions, the first letter N or W indicates the
normal or wavy cylinder. The next word points out the reduced velocity. The last letter A, B, C or D
indicates the mesh and temporal resolution. Specifically, compared with the case A, the time-step is
180 halved in the cases denoted B, and the cases denoted C are associated with the refined mesh in the x-y
plane. For the wavy cylinder, the D cases, in which the span-wise meshes are refined, are performed to
account for the span-wise waviness. Nc , Nr and Nz denote the grids in the circumferential, radial and
span-wise direction. Ymax is defined as the average of the top 10% peak amplitude. f is the primary
frequency that contains the highest energy content from the spectrum of the vibration response and
185 fn is the natural frequency in vacuum. The drag and lift coefficients are defined as

Cd = 2Fd / ρDm LU 2 ,

(8)

Cl = 2Fl / ρDm LU 2 ,

(9)
in which Fd and Fl are the total drag and lift forces calculated by integration of pressure and skin
friction stress over the whole cylinder surface, L is the span-wise length and Dm is the average diameter
of the cylinders. The averaged values as well as the root mean squared (rms) values of Cd and Cl are
190 included in the table for comparison. We compare the results based on the cases denoted A, which has
spatial resolution of 160 × 160 × 80 and time-step ∆t = 0.01. It could be seen that the results in terms
of maximum amplitude, primary frequency, fluid-induced forces are in reasonable agreement when the
spatial ro temporal resolutions are refined. Further more, as will be shown in the next chapter, the
current results of the normal cylinder in terms of maximum vibration amplitude, primary frequency,
195 drag and lift forces, etc., are in excellent agreement with the literature. Since the current work is
the first study on the VIV of the wavy cylinder, its results could not be validated through previous
literature. However, we have employed the same numerical model and the calculation method for the
wavy and normal cylinders, and the results of the latter have shown to be of high fidelity. Thus, it is
reasonable to believe that the results of the wavy cylinder is also of enough accuracy and trustworthy.
200

5. Results and discussion

5.1. Force mitigation of the static wavy cylinder


The drag, lift force coefficients and the Strouhal number of both the normal and wavy cylinders
are summarized in Table 2 at three subcritical Reynolds numbers, i.e., Re = 1000, 3000 and 5000.

6
Table 1: Mesh dependency test

Case Nc × Nr × Nz ∆t Ymax /D f /fn Cdmean Cdrms Clrms


N Static A 160 × 160 × 80 0.01 - - 0.95 0.021 0.075
N Static B 160 × 160 × 80 0.005 - - 0.97 0.034 0.082
N Static C 200 × 200 × 80 0.01 - - 0.96 0.036 0.080
N Ur5 A 160 × 160 × 80 0.01 0.98 0.88 2.45 0.71 0.89
N Ur5 B 160 × 160 × 80 0.005 0.94 0.93 2.53 0.69 0.81
N Ur5 C 200 × 200 × 80 0.01 0.92 0.94 2.48 0.59 0.73
N Ur7 A 160 × 160 × 80 0.01 0.62 1.09 1.51 0.24 0.29
N Ur7 B 160 × 160 × 80 0.005 0.60 1.09 1.51 0.25 0.32
N Ur7 C 200 × 200 × 80 0.01 0.60 1.07 1.49 0.22 0.26
W Static A 160 × 160 × 80 0.01 - - 0.80 0.023 0.013
W Static D 160 × 160 × 100 0.01 - - 0.80 0.017 0.014
W Ur5 A 160 × 160 × 80 0.01 0.83 0.90 2.26 0.45 0.76
W Ur5 D 160 × 160 × 100 0.01 0.79 0.91 2.26 0.45 0.69
W Ur7 A 160 × 160 × 80 0.01 0.55 1.06 1.40 0.16 0.22
W Ur7 D 160 × 160 × 100 0.01 0.56 1.09 1.42 0.17 0.20

Table 2: Drag, lift force coefficients and Strouhal number at different Reynolds number in the static configuration

Cd Cl St
Re
normal wavy normal wavy normal wavy
1000 1.06 0.85 0.18 0.010 0.21 -
3000 0.97 0.82 0.093 0.011 0.21 -
5000 0.95 0.80 0.075 0.013 0.21 -

205 The same mesh as Re = 5000 case is used for the simulations with the other two Reynolds numbers.
Compared with the normal cylinder, the wavy cylinder presents around 15% to 20% reduction in the
drag force coefficients. The mitigation in the lift coefficient is even more pronounced, ranging from
83% to 95%. In addition, the shedding frequency of the Kármán vortices behind the wavy cylinders
at the tested three Res could not be detected. This is manifested in Fig. 3, where the PSD of the lift
210 coefficients of the normal and wavy cylinders at Re = 5000 are presented. The other two Reynolds
numbers are not shown here since the behavior of the lift spectrum is similar. It could be observed
that the PSD of the normal cylinder is characterized by a sharp peak at the vortex shedding frequency
fcl = 0.21. However, such a well-marked vortex shedding frequency is not recognizable for the wavy
cylinder, rather, its frequency distribution resembles white noise for fcl < 0.2. The absence of the
215 primary frequency in the PSD of Cl indicates that the vortex shedding has been massively suppressed.
This is further supported by the two snapshots of axial vorticity ωz contour, as are presented in Fig.
4. For the normal cylinder, well-formed Kármán vortex loops are clearly displayed. However, in the
case of wavy cylinder, the free shear layers are distorted and roll up at much further wake than the
normal one, and the developed vortices appear rather irregular. The detected small-amplitude lift
220 force might be attributed to the 3-dimensional turbulence effect that is inherent in the cylinder wake
at Re = 5000.

5.2. Dynamic response


The vortex induced vibration responses in terms of both maximum amplitude and primary fre-
quency of the current study are plotted against the reduced velocity in Fig. 5, together with the data
225 from the experimental works of Khalak and Williamson [56, 57] (m∗ = 2.4, Re = 2000 ∼ 12000), Stap-
penbelt et al. [58] (m∗ = 2.36, subcritical Re) and the numerical investigation from Zhao et al. [44]
(m∗ = 2, Re = 1000). Efforts are made to convert some of the cited data that are based on the natural
frequency in water fnw to fn . We’ll firstly focus on the case of Re = 5000. Although the resolution of

7
1e-02 normal
wavy
1e-04

1e-06

PSD1e-08

1e-10

1e-12

0.1 1
fcl

Figure 3: Power spectrum of lift force coefficients of normal and wavy cylinders at Re = 5000

(a) Normal cylinder (b) Wavy cylinder

Figure 4: Contour of ωz = ±0.5 of the static normal and wavy cylinders at Re = 5000

reduced velocities in the current study is not fine enough to delineate the detailed boundaries between
230 different response regions, the vibration amplitude variation obtained from our numerical simulation
fits aptly to the classical initial-upper-lower branch type curve [24], and as far as the normal cylinder
is concerned, good agreement is achieved between our simulation with the literature. The maximum
displacement predicted by the current work occurs at Ur = 5 and is around 0.98, which is close to
the experiment of Khalak and Williamson [56] and much larger than the DNS of Zhao et al. [44] at
235 Re = 1000. The lower branch is characterized by a plateau with almost constant amplitude of around
0.6 persisting within Ur = 7 ∼ 10. At the same range the reduced frequency f /fn locks to a value
slightly larger than 1, indicating that rather than following the Strouhal law as in the lower reduced
velocity regimes, the vibration has come to be dominated by the structural frequency. The slight
deviation from unity of the frequency ratio is a result of the added mass effect [24].
240 The wavy cylinder exhibits a similar response curve as the normal cylinder, with noticeably smaller
vibration amplitude at most of the simulated reduced velocities. At Ur = 5, where the maximum
oscillation occurs, a 15% reduction in amplitude compared with the normal cylinder could be discerned.
Compared to the largely suppressed lift force when it’s fixed, wavy cylinder’s performance in alleviating
vortex induced vibration could be concluded disappointing. Large amplitude vibrations still occur
245 within quite a large range of reduced velocities. Frequency-wise, the reduced frequency of the wavy
cylinder almost overlap with the normal cylinder at the synchronization range. It is surprising to find
such coincidence in the response curves of the wavy and normal cylinders, since the former does not
have a leading shedding frequency, as is revealed in Fig. 3. Thus, it might experience vigorous VIV
at some reduced velocities other than 5. We believe that the occurrence of this coincidence might not
250 be fortuitous. As a matter of fact, it has been revealed by previous studies [8, 59] that even though

8
Present normal 5000
0.9 Present wavy 5000
Exp. Khalak et al.
Exp. Stappenbelt et al.
0.7 DNS Zhao et al.
Present normal 1000

Ymax/D
Present wavy 1000
0.5 Present normal 3000
Present wavy 3000

0.3

0.1
normal cylinder St Law at Re=5000
3
2.5
2
f / fn

1.5
1
0.5
0
2 3 4 5 6 7 8 9 10 11 12 13 14 15
Ur

Figure 5: VIV response of normal and wavy cylinders, with data from literature.

the vortex shedding intensity is reduced by the wavy cylinder, the frequency of the mitigated lift force
remains nearly the same with the normal cylinder. In the current case, the primary shedding frequency
of the wavy cylinder has been suppressed to such a level that is not able to be detected. It seems that
flexible mounting unveils the concealed primary frequency, which then plays its role in determining
255 the vibration response.
To further check the Reynolds number dependence of the wavy cylinders’ dynamic characteristics,
the maximum vibration amplitude and the normalized frequency at Re = 1000 and 3000 are also
included in Fig. 5 at Ur = 3, 4, 5 and 7. Except that at Re = 1000 the largest vibration of the normal
cylinder occurs at Ur = 4 rather than 5, the responses of both the normal and wavy cylinders resemble
260 each other with minor difference in the values, and the initial, upper and lower branches could also be
vaguely discerned from the amplitude plot. The frequency responses of the normal and wavy cylinders
at the said two Res are also very close to each other. No significant difference could be observed in the
vibration amplitude at the said two Reynolds numbers compared to Re = 5000, although the largest
vibration amplitude occurring at Ur = 5 or 4 decreases as the Reynolds number decreases, which is
265 in accordance with the result of Lucor et al. [39]. This fact suggests that the different flow control
behavior of the wavy cylinder at fixed and flexibly mounted configurations is associated with a large
range of Reynolds numbers. Since at the three test Reynolds numbers the vibration responses appear
qualitatively similar, in the texts that follow, only the results of case Re = 5000 will be presented.

5.3. Force coefficients


270 The span-wise averaged mean drag and rms lift coefficients are summarized in Fig. 6. Experimental
data from Stappenbelt et al. [58], Khalak and Williamson [24] and numerical study from Zhao et al.
[44] are also included in the figure for comparison. Although the motion of the cylinder mainly
couples with the lift force, a significant amplification of drag is seen in both the mean value and its
fluctuation during the lock-in range. The mean Cd predicted by our simulations is in considerably
275 good agreement with the experiment of Stappenbelt et al. [58]. For the normal cylinder, the mean
Cd reaches a peak of around 2.4 at Ur = 5, at which its fluctuation also maximizes. This corresponds
to maximum instantaneous Cd of around 4, which is four times larger than a fixed cylinder. The
mean and fluctuation Cd of the wavy cylinder are generally smaller than that of the normal cylinder.

9
2.6
Present normal
Present wavy
2.2 Exp. Stappenbelt et al.

Cd 1.8
mean

1.4

0.6

Cd 0.4
rms

0.2

2.4 Exp. Khalak et al.


DNS Zhao et al.
2
1.6
Cl
rms 1.2
0.8
0.4
0
2 3 4 5 6 7 8 9 10 11 12
Ur

Figure 6: Mean drag, rms drag and lift coefficients

The maximum instantaneous Cd occurring at Ur = 5 of the wavy cylinder is around 3, which is also
280 approximately four times larger than the fixed configuration.
As far as the rms lift coefficients are concerned, the discrepancy between the wavy and normal
cylinders at each reduced velocity is minor. Both Clrms go through massive magnification during the
synchronization. Our numerically predicted maximum rms Cl , which occurs at Ur = 4, agrees well
with the DNS of Zhao et al. [44], but falls far short of the experimental investigation by Khalak and
285 Williamson [24] occurring at Ur = 4.4. It is highly likely that, as noted by Zhao et al. [44], the rms lift
coefficient may reach a peak between Ur = 4 ∼ 5 if the simulation in the said range were conducted
with smaller reduced velocity intervals.
Moreover, the span-wise distributions of the sectional temporally averaged drag and the rms lift
force coefficients are presented in Fig. 7. The sectional coefficients are defined based on the local
290 diameters, i.e., by replacing the average diameter Dm and total length L in Eq. 8 and 9 with the local
diameter D (z) and the corresponding sectional width ∆L. Previous studies have shown that in the
static configuration, the sectional drag force coefficient Cd at the geometric nodes is greater than that
at the geometric saddle, while there is not much variation in the sectional lift force coefficient Cl in the
span-wise direction [6, 59]. These conclusions are generally true for the current work when the wavy
295 cylinder is static, except that the Cd reaches the minimum in between the node and saddle section.
This could be attributed to the large geometric gradient at the middle sections (where D (z) = Dm ).
The force coefficients are actually calculated as Ci = ΣCp · ni · A, where i = x, y or z denotes the
Cartesian coordinate, ni is the i component of the surface normal unit vector, for which the identity
n2x + n2y + n2z = 1 holds, Cp is the pressure coefficient defined as Cp = 2 (p − p∞ ) /ρU 2 and A is the
300 surface area that is proportional to the local radius of the cylinder. In the middle section, the span-

10
wise geometric gradient is as large as nz = 0.5 × dD (z) /dz|0.5 = 0.55. It consumes quite a portion of
the unit vector, yielding smaller values of nx and ny . Although the integration area A at the middle
section is greater than the saddle, this effect is offset by the massive decreases in the values of nx and
ny , the minimum values of the force coefficients at the middle section are thus engendered.
305 The motions of the wavy cylinders influence the sectional forces differently depending on the Ur .
Apart from the increased values, the undulations in the span-wise distributions of the sectional force
coefficients are slightly exaggerated at most of the reduced velocities. This is particularly true for the
lift coefficients as span-wise variations, with maximum at the node and minimum at the saddle, has
come to be noticeable in Fig. 7(b). However, compared to the Cdmean , these span-wise fluctuations
310 in the lift coefficients are still insignificant. Glaring exceptions are observed at Ur = 4 and 5, at
which the drag coefficients at the saddle exceed that at the node. The same conclusion also holds
for the lift coefficients, although their minimum values are found at the middle sections at the said
two reduced velocities. As is recorded in Fig. 5, both reduced velocities correspond to the top two
maximum vibration amplitude, which is thought to be responsible for the observed abnormality. This
315 is confirmed by an additional case with structural damping c = 4.8 at Ur = 5, denoted as U r5 c4.8 in
Fig. 7. In this case, the vibration amplitude has been suppressed to 0.3D. It could be observed in the
damped case that the force coefficients at the saddle has been suppressed to be lower than that at the
node. More detailed results about the cases with non-zero damping will be presented in Section 5.5.

2.6
Ur2
2.4 Ur3
Ur4
2.2 Ur5
2 Ur6
Ur7
1.8 Ur8
Cd 1.6 Ur9
mean
Ur10
1.4 Ur11
1.2 Ur12
static
1 Ur5_c4.8
0.8
0.6
0 1 2 3 4
z/D

(a) Time-averaged sectional drag force coefficient Cdmean


1.6
Ur2
1.4 Ur3
Ur4
1.2 Ur5
Ur6
1 Ur7
Ur8
Cl 0.8 Ur9
rms
Ur10
0.6 Ur11
Ur12
0.4 static
Ur5_c4.8
0.2
0
0 1 2 3 4
z/D

(b) rms sectional lift force coefficient Clrms

Figure 7: Span-wise distribution of the mean drag and the rms lift force coefficients

The span-wise correlation of the lift force serves as an accurate measurement of the three dimen-

11
320 sionality in the near wake of the body. Bearman [60] noted that the correlation length experiences
a significant increase when the oscillation magnitude exceeds Ymax /D = 0.05. However, more recent
studies indicated that there exists certain ranges in the lock-in region where the span-wise correlation
suffers severe drop [39, 44, 61, 62]. In the current work, we define the correlation coefficient of the lift
force coefficients as follows:
P nh ih io
Cl (0, t) − Cl (0) Cl (z, t) − Cl (z)
r (0, z) = s t=0 s , (10)
Ph i2 P h i2
Cl (0, t) − Cl (0) Cl (z, t) − Cl (z)
t=0 t=0

325 in which Cl (z, t) is the lift force coefficient at span-wise location z and time t, the overline of which
is its time-averaged value. Figure 8 presents the span-wise correlations at different reduced velocities
of both the normal and wavy cylinders. It could be seen in Fig. 8(a) that the lift force of the normal
cylinder at Ur = 3 and 4, which are in the initial branch, are almost fully correlated. The correlation
stays stably high before it drops to minus again at the end of the synchronization range. The above
330 descriptions are generally consistent with the findings of the previous publications that the correlations
deteriorate at the ends of the upper and lower branches.

1
0.8
0.6
Ur2
0.4 Ur3
Ur4
r(0,z) 0.2 Ur5
Ur6
Ur7
0 Ur8
Ur9
-0.2 Ur10
Ur11
-0.4 Ur12
static
-0.6
0 1 2 3 4
z/D

(a) Normal cylinder


1
0.9
0.8
0.7
Ur2
0.6 Ur3
Ur4
r(0,z) 0.5 Ur5
0.4 Ur6
Ur7
0.3 Ur8
Ur9
0.2 Ur10
Ur11
0.1 Ur12
static
0
0 1 2 3 4
z/D

(b) Wavy cylinder

Figure 8: Span-wise correlation of lift force coefficients of normal and wavy cylinder

The correlation curves for the wavy cylinder are complicated by the span-wise waviness, as could be
observed in Fig. 8(b). By comparing the static cases of the wavy and normal cylinders, it is revealed

12
that the geometric waviness substantially decreases the span-wise correlation length. However, it is
335 greatly enhanced once the wavy cylinder is excited to vibrate. Another remarked feature is that for
most of the reduced velocities, and inclined undulated pattern that conforms to the wavy geometry
is displayed. Since we calculate the correlation based on a fixed end node location, a recovery of
the coefficient is observed at the other two geometric nodes after the dip at the saddle locations.
This is significantly contrasted to the exceptions made at Ur = 2, 12 and the static case, in which such
340 undulated patterns are not found. We could infer from this discrepancy that, when the wavy cylinder is
static or undergoing small amplitude vibrations, the near wake flow presents high three dimensionality
because of the span-wisely protruded shape. When large amplitude oscillation occur, the control of
the near wake is taken over by the motion of the structure, as is reflected by the coherence between
the geometry and the correlation curves. The poorly correlated region at the end of the upper branch
345 found in the normal cylinder is not detected for the wavy one. Again, we attribute this to the coarse
resolution of the reduced velocity.
The decreased span-wise correlation length has been considered to contribute favorably to the
suppression of the vortex induced vibrations of the straked cylinder [63–65], which are used widely in
the engineering applications [1]. In our study of the wavy cylinders, the deteriorated correlation is also
350 manifested in the static case, although it clearly fails to suppress the oscillations as the reduced velocity
enters the lock-in region. Thus, it seems to be problematic to universally regard the decreased axial
correlation as an indicator of a successful attempt of VIV mitigation. As a matter of fact, the reduced
correlations are resulted from different mechanisms in the two cases. In the straked cylinder, the helical
add-ons chop up the flow and the vortex dislocations occur along the cylinder. This leads to irregular
355 shedding of the shear layers and consequently annihilation of the conventional Kármán vortices. While
for the wavy cylinder, the lift force is quite small and its primary frequency is missing. As mentioned
in Section 5.1, it is likely that the deteriorated lift coefficient correlation is a result of the unsteady
turbulence effect. Albeit contorted in a 3-dimensional fashion, the shear layers emanated from the
wavy cylinder are well-organized. As the initial disturbance in the cylinder motion is introduced, the
360 efficacy of the 3-dimensional shear layer in stabilizing the wake is lost. Rather, the shear layers interact
with each other in a way that is similar to the normal cylinder, large amplitude of vibration is thus
engendered.

5.4. Phase and vortex modes

1
Normal Ur=4 Ur=5 Ur=6 Ur=7
0.5
Y/D 0
-0.5
-1
1
Wavy Ur=4 Ur=5 Ur=6 Ur=7
0.5
Y/D 0
-0.5
-1
-2 -1 0 1 2 -2 -1 0 1 2 -2 -1 0 1 2 -2 -1 0 1 2
Cl Cl Cl Cl

Figure 9: Variation in the phase portrait of the lift force coefficient, Cl , relative to the transverse vibration amplitude,
Y /D of the normal (first row) and wavy (second row) cylinders.

The Lissajous plots of the lift force coefficient and transverse vibration amplitude at Ur = 4 ∼ 7 are
365 presented in Fig. 9 to identify the relative phase lags in different branches. Judging from the shapes
of these phase portraits, a clear change of relative phase angle from nearly 0◦ to nearly 180◦ could be
noticed between the initial branch of Ur = 4 and the lower branch of Ur = 7, in both the normal and
wavy cylinders. The phase diagrams of the normal cylinder at the Ur = 5 and 6 are characterized

13
by significantly larger vibration amplitude and appear less organized than the other two branches.
370 This chaotic behavior suggests that different states reside in this regime and intermingle with each
other. The phase portrait of the wavy cylinder at Ur = 5, however, does not exhibit such phenomenon.
It is revealed by Khalak and Williamson [24] that while the initial-upper transition is hysteric, the
upper-lower transition is associated with intermittent switches between the upper and lower branches.
Navrose and Mittal [66] also confirmed this unstable phenomenon via CFD simulations. A deeper
375 knowledge of this behavior is provided by the instantaneous phase lag between the lift and displacement
as well as the frequency, which could be obtained by performing the Hilbert transformation to the
recorded time histories of the lift force and displacement. For a signal s (t), given its Hilbert transform
ŝ (t), the analytic signal is defined as:

sA (t) = s (t) + jŝ (t) = A (t) ejφ(t) , (11)

in which φ (t) is the instantaneous phase. The instantaneous frequency could be obtained as

1 dφ (t)
f (t) = . (12)
2π dt
380 One may refer to the work of Khalak and Williamson [24] for further details on this technique. Fig.
10 shows the time histories of the lift coefficient, displacement, phase angle and frequency. At Ur = 5
(Fig. 10(a)), most of the time is spent with the upper branch, as is evidenced by the in-phase lift
and displacement. Phase jump is only observed at very small portion of time, e.g., t ≈ 50 and 160.
The instantaneous frequency at this Ur is quite stable, even in the presence of the phase jump. The
385 variation of the phase angle φ (t) becomes quite drastic when it comes to Ur = 6. Phase jumps are
discerned all along the time span. The time spent on the upper branch is comparable with that of on
the lower branch. Moreover, together with the jump in φ (t), the instantaneous frequency also undergo
a sudden shift. The upper branch is associated with lower frequency and at the lower branch the
frequency is higher.
390 It is generally known that the different response branches are associated with different vortex
shedding patterns [25]. While the initial branch corresponds to the 2S mode (two single vortices
per cycle), in the lower branch two pairs of vortices shed in each cycle, forming the 2P mode. The
shedding mode for the upper branch seems to be dependent on the Reynolds number. Govardhan and
Williamson [25] identified the 2P mode at Re ≈ 3100. However, 2S mode is found at Re = 1000 in
395 the numerical study by Navrose and Mittal [66]. Moreover, Zhao et al. [44] observed at Re = 1000
that the two modes coexist in the span-wise direction. In the current study, the axial component of
the vorticity fields ωz are plotted in Fig. 11 and 12 in an attempt to confirm the vortex shedding
patterns. As expected, the vortices behind the the normal cylinder shed in the 2S mode in the initial
branch (Ur = 4) and 2P mode in the lower branch (Ur = 7). The upper branch (Ur = 5) is observed
400 to be dominated by the 2P mode at Re = 5000. This is in agreement with the work of Govardhan
and Williamson [25] at higher Re and at variance with Navrose and Mittal [66] at lower Re. In the
upper-lower transition regime (Ur = 6), there exists significant span-wise variation in the shape of the
vorticity contour, which undermines the clear inspection of the shedding mode. This also corroborates
the deteriorated correlation of the lift forces, as is shown in Fig. 8(a). As for the wavy cylinder, the
405 vortex shedding patterns are in general consistent with the normal cylinder. The 2S and 2P modes
could be clearly observed at the initial (Ur = 4) and lower branch (Ur = 6 and 7). However, at the
upper branch (Ur = 5), in contrast to the well-organized vortex loops that are observed at the other
reduced velocities, the wake is dominated by many small-scale irregular vortices, and the shedding
mode can not be clearly identified.

410 5.5. The effect of structural damping


Additional simulations are conducted with non-zero structural damping, i.e., c 6= 0, at Ur = 5, at
which the maximum vibration amplitude is observed forboth the normal and wavy cylinders without

damping. The damping coefficient is defined as ζ = c/ 2π mk . The results are shown in Table 3.
Although only the damping is varied, the combined mass-damping parameter m∗ ζ is used since it has
415 been proven useful in collapsing the peak vibration amplitudes [24, 67, 68]. It could be seen that as

14
2 Y/D 2
Cl
1 1
Cl 0 0 Y/D
-1 -1
-2 -2
1.43 f / fn 270
φ
1.22 180
f / fn 1.01 90 φ
0.8 0
0.59 -90
40 60 80 100 t 120 140 160 180

(a) Ur = 5
2 Y/D 2
Cl
1 1
Cl 0 0 Y/D
-1 -1
-2 -2
1.43 f / fn 270
φ
1.22 180
f / fn 1.01 90 φ
0.8 0
0.59 -90
40 60 80 100 t 120 140 160 180

(b) Ur = 6

Figure 10: Instantaneous lift force coefficient, displacement, frequency and phase angle of the normal cylinder at Ur = 5
and 6.

Table 3: VIV responses with additional structural damping

m∗ ζ Case Ymax /D Cdmean Clrms


0 normal 0.98 2.45 0.89
0 wavy 0.83 2.26 0.76
0.61 normal 0.39 1.74 0.80
0.61 wavy 0.30 1.64 0.63
0.91 normal 0.25 1.39 0.64
0.91 wavy 0.012 0.86 0.030

the damping reaches a medium level at m∗ ζ = 0.61, although the three indicators (Ymax /D, Cdmean
and Clrms ) of wavy cylinder are smaller than the normal cylinder, the extent of reduction remains
at the same level as the zero-damping case. However, with further increase of the damping ratio
(m∗ ζ = 0.91), the vortex induced vibration of the wavy cylinder has shown to be almost eliminated,
420 together with massively mitigated drag and lift force coefficients. At the same value of damping, the
normal cylinder still experiences oscillations with considerable amplitude. This is in agreement with
the finding of Owen et al.[3], who devised the cylinders with wavy axis and with helical bumps in
attempt to reduce the hydrodynamic forces as well as VIV. In their work, a significant decrease in
the vibration amplitude was observed as the mass-damping parameter reaches around 0.8. Although
425 only Ur = 5, rather than the whole reduced velocity spectrum is investigated, we cautiously conclude
that the wavy cylinder could be used as a low VIV-responsive device if sufficient damping is supplied.
Additionally, it is surmised, based on the validity of the mass-damping parameter in predicting the
maximum response amplitude, that the wavy cylinders with larger mass ration, such as the case of

15
(a) Ur = 4

(b) Ur = 5

(c) Ur = 6

(d) Ur = 7

Figure 11: Contours of span-wise component of vorticity ωz = ±0.5 for the normal cylinder. The four time instants
the each Ur correspond to the cylinder at positive maximum position, the middle position with downward motion, the
negative maximum position, the middle position with upward motion. For Ur = 5 and 6, the snapshots are selected in
the time periods that the displacement is in and out of phase with the lift forces, respectively. The blue color represents
the minus value of ωz and the orange color represents the positive value.

the cable in air, might show better performance in mitigation of the VIV than the normal cylinder.
430 Further researches are under consideration to verify this conjecture.

6. Conclusion and Discussion

A series of numerical simulations have been carried out to study the flow past a spring-mounted
wavy cylinder. The aim is to clarify whether this hydrodynamically beneficial geometry is capable of
alleviating the flow induced vibrations. We study mainly the Reynolds number of 5000. The stiffness
435 k of the spring is varied to yield Ur = 2 ∼ 12, at an interval of 1. Another set of computations are
carried out for the normal cylinder in order to validate our computational model and serve as the basis
upon which the performance of the wavy cylinder could be evaluated. The wavy cylinder studied in the
current work exhibits impressive flow control efficacy in the static configuration. The mean drag and
the r.m.s lift forces are reduced by 17% and 83% respectively in comparison with the normal cylinder
440 at Re = 5000. In addition, no primary frequency is detected in the lift force of the wavy cylinder. This
indicates that the Kármán vortices that prevails the wakes of bluff bodies have been almost eliminated
by the span-wise waviness. Nevertheless, when the wavy cylinder is allowed to move in the transverse

16
(a) Ur = 4

(b) Ur = 5

(c) Ur = 6

(d) Ur = 7

Figure 12: Contours of span-wise component of vorticity ωz = ±0.5 for the wavy cylinder. The four time instants are
the same with Fig. 11.

direction, the typical ’lock-in’ phenomenon still occurs. Specifically, the main findings of the current
work are summarized as follows.
445 1. The vibration amplitude of the wavy cylinder is mitigated compared to the normal one, however,
the extent of the mitigation falls far short of that of the lift coefficients in the fixed configuration. The
amplitude response (Ymax /D vs Ur plot) of both cylinders exhibit the three-branch curve that is typical
for low mass-damping cases. The vibration frequency of the wavy and normal cylinders alike locks to
the natural frequency of the cylinder in the synchronization range. Similar to the normal cylinder, the
450 maximum vibration amplitude of the wavy cylinder increases with the Reynolds number. However, we
observe no qualitative change in the dynamic behavior of the wavy cylinder as the Reynolds number
varies from 1000 to 5000.
2. For the wavy cylinder, the magnification of the drag and lift forces during the lock-in is com-
parable with the normal cylinder. The span-wise correlation of the lift force for the wavy cylinder is
455 greatly enhanced once it is excited to vibrate. An undulated pattern, which conforms to the wavy
cylinder shape, is found in the Cl correlation curves, in contrast to the monotonic decrease in the
correlation of normal cylinder case. The poorly correlated regimes at the end of the upper branch is
detected for the normal cylinder, but not for the wavy cylinder.
3. Similar with the normal cylinder, the phase between the lift force and the displacement of the
460 wavy cylinder undergoes a change from 0◦ in the initial and upper branches to 180◦ in the lower branch.

17
The hysteric regime in between the upper and lower branches in not detected in the wavy cylinder.
Different vortex shedding modes are associated with the branches. The initial branch is found to
correspond to the 2S mode while the upper and lower branches to the 2P mode.
4. The performance of the normal and wavy cylinders are comparable in the case of small structural
465 damping. However, as enough damping (m∗ ζ ≈ 0.9) is supplied, the vortex induced vibration could
be reduced more efficiently than the normal cylinder.
Based on the above conclusions, it is revealed that the hydro-elastic behavior of the wavy cylinder
exhibits many features that are also found in the normal cylinder, although the former presents glaring
difference in the fixed configuration. This reminds us the discovery by Pastò [17], in which the vortex
470 induced vibration persists even at the critical Reynold number characterized by a cessation of coherent
vortex shedding in steady configuration. Both cases imply that the vortex induced vibrations may
not be initiated by the Kármán vortex shedding, and thus may defy the conventional view on the
mechanism of vortex induced vibration that VIV is a resonance effect with nonlinear feedback [22].
The authors further suspect that the different behavior of the wavy cylinder at fixed and flexibly
475 mounted configurations may be of similar mechanism with the finding that VIV of a circular cylinders
could occur at Reynolds numbers inferior to 47, below which the lift force is strictly zero [69–71]. In
future studies, linear stability analysis in the line of Zhang et al. [72] and Mittal [73] will be carried
out further explore the mechanism of the hydro-elastic behavior of the wavy cylinders and the role of
structural damping.

480 Acknowledgement

Support from the Major Program of the National Natural Science Foundation of China (No.
51490674), National Natural Science Foundation of China (No. 51278297), Doctoral Disciplinary
Special Research Project of Chinese Ministry of Education (No.20130073110096) and Research Pro-
gram of Shanghai Leader Talent (No.20) are acknowledged. The authors would also like to thank the
485 anonymous referees for their kind suggestions to improve this manuscript.

References

[1] R. A. Kumar, C.-H. Sohn, B. H. Gowda, Passive control of vortex-induced vibrations: an overview,
Recent Patents on Mechanical Engineering 1 (1) (2008) 1–11.
[2] H. Choi, W.-P. Jeon, J. Kim, Control of flow over a bluff body, Annual Review of Fluid Mechanics
490 40 (1) (2008) 113–139.
[3] J. C. Owen, P. W. Bearman, A. A. Szewczyk, Passive control of viv with drag reduction, Journal
of Fluids and Structures 15 (3) (2001) 597–605.
[4] J. Yoon, Control of flow over a circular cylinder using wake disrupter, Master’s thesis, Seoul
National University, Korea (2005).
495 [5] J. Kim, H. Choi, Distributed forcing of flow over a circular cylinder, Physics of Fluids (1994-
present) 17 (3) (2005) 033103.
[6] A. Ahmed, B. Bays-Muchmore, Transverse flow over a wavy cylinder, Physics of Fluids A: Fluid
Dynamics (1989-1993) 4 (9) (1992) 1959–1967.
[7] A. Ahmed, M. Khan, B. Bays-Muchmore, Experimental investigation of a three-dimensional bluff-
500 body wake, AIAA journal 31 (3) (1993) 559–563.

[8] K. Lam, F. Wang, J. Li, R. So, Experimental investigation of the mean and fluctuating forces of
wavy (varicose) cylinders in a cross-flow, Journal of fluids and structures 19 (3) (2004) 321–334.
[9] K. Lam, F. Wang, R. So, Three-dimensional nature of vortices in the near wake of a wavy cylinder,
Journal of Fluids and Structures 19 (6) (2004) 815–833.

18
505 [10] K. Lam, Y. Lin, Large eddy simulation of flow around wavy cylinders at a subcritical Reynolds
number, International Journal of Heat and Fluid Flow 29 (4) (2008) 1071–1088.
[11] K. Lam, Y. Lin, Effects of wavelength and amplitude of a wavy cylinder in cross-flow at low
Reynolds numbers, Journal of Fluid Mechanics 620 (2009) 195–220.
[12] C.-Y. Xu, L.-W. Chen, X.-Y. Lu, Large-eddy simulation of the compressible flow past a wavy
510 cylinder, Journal of Fluid Mechanics 665 (2010) 238–273.
[13] S.-J. Lee, A.-T. Nguyen, Experimental investigation on wake behind a wavy cylinder having
sinusoidal cross-sectional area variation, Fluid Dynamics Research 39 (4) (2007) 292–304.
[14] W. Zhang, S. J. Lee, et al., PIV measurements of the near-wake behind a sinusoidal cylinder,
Experiments in fluids 38 (6) (2005) 824–832.
515 [15] Y. Hwang, J. Kim, H. Choi, Stabilization of absolute instability in spanwise wavy two-dimensional
wakes, J. Fluid Mech 727 (2013) 346–378.
[16] S. Dong, G. Triantafyllou, G. Karniadakis, Elimination of vortex streets in bluff-body flows,
Physical Review Letters 100 (20) (2008) 204501.
[17] S. Pastò, Vortex-induced vibrations of a circular cylinder in laminar and turbulent flows, Journal
520 of Fluids and Structures 24 (7) (2008) 977–993.
[18] M. El-Gammal, H. Hangan, P. King, Control of vortex shedding-induced effects in a sectional
bridge model by spanwise perturbation method, Journal of Wind Engineering and Industrial
Aerodynamics 95 (8) (2007) 663–678.
[19] T. Sarpkaya, A critical review of the intrinsic nature of vortex-induced vibrations, Journal of
525 Fluids and Structures 19 (4) (2004) 389–447.
[20] R. Gabbai, H. Benaroya, An overview of modeling and experiments of vortex-induced vibration
of circular cylinders, Journal of Sound and Vibration 282 (3) (2005) 575–616.
[21] C. Williamson, R. Govardhan, A brief review of recent results in vortex-induced vibrations,
Journal of Wind Engineering and Industrial Aerodynamics 96 (6) (2008) 713–735.
530 [22] E. De Langre, Frequency lock-in is caused by coupled-mode flutter, Journal of Fluids and Struc-
tures 22 (6) (2006) 783–791.
[23] C. Feng, The measurement of vortex induced effects in flow past stationary and oscillating circular
and D-section cylinders, Retrospective Theses and Dissertations, 1919-2007.
[24] A. Khalak, C. Williamson, Motions, forces and mode transitions in vortex-induced vibrations at
535 low mass-damping, Journal of Fluids and Structures 13 (7) (1999) 813–851.
[25] R. Govardhan, C. Williamson, Modes of vortex formation and frequency response of a freely
vibrating cylinder, Journal of Fluid Mechanics 420 (2000) 85–130.
[26] R. Govardhan, C. Williamson, Resonance forever: existence of a critical mass and an infinite
regime of resonance in vortex-induced vibration, Journal of Fluid Mechanics 473 (2002) 147–166.
540 [27] N. Jauvtis, C. Williamson, The effect of two degrees of freedom on vortex-induced vibration at
low mass and damping, Journal of Fluid Mechanics 509 (2004) 23–62.
[28] T. He, Semi-implicit coupling of CS-FEM and FEM for the interaction between a geometrically
nonlinear solid and an incompressible fluid, International Journal of Computational Methods
12 (05) (2015) 1550025.
545 [29] Z. Han, D. Zhou, J. Tu, C. Fang, T. He, Flow over two side-by-side square cylinders by CBS finite
element scheme of spalart–allmaras model, Ocean Engineering 87 (2014) 40–49.

19
[30] T. He, On a partitioned strong coupling algorithm for modeling fluid-structure interaction, Inter-
national Journal of Applied Mechanics 7 (02) (2015) 1550021.
[31] T. He, Partitioned coupling strategies for fluid-structure interaction with large displacement:
550 Explicit, implicit and semi-implicit schemes, Wind and Structures 20 (3) (2015) 423–448.

[32] Z. Han, D. Zhou, T. He, J. Tu, C. Li, K. C. Kwok, C. Fang, Flow-induced vibrations of four circular
cylinders with square arrangement at low Reynolds numbers, Ocean Engineering 96 (2015) 21–33.
[33] S. Mittal, V. Kumar, Flow-induced vibrations of a light circular cylinder at Reynolds numbers
103 to 104 , Journal of Sound and Vibration 245 (5) (2001) 923–946.

555 [34] E. Guilmineau, P. Queutey, Numerical simulation of vortex-induced vibration of a circular cylinder
with low mass-damping in a turbulent flow, Journal of fluids and structures 19 (4) (2004) 449–466.
[35] H. Al-Jamal, C. Dalton, Vortex induced vibrations using large eddy simulation at a moderate
Reynolds number, Journal of Fluids and Structures 19 (1) (2004) 73–92.

[36] Z. Pan, W. Cui, Q. Miao, Numerical simulation of vortex-induced vibration of a circular cylinder
560 at low mass-damping using rans code, Journal of Fluids and Structures 23 (1) (2007) 23–37.
[37] J. B. Wanderley, G. H. Souza, S. H. Sphaier, C. Levi, Vortex-induced vibration of an elastic-
ally mounted circular cylinder using an upwind tvd two-dimensional numerical scheme, Ocean
Engineering 35 (14) (2008) 1533–1544.

[38] C. Evangelinos, G. E. Karniadakis, Dynamics and flow structures in the turbulent wake of rigid
565 and flexible cylinders subject to vortex-induced vibrations, Journal of Fluid Mechanics 400 (1999)
91–124.
[39] D. Lucor, J. Foo, G. Karniadakis, Vortex mode selection of a rigid cylinder subject to viv at low
mass-damping, Journal of Fluids and Structures 20 (4) (2005) 483–503.

[40] N. Kondo, Three-dimensional computation for flow-induced vibrations in in-line and cross-flow
570 directions of a circular cylinder, International Journal for Numerical Methods in Fluids 70 (2)
(2012) 158–185.
[41] S. Mittal, et al., Free vibrations of a cylinder: 3-d computations at re = 1000, Journal of Fluids
and Structures 41 (2013) 109–118.
[42] Y. Jus, E. Longatte, J.-C. Chassaing, P. Sagaut, Low mass-damping vortex-induced vibrations
575 of a single cylinder at moderate Reynolds number, Journal of pressure vessel technology 136 (5)
(2014) 051305.
[43] A. H. Lee, R. L. Campbell, S. A. Hambric, Coupled delayed-detached-eddy simulation and struc-
tural vibration of a self-oscillating cylinder due to vortex-shedding, Journal of Fluids and Struc-
tures 48 (2014) 216–234.
580 [44] M. Zhao, L. Cheng, H. An, L. Lu, Three-dimensional numerical simulation of vortex-induced
vibration of an elastically mounted rigid circular cylinder in steady current, Journal of Fluids and
Structures 50 (2014) 292–311.

[45] M. Germano, U. Piomelli, P. Moin, W. H. Cabot, A dynamic subgrid-scale eddy viscosity model,
Physics of Fluids A: Fluid Dynamics (1989-1993) 3 (7) (1991) 1760–1765.
585 [46] D. K. Lilly, A proposed modification of the germano subgrid-scale closure method, Physics of
Fluids A: Fluid Dynamics (1989-1993) 4 (3) (1992) 633–635.

[47] H. G. Weller, G. Tabor, H. Jasak, C. Fureby, A tensorial approach to computational continuum


mechanics using object-oriented techniques, Computers in physics 12 (6) (1998) 620–631.

20
[48] A. Dullweber, B. Leimkuhler, R. McLachlan, Symplectic splitting methods for rigid body molecu-
590 lar dynamics, The Journal of chemical physics 107 (15) (1997) 5840–5851.
[49] T. He, D. Zhou, Y. Bao, Combined interface boundary condition method for fluid-rigid body
interaction, Computer Methods in Applied Mechanics and Engineering 223 (2012) 81–102.

[50] T. He, D. Zhou, Z. Han, J. Tu, J. Ma, Partitioned subiterative coupling schemes for aeroelasticity
using combined interface boundary condition method, International Journal of Computational
595 Fluid Dynamics 28 (6-10) (2014) 272–300.
[51] T. He, A partitioned implicit coupling strategy for incompressible flow past an oscillating cylinder,
International Journal of Computational Methods 12 (02) (2015) 1550012.
[52] Y. Bao, D. Zhou, J. Tu, Flow interference between a stationary cylinder and an elastically mounted
cylinder arranged in proximity, Journal of fluids and structures 27 (8) (2011) 1425–1446.
600 [53] A. Placzek, J.-F. Sigrist, A. Hamdouni, Numerical simulation of an oscillating cylinder in a cross-
flow at low Reynolds number: Forced and free oscillations, Computers & Fluids 38 (1) (2009)
80–100.
[54] J. Tu, D. Zhou, Y. Bao, C. Fang, K. Zhang, C. Li, Z. Han, Flow-induced vibration on a circular
cylinder in planar shear flow, Computers & Fluids 105 (2014) 138–154.
605 [55] C. Lei, L. Cheng, K. Kavanagh, Spanwise length effects on three-dimensional modelling of flow
over a circular cylinder, Computer methods in applied mechanics and engineering 190 (22) (2001)
2909–2923.
[56] A. Khalak, C. Williamson, Dynamics of a hydroelastic cylinder with very low mass and damping,
Journal of Fluids and Structures 10 (5) (1996) 455–472.
610 [57] A. Khalak, C. Williamson, Fluid forces and dynamics of a hydroelastic structure with very low
mass and damping, Journal of Fluids and Structures 11 (8) (1997) 973–982.
[58] B. Stappenbelt, F. Lalji, G. Tan, Low mass ratio vortex-induced motion, in: 16th Australasian
Fluid Mechanics Conference, 2-7 December, Crown Plaza, Gold Coast, Australia, 2007, pp. 1491–
1497.

615 [59] K. Zhang, H. Katsuchi, D. Zhou, H. Yamada, Z. Han, Numerical study on the effect of shape
modification to the flow around circular cylinders, Journal of Wind Engineering and Industrial
Aerodynamics 152 (2016) 23–40.
[60] P. W. Bearman, Vortex shedding from oscillating bluff bodies, Annual review of fluid mechanics
16 (1) (1984) 195–222.

620 [61] F. Hover, A. Techet, M. Triantafyllou, Forces on oscillating uniform and tapered cylinders in
crossflow, Journal of Fluid Mechanics 363 (1) (1998) 97.
[62] F. S. Hover, J. T. Davis, M. S. Triantafyllou, Three-dimensionality of mode transition in vortex-
induced vibrations of a circular cylinder, European Journal of Mechanics-B/Fluids 23 (1) (2004)
29–40.

625 [63] T. Zhou, S. M. Razali, Z. Hao, L. Cheng, On the study of vortex-induced vibration of a cylinder
with helical strakes, Journal of Fluids and Structures 27 (7) (2011) 903–917.
[64] Y. Constantinides, O. H. Oakley, Numerical prediction of bare and straked cylinder VIV, in:
25th International Conference on Offshore Mechanics and Arctic Engineering, American Society
of Mechanical Engineers, 2006, pp. 745–753.

21
630 [65] I. Korkischko, J. R. Meneghini, R. S. Gioria, P. J. Jabardo, E. Casaprima, R. Franciss, An ex-
perimental investigation of the flow around straked cylinders, in: ASME 2007 26th International
Conference on Offshore Mechanics and Arctic Engineering, American Society of Mechanical En-
gineers, 2007, pp. 641–647.

[66] Navrose, S. Mittal, Free vibrations of a cylinder: 3-D computations at Re= 1000, Journal of
635 Fluids and Structures 41 (2013) 109–118.
[67] O. Griffin, Vortex-excited cross-flow vibrations of a single cylindrical tube, Journal of Pressure
Vessel Technology 102 (2) (1980) 158–166.

[68] M. Bahmani, M. Akbari, Effects of mass and damping ratios on VIV of a circular cylinder, Ocean
Engineering 37 (5) (2010) 511–519.
640 [69] E. Buffoni, Vortex shedding in subcritical conditions, Physics of Fluids (1994-present) 15 (3)
(2003) 814–816.
[70] S. Mittal, S. Singh, Vortex-induced vibrations at subcritical re, Journal of Fluid Mechanics 534
(2005) 185–194.
[71] J. Leoniti, M. Thompson, Cylinder wake destabilisation due to elastic mounting, in: IUTAM
645 Symposium on Bluff Body Wakes and Vortex-Induced Vibrations, 22-25 June, Capri, Italy, 2010,
pp. 221–224.

[72] W. Zhang, X. Li, Z. Ye, Y. Jiang, Mechanism of frequency lock-in in vortex-induced vibrations
at low Reynolds numbers, Journal of Fluid Mechanics 783 (2015) 72–102.
[73] S. Mittal, Lock-in in vortex-induced vibration, Journal of Fluid Mechanics 794 (2016) 565–594.

22

You might also like