You are on page 1of 13

energies

Article
Effects of Mass and Damping on Flow-Induced Vibration of a
Cylinder Interacting with the Wake of Another Cylinder at
High Reduced Velocities
Md. Mahbub Alam

Center for Turbulence Control, Harbin Institute of Technology (Shenzhen), Shenzhen 518055, China;
alamm28@yahoo.com or alam@hit.edu.cn

Abstract: Flow-induced vibration is a canonical issue in various engineering fields, leading to fatigue
or immediate damage to structures. This paper numerically investigates flow-induced vibrations of a
cylinder interacting with the wake of another cylinder at a Reynolds number Re = 150. It sheds light
on the effects of mass ratio m* , damping ratio, and mass-damping ratio m* ζ on vibration amplitude
ratio A/D at different reduced velocities Ur and cylinder spacing ratios L/D = 1.5 and 3.0. A couple of
interesting observations are made. The m* has a greater influence on A/D than ζ although both m* and ζ
cause reductions in A/D. The m* effect on A/D is strong for m* = 2–16 but weak for m* > 16. As opposed
to a single isolated cylinder case, the mass-damping m* ζ is not found to be a unique parameter for a
cylinder oscillating in a wake. The vortices in the wake decay rapidly at small ζ. Alternate reattachment
of the gap shear layers on the wake cylinder fuels the vibration of the wake cylinder for L/D = 1.5 while
the impingement and switch of the gap vortices do the same for L/D = 3.0.

 Keywords: vibration response; damping ratio; flow-induced vibration; mass ratio; reduced velocity;
Citation: Alam, M.M. Effects of Mass wake interaction; energy harvesting
and Damping on Flow-Induced
Vibration of a Cylinder Interacting
with the Wake of Another Cylinder at
High Reduced Velocities. Energies 1. Introduction
2021, 14, 5148. https://doi.org/
Flow over cylindrical structures is ubiquitous in engineering fields such as naval
10.3390/en14165148
engineering (submarines, ship propellers), offshore engineering (semisubmersibles, spar,
gravity platforms, and jackets), renewable energy engineering (offshore wind and tidal tur-
Academic Editor: Satoru Okamoto
bines), nuclear engineering (reactors, cooling tower, chimneys), civil engineering (skyscrap-
ers and cables of suspension bridges), electrical engineering (power lines), etc. These
Received: 8 July 2021
Accepted: 15 August 2021
cylindrical structures undergo undesirable flow-induced vibrations because of fluctuating
Published: 20 August 2021
forces induced by flow separation and alternate vortex shedding. When more than one
cylinder is in a group, the fluid force acting on a cylinder in the group is different from
Publisher’s Note: MDPI stays neutral
that on a single isolated cylinder. The difference arises from the mutual fluid–structure
with regard to jurisdictional claims in
interactions between the cylinders. The flow over a cylinder placed in the wake of another
published maps and institutional affil- cylinder is considered as the baseline to study fluid–structure interactions between the
iations. cylinders in a group. In this case, the wake cylinder (downstream cylinder) receives a
strong interaction from the wake-generating cylinder (upstream cylinder).
Flow-induced vibrations commonly include instability-induced excitation (vortex-
induced vibration) and movement-induced excitation (galloping). Vortex-induced vibra-
Copyright: © 2021 by the author.
tion (VIV) usually occurs over a limited range of reduced velocity, hence, is self-limiting
Licensee MDPI, Basel, Switzerland.
and is a kind of resonance, which occurs when the shedding frequency equals the oscil-
This article is an open access article
lation frequency of the structure. On the other hand, galloping is a movement-induced
distributed under the terms and excitation where the vibration amplitude grows with increasing reduced velocity. The
conditions of the Creative Commons galloping vibration is generated when the shedding frequency is greater or smaller than
Attribution (CC BY) license (https:// the oscillation frequency of the structure. Flow-induced vibration is a function of mass
creativecommons.org/licenses/by/ ratio m* , damping ratio ζ, natural frequency fn , Reynolds number Re, reduced velocity Ur
4.0/).

Energies 2021, 14, 5148. https://doi.org/10.3390/en14165148 https://www.mdpi.com/journal/energies


Energies 2021, 14, 5148 2 of 13

(=U/(fn D)), structure shape, turbulent intensity, etc., where U is the freestream velocity
and D is the cylinder diameter.
Several experimental investigations have been conducted on flow-induced vibrations of
two tandem cylinders, investigating cylinder vibration and frequency responses, lift forces,
vortex shedding, wake interference, vortex structures, vortex impingement and phase lag
between displacement and lift force [1–7]. When an elastic (e.g., spring-mounted) cylin-
der is placed in the wake of a fixed upstream cylinder, the elastic cylinder may undergo
both VIV and galloping. With the cylinder mass-damping ratio m* ζ = 0.109, Bokaian
and Geoola [1] found four distinct dynamic responses, including galloping vibration at
L/D = 1.09, VIV at L/D > 3, separated VIV and galloping vibration at 2 < L/D < 3, and com-
bined VIV and galloping vibration at L/D = 1.5, where L is the cylinder center-to-center
spacing, D is the cylinder diameter. Hover and Triantafyllou [3] examined flow-induced
vibrations of two tandem cylinders with L/D = 4.75 and Re = 3.05 × 104 . They observed VIV
and galloping to occur for Ur = 2.0–17.0. Assi et al. [8] at m* ζ = 0.013, L/ D = 2.0–5.6 and
Re = 0.3 × 104 –1.3 × 104 found that the vibration amplitude increases monotonically when
Ur is increased from 2 to 12. The amplitude curve did not show a VIV peak. Although termed
as galloping-like vibration, it was essentially combined VIV and galloping. Korkischko and
Meneghini [9] presented amplitude and frequency responses of the wake cylinder when
L/D = 2.0–6.0 and Re = 0.1 × 104 –1.0 × 104 . They also identified combined VIV and galloping.
When both cylinders arranged in tandem are elastic and allowed to oscillate in
the cross-flow direction, the fluid–structure coupling is more complicated, resulting in
galloping vibrations although a single cylinder does not experience galloping [7,10].
Zdravkovich [11] experimented with flow-induced vibrations of two cylinders at high
m* ζ = 50 with L/D = 4. The generation of galloping vibration was delayed to Ur ≈ 50. Kim
et al. [12] at m* ζ = 0.65, L/D = 1.1–4.2 and Ur = 1.5–26 reported that vibration response
is highly dependent on L/D. For L/D = 2.0–4.0 and m* ζ = 0.043, Huera-Huarte andand
Bearman [13] revealed that the wake-generating cylinder has a larger vibration ampli-
tude than the wake cylinder for L/D = 2–2.5 at Ur = 4–9 while the wake cylinder with
L/D = 3.0–4.0 experiences galloping vibrations at Ur > 9.
Compared to the experimental investigations, numerical studies on flow-induced vi-
brations of two tandem cylinders are small in number. Papaioannou et al. [14] numerically
investigated the effect of L/D on the vibrations of two cylinders at Re = 160, m* ζ = 0.127,
Ur = 3–10, and L/D = 2.5, 3.5 and 5. The Ur range of VIV of the wake-generating cylinder
was larger at a smaller L/D. Two- and three-dimensional numerical simulations were car-
ried out by Borazjani and Sotiropoulos [15] to examine the vibration responses of tandem
cylinders at Re = 200, L/D = 1.5, m* = 2.6 and ζ = 0. They proved that although the flow at
Re = 200 is three-dimensional for a single cylinder, two-dimensional simulations are enough
up to Re = 200 for two tandem cylinders. At small Ur (≤4), the wake-generating cylin-
der had a larger vibration amplitude than the wake cylinder. On the other hand, the
opposite scenario prevailed at large Ur (=7–14). Given that their investigated Ur range is
small in both Papaioannou et al. ([14], Ur = 3–10) and Borazjani and Sotiropoulos ([15],
Ur = 3–14), only VIV responses were observed in both investigations, albeit the VIV range
was wider in Borazjani and Sotiropoulos [15] than in Papaioannou et al. [14]. When the
wake-generating cylinder is fixed and the wake cylinder is allowed to oscillate, Carmo
et al. [2] examined vibration responses of the downstream cylinder at Re = 150 and 300,
m* = 2.6, ζ = 0.007, Ur = 3–30. They found that, compared with the single cylinder counter-
part, the maximum vibration amplitude of the wake cylinder is large, the VIV regime is
wide, and the vibration amplitudes at high Ur, outside the lock-in, are significant. They
ascribed these differences to the flow oscillation in the gap between the cylinders. Prasanth
and Mittal [16] examined flow-induced vibrations of two tandem cylinders at m* = 10,
Re = 100, L/D = 5.5, and Ur = 2–15. Each cylinder was allowed to vibrate in both trans-
verse and streamwise directions. Although L/D (=5.5) is large, the downstream cylinder
vibration considerably affected the upstream cylinder vibration response. Compared to a
single cylinder, the upstream cylinder had lock-in at smaller Ur, with the vortex shedding
affected the upstream cylinder vibration response. Compared to a single cylinder, the up-
stream cylinder had lock-in at smaller Ur, with the vortex shedding frequency away from
the natural frequency of the cylinder. Hysteresis was observed in the responses of both
Energies 2021, 14, 5148 cylinders. The review of the literature suggests that numerical investigations at 3low of 13Re are
scarce for the case where the wake cylinder is free to oscillate but the wake-generating
cylinder is fixed. There are a few key issues to be resolved. For example, what are the
effects of m* and ζ on VIV responses and wake topology for the wake cylinder interacting
frequency away from the natural frequency of the cylinder. Hysteresis was observed
with a wake generated by another cylinder? Are the effects dependent on L/D? Is mass-
in the responses of both cylinders. The review of the literature suggests that numerical
damping ratio mat*ζlow
investigations a unique parameter
Re are scarce for thetocase
characterize the vibration?
where the wake cylinder isThe
free objective
to oscillateof this
work
but to
theinvestigate the effects
wake-generating of m
cylinder
* (=2–32),
is fixed. ζ (=0–0.5),
There are a fewmkey
* ζ (=0–8)
issues and
to beL/D (=1.5 and
resolved. For 3) on
flow-induced
example, what vibration
are the effects of m* andsubmerged
of a cylinder in the and
ζ on VIV responses wake of another.
wake topology Numerical
for the wake simu-
lations areinteracting
cylinder conducted at aRewake
with = 150 for the by
generated parametric ranges Are
another cylinder? mentioned
the effectsabove. The Ur is
dependent
on L/D? Is mass-damping * ζ a unique parameter to characterize the vibration? The
ratio mflow
varied from 2.5 to 30. A low-Re provides a better understanding of flow physics, and
oneobjective of this
can clearly work tothe
observe investigate
development the effects of m*layers,
of shear (=2–32), formation
* ζ (=0–8) and
ζ (=0–0.5),ofmvortices, evolution
of L/D (=1.5 and 3) on flow-induced vibration of a cylinder submerged in the wake of another.
vortices, etc. Although major flow phenomena at both low- and high-Re flows are the
Numerical simulations are conducted at Re = 150 for the parametric ranges mentioned
same (e.g., vortex shedding, lock-in, vibration nature), quantitative magnitudes of forces,
above. The Ur is varied from 2.5 to 30. A low-Re flow provides a better understanding
Strouhal
of flow number,
physics, andlock-in range,
one can maximum
clearly vibration
observe the amplitude,
development of shearetc. differ
layers, between the
formation
low- and high-Re
of vortices, flows.
evolution The results
of vortices, etc. from a low-Re
Although major flow can, thus, not
flow phenomena be extrapolated
at both low- and to
a high-Re flow.
high-Re flows are the same (e.g., vortex shedding, lock-in, vibration nature), quantitative
magnitudes of forces, Strouhal number, lock-in range, maximum vibration amplitude, etc.
differ between
2. Numerical the low- and high-Re flows. The results from a low-Re flow can, thus, not
Methodology
be extrapolated to a high-Re flow.
2.1. Computational Domain
2. Numerical
The flow isMethodology
given in a rectangular computational domain where two circular cylin-
2.1.are
ders Computational
arranged in Domain
tandem at the horizontal centerline of the domain (Figure 1). The
wake-generating cylinder
The flow is given is fixed, and
in a rectangular the wake cylinder
computational is spring-mounted.
domain where The latter
two circular cylinders
are arranged in tandem at the horizontal centerline of the domain (Figure 1).
cylinder is allowed to oscillate in the transverse direction only. The total spring stiffness The wake-
generating
is k. cylinder
The Cartesian is fixed, and
coordinate the wake
system cylinder
(x–y) has itsis origin
spring-mounted. Theof
at the center latter
thecylinder
wake-gener-
is allowed to oscillate in the transverse direction only. The total spring
ating cylinder. The inlet and outlet boundaries of the computational domain are stiffness is k. The
placed at
Cartesian coordinate system (x–y) has its origin at the center of the wake-generating cylinder.
x = −15 D and L + 45 D, respectively, where L is the distance between the centers of the
The inlet and outlet boundaries of the computational domain are placed at x = −15 D and
cylinders,
L + 45 D,and D is thewhere
respectively, diameter. The
L is the upperbetween
distance and lower boundaries
the centers of the are symmetrically
cylinders, and D is sep-
arated by 30 D from each other [17,18]. The corresponding cylinder
the diameter. The upper and lower boundaries are symmetrically separated by 30 D blockage ratio
fromis 3.3%
[19].
each other [17,18]. The corresponding cylinder blockage ratio is 3.3% [19].

∂u/∂y = 0, v = 0

15D
k
y
u=U x ∂u/∂x = 0
D
v=0 ∂v/∂x = 0
Wake-generating cylinder Wake cylinder

15D

∂u/∂y = 0, v = 0

15D L 45D

Figure 1. A
Figure schematic
1. A schematicofofcylinder configurationand
cylinder configuration and computational
computational domain.
domain.
Energies 2021, 14, x FOR PEER REVIEW 4 of 14

Energies 2021, 14, 5148 4 of 13


2.2. Governing Equations and Numerical Technique
The continuity and Navier–Stokes equations are the governing equations that can be
written as:
2.2. Governing Equations and Numerical Technique
¶u * ¶v *
The continuity and Navier–Stokes equations + =are
0, the governing equations that can(1) be
¶x * ¶y *
written as:
∂u* ∗ ∂v∗*
∂u * ∂ u *
∂ u + ∂p = 10,  ∂ 2 u * ∂ 2 u *  (1)
*
+ u * * + v * ∂x* ∗= − ∂y∗* +  *2 + *2  , and (2)
∂t ∂x ∂y ∂x  ∂x
Re  ∂y  
∂u∗ ∗ ∂u

∗ ∂u
∗ ∂p∗ 1 ∂2 u ∗ ∂2 u ∗
+ u + v = − + + , and (2)
∂t∗ ¶v ∂x ∗ * ¶v ∂y∗* ¶v ∂x ∗ ¶p Re 1∂xçæ ¶
* * * * ∗ 2 v ∂y¶∗22v * ö
2 *
÷÷ .
+ u + v = - + ç + *2 ÷ (3)
* * * * ç *2 ÷
∂v∗ ¶t ∗ ∂v∗¶x ∗ ∂v¶ ∗y ∂p¶∗y 1Re è∂¶2xv∗ ¶∂y2 v∗ø
+u +v =− ∗ + + ∗2 . (3)
The parameters∂t∗in the equations
∂x ∗ ∂y ∗ ∂y
are non-dimensionalRe ∂x ∗2
and have ∂y the following expres-
sions.The parameters in the equations are non-dimensional and have the following expressions.
x u v
x * = x , ∗y * = yy , u∗* = u, v*∗ = v , p∗* = p p2 , Re = ρUD ρUD , and t * = tU tU .

x = D, y = D, u =U , v =U , p =ρU 2 , Re = μ , and t∗ = D .
D D U U ρU µ D
Here, u and ν are the streamwise and transverse components of the flow velocity,
Here, u pand
respectively, ν are
is the thepressure,
static streamwise μ isand
the transverse components
fluid viscosity, of the flowvelocity,
U is the freestream velocity,
respectively,
and p is the static pressure, µ is the fluid viscosity, U is the freestream velocity, and
t is the time.
t is the
Thetime.
cylinders are surrounded by an O-xy grid system while a rectangular grid system
The cylinders
is provided away fromare surrounded
the cylinders,by anas O-xy
shown grid
in system
Figure while
2. Theameshes
rectangulararoundgridthe
system is
cylin-
provided
ders awayafrom
are given the density.
greater cylinders,We as shown
set the in Figure
first grid2.level
The0.009D
meshesawayaround the the
from cylinders are
cylinder
given a greater
surface, density.
with a mesh We set the
expansion firstofgrid
ratio lesslevel
than0.009D
1.1. Theaway from theuses
simulation cylinder surface,mesh
a dynamic with
a mesh where
scheme expansion ratio of less than151.1.
the ANSYS-Fluent The is
solver simulation
used to moveuses boundaries
a dynamic mesh and/orscheme
objectswhere
and
the ANSYS-Fluent 15 solver is used to move boundaries and/or objects
to adjust the mesh accordingly. The grid box around the cylinder moves with the cylinder and to adjust the mesh
accordingly. The
displacement. Thegrid box aroundfunction
user-defined the cylinder moves
is feed into with the cylinder
the solver displacement.
to estimate The
the cylinder
user-defined function is feed into the solver to estimate the cylinder
displacement. At each time step, the domain deformation is handled by the dynamic displacement. At each
time step,tool
meshing the in
domain deformation
ANSYS-Fluent 15,isand
handled
the meshby the dynamic meshing
is updated using thetool in ANSYS-Fluent
Laplace smoothing
15, and the mesh is updated using the Laplace smoothing method.
method.

Figure 2.
Figure (a) Grid
2. (a) Grid system
system in
in the
the computational
computational domain
domain and
and (b)
(b) close-up
close-up view
view of
of grids
grids around
around the
the cylinders. L/D == 1.5.
cylinders. L/D 1.5.

The boundary conditions are given as


The boundary conditions are given as
u* = 0 and v* = 0 at the surfaces of the upstream and downstream cylinders,
u**= 0 and v* *= 0 at the surfaces of the upstream and downstream cylinders,
u = 1 and v* = 0 at the inlet,
u* =* 1 and v = 0 at the inlet,
∂u / ∂y* * = 0 and *v* = 0 at the lateral sides, and
∂u*/∂y = 0 and v = *0 at *the lateral sides, and
∂u** / ∂x* * = 0 and ∂v / ∂x = 0 at the outlet.
∂u /∂x = 0 and ∂v*/∂x* = 0 at the outlet.* *
In the governing Equations (1)–(3), p* , u* and v** are the unknown parameters, solved
In the governing Equations (1)–(3), p , u and v are the unknown parameters, solved
by coupling the governing equations. They are solved for the unsteady and incompressible
by coupling the governing equations. They are solved for the unsteady and incompressi-
flow with constant fluid properties. The computations are conducted using the finite vol-
ble
ume flow with constant
method. fluid properties.
The convective components Theare
computations are conducted
discretized using using the
the second-order finite
upwind
volume method. The convective components are discretized using the second-order
scheme. The coupling between the velocity and pressure fields is done by the pressure up-
correction-based iterative algorithm SIMPLE (Semi-implicit method for pressure linked
equations) proposed by Patankar [20]. We used a first-order implicit formulation for the
time discretization.
wind scheme. The coupling between the velocity and pressure fields is done by the pres-
Energies 2021, 14, 5148
sure correction-based iterative algorithm SIMPLE (Semi-implicit method for pressure 5 of 13
linked equations) proposed by Patankar [20]. We used a first-order implicit formulation
for the time discretization.
The governing equation of the cylinder motion in the dimensionless form can be writ-
The governing equation of the cylinder motion in the dimensionless form can be
ten as:
written as:
Y4πF Y +(2πF
.. .
Y+ + 4πnF
ζnYζ + (2πnF)n2)Y2Y==CCLiLi /2m
/ 2m∗*, (4)
.
where Y
where Y is
is the
the cylinder
cylinderdisplacement
displacementmeasured
measuredfrom fromyy==0,0,𝑌Y is the the cylinder
cylinder velocity,
velocity, and
..
𝑌Y is the cylinder acceleration.
acceleration. The The CCLiLiisisthe
theinstantaneous
instantaneouslift liftcoefficient
coefficientof ofthe
thecylinder.
cylinder.
The
The FFnn==fnfD/U,
D/U, where
where fn is
f the
is natural
the natural frequency
frequency of the
of cylinder.
the cylinder. The
The m m * (=4
* (=4 m/ρπD
m/ρπD 2) is2the
) is
n n
cylinder massmass
the cylinder ratio, where
ratio, m stands
where m standsfor the
for mass of the
the mass cylinder.
of the We We
cylinder. solved Equation
solved Equation (4)
using the fourth-order
(4) using the fourth-order Runge-Kutta
Runge-Kuttamethod for every
method time step.
for every timeThestep.Re The
= 150Refor= all
150sim-
for
ulations including
all simulations validation.
including The flowThe
validation. around
flow aaround
single acylinder transits from
single cylinder transitstwo-fromto
three-dimensional at Re ≈ 190
two- to three-dimensional Re ≈When
at [21]. 190 [21].twoWhen
cylinders
twoare placed are
cylinders in tandem,
placed in thetandem,
transi-
the transition
tion Re (Re
Re (Recr2) from cr2 ) from two-dimensional
two-dimensional to three-dimensional
to three-dimensional flows is
flows is delayed L/D ≤ 3.0,
fordelayed for
see ≤ 3.0,3see
L/DFigure Figure 3 reproduced
reproduced from Rastan from and AlamRastan and
[22]. TheAlam
flow[22].
aroundThetwo flow around
tandem two
cylin-
tandem
ders cylinders
at L/D at L/D
= 1.5 and = 1.5 and 3.0
3.0 examined examined
is thus assumedis thus
to beassumed to be two-dimensional.
two-dimensional.

100000
105
10000
104 Re →
Re ← Recr1
1000
103
1SB 2 AR 3 4
BS 5 6 7
CS 8 9 10
SB AR HS CS Recr1o , single cylinder
500
450
400 Recr2
3D vortex shedding
350 Mode A

300
Re

Mode B
250
200 Mode A, Recr2o
Single cylinder
150 Mode B
Laminar unsteady flow
100 – 2D vortex shedding
Flow regime border
50
Laminar steady flow (2D) 2D/3D hysteresis
0
1 2 3 4 5 6 7 8 9 10 Steady/unsteady hysteresis
L/D

Marginal flow
Figure 3. Marginal flow map
map (colored
(colored background)
background) andand flow
flow regimes
regimes bordered
bordered by
by purple
purple lines
lines for
for two
two tandem
tandem circular
circular
cylinders. SB,
SB, slender/extended-body
slender/extended-bodyflow;
flow;AR,
AR,alternate
alternatereattachment
reattachmentflow;
flow;HS,
HS,hysteresis
hysteresiszone;
zone;BS,
BS,bi-stable
bi-stableflow;
flow; CS,
CS,
co-shedding
co-shedding flow.
flow. Reproduced
Reproduced from
from Rastan
Rastan and
and Alam
Alam [22].
[22]. For
For the
the definitions
definitions of
of the
the legends,
legends, please
please refer
refer to
to Rastan
Rastan and
and
Alam
Alam [22].
[22]. Re
Recr1 represents the flow transition from steady to two-dimensional unsteady while Recr2 indicates the flow
cr1 represents the flow transition from steady to two-dimensional unsteady while Recr2 indicates the flow
transition from two-dimensional unsteady to three-dimensional unsteady.
transition from two-dimensional unsteady to three-dimensional unsteady.

2.3. Validation
2.3. Validation
A mesh independence test was conducted based on the mesh system adopted in
A mesh independence test was conducted based on the mesh system adopted in Zafar
Zafar and Alam [23] and Abdelhamid et al. [24]. Three different meshes (M1, M2, and M3
and Alam [23] and Abdelhamid et al. [24]. Three different meshes (M1, M2, and M3 consisting
consisting of 76,350, 80,050, and 9400 elements, respectively) are tested for a single fixed
of 76,350, 80,050, and 9400 elements, respectively) are tested for a single fixed cylinder at
cylinder at Re = 150. The results of global parameters including time-mean drag coefficient
Re = 150. The results of global parameters including time-mean drag coefficient C D , fluctuating
C D coefficient
lift , fluctuating
CL0 ,lift
andcoefficient C L¢ , and
Strouhal number Strouhal
St for numbermeshes
three different St for three differentinmeshes
are presented Table 1
are
andpresented
comparedinwith Table 1 and
those in compared withThe
the literature. those in theforliterature.
results the threeThemesh results for the
systems are
three mesh converged
essentially systems areasessentially
these mesh converged
systems as these
were mesh systems
decided based on were
the decided based
mesh systems
in [23–26].
on the mesh The maximum
systems deviation
in [23–26]. C D is found
Theinmaximum to be less
deviation C D 2.5%.
in than The to
is found present results
be less than
overall show quite good agreement with those from the literature. Mesh
2.5%. The present results overall show quite good agreement with those from the litera- M2 system, with the
sameMesh
ture. grid distributions
M2 system,around thesame
with the two cylinders, was, however,
grid distributions around usedthe
fortwo
the computations
cylinders, was,of
two tandem cylinders when the wake cylinder is free to oscillate, and the results are validated
again in Figure 4. To compare our results with the numerical results of Carmo et al. [2], we
first simulated vibration responses for m* = 2, ζ = 0.003, and L/D = 3 which are the same
geometrical and physical conditions used by Carmo et al. [2]. A comparison between the
however, used for the computations of two tandem cylinders when the wake cylinder is
free to oscillate, and the results are validated again in Figure 4. To compare our results
with the numerical results of Carmo et al. [2], we first simulated vibration responses for
m* = 2, ζ = 0.003, and L/D = 3 which are the same geometrical and physical conditions used
Energies 2021, 14, 5148 6 of 13
by Carmo et al. [2]. A comparison between the present and Carmo et al.’s results is made
in Figure 4 where the vibration amplitude ratio A/D is presented against Ur. Here, A rep-
resents the cylinder vibration amplitude obtained from displacement signal Y. First, root-
present (rms)
mean-square and Carmo
valueetofal.’s
Y (Yresults is made inand
rms) is obtained, Figure
then4 where the vibration
A is calculated A = Yrms ´ ratio
as amplitude 2 A/D
presented against Ur.
. The present results agree well with those by Carmo et al. [2], with a maximum deviation from
is Here, A represents the cylinder vibration amplitude obtained
displacement
of 5% occurring = 10.Y. First,√
signal
at Ur root-mean-square (rms) value of Y (Yrms ) is obtained, and then A
is calculated as A = Yrms × 2. The present results agree well with those by Carmo et al. [2],
with
Table 1. Mesha maximum
independence deviation of 5%
test results foroccurring at Ur
a single fixed = 10. at Re = 150.
cylinder

Table 1. Mesh independence testMesh


results for Elements C D at Re = C
a single fixed cylinder L¢
150. St
M1 76,350 1.3487¯ 0.3529 0 0.1850
Mesh Elements St
Present M2 80,050 1.3495CD 0.3530CL 0.1851
M3 M1 94,400 76,350 1.3499
1.3487 0.3532
0.3529 0.18510.1850
Present
Sucker and Brauer [27] M2 80,050 1.361.3495 0.3530 0.1851
M3 94,400 1.3499 0.3532 0.1851
Norberg [28] -- 0.3555 0.1830
Sucker and Brauer [27] 1.36
Silva etNorberg
al. [29] [28] 1.37 – -- 0.3555 -- 0.1830
Posdziech andSilva
Grundmann
et al. [29] [30] 1.32241.37 -- – 0.1841 –
Posdziech and
Qu et al. [31]Grundmann [30] 1.3224
1.3152 0.3560 – 0.18450.1841
Abdelhamid Quetet al.
al. [24]
[31] --1.3152 0.35370.3560 0.18280.1845
Abdelhamid et al. [24] – 0.3537 0.1828

Figure 4. Validation of wake-cylinder vibration amplitude A/D plotted against reduced velocity
Figure 4. Validation of wake-cylinder vibration amplitude A/D plotted against reduced velocity Ur.
Ur.* L/D = 3, m* = 2, ζ = 0.003, and Re = 150.
L/D = 3, m = 2, ζ = 0.003, and Re = 150.
3. Results and Discussion
3. Results and Discussion
3.1. Vibration Response at Small m* and ζ
3.1. Vibration Response
Figure at Small
5 shows them* and ζ
dependence of vibration amplitude ratio A/D on Ur for m* = 2,
Figure 5 shows
ζ = 0.005 the dependence
and L/D of vibration
= 3.0. The A/D amplitude
firstly increases withratio A/D on Ur,
increasing Ur for m* = 2, aζ peak
reaching = at
0.005 and
Ur =L/D = 3.0. The
7 before A/D firstly
declining with aincreases with increasing
further increase Ur, decline
in Ur. The reachingisarapid
peak for
at Ur
Ur== 7–15
but
7 before mild forwith
declining Ur >a15. The vorticity
further increase instructures shown is
Ur. The decline inrapid
Figure
for5b,c
Ur illustrate
= 7–15 butthat
mildeach of
the wakes at Ur = 7 and 15 is characterized by two rows of opposite sign vortices. Yet, the
two wakes differ in different aspects. Firstly, the lateral separation between the two vortex
rows is wider for Ur = 7 than for Ur = 15. Secondly, the streamwise distance between two
consecutive vortices is large for Ur = 15, compared to that for Ur = 7. As the vibration
amplitude gently decreases with Ur for Ur > 15 (Figure 5a), there is no galloping. In a
water tunnel test, Bokaian and Geoola [1] observed galloping vibration at 1.09 ≤ L/D ≤ 3
in the range of Re = 600–6000 with m* ζ = 0.109. When both cylinders are allowed to
for Ur > 15. The vorticity structures shown in Figure 5b,c illustrate that each of the wakes
at Ur = 7 and 15 is characterized by two rows of opposite sign vortices. Yet, the two wakes
differ in different aspects. Firstly, the lateral separation between the two vortex rows is
wider for Ur = 7 than for Ur = 15. Secondly, the streamwise distance between two consec-
utive vortices is large for Ur = 15, compared to that for Ur = 7. As the vibration amplitude
Energies 2021, 14, 5148 7 of 13
gently decreases with Ur for Ur > 15 (Figure 5a), there is no galloping. In a water tunnel
test, Bokaian and Geoola [1] observed galloping vibration at 1.09 ≤ L/D ≤ 3 in the range of
Re = 600–6000 with m* ζ = 0.109. When both cylinders are allowed to vibrate, Kim et al. [12]
in a wind
vibrate, Kimtunnel test found
et al. [12] thetunnel
in a wind occurrence of galloping
test found vibrations
the occurrence for 1.2 ≤ L/D
of galloping ≤ 1.6 atfor
vibrations Re
= 4365–74,200
1.2 ≤ L/D ≤ 1.6and m ζ= =4365–74,200
at Re * 0.64. It suggests
and m thatζ the
* occurrence
= 0.64. of galloping
It suggests that the is highly sensi-
occurrence of
tive to L/D,
galloping Re, and
is highly m* ζ. Interestingly,
sensitive to L/D, Re, and as m * ζ. Interestingly,
seen in Figure 5, the A/D in
as seen value
Figureremains high
5, the A/D
(A/D remains
value > 0.4) even
high at (A/D
high >Ur0.4)
(>15).
evenIt at
is high
worth Urinvestigating whether
(>15). It is worth galloping whether
investigating occurs at
other m* and
galloping ζ values.
occurs at otherWe m* will
and therefore focus
ζ values. We ontherefore
will the vibration
focus response at largeresponse
on the vibration Ur (>10)
only.
at large Ur (>10) only.

Figure (a) Variation


Figure5.5. (a) Variation ininvibration
vibrationamplitude A/Dwith
amplitudeA/D with reduced
reduced velocity Ur.Ur.
velocity Vorticity
Vorticity structures
structures (b)=Ur
for Ur
for (b) = 7(c)
7 and andUr
(c) Ur = 15. L/D = 3, m * = 2, and ζ = 0.005.
= 15. L/D = 3, m* = 2, and ζ = 0.005.

3.2.
3.2.Effect
EffectofofDamping
DampingRatio Ratioon
onVibration
VibrationAmplitude
Amplitude
The
Thevibration
vibrationresponse
response isisinvestigated for Ur
investigated for Ur >> 10
10 when
when ζζ is is increased
increased from
from 00 toto 0.5.
0.5.
A/D * = 2 and L/D = 3.
Figure
Figure 66 depicts the effecteffectof
ofζζ(=0–0.5)
(=0–0.5)onon
A/D UrUr
forfor = 10–30
= 10–30 with m* =m2 and
with L/D = 3. There
There
is no is no apparent
apparent difference
difference in A/D
in A/D between
between ζ =ζ 0= and
0 and0.0050.005while
whileA/DA/D for
for both ζζ values
values
diminishes, particularly at Ur
diminishes, particularly at Ur = 10–15, with increasing ζ for ζ > 0.005. The A/Dat
= 10–15, with increasing ζ for ζ > 0.005. The A/D atthe
thelargest
largest
ζζ== 0.5
0.5 examined
examined isis not
notvery
verysensitive
sensitivetotoUr.
Ur.ItItshould
shouldbebenotednotedthatthatA/DA/Disisstill
stilllarger
largerthan
than
0.4
0.4even
evenatatthethelargest
largestζ ζ==0.5
0.5examined.
examined.FromFromFigure
Figure6,6,it itis is
understood
understood that
thatthethe
degree
degreeof of
ζ
effect on on
ζ effect A/DA/Dis not the the
is not same at different
same Ur values.
at different Then,Then,
Ur values. what what
is theisrelationship between
the relationship be-
and A/D
ζtween ζ andat A/D
different Ur values?
at different A further
Ur values? question
A further arises,arises,
question is theisζ the
effect on A/D
ζ effect on is
A/Dtheis
same at other m * values?
the same at other m values?
*
Figure 7 answers the above-mentioned two questions where A/D is presented against
ζ for m* = 2 and 16 in the case of Ur = 10, 15 and 21. As seen in the figure, an increase in ζ
causes a reduction in A/D for both m* = 2 and 16 with Ur = 10 (Figure 7a). The reduction is,
however, large at small m* , the gradient (d(A/D)/dζ) in the range of ζ = 0–0.5 being −0.42
and −0.15 for m* = 2 and 16, respectively. When m* is increased from 2 to 16, dramatic
reductions in A/D occur for all ζ values (Figure 7a). Although a similar observation is made
for Ur = 15 (Figure 7b), the gradient magnitude is smaller for Ur = 15 than for Ur = 10. The
A/D becomes almost insensitive to ζ for Ur = 21. Overall, the effect of ζ on A/D is large at
small m* and small Ur. The m* effect on A/D is significant for all ζ values, albeit larger at a
smaller ζ. Given that the ζ values of engineering structures appear less than 0.1 and the m*
Energies 2021, 14, 5148 8 of 13

Energies 2021, 14, x FOR PEER REVIEW 8 of 14


may vary up to the order of 3 [32–35], it can be said the effect of m* on A/D is much more
than that of ζ.

Figure 6. Effect of damping ratio ζ on vibration amplitude A/D. L/D = 3 and m* = 2.

Figure 7 answers the above-mentioned two questions where A/D is presented against
ζ for m* = 2 and 16 in the case of Ur = 10, 15 and 21. As seen in the figure, an increase in ζ
causes a reduction in A/D for both m* = 2 and 16 with Ur = 10 (Figure 7a). The reduction
is, however, large at small m*, the gradient (d(A/D)/dζ) in the range of ζ = 0–0.5 being −0.42
and −0.15 for m* = 2 and 16, respectively. When m* is increased from 2 to 16, dramatic
reductions in A/D occur for all ζ values (Figure 7a). Although a similar observation is
made for Ur = 15 (Figure 7b), the gradient magnitude is smaller for Ur = 15 than for Ur =
10. The A/D becomes almost insensitive to ζ for Ur = 21. Overall, the effect of ζ on A/D is
large at small m* and small Ur. The m* effect on A/D is significant for all ζ values, albeit
larger at a smaller ζ. Given that the ζ values of engineering structures appear less than 0.1
and the m* may vary up to the order of 3 [32–35], it can be said the effect of m* on A/D is
Figure6.
Figure 6. Effect
Effect of
of damping
dampingratio
ratioζζon
onvibration amplitudeA/D.
vibrationamplitude A/D.L/D
L/D== 33 and m** = 2.
and m
much more than that of ζ.
Figure 7 answers the above-mentioned two questions where A/D is presented against
ζ for m* = 2 and 16 in the case of Ur = 10, 15 and 21. As seen in the figure, an increase in ζ
causes a reduction in A/D for both m* = 2 and 16 with Ur = 10 (Figure 7a). The reduction
is, however, large at small m*, the gradient (d(A/D)/dζ) in the range of ζ = 0–0.5 being −0.42
and −0.15 for m* = 2 and 16, respectively. When m* is increased from 2 to 16, dramatic
reductions in A/D occur for all ζ values (Figure 7a). Although a similar observation is
made for Ur = 15 (Figure 7b), the gradient magnitude is smaller for Ur = 15 than for Ur =
10. The A/D becomes almost insensitive to ζ for Ur = 21. Overall, the effect of ζ on A/D is
large at small m* and small Ur. The m* effect on A/D is significant for all ζ values, albeit
larger at a smaller ζ. Given that the ζ values of engineering structures appear less than 0.1
and the m* may vary up to the order of 3 [32–35], it can be said the effect of m* on A/D is
much more than that of ζ.

Figure 7.
Figure 7. Effect of damping ratio ζζ on vibration
vibration amplitude
amplitudeA/D
A/Dat differentmm* *and
atdifferent andUr.
Ur.(a)
(a)Ur
Ur= =10,
10,
(b) Ur
(b) Ur == 15, and (c) Ur = 21. L/D == 3.
3.

3.3. Effect of Mass Ratio on Vibration Amplitude


More detail of the effect of m* on A/D is presented in Figure 8 for L/D = 3.0 and 1.5
with ζ = 0. For both L/D values, A/D decreases with increasing m* , the decrease being
large at small Ur. Considering the steps of the increase in m* , the decrease in A/D is largest
between m* = 2 and 4. When m* is large enough (i.e., m* = 32 and 16 for L/D = 3.0 and 1.5,
respectively), A/D becomes very small and almost independent of Ur. There is a large drop
in A/D for m* = 4 from Ur = 10 to 15. As seen, A/D is highly sensitive to m* for m* = 2–8.
At m* = 4, it seems that A/D at Ur = 10 and Ur ≥ 15 has similar characteristics to that at

Figure 7. Effect of damping ratio ζ on vibration amplitude A/D at different m* and Ur. (a) Ur = 10,
(b) Ur = 15, and (c) Ur = 21. L/D = 3.
between Ur = 10 and 15, which requires further investigation.
Figure 9 shows the dependence of A/D on m*, Ur and ζ for L/D = 1.5. At Ur = 10 (Figure
9a), with increasing m*, the A/D exponentially drops between m* = 2 and 16 and gently for
m* > 16. The same figure (Figure 9a) further proves that there is an effect of ζ on A/D, the
Energies 2021, 14, 5148 effect being small between ζ = 0 and 0.05 but strong between ζ = 0.05 and 0.5. At Ur9=of15 13
(Figure 9b), again the effect of m* on A/D is very strong for m* = 2–16 and weak for m* > 16.
Here, the effect of ζ on A/D is very small. Figure 9c further describes the relationship be-
tween A/D, m* and Ur. For all Ur values, A/D exponentially declines with increasing m* up
m*16.
to = In
2 and 8, respectively.
addition, There mightbetween
A/D drops significantly be a transition from
Ur = 10 and 15high- to low-amplitude
and mildly between Ur
vibration between Ur = 10 and 15, which requires further investigation.
= 15 and 30. It can be generalized that A/D is inversely linked to m , ζ and Ur.
*

Figure
Figure8.8.Effect
Effectof
ofmass ratiomm* *on
massratio onvibration
vibrationamplitude A/D.(a)
amplitudeA/D. L/D==3.0
(a)L/D 3.0and
and(b) L/D==1.5.
(b)L/D 1.5.ζ ζ= =0.0.

Figure 9 shows the dependence of A/D on m* , Ur and ζ for L/D = 1.5. At Ur = 10


(Figure 9a), with increasing m* , the A/D exponentially drops between m* = 2 and 16 and
gently for m* > 16. The same figure (Figure 9a) further proves that there is an effect of ζ
on A/D, the effect being small between ζ = 0 and 0.05 but strong between ζ = 0.05 and
0.5. At Ur = 15 (Figure 9b), again the effect of m* on A/D is very strong for m* = 2–16 and
weak for m* > 16. Here, the effect of ζ on A/D is very small. Figure 9c further describes the
relationship between A/D, m* and Ur. For all Ur values, A/D exponentially declines with
increasing m* up to 16. In addition, A/D drops significantly between Ur = 10 and 15 and
mildly between Ur = 15 and 30. It can be generalized that A/D is inversely linked to m* ,
ζ and Ur.

3.4. Effect of Mass-Damping Ratio on Vibration Amplitude


In the previous section, both m* and ζ causes a reduction in A/D. For a single isolated
cylinder, mass-damping ratio (m* ζ) has been proven to be another characteristic parameter
that has a connection with A/D [36]. The A/D is found to monotonically decrease with
increasing m* ζ for a single isolated cylinder [37]. To see the relationship between A/D and
m* ζ for the cylinder interacting with the wake, A/D is plotted against m* ζ in Figure 10 for
Ur = 10 and 15. For these data, ζ varies from 0 to 0.5 for both m* = 2 and 16 (see the legends
in the figure). Interestingly, although A/D decreases with increasing m* ζ for a given m* , the
decrease rate is contingent on m* , high at small m* , and vice versa. The A/D data fail to
collapse onto a single line against m* ζ, which suggests that m* ζ is not a unique parameter
to characterize the vibration for a cylinder interacting with the wake of another cylinder.
s 2021, 14, xEnergies
FOR PEER 14, 5148
2021,REVIEW 10 of 14 10 of 13

0.7 0.7
ζ =0
0.6 Ur = 10 ζ = 0.005 0.6 Ur = 15 ζ =0
ζ = 0.05 ζ = 0.005
0.5 ζ = 0.10 0.5 ζ = 0.05
ζ = 0.50 ζ = 0.10
0.4 0.4 ζ = 0.50
A/D
0.3 0.3

0.2 0.2

0.1 0.1
0
0 5 10 15 20 25 30 35 0
(b) 0 5 10 15 20 25 30 35
(a) m*
0.7 m*

0.6 ζ =0 Ur = 10
Ur = 15
0.5 Ur = 30

0.4
A/D
0.3

0.2

0.1

0
(c) 0 5 10 15 20 25 30 35
m*
EnergiesFigure
2021, 14,9.xEffect
FOR PEER
Figure 9.REVIEW
Effect
of mass ratioofmmass ratio m* on
* on vibration vibrationA/D
amplitude amplitude A/Dζ at
at different different
and ζ and
Ur values Ur (a)
when values
Ur =when Ur = 10 11
10, (b)(a) andof 14
Ur= =1.5.
(c) ζ = 0. L/D 10, (b) Ur = 10 and (c) ζ = 0. L/D = 1.5.

3.4. Effect of Mass-Damping Ratio on Vibration Amplitude


In the previous0.6section, both m* and ζ causes a reduction in A/D. For a single isolated
cylinder, mass-damping ratio (m*ζ) has been proven to be another characteristic parameter
m* = 2
that has a connection with A/D [36]. The A/D is found
Ur =to
10 monotonically decrease with
 m* = 16
increasing m*ζ for a0.4
single isolated cylinder [37]. To see the relationship between A/D and
 m* = 2
m*ζ for the cylinder interacting with the wake, A/D is Ur
 m* = 16
= 15 against m*ζ in Figure 10 for
plotted
A/D
Ur = 10 and 15. For these data, ζ varies from 0 to 0.5 for both m* = 2 and 16 (see the legends
in the figure). Interestingly,
0.2
although A/D decreases with increasing m*ζ for a given m*,
the decrease rate is contingent on m*, high at small m*, and vice versa. The A/D data fail to
collapse onto a single line against m*ζ, which suggests that m*ζ is not a unique parameter
to characterize the vibration for a cylinder interacting with the wake of another cylinder.
0
(a) 0 2 4 6 8

0.6

0.4
A/D

0.2

0
0 2 4 6 8
(b) m*ζ

Figure 10.10.
Figure A/D vs.vs.
A/D mass-damping ratio
mass-damping (m*ζ).
ratio (a)(a)
(m*ζ). L/D = 3.0,
L/D and
= 3.0, (b)(b)
and L/D = 1.5.
L/D = 1.5.

3.5. Effects of L/D, m*, and ζ on Wake Structure


It is worth investigating how the wake structure is modified when L/D, m*, and ζ are
changed. Figure 11 shows vorticity structures for L/D = 3.0 and 1.5 for different values of
m* and ζ. The first- and second-column snapshots correspond to Y ≈ 0 (i.e., mean position)
Energies 2021, 14, 5148 11 of 13

3.5. Effects of L/D, m* , and ζ on Wake Structure


It is worth investigating how the wake structure is modified when L/D, m* , and ζ are
changed. Figure 11 shows vorticity structures for L/D = 3.0 and 1.5 for different values of
m* and ζ. The first- and second-column snapshots correspond to Y ≈ 0 (i.e., mean position)
and Y ≈ Ymax (i.e., maximum displacement), respectively. In the first- and second-row
wake structures (Figure 11a,b), the effect of increasing m* from 2 to 16 on the wake structure
is displayed. For m* = 2 and ζ = 0 with L/D = 3.0 (Figure 11a), the wake behind the wake
cylinder is characterized by 2S mode (i.e., two vortices shed in one oscillation cycle). The
shear layers emanating from the wake-generating cylinders roll up into the gap between
the cylinders. When the vibrating cylinder is close to the mean position (Y ≈ 0), the vortices
from the wake-generating cylinder impinge on the vibrating cylinder. On the other hand,
they both pass over the same side of the vibrating cylinder when the vibrating cylinder
is at its maximum displacement (Y ≈ Ymax ). When the m* is increased to 16 (Figure 11b),
the shear layers from the wake-generating cylinder do not roll up in the gap between the
cylinders but alternately reattach on the vibrating cylinder. Another remarkable difference
in the wake structures between m* = 2 and 16 (Figure 11a,b) is that a greater number of
vortices appear in the same downstream distance (x* = 4–26) for m* = 2 than for m* = 16.
Energies 2021, 14, x FOR PEER REVIEW 12 of 14
Given the 2S mode for both cases, it can be argued that the cylinder oscillation frequency
*
reduces when m is increased from 2 to 16, which is consistent with our intuition.

Figure
Figure 11. Vorticity contours
11. Vorticity for (a)
contours for m* == 2,
(a) m* 2, ζ= L/D= =3.0;
ζ =0,0,L/D (b)m*m*= =
3.0;(b) 16,16, = 0,
ζ =ζ 0, L/D
L/D = 3.0;
= 3.0; m*m*
(c)(c) = 2,= ζ2,=ζ0.5, L/DL/D
= 0.5, = 3.0;
= 3.0; andand
(d)
(d) m* = 2, = 0, L/D = 1.5. Ur
m* = 2, ζ = 0, L/D = 1.5. Ur = 10.
ζ = 10.

AA comparison
comparisonofofsnapshots
snapshotsbetween
betweenthe the first
first andand third
third rows
rows gives
gives out out the effect
the effect of ζ
of ζ (=0 and 0.5) on the wake structure. The arrangement and structure
(=0 and 0.5) on the wake structure. The arrangement and structure of vortices in the wake of vortices in
the wake do not differ much between ζ = 0 and 0.5. The vortices in the
do not differ much between ζ = 0 and 0.5. The vortices in the wake, however, decay more wake, however,
decay more
rapidly for ζ rapidly
= 0 thanforforζ ζ==00.5.
than foreffect
The ζ = 0.5.
of L/DThecaneffect of L/D can be
be understood by understood
comparing theby
comparing
vortex the vortex
shedding shedding
between 3.0 andL/D
L/D =between 1.5 =(Figure
3.0 and11a,d),
1.5 (Figure
both11a,d),
for theboth formthe
same same
* and ζ.
Although the wake structure is the same for both L/D values, the flow structure around
the cylinders is completely different in the two cases. While vortex shedding occurs from
both cylinders for L/D = 3.0 (Figure 11a), only the wake cylinder sheds vortices for L/D =
1.5 (Figure 11d). Vortex shedding does not take place in the gap between the cylinders for
L/D = 1.5. The alternate reattachment of the shear layer from the wake-generating cylinder
Energies 2021, 14, 5148 12 of 13

m* and ζ. Although the wake structure is the same for both L/D values, the flow structure
around the cylinders is completely different in the two cases. While vortex shedding occurs
from both cylinders for L/D = 3.0 (Figure 11a), only the wake cylinder sheds vortices for
L/D = 1.5 (Figure 11d). Vortex shedding does not take place in the gap between the cylinders
for L/D = 1.5. The alternate reattachment of the shear layer from the wake-generating
cylinder sustains the wake cylinder vibration for L/D = 1.5 while alternate impingement
and switch of the gap vortices do the same for L/D = 3.0.

4. Conclusions
Flow-induced vibrations of a cylinder interacting with the wake of another cylinder
is investigated at Re = 150. The focus is given on the effect of m* (=2–32), ζ (=0–0.5), and
m* ζ (=0–8) on A/D at different Ur and L/D values. While investigations at high Re found
galloping vibration for the wake cylinder (e.g., [1]), no galloping occurs at this low Re.
Following VIV peak, A/D remains high (A/D > 0.4) even at high Ur (>15).
The A/D is more sensitive to m* than to ζ. Both m* and ζ reduce A/D. The effect of ζ
on A/D is larger at smaller m* . On the other hand, the m* has a considerable effect on A/D
for all ζ values. The effect is, however, strong for m* = 2–16 and relatively weak for m* > 16.
Generally, the A/D is inversely linked to m* and ζ.
Although m* ζ is believed to be a unique parameter determining the vibration ampli-
tude for a single isolated cylinder, this is not the case for the cylinder interacting with the
wake generated by another. Although A/D decreases with m* ζ, the decrease rate is high
at small m* . As such, the A/D data fail to collapse onto a single line against m* ζ, which
signifies that m* ζ is not a parameter that uniquely characterizes interaction of a cylinder
with the wake of another cylinder.
The shear layers of the wake-generating cylinder do roll up in the gap between the
cylinders at m* = 2 for ζ = 0, Ur = 10, and L/D = 3. With the same initial conditions (ζ = 0,
Ur = 10, and L/D = 3), when m* is increased to 16, the shear layers do not roll up in the gap
but alternately reattach on the vibrating cylinder. Vortices in the wake appear more for
m* = 2 than for m* = 16, decaying more rapidly at small ζ. While the wake-generating-
cylinder shear layers reattaching alternately on the wake cylinder fuels the vibration of
the wake cylinder for L/D = 1.5, the gap-vortex impingement and switch do the same for
L/D = 3.0. The results are useful to riser designers to understand the role of m* and ζ in
flow-induced vibrations.

Funding: This research was funded by the Khalifa University of Science and Technology through
Grants CIRA-2020-057.
Data Availability Statement: The data that support the findings of this study are available within
the article.
Acknowledgments: The author also greatly acknowledges the contribution of Liu Haifeng to simulation.
Conflicts of Interest: The author declares no conflict of interest.

References
1. Bokaian, A.; Geoola, F. Wake-induced galloping of two interfering circular cylinders. J. Fluid Mech. 1984, 146, 383–415. [CrossRef]
2. Carmo, B.; Sherwin, S.; Bearman, P.; Willden, R. Flow-induced vibration of a circular cylinder subjected to wake interference at
low Reynolds number. J. Fluids Struct. 2011, 27, 503–522. [CrossRef]
3. Hover, F.; Triantafyllou, M. Galloping response of a cylinder with upstream wake interference. J. Fluids Struct. 2001, 15, 503–512.
[CrossRef]
4. Assi, G.R.S.; Bearman, P.W.; Meneghini, J.R. On the wake-induced vibration of tandem circular cylinders: The vortex interaction
excitation mechanism. J. Fluid Mech. 2010, 661, 365–401. [CrossRef]
5. Assi, G.R.D.S.; Bearman, P.W.; Carmo, B.; Meneghini, J.R.; Sherwin, S.; Willden, R.H.J. The role of wake stiffness on the
wake-induced vibration of the downstream cylinder of a tandem pair. J. Fluid Mech. 2013, 718, 210–245. [CrossRef]
6. Qin, B.; Alam, M.; Zhou, Y. Two tandem cylinders of different diameters in cross-flow: Flow-induced vibration. J. Fluid Mech.
2017, 829, 621–658. [CrossRef]
Energies 2021, 14, 5148 13 of 13

7. Qin, B.; Alam, M.M.; Zhou, Y. Free vibrations of two tandem elastically mounted cylinders in cross-flow. J. Fluid Mech. 2019, 861,
349–381. [CrossRef]
8. Assi, G.R.S.; Meneghini, J.R.; Aranha, J.; Bearman, P.; Casaprima, E. Experimental investigation of flow-induced vibration
interference between two circular cylinders. J. Fluids Struct. 2006, 22, 819–827. [CrossRef]
9. Korkischko, I.; Meneghini, J.R. Experimental investigation of flow-induced vibration on isolated and tandem circular cylinders
fitted with strakes. J. Fluids Struct. 2010, 26, 611–625. [CrossRef]
10. Qin, B.; Alam, M.; Ji, C.; Liu, Y.; Xu, S. Flow-induced vibrations of two cylinders of different natural frequencies. Ocean Eng. 2018,
155, 189–200. [CrossRef]
11. Zdravkovich, M. Flow induced oscillations of two interfering circular cylinders. J. Sound Vib. 1985, 101, 511–521. [CrossRef]
12. Kim, S.; Alam, M.; Sakamoto, H.; Zhou, Y. Flow-induced vibrations of two circular cylinders in tandem arrangement. Part 1:
Characteristics of vibration. J. Wind. Eng. Ind. Aerodyn. 2009, 97, 304–311. [CrossRef]
13. Huera-Huarte, F.; Bearman, P. Vortex and wake-induced vibrations of a tandem arrangement of two flexible circular cylinders
with near wake interference. J. Fluids Struct. 2011, 27, 193–211. [CrossRef]
14. Papaioannou, G.; Yue, D.; Triantafyllou, M.; Karniadakis, G. On the effect of spacing on the vortex-induced vibrations of two
tandem cylinders. J. Fluids Struct. 2008, 24, 833–854. [CrossRef]
15. Borazjani, I.; Sotiropoulos, F. Vortex-induced vibrations of two cylinders in tandem arrangement in the proximity–wake
interference region. J. Fluid Mech. 2009, 621, 321–364. [CrossRef] [PubMed]
16. Prasanth, T.; Mittal, S. Flow-induced oscillation of two circular cylinders in tandem arrangement at low Re. J. Fluids Struct. 2009,
25, 1029–1048. [CrossRef]
17. Zheng, Q.; Alam, M. Intrinsic features of flow past three square prisms in side-by-side arrangement. J. Fluid Mech. 2017, 826,
996–1033. [CrossRef]
18. Zheng, Q.; Alam, M. Evolution of the wake of three inline square prisms. Phys. Rev. Fluids 2019, 4, 104701. [CrossRef]
19. Bhatt, R.; Alam, M. Vibrations of a square cylinder submerged in a wake. J. Fluid Mech. 2018, 853, 301–332. [CrossRef]
20. Patankar, S.V. Numerical Heat Transfer and Fluid Flow; CRC Press: Boca Raton, FL, USA, 1980; ISBN 9781315275130.
21. Barkley, D.; Henderson, R. Three-dimensional Floquet stability analysis of the wake of a circular cylinder. J. Fluid Mech. 1996, 322,
215–241. [CrossRef]
22. Rastan, M.R.; Alam, M.M. Transition of wake flows past two circular or square cylinders in tandem. Phys. Fluids 2021, 33, 081705.
[CrossRef]
23. Zafar, F.; Alam, M.M. Flow structure and heat transfer from cylinders modified from square to circular. Phys. Fluids 2019, 31,
083604. [CrossRef]
24. Abdelhamid, T.; Alam, M.M.; Islam, M. Heat transfer and flow around cylinder: Effect of corner radius and Reynolds number.
Int. J. Heat Mass Transf. 2021, 171, 121105. [CrossRef]
25. Alam, M.; Abdelhamid, T.; Sohankar, A. Effect of cylinder corner radius and attack angle on heat transfer and flow topology. Int.
J. Mech. Sci. 2020, 175, 105566. [CrossRef]
26. Zafar, F.; Alam, M. A low Reynolds number flow and heat transfer topology of a cylinder in a wake. Phys. Fluids 2018, 30, 083603.
[CrossRef]
27. Sucker, D.; Brauer, H. Fluiddynamik bei der angestromten Zilindern. Warme Stoffubertrag. 1975, 8, 149. [CrossRef]
28. Norberg, C. Flow around a circular cylinder: Aspects of fluctuating lift. J. Fluids Struct. 2001, 15, 459–469. [CrossRef]
29. Silva, A.L.E.; Silveira-Neto, A.; Damasceno, J. Numerical simulation of two-dimensional flows over a circular cylinder using the
immersed boundary method. J. Comput. Phys. 2003, 189, 351–370. [CrossRef]
30. Posdziech, O.; Grundmann, R. A systematic approach to the numerical calculation of fundamental quantities of the two-
dimensional flow over a circular cylinder. J. Fluids Struct. 2007, 23, 479–499. [CrossRef]
31. Qu, L.; Norberg, C.; Davidson, L.; Peng, S.-H.; Wang, F. Quantitative numerical analysis of flow past a circular cylinder at
Reynolds number between 50 and 200. J. Fluids Struct. 2013, 39, 347–370. [CrossRef]
32. Dong, R.G. Effective Mass and Damping of Submerged Structures; UCRL-52342; Lawrence Livermore Laboratory, California
University: Oakland, CA, USA, 1978.
33. Tabatabai, H.; Mehrabi, A. Design of Mechanical Viscous Dampers for Stay Cables. J. Bridg. Eng. 2000, 5, 114–123. [CrossRef]
34. Alam, M.; Kim, S. Free vibration of two identical circular cylinders in staggered arrangement. Fluid Dyn. Res. 2009, 41, 17.
[CrossRef]
35. Alam, M.; Meyer, J. Global aerodynamic instability of twin cylinders in cross flow. J. Fluids Struct. 2013, 41, 135–145. [CrossRef]
36. Khalak, A.; Williamson, C. Dynamics of a hydroelastic cylinder with very low mass and damping. J. Fluids Struct. 1996, 10,
455–472. [CrossRef]
37. Williamson, C.; Govardhan, R. Vortex-induced vibrations. Annu. Rev. Fluid Mech. 2004, 36, 413–455. [CrossRef]

You might also like