You are on page 1of 49

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/370820285

Vortex-induced vibrations of tandem diamond cylinders: A novel lock-in


behavior

Article  in  International Journal of Mechanical Sciences · May 2023


DOI: 10.1016/j.ijmecsci.2023.108463

CITATION READS

1 62

2 authors:

Deepak Kumar Kumar Sourav


Indian Institute of Technology (ISM) Dhanbad Clemson University
19 PUBLICATIONS   143 CITATIONS    18 PUBLICATIONS   150 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Fluid-Structure Interactions View project

Wake-induced-vibrations of tandem circular cylinders at low Reynolds number View project

All content following this page was uploaded by Kumar Sourav on 19 May 2023.

The user has requested enhancement of the downloaded file.


Journal Pre-proof

Vortex-induced vibrations of tandem diamond cylinders: A novel lock-in


behavior

Deepak Kumar, Kumar Sourav

PII: S0020-7403(23)00365-X
DOI: https://doi.org/10.1016/j.ijmecsci.2023.108463
Reference: MS 108463

To appear in: International Journal of Mechanical Sciences

Received date : 13 March 2023


Revised date : 23 April 2023
Accepted date : 10 May 2023

Please cite this article as: D. Kumar and K. Sourav, Vortex-induced vibrations of tandem diamond
cylinders: A novel lock-in behavior. International Journal of Mechanical Sciences (2023), doi:
https://doi.org/10.1016/j.ijmecsci.2023.108463.

This is a PDF file of an article that has undergone enhancements after acceptance, such as the
addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive
version of record. This version will undergo additional copyediting, typesetting and review before it
is published in its final form, but we are providing this version to give early visibility of the article.
Please note that, during the production process, errors may be discovered which could affect the
content, and all legal disclaimers that apply to the journal pertain.

© 2023 Elsevier Ltd. All rights reserved.


Revised Manuscript - MARKED PDF Journal Pre-proof Click here to view linked References

Vortex-Induced Vibrations of Tandem Diamond Cylinders: A Novel


Lock-in Behavior
Deepak Kumara , Kumar Souravb,∗

of
a
School of Mechanical Engineering, Dr. Vishwanath Karad MIT World Peace University, Pune 411038,
India
b
Department of Mechanical Engineering, Clemson University, Clemson 29632, SC

pro
Abstract
This paper presents a comprehensive numerical investigation of the undamped, simultaneous
in-line, and transverse flow-induced vibrations of two identical diamond cylinders (of cross-
stream length D) in a tandem arrangement at a low Reynolds number (= 100) in uniform
flow, using a stabilized space-time finite-element formulation. The study focuses on the
effects of cylinder proximity and shape, considering diamond cylinders with a mass ratio of

re-
10 and a separation distance of 5D and varying the reduced velocity from 1 to 20. The results
reveal distinct lock-in behavior for the upstream and downstream cylinders, with a low-
frequency extended lock-in phenomenon observed for the downstream cylinder. The motion
trajectories of both cylinders exhibit a distinctive “raindrop-shaped” pattern. Notably, the
synchronization zone shows a non-zero mean lift for a narrow range of reduced velocity. The
study also identifies three distinct modes of vortex arrangement in the gap region between
the cylinders. Comparing the findings for tandem diamond cylinders with those for single
lP
diamond and tandem square cylinders provides further evidence of the significant impact of
proximity and shape on the vibration characteristics of the cylinders.
Keywords: Vortex-induced vibrations, Tandem diamond cylinders, Stabilized space-time
finite-element method, Low frequency extended lock-in, Non-zero mean lift
rna

Nomenclature
B Blockage
Cd Drag coefficient
Cdrms r.m.s. of Cd
Cdavg Average of Cd
Cl Lift coefficient
Clrms r.m.s. of Cl
Clavg Average of Cl
Jou

Cp Pressure coefficient
D Oscillator diameter
f Body force vector per unit volume
fN Dimensional natural frequency of the oscillator
FN Reduced natural frequency of the oscillator
Fx Dimensionless in-line oscillation frequency
Fy Dimensionless transverse oscillation frequency


Corresponding author: krsourav678@gmail.com

Preprint submitted to International Journal of Mechanical Sciences


Journal Pre-proof

Fd Drag force
Fl Lift force
FCd Dimensionless frequency of drag
FCl Dimensionless frequency of lift
h Surface traction vector
I Identity tensor

of
k Spring stiffness
m Mass of an oscillator
m∗ Mass ratio
md Mass of the displaced fluid

pro
n Unit normal vector
p Pressure
P () Power spectral density of () signal
Re Reynolds number
S/D Horizontal gap between the cylinders
St Strouhal number
St0 Dimensionless vortex-shedding frequency from a stationary cylinder
StD Dimensionless vortex-shedding frequency from tandem cylinders
t
t∗ = tU
T /D
T
D
re- Dimensional time
Non-dimensional time
Transverse gap between the cylinders
Dimensional time domain
u Velocity vector
u0 Initial velocity field
lP
U∗ Reduced velocity
X In-line displacement of a vibrating cylinder
X∗ = X D
Normalized axial displacement
Xmax Peak Amplitude of in-line response
Xrms r.m.s. of in-line displacement
x Dimensional spacial coordinates vector
rna

Y Transverse displacement of vibrating cylinder


Y
Y∗ = D Normalized transverse displacement
Ymax Peak Amplitude of transverse response
Yrms r.m.s. of transverse displacement
Acronyms
DS I First regime of desynchronization
DS II Second regime of desynchronization
Jou

EIB Extended Initial Branch


FFT Fast Fourier transform
IB Initial branch
LB Lower branch
UB Upper branch
Greek symbols
µ Dynamic viscosity
∇ Del operator
ρ Density of fluid

2
Journal Pre-proof

σ Cauchy Stress tensor

1. Introduction
Flow-induced vibration (FIV) is a significant concern in engineering due to its practical
implications for structural integrity. FIV refers to the phenomenon whereby fluid flow around

of
a solid structure induces vibrations at its natural frequency. When fluid flows past a bluff
structure, it generates a wake with alternating shedding vortices due to the fluctuating
pressure force near the structure’s rear. If the structure’s oscillation frequency is in close
proximity to the system’s natural frequency, it can enter a lock-in state, whereby it undergoes

pro
a highly periodic vibration that can persist for an extended duration. The interaction of shed
vortices with the structure’s oscillations can cause them to amplify, resulting in high levels
of stress and fatigue that can compromise the structure’s integrity and reduce its lifespan.
In recent decades, research in the field of FIV has advanced through both experimental
and computational studies. The underlying principles of FIV have been revealed through
investigations of basic geometries. Numerous studies have investigated the underlying me-
chanics of FIV, with particular attention given to the impact of lock-in states on structural
performance. The literature has extensively covered these topics in numerous articles pub-

Wang et al. (2020).


re-
lished on the subject, including detailed analyses by Bearman (1984), Sarpkaya (2004), and

Innovations in engineering structures often require significant geometric modifications,


which can result in structural instability. Therefore, the study of non-circular cross-sections
is particularly relevant in the current context. FIVs are generally considered undesirable due
to safety concerns. Therefore, its understanding, and control remain crucial for designing
lP
various engineering systems, including bridges, chimney stacks, offshore structures, aircraft,
and submarines. However, with the depletion of non-renewable energy sources, there has
been a growing interest in exploring renewable energy sources such as wind and ocean.
Researchers have successfully developed converters that utilize FIV to harness energy from
these sources (Bernitsas et al., 2008; Ding et al., 2016; Zhang et al., 2018, Tang et al.
2022). The amount of energy extracted from FIV depends on the oscillation amplitude,
rna

with triangular cylinders exhibiting more oscillations and thus producing more energy than
their circular counterparts (Seyed-Aghazadeh et al. 2017). As such, continued research
efforts to improve our understanding of FIV and its control methods are imperative.
The available research on FIV has predominantly focused on circular cylinders due to
their axial symmetry (Prasanth and Mittal, 2008, Aswathy and Sarkar 2019); however,
recent studies have shown that the response to FIV is sensitive to geometric shapes (Sen
and Mittal, 2011; Sourav and Sen, 2020) and uncontrolled high-amplitude, low-frequency
oscillations (referred to as ‘galloping’) may occur for aerodynamically unstable cross-sections
Jou

(Seyed-Aghazadeh et al. 2017; Sourav and Sen 2019, 2020). These studies have identified
the critical non-dimensional parameters that govern FIV, including the Reynolds number
(Re = UνD ), mass ratio (m∗ = mmd ), and reduced velocity (U ∗ = fNUD ). In these relations,
‘U ’ is the free-stream flow velocity, ‘D’ is the cross-stream dimension of the rigid body, ‘ν’
is the kinematic viscosity of the fluid, ‘m’ is the mass of the body, ‘md ’ is the mass of the
fluid displaced by the volume of the body, and, ‘fN ’ is the dimensional natural frequency of
the rigid body.
The use of multi-cylinder systems in engineering applications is increasing, which has
resulted in a demand for a comprehensive analysis of the vibration mechanisms within such

3
Journal Pre-proof

systems. The most basic model for a multi-cylinder system is one with two cylinders. In addi-
S
tion to the vibration parameters mentioned above, the linear and transverse spacing ratios ( D
T
and D ) between cylinder centers significantly affect the flow and Vortex-Induced Vibration
(VIV) characteristics of multi-cylinder arrangements (Zdravkovich 1977, 1985). Zdravkovich
(1977) introduced the concept of critical spacing for vortex shedding from the upstream
cylinder for linear cylinder arrangements. Zdravkovich (1985) later classified three regimes

of
of wake interactions, i.e., proximity, interference, and wake interference or co-shedding, while
S T
considering both D and D . Later, Sumner (2010) reported further refinements in the inter-
ference regime. In the co-shedding regime, vortices shed from the upstream cylinder impinge
with the rear cylinders, leading to a sudden increase in fluid loading on the downstream

pro
cylinder. The initial research on multi-cylinder arrangements predominantly focused on a
tandem, staggered, and side-by-side configurations of two circular cylinders. Several studies
have investigated the effect of spacing ratios on flow over stationary tandem circular cylin-
ders, reporting critical spacing values between 3 to 4 (Mittal et al., 1997; Meneghini et al.,
2001; Sharman et al., 2005, Haider and Sohn 2018). A comprehensive review of the wake
dynamics for interacting cylinders can be found in Zhou and Alam (2016). Due to their ex-
tensive engineering applications, vortex-induced vibrations of tandem circular cylinders have
also attracted significant research attention (Prasanth and Mittal, 2009a, 2009b; Carmo et

(D
re-
al., 2011; Mysa et al., 2016). To ensure vortex shedding in the gap region, the spacing ratio
S
) used in these studies was 5.5 for Prasanth and Mittal (2009a, 2009b), 1.5-8 for Carmo
et al. (2011), and 5 for Mysa et al. (2016). Considering coupled oscillations of upstream
and downstream cylinders at a Reynolds number of 150, Zhao et al. (2013) observed vortex
shedding from the upstream cylinder at a spacing ratio as low as two within the lock-in
regime.
lP
Although square cylinders are commonly found in engineering applications, research on
flow over tandem square cylinders gained attention later than circular cylinders. Analogous
to the vortex-shedding regimes reported by Zdravkovich (1985) for tandem circular cylinders,
Tatsutani et al. (1993) identified three vortex-shedding/wake structure regimes for tandem
square cylinders. They observed vortex-shedding and intense mixing in the gap region beyond
a critical spacing ratio for flow over stationary tandem square cylinders. Subsequently,
rna

Sohankar and Etminan (2009), Etminan et al. (2011), and Sohankar (2012) investigated
steady and unsteady flows over identical tandem square cylinders, using a fixed S/D = 6 to
ensure vortex-shedding in the gap region. Chatterjee and Mondal (2012) examined forced
convection heat transfer from tandem square cylinders over a Reynolds number range of 50
to 150, varying the spacing ratio between S/D = 1 and 10. They found that the critical
spacing for vortex-shedding in the gap region depends on the Reynolds number. For the
lower range of Reynolds numbers, the spacing ratio was within 5-7, narrowing to 3-4 as the
Reynolds number approached 150. Thus, the choice of S/D = 6 by Sohankar and Etminan
Jou

(2009), Sohankar (2012), and Etminan et al. (2011) for their numerical simulations appears
to be reasonable. Similarly, Zhao et al. (2016) reported a critical spacing of S/D = 4 at
Re = 100. Consequently, the present study employs a spacing ratio of S/D = 5 to ensure
vortex interaction in the co-shedding regime.
Unlike circular cylinders, square cylinders with zero-degree incidence are prone to gallop-
ing. Zhao et al. (2015) investigated wake-induced vibrations of downstream square cylinder
free to oscillate along and across the flow in tandem arrangements at Re = 200, S/D =
4, and m∗ = 5. They identified initial, lower desynchronization, and galloping branches in
the U ∗ range [1, 40], with galloping observed beyond U ∗ = 25. Sen and Mittal (2011) also

4
Journal Pre-proof

reported vortex-induced excitation and galloping for a freely vibrating square cylinder at
Re ≥ 160 and m∗ = 10.
Using a stable second-order partitioned iterative scheme, Jaiman et al. (2016) observed
galloping along with vortex-induced excitation for wake-induced vibrations of square cylin-
ders downstream of a stationary identical square cylinder at Re = 200, S/D = 4, and m∗ = 5.
Bhatt and Alam (2018) reported wake-induced vibrations of square cylinders with m∗ = 3

of
and executing y−only motion. They varied the spacing ratio from 2 to 6 and observed
vortex-shedding in the gap region for 4.5 < S/D < 5. Only vortex-induced excitation was
observed at Re = 100 and 200. Han et al. (2018) numerically investigated y-only wake-
induced vibrations of square cylinders at m∗ = 2 and S/D = 5, with results computed over a

pro
U ∗ range of [3, 18] at five Re values (40, 80, 120, 160, and 200). They found the downstream
cylinder’s response to comprise only vortex-induced excitation. The differences in transverse
response and branching behavior between Jaiman et al. (2016), Bhatt and Alam (2018),
and Han et al. (2018) may be attributed to the low m∗ (2 and 3) used by the latter two
studies. This highlights the importance of m∗ in resolving various response branches in FIV
of multi-cylinder arrangements.
Tamimi et al. (2020) experimentally examined transverse-only wake-induced vibrations
of square and diamond cylinders at high Reynolds numbers over an S/D ratio range of
re-
3–11. They reported that the diamond cylinder exhibited a higher transverse response and
greater potential for hydro energy harvesting. Nepali et al. (2020) reported free vibrations
of both upstream and downstream square cylinders with m∗ = 2 over a Re range of 40–200
and S/D = 5. Subsequently, Kumar and Sen (2021) investigated the effect of hysteresis
on FIV of two square cylinders with m∗ = 10 and simultaneous X − Y free vibrations at
Re = 100. The effect of the order of magnitude difference in m∗ (2 and 10) was evident
lP
in response amplitudes. For example, Kumar and Sen (2021) observed a peak transverse
response amplitude of the upstream cylinder of approximately 0.2D at Re = 100, while
Nepali et al. (2020) reported a value of approximately 0.8D at Re = 80.
Qiu et al. (2021a) investigated the spacing effect (S/D) on undamped two degrees-of-
freedom VIV of tandem square cylinders at Re = 100 and U ∗ = 3 − 12. The spacing ratios
were categorized into small (2, 2.5), moderate (3, 4), and large (5, 6). The upstream cylinder
rna

exhibited no lock-in for small spacing ratios, while the downstream cylinder showed a narrow
lock-in region. Both cylinders displayed a wide soft lock-in zone at high reduced velocities
(U ∗ ≥ 9.5). The upstream cylinder showed no lock-in or soft lock-in for moderate and
large spacing ratios, while the downstream cylinder exhibited varying lock-in regions. In a
follow-up study, Qiu et al. (2021b) examined the mass ratio effect on two degrees-of-freedom
VIV of tandem square cylinders separated by 4D. They reported soft lock-in for both m∗
cylinders, with lock-in observed only for the downstream cylinder.
Recently, researchers have focused on the VIV of tandem diamond cylinders. Xu et
Jou

al. (2021) studied the Reynolds number effect on the VIV of tandem diamond cylinders
(each with m∗ = 3) separated by 5D. The lock-in region remained unchanged for the
upstream cylinder with increasing Re but expanded for the downstream cylinder. At lower
Re values, figure-eight cylinder trajectories were observed. In contrast, at other Re values,
the upstream cylinder exhibited a closed ring-shaped trajectory within the lock-in region, and
the downstream cylinder displayed a more complex trajectory. Gu et al. (2022) investigated
the effect of varying spacing ratios on wake-induced vibrations of tandem diamond cylinders
with a mass ratio of 3. They reported a non-zero mean lift for the downstream diamond
cylinder and a consistent wake mode within the lock-in zone for all spacing ratios of S/D ≥ 3.

5
Journal Pre-proof

Gu et al. (2023) further examined the impact of spacing ratios (2−6) on VIV of tandem
diamond cylinders (each with a mass ratio of 3) at Re = 100 and reduced velocities of 3−18.
They identified two high-amplitude regimes at U ∗ ≈ 5 − 9 and 12−14 for S/D = 2 − 3, with
the former considered lock-in. For other spacing ratios (S/D = 4−6), the lock-in regime was
narrower (U ∗ = 5 − 7), and the cylinder trajectory was single-looped within the lock-in zone
while more complex and diverse outside the lock-in zone. Table 1 summarizes the recent

of
studies on FIV of square and diamond cylinders.
Table 1: Summary of recent literature on FIV of twin-cylinders with various shapes and degrees of freedom.
S−S stands for square-square cylinders in tandem arrangement. D−D refers to diamond-diamond cylinders
in tandem. Acronyms: UC: Upstream cylinder, DC: Downstream cylinder, DS: Desynchronization, IB:

pro
Initial branch, LB: Lower branch, UB: Upper branch, GB: Galloping branch, EIB: Extended initial branch,
BNR: Branch not resolved.

S
Geometry Study Re m∗ D
Degree of Freedom Response Branches
UC DC UC DC
S−S Zhao et al. (2015) 200 5 4 Fixed x−y − IB, LB, DS, GB
S−S Jaiman et al. (2016) 200 5 4 Fixed x−y − DS, lock-in, GB
S−S Bhatt and Alam (2018) 100, 200 3 2, 6 Fixed y−only − IB, LB, DS
S−S Nepali et al. (2020) 40−200 2 5 x−y x−y Lock-in, BNR Lock-in, BNR
S−S Kumar and Sen (2021) 100 10 5 x−y x−y DSI, LB, DSII DSI, LB, DSII
S−S Qui et al. (2021a) 150 10 2−6 x−y x−y lock-in, soft lock-in, BNR lock-in, soft lock-in, BNR
S−S
S−S
S−S
D−D
D−D
D−D
D−D
Qui et al. (2021b)
Zhang et al. (2022)
Zhou et al. (2023)
Xu et al. (2021)
Gu et al. (2022)
Gu et al. (2023)
Present
150
100, 200
100
40-200
100
100
100
re-
3, 10, 20
3
3
2
3
3
10
4

6
5
2−6
2−6
5
x−y

x−y
Fixed
x−y
x−y
x−y
2, 6 y−only y−only
y−only y−only
x−y
y-only
x−y
x−y
lock-in, soft lock-in, BNR
DB, IB, LB, GB
Lock-in, BNR
Lock-in, BNR

Lock-in, BNR
DSI, IB, UB, LB, DSII
lock-in, soft lock-in, BNR
DB, IB, LB, GB
Lock-in, BNR
Lock-in, BNR
Lock-in, BNR
Lock-in, BNR
DSI, IB, EIB, LB, DSII
lP
Research gap, motivation, and objectives
For an isolated cylinder, it can be established from the literature that the flow and vibra-
tion characteristics of a rigid body placed in a moving fluid are susceptible to modifications in
cross-section. Furthermore, adding another interfering cylinder(s) near an isolated cylinder
modifies its flow field (Inoue et al., 2006; Zhao et al., 2016), resulting in complex dynamic
responses, fluid forces, and flow fields. It is thus important to study the multi-cylinder sys-
rna

tems.
A diamond cylinder, in particular, has been established as an interesting geometric cross-
section to investigate (Zaki et al., 1994; Sohankar, 2012). A diamond cylinder is essentially
a square cylinder rotated by 45◦ , but interestingly, it displays unique flow features. For a
stationary diamond cylinder, in the steady flow regime, Kumar et al. (2018b) observed a
previously unreported flow structure: laminar separation bubbles (LSB) on both top and
bottom rear edges. LSB was found to be connected with the surface pressure gradient in a
Jou

later study by Kumar et al. (2019). Lately, Yadav et al. (2021) presented diverse separation
topologies for a diamond cylinder that has not been resolved for a circular or square cylinder.
When allowed to oscillate, a diamond cylinder demonstrates response features closer to a
circular cylinder than a square (Nemes et al., 2012; Sourav et al., 2022). Notably, the
galloping phenomenon observed for a square cylinder (Sourav and Sen, 2019; Li et al., 2019)
is suppressed by rotating the geometry by 45◦ . The oscillation amplitude in the lock-in
regime also increases significantly (Sourav et al., 2020). It is thus expected that tandem
diamond cylinders will display several interesting flow and vibration characteristics. It will
also be interesting to see if an upstream oscillating cylinder can trigger galloping. For a
transversely oscillating diamond cylinder, Sourav et al. (2020) reported a non-zero mean

6
Journal Pre-proof

lift at one value of reduced velocity (U ∗ = 3.5), accounting for an asymmetric wake mode.
Later, Sourav et al. (2022) observed a band of positive and negative asymmetric lifts (and
thus asymmetric wake modes) for two degrees of motion of an isolated diamond cylinder.
They found that the band of asymmetric lifts is a characteristic of a newly resolved response
branch: the ‘upper branch’ for a diamond cylinder. It is thus intriguing to explore the effect
of an additional oscillating diamond cylinder on the fluid forces, dynamic response, flow

of
periodicity, vortex-shedding modes, etc.
Research on fluid-induced vibration (FIV) of multi-cylinder non-circular systems has
predominantly concentrated on stationary and vibrating square-shaped cylinders. Studies
examining tandem diamond cylinders with two degrees of motion are comparatively limited.

pro
Notably, only a few publications, such as Xu et al. (2021) and Gu et al. (2022, 2023), have
investigated tandem diamond cylinders. Gu et al. (2022) examined the effect of spacing
ratio on wake-induced vibrations for a diamond cylinder with a mass ratio (m∗ ) of 3 at a
Reynolds number (Re) of 100. Xu et al. (2021) and Gu et al. (2023) investigated the
free oscillations of both diamond cylinders. Xu et al. (2021) analyzed the impact of Re on
FIV for cylinders with a mass ratio of 2, while Gu et al. (2023) focused on the influence
of varying spacing ratios on FIV for cylinders with an identical mass ratio of 3 at Re=100.
Both studies employed lower mass ratios (< 6). Previous research by Jauvtis and Williamson
re-
(2004) revealed significant differences in FIV characteristics between mass ratios of less than
6 and greater than 6, suggesting that diamond cylinders with a mass ratio of 10 would exhibit
different behavior compared to those with mass ratios of 2 or 3. Furthermore, the analysis of
FIV for tandem square cylinders demonstrated that the presence of the galloping branch and
the overall response branching behavior are sensitive to variations in mass ratio at a given
spacing ratio and Re (Jaiman et al., 2016), emphasizing the importance and motivation
lP
of the current study to explore FIV of tandem diamond cylinders. In Xu et al.’s (2021)
study, the vortex-induced vibration of tandem diamond cylinders was examined at various
Reynolds numbers, ranging from 40 to 200, but within reduced velocities 3 and 12. However,
this reduced velocity range was insufficient to capture all FIV features of the downstream
cylinder, such as the post-lock-in desynchronization zone. Additionally, Xu et al. (2021)
did not identify transition locations, omitting intriguing coupled dynamics that may occur
rna

at those points. In the present numerical work, we aim to investigate the bidirectional
FIV characteristics of freely vibrating tandem diamond cylinders (with a mass ratio of 10
each) along and across the flow. We have addressed the limitations of Xu et al. (2021) by
resolving all transitions and focusing on coupled dynamics in our analysis. To achieve this,
computations are performed using the stabilized space-time finite-element formulation.
The remaining part of the manuscript is organized as follows: methodology is described
in Section 2; the problem definition and mesh are discussed in Section 3; key findings are de-
tailed in Section 4, followed by concluding remarks in Section 5. The non-dimensionalization
Jou

of the rigid body equation is discussed in Appendix A. Detailed convergence and validation
studies for the isolated and tandem cylinder arrangements are reported in Appendix B.

2. Methodology
This section presents the numerical model and technique employed for solving the gov-
erning equations related to fluid flow and rigid body dynamics. It begins by introducing the
governing equations that describe the fluid flow behavior and the rigid body equation, elu-
cidating the key assumptions and simplifications made to establish the mathematical frame-
work. Following this, the numerical details and computational approaches are outlined for

7
Journal Pre-proof

solving these equations, including the discretization methods and space-time finite-element
formulation. This comprehensive discussion of the methodology provides a solid foundation
for the analysis and results presented in the subsequent sections.

2.1. The governing equations


The Navier-Stokes equations in the primitive variables form (Equations 1 and 2) are

of
used to model the incompressible fluid flow.

 
∂u
ρ + u · ∇u − f −∇·σ = 0 (1)

pro
∂t
∇ · u = 0. (2)

Here, ρ,u,f , and σ are fluid density, instantaneous velocity vector, body force per unit
volume, and the Cauchy stress tensor, respectively. In the present study, the contribution
from the body force term is zero. The Cauchy stress tensor for a Newtonian fluid is written
in terms of its pressure and viscous components:
1
σ = −pI + T,
re-
T = 2µεε(u), ε(u) = ((∇
2
∇u)T ),
∇u) + (∇ (3)

where p, I, µ, and ε are the pressure, identity tensor, the dynamic viscosity of the fluid, and
strain rate tensor, respectively.
The rigid-body structure is modeled as a diamond cylinder of diagonal D mounted on
an elastic support in the cross-stream direction and experiences time-varying flow-induced
forces. The streamwise and transverse motion of the diamond cylinder mounted on elastic
lP
support is governed by the following equations in Cartesian coordinates:

d2 X dX
m 2
+c + kx X = Fd , (4)
dt dt
d2 Y dY
rna

m 2 +c + ky Y = Fl . (5)
dt dt
Here, m is the mass of the oscillator, and k is the spring stiffness. The spring is assumed to
be linear, i.e., k is constant. Fd and Fl are the time-varying drag and lift forces experienced
by the cylinder. X and Y are the displacements of the cylinder center along and across
the flow, while corresponding velocity and acceleration are given by the first and second
derivatives of displacement with time (t). The non-dimensional form of Equations 4 and 5
are:
Jou

 2
d2 X ∗ 1 dX ∗ 2π Cd
+ 4πζ + X∗ = , (6)
dt∗2 U ∗ dt∗ U∗ m∗
 2
d2 Y ∗ 1 dY ∗ 2π Cl
+ 4πζ + Y∗ = .. (7)
dt∗2 U ∗ dt∗ U∗ m∗

In Equations 6 and 7, FN is the non-dimensional natural frequency of the oscillator,


ζ is the structural damping ratio and Cd , Cl are instantaneous drag and lift coefficients.
FN is defined as fNUD where fN is the natural frequency of the spring-mass system in a

8
Journal Pre-proof

vacuum, D is the cylinder diagonal, and U is the free-stream velocity. The reciprocal of
FN is U ∗ , the reduced velocity. It is defined as U ∗ = F1N = fNUD . Another parameter of
interest is the frequency ratio, f ∗ . It is defined as the ratio of cylinder oscillation frequency
2 ∗ ∗
to the cylinder’s natural frequency in a vacuum. ddtX∗2 , dX dt∗
and X ∗ denote the normalized
acceleration, velocity, and displacement of the body, respectively, in the direction of the flow.
Similar notations are also used to define corresponding normalized acceleration, velocity, and

of
displacement of the body in the transverse direction. The characteristic scales to normalize
length, velocity, and time are D, U , and D U
, respectively. To encourage high oscillation
amplitude, ζ is set to zero.
The force coefficients Cd and Cl are calculated by integrating elementwise contributions

pro
of pressure and viscous stresses for elements located on the cylinder surface. Cd and Cl are
defined as:

Z
1
Cd = 1 (σ · n) · nx dΓ, (8)
2
ρU 2 D Γ
Z
1
Cd = 1 2 (σ · n) · ny dΓ. (9)
ρU D Γ
2

re-
Here, the Cartesian components of the unit normal vector n at the interface are denoted
by nx and ny .

2.2. Numerical details


The discretization of the governing equations for incompressible fluid flow (Equations 1
and 2) is achieved using a stabilized space-time finite-element formulation in conjunction
lP
with Galerkin Least-Squares (GLS) stabilization, as originally proposed by Tezduyar et al.
(1992a, 1992b). It is well known that the Galerkin formulation is unstable with respect
to the advection operator as the cell Reynolds number (based on the local flow velocity
and mesh size) becomes larger. Also, not all combinations of the velocity and pressure
interpolations are admissible in the Galerkin formulation. Elements that do not satisfy the
Babuska–Brezzi condition lead to oscillatory solutions and, sometimes, no solution at all. To
rna

give stability to the basic formulation, a series of element-level integrals are added. To this
end, SUPG and PSPG stabilization terms added to the variational formulations (Tezduyar
et al. 1992c; Mittal and Tezduyar 1992). The SUPG terms make the formulation stable
in the presence of advection operator. The PSPG terms allow to use any combination of
velocity and pressure interpolation, including equal-order interpolation. The details of the
stabilization terms are available in (Tezduyar et al., 1992a,b,c). The nonlinear system of
equations that arise from the finite element discretization of the flow equations are solved
Jou

using the matrix-free Generalized Minimal RESidual (GMRES) iterative solver, as proposed
by Saad and Schultz(1986), in conjunction with diagonal preconditioners.
In the present FIV study, the fluid flow and structure motion are strongly coupled via
time-dependent no-slip boundary conditions and time-varying fluid forces. With the finite-
element framework used in the study, a partitioned or staggered approach is employed to
solve the fluid flow equations. In the partitioned approach, the fluid and solid equations
are solved separately (but not independently) in each time step. Implicit time stepping
is used to integrate the equations of fluid motion. A non-dimensional time step size of
0.03 is used for all computations. The rigid body ordinary differential equations (ODEs)
are integrated numerically. Furthermore, the rigid body equations are also casted into the

9
Journal Pre-proof

finite-element formulation (Tezduyar et al., 1992c). Further details on the space-time finite-
element formulation can be found in the works of Tezduyar et al. (1992a, 1992b, and 1992c).
In this approach, a stationary reference frame is utilized, which requires remeshing in every
time step to account for the domain deformation caused by the motion of rigid bodies.
The displacements of the rigid body are obtained by numerically integrating the normalized
ordinary differential equations (Mittal and Tezduyar, 1994). A detailed derivation of the

of
non-dimensional forms of rigid body equations of motion is presented in Appendix A.

3. Problem description

pro
In this section, the problem setup, computational domain, and boundary conditions
applied are discussed. Additionally, the rationale behind the selection of parameters used in
the study is provided. The section also delves into the specifics of the finite-element mesh,
highlighting its key features and the mesh moving scheme employed to accommodate the
oscillations of the tandem diamond cylinders. Through this comprehensive exposition of
the problem description, a thorough understanding of the computational framework and its
associated parameters is established.

re-
3.1. Computational domain and boundary conditions
Figure 1 depicts a schematic (not to scale) of uncoupled tandem diamond cylinders,
each with a cross-stream dimension, D, and m∗ = 10 in a uniform flow field. The springs
attached along and across the flow are linear, allowing upstream (UC) and downstream (DC)
cylinders to vibrate simultaneously along and across the free-stream flow. The Reynolds
number Re for all simulations is fixed at 100, and U ∗ is varied between 1 and 20. The
lP
cylinders are placed in a rectangular computational domain whose upstream and downstream
boundaries are located at 25D and 105D, respectively, from the center of the reference frame,
which overlaps with the center of the upstream cylinder when stationary. The center-to-
center distance between the tandem cylinders is kept constant at 5D. The lateral boundaries
are each located at a distance of 10D from the center. Thus, the blockage ratio (B) offered
by the cylinder to the flow is given by the ratio of the cross-stream projected dimension of
rna

D
the cylinder to the width of the computational domain, i.e., B = 20D = 0.05.
The velocity on the cylinder’s surface is subjected to the no-slip condition (u = Ẋ,
v = Ẏ ). The velocity at the upstream boundary is given free-stream values (u = U , v =
0). At the downstream boundary, the stress vector (σxx = 0, σyx = 0) is set to zero. The
normal velocity and the stress vector’s tangential component are zero (v = 0, σxy = 0) on
the lateral boundaries. At each non-linear iteration of the time-marching solution process,
the cylinder’s location, velocity, and boundary conditions are updated.
Jou

3.2. Choice of parameters


Low Reynolds number: For better development of the fundamentals, studies are
performed at low Re partly due to the low computational cost associated with low Reynolds
number simulations. Primarily, it has been established by the researchers that the flow and
vibration characteristics can be reliably predicted at low Reynolds numbers (Zhao et al.,
2013, Jaiman et al., 2016). Therefore, a low Reynolds number (= 100) is selected for the
present study.
Reduced velocity: The reduced velocities are selected to efficiently capture the lock-in
for both cylinders. The typical range of U ∗ considered by various researchers in their study

10
Journal Pre-proof

free−slip lateral walls

10 D
no−slip
uniform flow inlet

stress−free outlet
y

of
k k Symmetry
O x D

k k

pro
25 D 5D 100 D

free−slip lateral walls

Figure 1: (Color online) Schematic of the problem setup for simultaneous streamwise and cross-stream
oscillations of undamped tandem diamond cylinders with cross-stream length ‘D’ and blocking 5% of the
flow. The linear springs are located in both streamwise and cross-stream directions, with the origin of the
Cartesian coordinate axes coinciding with the geometric center of the upstream diamond cylinder.

U ∗ between 1 and 20.


re-
is 3 − 12 (Nepali et al., 2020; Xu et al., 2021). In the present simulations, we thus vary the

Mass ratio: Mass ratio (m∗ ) is a key parameter that governs the flow and vibration
characteristics. Xu et al. (2021) recently performed 2-DOF vibrations of tandem diamond
cylinders of m∗ = 2. This value of mass ratio corresponds to a lighter body. The response
lP
and vibration characteristics are expected to alter for bodies with a moderate mass ratio
significantly. Thus, we are selecting m∗ = 10 as a choice of moderate mass ratio. Qui et al.
(2021b), and Kumar and Sen (2021) also considered m∗ = 10 in their studies with tandem
cylinders.
Gap between the cylinders:In the present work, we intend to study the FIV in a
co-shedding regime. It is found from the literature that co-shedding exists for S/D ≥ 3.5.
rna

Thus, we choose S/D = 5 for the present work.


Damping ratio: The damping ratio is set to zero to encourage maximum oscillations
of the cylinder (Zhao et al., 2013; Kumar et al., 2018a; Sourav et al., 2022). The cylinder
vibrates faster at zero damping, which can help conserve computational resources.

3.3. Finite element mesh and mesh moving scheme

The multi-block, structured, non-uniform finite-element mesh (Figure 2) comprising bi-


Jou

linear quadrilateral elements is used to discretize the computational domain (Figure 1). It
consists of 68947 nodes and 68160 elements, respectively. The cylinders are accommodated
in two square (2D × 2D) central blocks generated via families of radial and tangential grid
lines. Figure 2(b) illustrates the mesh between and close to the cylinders. A total of 240
nodes define the surface of each cylinder. A minimum radial thickness of 0.005D on the
cylinder surface generates the non-uniform segments of a radial grid line. The number of
nodes and elements within each square block are 12240 and 12000, respectively. The remain-
ing eight mesh blocks are rectangular. These mesh blocks are non-uniform and of Cartesian
type. The mesh moving scheme has been designed such that the mesh near the cylinder

11
Journal Pre-proof

of
pro
Figure 2: (Color online) (a) Computational domain of size 130D × 20D with multi-block, structured, and
non-uniform finite-element mesh around tandem diamond cylinders. (b) Zoomed-in view of the mesh near
re-
the cylinders and in the gap region between them. x/D and y/D represents the normalized coordinate axes.

moves along with it like a rigid body while the outer boundary remains fixed. It has been
used in our earlier works with isolated (Kumar et al., 2018a, 2018c; Yadav et al., 2023;
Sourav et al., 2022) and tandem (Kumar et al. 2019a; Kumar and Sen, 2021; Mishra et al.,
2022) cylinders. The details of the mesh moving scheme can be found in Sourav and Sen
lP
(2017). This mesh movement method is expected to give almost no projection errors in the
solution near the cylinder (Tezduyar et al., 1992a, 1992b).
For completeness, the detailed verification of the numerical method and mesh and time
convergence studies are presented in Appendix B. We next consider the tandem diamond
cylinders free to oscillate in and across the direction of flow for investigation. The effect on
the response, frequency behavior, and dynamic fluid loading is investigated systematically.
rna

The results are also compared with tandem square cylinders under identical situations.

4. Results and Discussions


A stabilized space-time finite-element computation in two dimensions has been em-
ployed to generate results for undamped 2-DOF VIV of rigid diamond oscillators, each of
m∗ = 10 at Re = 100 and U ∗ = 1-12. A blockage of 5% is used for all computations. This
Jou

section discusses the frequencies, oscillation amplitudes, cylinder trajectories, dynamic fluid
loading, and vortex-shedding.

4.1. Frequency response of oscillators


An oscillator’s frequency characteristics are essential to investigate in FIV research since
they are directly related to various physical phenomena. It is well understood that when the
oscillation frequency (Fy ) closely approaches or coincides with the natural frequency (FN ) of
the cylinder, ‘lock-in’ occurs, and the oscillation amplitude increases (Bearman, 1984). Fig-
ure 3 depicts both cylinders’ oscillation frequency (Fy ) characteristics. The natural frequency

12
Journal Pre-proof

(FN = U1∗ ) is also shown. We also show the vortex-shedding frequencies of stationary iso-
lated and tandem diamond cylinders at Re = 100, abbreviated as St0 and StD , respectively,
to help us better comprehend the connection between the oscillation and vortex-shedding
frequencies. It is worth noting that for stationary tandem diamond cylinders, the vortices
are shed with the same frequency by both the upstream and downstream cylinders, i.e.,
StDU C = StDDC = StD = 0.1671 at Re = 100. A similar observation has also been made for

of
stationary tandem square (Sohankar, 2012; Kumar and Sen, 2021) and circular (Sharman et
al., 2005; Borazjani and Sotiropoulus, 2009) cylinders.

4.1.1. Interesting features of oscillation frequency

pro
Prasanth and Mittal (2009a, 2009b) and Borazjani and Sotiropoulos (2009) reported the
frequency characteristics of both cylinders for vibrating tandem circular cylinders. Except
for a couple of U ∗ values near the onset of lock-in, Prasanth and Mittal (2009a, 2009b) found
the oscillation frequency for upstream and downstream cylinders to overlap. Kumar and Sen
(2021) also reported a similar overlapping trend of Fy for the two degrees of freedom FIV of
two identical square cylinders in tandem arrangements at Re = 100 and m∗ = 10. In their
frequency ratio (F ∗ = FFNy = Fy × U ∗ ) representation, Nepali et al. (2020) discovered that
the frequencies of tandem square cylinders (m∗ = 2) are the same for Re between 40 and
re-
160. However, a shift from the commonly observed overlapping trend of Fy at higher Re
values was reported, with a divergence in Fy values for U ∗ > 10, which may be attributable
to enhanced flow field instability induced by high-speed flows. The present computations
also display a mixed overlapping and diverging trend of Fy of upstream and downstream
cylinders.
lP
0.28
UC Diamond
DC Diamond
FN
StD
0.22 St0
0.1812
Fy

0.16
rna

0.1671

0.10

0.04
0 4 8 12 16 20
*
U
Jou

Figure 3: (Color online) Non-dimensional oscillation frequency (Fy ) plotted against reduced velocity (U ∗ ),
with FN denoting the non-dimensional natural frequency of the oscillator (FN = U1∗ ). The figure also
displays the vortex-shedding frequencies (St0 (= 0.1812) and StD (= 0.1671)) of an isolated and tandem
stationary diamond cylinders at Reynolds number (Re) = 100. Here, UC and DC refer to the upstream and
downstream cylinders, respectively.

As shown in Figure 3, Fy for upstream and downstream cylinders perfectly overlap at


the outset of the frequency characteristics curve. A sudden jump in Fy values at U ∗ = 3.5
demarcates the homogeneity in the onset of lock-in for both cylinders. Note that the jump
in Fy at U ∗ = 3.5 brings it close to its natural frequency. The overlapping of Fy extends over

13
Journal Pre-proof

the upstream cylinder’s entire lower branch, LB (definitions of different response branches
are made in Section 4.1.2). Fy for both cylinders follow the FN curve demonstrating the
persistence of lock-in. Surprisingly, Fy of the downstream cylinder continues to follow the
FN curve beyond U ∗ = 6.8, thus ensuring a divergence in Fy values for upstream and
downstream cylinders. According to the authors, this prolonged lock-in of the downstream
cylinder is a unique behavior demonstrated by diamond cylinders in tandem configuration.

of
It should be noted that such phenomena have not been documented for square and circular
cylinders. The extended lock-in of the downstream cylinder continues up to ≈ U ∗ = 14
followed by a jump in Fy values at U ∗ = 15, demarcating the closure of lock-in.
Another exciting feature of Fy for the downstream cylinder is their values in the prolonged

pro
lock-in zone. It attains values as low as ≈ 0.08, much lower than StD . Low values of Fy
and high oscillation amplitude are characteristic of another type of flow instability known
as ‘galloping’ (Sen and Mittal, 2015; Sourav and Sen, 2020). However, galloping is only
observed in square cylinders at 0◦ incidence among symmetric geometries. As a result,
even if the Fy values for the downstream cylinder in the prolonged lock-in zone appear to
satisfy the low-frequency criteria of galloping instability, labeling the extended lock-in zone
as galloping instability is an exaggeration. Once set, galloping amplitudes don’t settle. Also,
the oscillation frequency in galloping keeps decreasing (Sourav and Sen, 2020). As stated
re-
above, in Figure 3, a jump in the Fy value of the downstream cylinder from its natural
frequency to the value of Fy for the upstream cylinder confirms the presence of second
desynchronization zone for the downstream cylinder. This further ensures the absence of
galloping.
It is also discernible from Figure 3 that the vortex-shedding frequency of stationary
tandem diamond cylinders (StD = 0.1671) is lower than that of a stationary isolated dia-
lP
mond cylinder (St0 = 0.1812). Except for the self-excitation zone of frequency lock-in, the
oscillation, and vortex-shedding frequencies are expected to match the corresponding sta-
tionary vortex-shedding frequency of the oscillators closely. The low amplitude oscillation
of the cylinders in the desynchronization zones is responsible for the close frequency match.
Because the flow field is relatively unaffected by oscillator motion at low amplitudes, it re-
sembles the flow over stationary cylinders. The near-perfect overlap of StD and Fy in the
rna

desynchronization regime for both cylinders support the preceding discussion.

0.28
UC Diamond 3 UC Diamond
DC Diamond DC Diamond
Single Diamond Single Diamond
0.22 UC Square UC Square
DC Square 2 DC Square
FN
*
Fy.U
Fy

0.16

1
0.10
Jou

0.04 0
0 4 8 12 16 20 0 4 8 12 16 20
*
U U*
(a) (b)

Figure 4: (Color online) Figure (a) compares the non-dimensional oscillation frequency (Fy ) of tandem
diamond cylinders with isolated diamond, and tandem square cylinders. In Figure (b), the frequency ratio
(= Fy /FN ) is presented. The equivalence of the oscillation and natural frequencies is demonstrated by a
continuous line passing through Fy U ∗ = 1 and parallel to the U ∗ axis in Figure (b).

The oscillation frequency of tandem diamond cylinders is now compared with the cor-

14
Journal Pre-proof

responding oscillation frequency for isolated diamond and tandem square cylinders. The
comparisons are made in terms of oscillation frequency (Fy ) and frequency ratio (F ∗ ) in
Figures 4(a) and 4(b), respectively. The frequency ratio is defined as the ratio of the oscil-
lation frequency (Fy ) to the natural frequency (FN ). The reciprocal of FN is U ∗ , as defined
in Section 2. As a result, the frequency ratio is represented as Fy .U ∗ . In Figure 4(b), the
horizontal line at Fy .U ∗ = 1 represents the oscillator’s oscillation and natural frequencies

of
equivalence. The figures clearly show a larger zone of synchronization for the downstream
diamond cylinder. Interestingly, the extent of the synchronized zone for both the single
diamond and the upstream cylinder of the tandem diamond configuration is the same. Tan-
dem square cylinders also have the same oscillation frequency (Figures 4a and 4b). The Fy

pro
values for tandem square cylinders are significantly lower than those for diamond cylinders.
As a direct consequence, diamond cylinders always exhibit a higher oscillation amplitude
than square cylinders. This also holds for isolated square and diamond cylinders (Sourav et
al., 2020). A higher Fy for the diamond cylinder boosts the chances of early lock-in. While
square cylinders experience lock-in at U ∗ = 6.3, diamond cylinders experience it at U ∗ = 3.4.
Strikingly, the presence of an identical square cylinder upstream did not affect the frequency
characteristics of the downstream square cylinder; however, this is not true for tandem dia-
mond cylinders. The frequency characteristics of the downstream cylinder are found to be
re-
significantly influenced by the upstream cylinder. The evidence is an extended initial and
lower branch for the downstream diamond cylinder, which results in an elongated lock-in
zone. Next, the classification of various response branches for upstream and downstream
diamond cylinders in the tandem arrangement is discussed.

4.1.2. Classification of response branches


lP
Khalak and Williamson (1996) first conceptualized the branching of the cylinder response.
They examined the response and frequency characteristics to determine the initial excitation
branch (IB), upper branch (UB), and lower branch (LB). The upper branch is unresolved
for cylinders oscillating in a laminar flow regime. Collectively, IB and LB make the lock-
in zone. The lock-in zone is generally bounded at both ends by a desynchronization zone
(Navrose et al., 2014). The first desynchronization zone (DS I) appears before IB, and the
rna

second desynchronization zone (DS II) appears after LB. Identifying new response branches
for various shapes of cylinders is evident in the literature. Kumar et al. (2018a) recently
proposed an oscillation frequency-based technique for identifying response branches. They
demonstrated the technique’s applicability on cylinders of various shapes. Later, Sourav et
al. (2020) resolved new response branches for an elliptic cylinder with an aspect ratio of 1.11
and m∗ = 1 using the technique proposed by Kumar et al. (2018a). In this study, we also
use the frequency-based method of Kumar et al. (2018a) to identify the response branches
for upstream and downstream cylinders. Figures 5(a) and 5(b) show the identified response
Jou

branches for upstream and downstream cylinders, respectively. The colored zones in both
figures refer to various response branches identified for tandem diamond cylinders. The
initial few points at low U ∗ belong to the desynchronization (DSI) regime. The frequency
curve of upstream and downstream cylinders (Figures 5a and 5b) makes its first jump at
U ∗ = 3.5. This marks the end of DSI and the beginning of IB for both cylinders. A change
in slope at U ∗ = 4 confirms the termination of IB and the beginning of another response
branch in both cylinders. While the upstream cylinder shifts to an upper branch UB, the
downstream cylinder witnesses an ‘extended initial branch’ (EIB). Both cylinders descend
to the lower branch (LB) at U ∗ = 5.8 for the upstream and 6.8 for the downstream cylinder.

15
Journal Pre-proof

A close examination of the frequency values in IB−LB reveals the degree of similarity of
IB LB 0.28 IB 1.2
0.30 0.64
Ymax/D Ymax/D
Fy
FN Fy
0.26 0.48 0.22 0.9
FN

Ymax/D
Ymax/D

Fy
0.16 0.6
Fy

0.22 0.32

of
DSI DSI EIB LB DSII
0.18 DSII 0.16 0.10 0.3

UB
0.14 0.00 0.04 0.0
0 4 8 12 16 20 0 4 8 12 16 20
* *
U U

pro
(a) (b)

Figure 5: (Color online) Branching behavior of (a) upstream and (b) downstream diamond cylinders in
tandem. The desynchronization regimes DS I and DS II are indicated, along with the initial branch (IB),
upper branch (UB), lower branch (LB), and extended initial branch (EIB), which are identified based on the
slope change of the oscillation frequency (Fy ) variation with reduced velocity (U ∗ ). Lock-in sets in when Fy
closely matches FN . Additionally, the non-dimensional maximum oscillation amplitude (Ymax /D) for each
cylinder is plotted.

re-
Fy values with the FN curve in IB−LB. Fy is closer to FN in LB than in IB, resulting in
higher oscillation amplitudes in LB. This is demonstrated in Figure 6. The departure of the
upstream cylinder’s Fy from the proximity of the FN curve at U ∗ = 7.1 ensures the end of
lock-in/LB and the onset of the second desynchronization regime (DS II). The lower branch,
for the downstream cylinder, is quite long and terminates at U ∗ = 14, followed by DS II.
Table 2) summarizes the extent of various response branches for isolated diamond (1- and
2-DOF), tandem square (two-DOF), and tandem diamond (2-DOF) cylinders.
lP
Table 2: Summary of various response branches for square and diamond cylinders.

Response branches −→ DSI IB EIB UB LB DSII


Oscillator Studies
1-DOF Diamond Sourav et al. (2022) 1–3.3 3.4–4.1 Absent Absent 4.2–6.8 6.9 onward
rna

2-DOF isolated Diamond Sourav et al. (2022) 1–3.3 3.4–4.1 Absent 4.2-5.7 5.8-6.8 6.9 onward
2-DOF Square UC Kumar and Sen (2021) 1-6.5 Absent Absent Absent 6.6-7.9 8 onward
2-DOF Square DC Kumar and Sen (2021) 1-6.5 Absent Absent Absent 6.6-7.9 8 onward
2-DOF Diamond UC Present 1-3.5 3.6-4 Absent 4.1-5.7 5.8-7.1 7.2 onward
2-DOF Diamond DC Present 1-3.5 3.6-4 4.1-6.8 Absent 6.8-14 14 onward
Jou

4.2. Oscillator amplitude response


4.2.1. across the flow
Figure 6 depicts the maximum non-dimensional transverse displacement of diamond
cylinders upstream (UC) and downstream (DC). Figure 6(a) illustrates a comparison with
the oscillation amplitudes of a single diamond cylinder (adapted from Sourav et al., 2022).
As expected, the upstream cylinder has oscillation amplitudes similar to an isolated diamond
cylinder, whereas the downstream cylinder has higher oscillation amplitudes. Because the
flow conditions are identical, the oscillation amplitudes of upstream and isolated diamond
cylinders are similar. Furthermore, the gap between the cylinders is important in determining

16
Journal Pre-proof

cylinder response (Zhao, 2013; Mysa et al., 2016; Bhatt and Alam, 2018; Qui et al., 2021).
The downstream cylinder lies in the wake of the upstream cylinder and is thus subject to the
upstream cylinder’s alternately shedding vortices. Not only are the oscillation amplitudes
similar for the upstream and isolated diamond cylinders, but so are the extents of various
response branches. Sourav et al. (2020) resolved an upper branch (UB) for an isolated
diamond cylinder performing the 2-DOF motion. Since the upstream cylinder’s oscillation

of
amplitude follows the single cylinder’s oscillation amplitude, and all other features are the
same, an upper branch is thus also resolved in the present research.

1.2 1.2

pro
UC Diamond UC Diamond
DC Diamond DC Diamond
Single Diamond UC Square
0.9 0.9 DC Square
Ymax/D

Ymax/D
0.6 0.6

0.3 0.3

0.0
0.0
4 8 12 16 20 0 4 8 12 16 20
*
U U*
(a)
re- (b)

Figure 6: (Color online) Comparison of the non-dimensional peak transverse oscillation amplitude (Ymax /D)
of upstream and downstream diamond cylinders in tandem with that of (a) an isolated diamond, and (b)
tandem square cylinders at Re = 100 and B = 0.05.

The downstream cylinder’s oscillation behavior is quite distinct. Its oscillation ampli-
lP
tude in DS I is the same as that of the upstream cylinder but slightly higher in DS II.
Both cylinders are locked in at the same U ∗ value. The upstream cylinder begins high-
amplitude oscillations soon after synchronization (IB), but the downstream cylinder does
not. It increases and decreases until IB is closed at U ∗ = 4 for both cylinders. This is
due to the vortices coalescing in the gap region between the cylinders (explained in Section
4.3.1 in detail). When switching to the next response branch (UB for upstream and EIB
rna

for downstream cylinder), the oscillation for both cylinders begins to grow. It is so because
the cylinders are vibrating close to their natural frequencies (see Figures 3 and 4). They
peak towards the closure of UB and EIB, respectively, and then gradually fall. While the
upstream cylinder desynchronizes at U ∗ = 7.1, the downstream cylinder continues to vibrate
with high amplitude near its natural frequency for a long range of U ∗ before locking out at
U ∗ = 15. The lower branch for the downstream cylinder extends till U ∗ = 14 with a local
increment for U ∗ = [11 − 12]. LB is followed by DS II and has approximately a constant
oscillation amplitude (≈ 0.15D).
Jou

Figure 6(b) compares the oscillation amplitudes for tandem diamond cylinders with tan-
dem square cylinders performing 2-DOF motion under identical flow conditions. The effect
of changing the angle of incidence is quite noticeable in the oscillation amplitudes of the
upstream and downstream cylinders. They are noticeably different from their square coun-
terparts. Figures 3 and 4 show that both diamond cylinders have higher oscillation frequen-
cies than square cylinders and thus lock in with the oscillator’s natural frequency earlier
than its square counterpart. As a result, both upstream and downstream diamond cylinders
experience high-amplitude oscillations. However, an extended initial and lower branch was
not observed for square cylinders (see Figure 3 of Kumar and Sen, 2021).

17
Journal Pre-proof

4.2.2. along the flow


In general, the trend shows a rise in the variability of Xmax /D as U ∗ increases, with
Xmax /D values consistently surpassing those of the previous U ∗ values. Nevertheless, a
departure from this increasing trend is detected within the lock-in zone (as noted by Sen
and Mttal, 2011; Navrose et al., 2014; Nepali et al., 2020; Kumar and Sen, 2021). For the
present case, the lock-in zone of each cylinder is distinct. The upstream cylinder experiences

of
lock-in in the U ∗ range of [3.5-7.1], whereas the downstream cylinder remains in lock-in in
the range of U ∗ [3.5-15]. Figure 7(a) illustrates Xmax /D−U ∗ variation for the present case.
The enlarged view of the lock-in zone (included as a sub-figure in Figure 7a) shows a similar
trend of local fluctuations in Xmax /D for the upstream cylinder. Notably, the lock-in of the

pro
upstream cylinder also affects the fluctuations of Xmax /D in the downstream cylinder. With
further increase in U ∗ , the upstream cylinder locks out, whereas the downstream cylinder
remains locked in, leading to local fluctuations in Xmax /D. In contrast, the uptrend of
Xmax /D for the upstream cylinder remains unaffected in this zone. To further strengthen
our understanding, the Xmax /D-U ∗ variations of an isolated diamond cylinder (Sourav et
al., 2022) and two square cylinders in tandem arrangements (Kumar and Sen, 2021) are
compared with those of two diamonds cylinders under identical conditions in Figures 7(a)
and 7(b).

1.8

0.35
UC Diamond
DC Diamond
Single Diamond
re- 1.8
UC Diamond
DC Diamond
UC Square
DC Square
1.2 1.2
Xmax/D

Xmax/D
lP
0.6 0.6
0.00
3 8

0.0 0.0
0 4 8 12 16 20 0 4 8 12 16 20
*
U U*
(a) (b)
0.15 0.15
UC Diamond UC Diamond
rna

DC Diamond DC Diamond
Single Diamond UC Square
DC Square
0.10 0.10
Xrms/D

Xrms/D

0.05 0.05

0.00 0.00
0 4 8 12 16 20 0 4 8 12 16 20
*
U U*
Jou

(c) (d)

Figure 7: (Color online) A comparison of non-dimensional peak in-line oscillation amplitude (Xmax /D) and
its root-mean-square (r.m.s) (Xrms /D) for tandem diamond cylinders with isolated diamond cylinder (a, c),
and tandem square cylinders (b, d). A zoomed-in view of the lock-in zone of the upstream diamond cylinder
is shown as an inset in Figure (a).

As hypothesized, any local variations in the Xmax /D-U ∗ curve from the monotonous
uptrend are only observed in the lock-in regime of respective cylinders. Figure 7(a) indicates
the qualitative and quantitative similarity of Xmax /D-U ∗ variations for the isolated diamond
and the upstream diamond cylinder. In contrast, the Xmax /D-U ∗ curve for tandem diamond

18
Journal Pre-proof

and square cylinders portrayed in figure 7(b) are significantly different. This distinction of
Xmax /D-U ∗ curve is attributed to two major factors, i.e., (i) the difference in the shape of
the cylinders and (ii) the lock-in range of the cylinders. Both square cylinders have the same
lock-in range, whereas the upstream and downstream diamond cylinders have different ones.
It should be noted that the above findings have significant implications for fluid dynamics.
A comparison of the fluctuations in inline oscillations (Xrms /D) for isolated diamond

of
(Figure 7c), tandem square (Figure 7d), and tandem diamond cylinders reveals distinct
behavior. Unlike the equivalent Xmax /D for upstream and isolated diamond cylinders within
the lock-in zone, the Xrms /D for the upstream diamond cylinder is marginally larger up to
U ∗ = 6.7. Subsequently, a localized increase in fluctuation intensity is observed around U ∗ =

pro
12, which coincides with the U ∗ value at which the downstream cylinder exhibits increased
inline oscillation fluctuations for the second time. This finding suggests that the downstream
diamond cylinder generates wake or flow disturbances that propagate upstream, influencing
the flow around the upstream cylinder. As a result, the vortex shedding patterns and forces
acting on the upstream cylinder are altered, ultimately affecting the Xrms . Furthermore,
for the downstream cylinder, two peaks are visible in the Xrms /D − U ∗ variation. Notably,
the start of the second upward trend in Xrms values corresponds with the onset of the
downstream cylinder’s lower branch or extended lock-in zone. It is worth mentioning that
re-
the amplitude of in-line fluctuations in the downstream cylinder is considerably greater than
that of the upstream cylinder. This higher amplitude of the downstream cylinder can be
attributed to the continuation of lock-in. In contrast, the upstream cylinder exits the lock-in
stage at this point, resulting in a lower in-line response in the desynchronization zone, which
aligns with the typical expectations.
In FIV studies, both the root-mean-square (r.m.s) of inline oscillations and maximum
lP
oscillation amplitude are important, as they provide different insights into the system’s be-
havior. The r.m.s of inline oscillations indicate the average energy content of the vibrations,
which is useful for understanding the system’s overall response to vortex shedding and for as-
sessing the fatigue life of structures. On the other hand, the maximum oscillation amplitude
provides insight into the peak values of oscillatory motion, which is crucial when evaluat-
ing a system’s structural safety and stability under extreme conditions. Considering the
rna

complexity of FIV interactions, further research is required to fully comprehend the mecha-
nisms underlying these observed trends and how they can improve fluid systems’ design and
efficiency.

4.2.3. Motion Trajectory


Motion trajectories can help determine various wake modes, evaluate mitigation strate-
gies, and develop FIV prediction models. The trajectory of the upstream and downstream
cylinder motion in two dimensions is presented in Figure 8. Both cylinders traverse single-
Jou

looped, double-looped, and complex paths. The number of loops in the X − Y plots is
directly related to the frequency of the in-line and transverse oscillations. The single-looped
X − Y plot suggests that the cylinder oscillates along and across the flow with the same
frequency. On the other hand, the double-looped path traversed by the cylinder means that
the frequency of in-line oscillation is twice that of transverse oscillation. X − Y plots with
many loops indicate the cylinder’s complex periodic motion, i.e., quasi-periodic. FFTs of X
and Y signals for each motion type at representative U ∗ values are shown in Figure 9.
It can be established from Figure 8 that the DS I response branch oscillations are periodic
for both cylinders. However, on this branch, the motion of the upstream cylinder is domi-

19
Journal Pre-proof

0.05 0.4

U* = 3.5
U* = 3

U* = 4

U* = 4.5
0.6 1.5
Y/D

Y/D

Y/D

Y/D

Upstream cylinder
-0.05 -0.4 -0.6 -1.5
0.065 0.089 0.09 0.12 0.1225 0.1265 0.02 0.42
X/D X/D X/D X/D

of
U* = 7
1 0.6

U* = 10

U* = 15
0.25 0.2
U* = 6.5

Y/D
Y/D

Y/D
Y/D
-1 -0.6 -0.25 -0.2
0.42 0.48 0.45 0.48 0.864 0.890 1.916 1.926
X/D X/D X/D

pro
X/D

U* = 3.5

U* = 4
0.1

U* = 4.5
0.25 0.2 0.25
U* = 3

Y/D
Y/D
Y/D

Y/D

Downstream cylinder
-0.1 -0.25 -0.2 -0.25
0.005 0.051 -0.02 0.10 0.000 0.008 -0.15 0.25
X/D X/D X/D X/D
U* = 6.5

U* = 7

1.7 2.1

U* = 10

U* = 15
1.5 0.32
Y/D

re-
Y/D

Y/D

Y/D
-1.7 -2.1 -1.5 -0.32
0.08 0.28 0.22 0.46 0.19 0.65 0.512 0.526
X/D X/D X/D X/D

Figure 8: (Color online) The X − Y motion trajectories for upstream and downstream cylinders at represen-
tative U ∗ values (= 3, 3.5, 4, 4.5, 6.5, 7, 10, and 15) across multiple response branches. The top two rows
are for the upstream and bottom two rows are for the downstream cylinder.
lP
nated along the flow while the downstream cylinder experiences more oscillations across the
flow. The exposure of the downstream cylinder to non-uniform flow (wake of the upstream
cylinder) can be attributed to increased transverse oscillations. Both cylinders demonstrate
quasi-periodic motion at the onset of lock-in (U ∗ = 3.5), which becomes periodic for the
rna

upstream cylinder near the closure (U ∗ = 4) of the initial branch IB. However, the down-
stream cylinder periodicity is revoked at a later U ∗ value. At U ∗ = 4.5, which belongs to
the upper branch (UB) for the upstream and extended initial branch (EIB) for the down-
stream cylinder, the cylinder oscillates with equal frequency along and across the flow. The
entire UB and a part of the LB of the upstream cylinder showcase this behavior. Sourav
et al. (2022) also reported similar cylinder trajectories for the isolated diamond cylinder,
but only on UB and not LB. Due to the single-looped plot, they called the such motion a
‘raindrop-shaped’ cylinder motion. Kang et al. (2016), and Dorogi and Baranyi (2019) also
Jou

reported ‘raindrop-shaped’ cylinder motion in their studies. Both cylinders seize to oscillate
with equal frequencies along and across the flow at U ∗ = 6.3. This is a point on LB for the
upstream and EIB for the downstream cylinder. Notably, the X −Y trajectory is asymmetric
on the UB for the upstream cylinder, while it’s periodic for the downstream cylinder on EIB.
Sourav et al. (2022) also observed this phenomenon. For both cylinders, the plot is oriented
in the 2nd and 4th quadrants at U ∗ = 4.5 and in the 1st and 3rd quadrants at U ∗ = 6.3.
This essentially means that the in-line oscillations lag behind the transverse oscillations at
U ∗ = 4.5.
On the other hand, at U ∗ = 6.3, the in-line oscillations lead to the transverse oscillations

20
Journal Pre-proof

1.0 1.0 1.0 1.0

U* = 3

U* = 4.5
Fx

U* = 3.5

U* = 4
Fx
P(X), P(Y)

P(X), P(Y)
Fx Fx

P(X), P(Y)

P(X), P(Y)

Upstream cylinder
Fy Fy Fy Fy
0.5 0.5 0.5 0.5

0.0 0.0 0.0 0.0


0.0 0.5 1.0 0.0 0.5 1.0 0.0 0.5 1.0 0.0 0.5 1.0
F F F F

of
1.0 1.0 1.0 1.0

U* = 15
U* = 7
U* = 6.5
Fx Fx

U* = 10
Fx

P(X), P(Y)
P(X), P(Y)

Fx

P(X), P(Y)
P(X), P(Y)
Fy Fy Fy Fy
0.5 0.5 0.5 0.5

0.0 0.0 0.0 0.0


0.0 0.5 1.0 0.0 0.5 1.0 0.0 0.5 1.0 0.0 0.5 1.0
F F F

pro
F
1.0 1.0 1.0 1.0

U* = 3.5

U* = 4.5
U* = 3

U* = 4
Fx Fx Fx Fx
P(X), P(Y)
P(X), P(Y)

P(X), P(Y)

P(X), P(Y)
Fy Fy Fy Fy

Downstream cylinder
0.5 0.5 0.5 0.5

0.0 0.0 0.0 0.0


0.0 0.5 1.0 0.0 0.5 1.0 0.0 0.5 1.0 0.0 0.5 1.0
F F F F

1.0 1.0 1.0 1.0


U* = 7
U* = 6.5

U* = 10

U* = 15
Fx Fx
P(X), P(Y)

Fx Fx
P(X), P(Y)

P(X), P(Y)

P(X), P(Y)
Fy Fy Fy Fy
0.5

0.0
0.0 0.5
F
1.0
0.5

0.0
re-
0.0 0.5
F
1.0
0.5

0.0
0.0 0.5
F
1.0
0.5

0.0
0.0 0.5
F
1.0

Figure 9: (Color online) Fast Fourier Transforms (FFTs) of in-line and transverse response fluctuations
plotted together for upstream and downstream cylinders at representative U ∗ values (= 3, 3.5, 4, 4.5, 6.5, 7,
10, and 15) across multiple response branches. The top two rows are for the upstream, and the bottom two
lP
are for the downstream cylinder. In all FFT plots, F refers to the non-dimensional frequency, and P(X) and
P(Y ) are the power spectral density of in-line and transverse displacements of the cylinder center. Similarly,
Fx and Fy are the corresponding non-dimensional oscillation frequencies.

(Figure 9). This change in the relative position of X/D from “lagging” to “leading” Y /D
in time traces can be attributed to the change in the orientation of the cylinder trajectory.
rna

The cylinder motion for the upstream cylinder at other U ∗ values on LB and DS II is mostly
periodic except at U ∗ = 10. On the contrary, the cylinder motion on LB and DS II of the
downstream cylinder is quasi-periodic.

4.3. Fluid dynamic loading


The fluid loading characteristics of upstream and downstream diamond cylinders are
shown in Figure 10. A comparison with corresponding values for isolated diamond cylinders
Jou

and tandem square cylinders is also presented. The first (Figures 10a & 10b) and second
(Figures 10c & 10d) rows of the figure present the variation of mean drag and r.m.s. of lift
coefficients, respectively with U ∗ . The upstream diamond cylinder’s drag is compared with
the corresponding square and isolated diamond cylinder in Figure 10(a). The horizontal
dashed line in the figure is the value of Cdavg for stationary tandem diamond cylinders
S
under identical flow conditions and cylinder arrangement (i.e., D = 5). It is apparent from
the figure that the upstream diamond cylinder always experiences more drag than its square
counterpart. An isolated diamond cylinder experiences a drag close to its corresponding value
for a stationary diamond cylinder (= 1.7880) in the desynchronization regimes (Sourav et al.,
2022) with a dip on the initial branch. A similar variation of drag is seen for the upstream

21
Journal Pre-proof

diamond cylinder. It follows the drag (= 1.6411) for stationary tandem diamond cylinders
in DS I and DS II. It is also noteworthy to report that an upstream diamond cylinder
always experiences a smaller drag than an isolated diamond cylinder. A higher value of
vortex-shedding frequency for isolated cylinders can be attributed to this variation.
The drag experienced by the downstream diamond cylinder and its deviation from the
corresponding square counterpart is shown in Figure 10(b). It is discernible from the figure

of
that the downstream diamond cylinder experiences a smaller drag than the corresponding
square cylinder. This is opposite to what was observed for upstream diamond cylinders. Like
the upstream cylinder, the downstream diamond cylinder also experiences a drag similar to
its stationary counterpart in the desynchronization regime. Physically, in desynchronization

pro
regimes DS I and DS II, the cylinders execute low amplitude oscillations at a frequency close
to its stationary vortex-shedding frequency. These low amplitude oscillations in DS I and
DS II seem to have negligible effects on the drag. Compared to the drag for the upstream
diamond cylinder, a smaller drag is experienced by the downstream diamond cylinder. This
finding is in line with the findings of Kumar and Sen (2021) for tandem square cylinders
and Mysa et al. (2016) for tandem circular cylinders. The reason could be the pressure

1.5
2.6
2.4

2.0
UC Diamond
UC Square
Single Diamomd
re- avg

1.0
DC Diamond
DC Square
avg

Cd

1.7880
Cd

0.5
1.6 1.6411
0.4245
lP
1.2 0.0
0 4 8 12 16 20 0 4 8 12 16 20
*
U
* U
(a) (b)
1.2 1.4
UC Diamond
UC Square 1.2
Single Diamomd
rna

0.9 0.9109

0.8
rms
rms

0.6 0.4791
Cl
Cl

0.4
0.3
DC Diamond
DC Square
0.0 0.0
0 4 8 12 16 20 0 4 8 12 16 20
* *
U U
Jou

(c) (d)

Figure 10: (Color online) The variation of time-averaged coefficient of drag Cdavg of upstream diamond,
isolated diamond, and upstream square cylinders are compared in Figure (a). It also shows the mean drag
coefficients for a stationary isolated diamond cylinder (1.7880) and an upstream diamond cylinder (1.6411)
in tandem arrangement. The variation of Cdavg for downstream diamond and square cylinders is presented
in Figure (b). In the Figure, 0.4245 is the average drag coefficient (Cdavg ) of stationary downstream diamond
cylinder in the tandem arrangement. Corresponding comparisons for r.m.s of lift coefficient (Clrms ) are show
in Figures (c) and (d), respectively. The values of 0.4791 and 0.9109 in Figures (c) and (d) represent the
root-mean-square lift coefficients (Clrms ) for the stationary upstream and downstream diamond cylinders in
tandem arrangement, respectively.

22
Journal Pre-proof

difference between the front and rear surfaces of the cylinder. In a co-shedding regime, the
gap region between the cylinders is the wake for the upstream cylinder, a low-pressure zone.
As a result, the space between the cylinders will have a lower pressure than the pressure at
the front surface of the upstream cylinder. The pressure is highest at the forward stagnation
point on the frontal surface of the body. This pressure difference between the front and
rear surfaces of the cylinder greatly influences the drag experienced by the cylinder. This

of
also holds for the downstream cylinder, but note that the downstream cylinder is already
in the low-pressure zone. This implies that the pressure difference between the front and
rear surfaces of the downstream cylinder is even smaller, resulting in lower drag than the
upstream cylinder.

pro
The transverse oscillations are the outcome of the lift force acting on the cylinder, which
varies with time as vortices shed and interact with the cylinder. In FIV studies, the r.m.s.
of the lift is often considered to be more important than its mean value because it provides
information about the fluctuation intensity of the lift force acting on the object. This is
significant because the intensity of the fluctuation, rather than the mean value, decides the
level of fatigue damage that a structure will incur. Even though the mean lift coefficient is
low, high variation intensity can cause severe fatigue damage. Taking this into account, the
variation of Clrms for upstream and downstream diamond cylinders are presented in Figures
10(c) and 10(d), respectively. re-
It is depicted from Figure 10(c) that the upstream diamond cylinder experiences more or
less the same fluctuations in lift as that of the isolated diamond cylinder. This establishes
that the presence of the downstream cylinder has negligible effect. Except for the synchro-
nized zone, the lift experienced by the upstream diamond cylinder is close to the Clrms of the
tandem stationary diamond cylinders. The differences between the Clrms for the upstream
lP
square cylinder are also established in Figure 10(c). The lift fluctuations for the upstream
diamond cylinder are higher than their square counterpart. The jumps and falls in the
lift fluctuations of the diamond and square cylinders can be attributed to the difference in
natural and vortex-shedding frequencies of the cylinders resulting in different lock-in zones.
But the variation of Clrms for the downstream cylinder (Figure 10d) is different from that
of the upstream cylinder. It experiences lower fluctuations in the lift than the downstream
rna

square cylinder. Like the upstream cylinder, except for the synchronized zone, the down-
stream cylinder also experiences a lift close to its tandem stationary diamond cylinder (=
0.9109). The differences in the lift fluctuations could result from the vortex-shedding from
the upstream cylinder.
The fluid forces on the upstream cylinders are solely governed by the vortex-shedding
from its surfaces. However, for the downstream cylinder, in addition to the vortices shed by
the downstream cylinder, the vortices shed by the upstream cylinder also play a role. These
vortices impinge on the downstream cylinder causing large fluctuations in the flow field. As
Jou

a consequence, larger fluctuations in the lift are observed for the downstream cylinder. The
r.m.s lift can also measure vortex-shedding strength (Yang and Wu, 2013) because the lift
force is proportional to the vortex-shedding frequency and the strength of the vortices. It is a
well-accepted fact that higher lift fluctuations imply a vortex-shedding with higher strength.
Thus, in Figures 10(c) and 10(d), for upstream and downstream diamond cylinders, the
vortex-shedding will be the strongest at a value of U ∗ with maximum Clrms and weakest
at U ∗ with minimum Clrms . A lift fluctuation close to zero signifies that the amplitude of
the lift force acting on the cylinder is also close to zero, and there is very little variation in
the force over time. Note that it doesn’t necessarily infer that the mean lift is zero. In the

23
Journal Pre-proof

context of practical significance, a low lift fluctuation is desirable in engineering applications


where minimizing the fluctuation is important to reduce the risk of fatigue damage to the
structure.

4.3.1. Asymmetric mean lift


Next, the mean lift (Clavg ) for upstream and downstream diamond and square cylinders,

of
along with an isolated diamond cylinder, is presented in Figure 11. The symmetric obstacle
in the flow results in the alternate shedding of inherently unsteady and symmetric vortices,
generating a small or zero time-averaged lift on the obstacle. As presented in Figure 11,
the upstream and downstream square cylinders experience a lift force close to zero, implying

pro
that the vortex-shedding for these cases is symmetric. However, for a small range of U ∗ ,
the mean lift generated on the upstream and downstream cylinders is non-zero. For an

0.4
UC Diamond
DC Diamond
UC Square
DC Square
Single Diamond
0.2

re-
avg

0.0
Cl

-0.2
lP
-0.4
0 4 8 12 16 20
*
U

Figure 11: (Color online) A comparison of the time-averaged lift coefficient (Clavg ) for tandem diamond,
rna

isolated diamond, and tandem square cylinders is presented. A non-zero mean lift is primarily observed for
the isolated and upstream diamond cylinders.

isolated cylinder executing 2-DOF FIV, Sourav et al. (2022) reported a non-zero mean lift
for a range of U ∗ values corresponding to the upper branch (UB). The upstream diamond
cylinder also experiences a non-zero mean lift on the upper branch and a part of the lower
branch (U ∗ = 4.1 − 6.8). However, the downstream cylinders don’t experience a large non-
zero mean lift, but small deviations (≈ 0.09) from the zero mean lift line are obtained. This
Jou

indicates that the vortex-shedding will be asymmetric for such cases. A detailed discussion
on various modes of vortex-shedding is presented next.

4.3.2. Transition from negative to positive non-zero mean lift


Another significant and intriguing observation in Figure 11 that necessitates explanation
is the change in mean lift direction. Two representative values of U ∗ (= 4.7 and 5) corre-
sponding to negative and positive non-zero mean lift are chosen for this discussion. The lift
and drag characteristics are investigated in Figure 12 using temporal variations and phase
portraits. The zero-lift line in Figure 12 is used as a reference to depict negative and positive
biases for the cylinders. The asymmetry in the lift experienced by the upstream cylinder is

24
Journal Pre-proof

U* = 4.7 Upstream cylinder


(a) 3.5
Cl Cd (b) 1.0
3.0

1.5 0.0
Cl, Cd

Cl

of
0.0
-1.0

-1.5
-1.5 1.8 2.4 3.0
360 380 400 420 430
tU/D Cd
U* = 5

pro
(c) 3.5
Cl Cd (d) 1.2
0.8
Cl, Cd

1.5

Cl
0.0

0.0

-0.8
-1.0 1.72 2.25 2.78

(e) 0.5
600

U* = 4.7
620

re-
tU/D

Downstream cylinder
640

Cl Cd
650

(f) 0.3
Cd

0.3
Cl, Cd

Cl

0.0
lP
0.0

-0.3
-0.3 0.22 0.30 0.38
360 380 400 420
tU/D Cd
U* = 5
(g) 0.8
Cl Cd
(h) 0.3
rna

0.6
Cl
Cl, Cd

0.3 0.0

0.0

-0.3
-0.3
600 620 640 660 0.15 0.40 0.65
tU/D Cd
Jou

Figure 12: (Color online) Time histories of drag and lift and corresponding Cd − Cl phase portraits at
representative U ∗ values of 4.7 and 5 for upstream (a−d) and downstream (e−h) diamond cylinders. The
dashed line in the time histories and continuous line in the Cd − Cl phase portraits corresponds to zero lift.

established by the time traces and Cd − Cl phase portraits in Figure 12(a−d). The lift is
negatively biased at U ∗ = 4.7 but positively biased at U ∗ = 5. The degree of lift asymmetry
can be estimated by comparing the upper and lower bounds of the phase portraits. While the
upstream cylinder has a non-zero mean lift, the downstream cylinder has a lesser degree of
non-zero mean lift (Figures 12e−h). This is also visible in the corresponding phase portraits.

25
Journal Pre-proof

4.3.3. What causes such a transition?


Figure 13 depicts the vortex-shedding mechanisms for one oscillation cycle of the up-
stream cylinder at U ∗ = 4.7 and 5 to understand better the asymmetric mean lifts and
their transition from negative to positive values. The first and second rows of the figure
have a distinct feature: they are nearly identical but opposite. The vortex-shedding process
goes through three stages for both U ∗ values: Stage I, II, and III. The first stage (Figures

of
13a and 13d) determines the difference in vortex strengths/sizes from the top and bottom
shoulders for clockwise and anticlockwise vortices, respectively. This difference in strength
and the arrangement of clockwise and anticlockwise vortices in the wake of the upstream
cylinder accounts for asymmetry in the pressure distribution near the cylinder, resulting in a

pro
non-zero mean lift. Sourav et al. (2022) also observed negative and positive mean lifts with
a similar arrangement and varying strengths of vortices in the wake of an isolated diamond
cylinder. Progressing forward in time, the stretched vortices in Stage I stretch even more.
The elongated portion aligns with the oppositely signed vortex in Stage II (Figures 13b and
13e) before the stretched vortex notices the necking phenomenon and breaks (Figures 13c
and 13f).

4 4 4

2
U *= 4.7
2 re- U *= 4.7
2
U *= 4.7
Y/D

Y/D

Y/D
0 0 0

-2 -2 -2
stretching stretching Breaking
Stage I Stage II Stage III
-4 -4 -4
-2 0 2 4 6 8 10 12 -2 0 2 4 6 8 10 12 -2 0 2 4 6 8 10 12
X/D X/D X/D
lP
(a) (b) (c)
4 4 4
U *= 5 stretching U *= 5 Breaking U *= 5
stretching
2 2 2
Y/D

Y/D

Y/D

0 0 0

-2 -2 -2
Stage I Stage II Stage III
-4 -4 -4
rna

-2 0 2 4 6 8 10 12 -2 0 2 4 6 8 10 12 -2 0 2 4 6 8 10 12
X/D X/D X/D
(d) (e) (f)

Figure 13: (Color online) Instantaneous vortex-shedding plots in the gap region between cylinders for U ∗ =
4.7 (a-c) and 5 (d-f), showing negative and positive non-zero mean lifts of comparable magnitude. Three
Stages (I, II, and III) are identified, with Stages I and II associated with vortex stretching and Stage III
depicting the break of stretched vortices.
Jou

4.3.4. Cylinder displacement: symmetric or asymmetric?


A positive or negative bias in the lift results in a stronger lift force acting on the cylinder’s
top or bottom surface. Lift asymmetry may cause cylinders to oscillate asymmetrically. To
demonstrate this, Figure 14 shows the Y − Cl phase portraits of upstream and downstream
cylinders for U ∗ = 4.7 and 5. The abscissa’s extremities are marked at symmetric distances
from the origin. The gaps between the extreme right and left points of the Y − Cl phase
portraits and the abscissa’s right and left boundaries are filled with green color. The thickness
of the colored patches near the left and right boundaries of Figures 14(a) and 14(c) (for the
upstream cylinder at U ∗ = 4.7 and 5, respectively) indicates asymmetry in the cylinder’s
transverse displacement. However, for the downstream cylinder, these colored patches are

26
Journal Pre-proof

either symmetric or negligibly asymmetric, indicating slight or no asymmetry in its transverse


response (see Figures 14c and 14d).

0.8
UC 0.3
DC 1.2
UC DC
0.3
U* = 4.7 U* = 4.7 U* = 5 U* = 5

0.0

of
Cl

Cl

Cl
0.0

Cl
0.0
0.0

-1.4 -0.3 -0.8 -0.3


-1.2 0.0 1.2 -0.3 0.0 0.3 -1.2 0.0 1.2 -0.3 0.0 0.3
Y/D Y/D Y/D Y/D

pro
(a) (b) (c) (d)

Figure 14: (Color online) Phase portraits of Y /D − Cl at U ∗ = 4.7 (a, b) and 5 (c, d) for upstream (U C)
and downstream (DC) cylinders, respectively. The difference in the width of the green zone represents the
asymmetry in cylinder displacement between upstream and downstream diamond cylinders.

Despite the large and significant variations in lift coefficients, the small asymmetry ob-
served in transverse displacement may be questioned. One possible explanation for this can
be the following: The cylinder’s lift and transverse displacement are related by Equation 7.
It can be rewritten as:

Y ∗ |m =
 ∗ 2 
U

Cl |m
m∗

 2 ∗
d Y |m
dt∗2
re-
+ 4πζ
1 dY ∗ |m
U ∗ dt∗

. (10)

Here, the variables with (-)m represent the corresponding means. At U ∗ = 5, the mean for
∗ 2 ∗
all other time-varying variables is Cl |m = 0.3680, dYdt∗|m = −0.0050, and d dtY ∗2|m = 0.0032 for
lP
Y ∗ |m of the upstream cylinder of m ∗ = 10. Since the study sets the structural damping to
zero, the term with ζ has no contribution. Substituting the relevant values into Equation
10 results in: Y ∗ |m = 0.0212. This is a low value compared to the transverse displacement of
the oscillator, and thus we don’t observe significant non-zero mean transverse displacement.
On closely examining the values of the time-varying parameters, we observe that the
contribution of both acceleration and velocity terms is negligible. As a result, the only
rna

term controlling the asymmetry for a given U ∗ is m∗ . Assuming that the contribution of
all other terms is negligible, the mean displacement Y ∗ |m ∝ Cml |∗m for a given U ∗ . The mean
displacement will be even smaller for higher m∗ values. It may, however, be considerable for
lighter structures (m∗ ≈ O(1)).

4.4. Quasi-periodicity in flow


Periodicity and quasi-periodicity of flow are important to study because they affect the
Jou

dynamics of fluid-structure interactions and the resulting vibration patterns. Periodic flow
refers to a pattern that repeats after a fixed interval, while quasi-periodic flow patterns do
not repeat exactly but have some regularity or periodicity. The time traces of lift and drag
coefficients and/or Cd − Cl phase portraits can be analyzed to establish periodicity/quasi-
periodicity in flow. Figure 15 presents the Cd − Cl phase portraits for upstream and down-
stream cylinders. These plots are created by plotting approximately 20 cycles of the lift and
drag signals simultaneously.
It is evident from the figure that, for the upstream cylinder exposed to uniform flow,
except for the initial branch (U ∗ = [3.5 − 4]), the drag and lift coefficient values overlap. The
resulting Cd − Cl phase portraits exhibit a clear and regular closed-loop pattern establishing

27
Journal Pre-proof

0.8 1.5 1.5 1.5

U* = 3.5
U* = 3

U* = 4

U* = 4.5
Cl

Cl
0.0 0.0

Upstream cylinder
Cl

Cl
0.0 0.0

-0.8 -1.5 -1.5 -1.5


1.622 1.630 1.638 1.4 1.6 1.8 1.44 1.49 1.54 1.8 2.4 3.0
Cd Cd Cd Cd

of
0.6 0.6 0.8

U* = 6.5

U* = 7
0.8

U* = 10

U* = 15
Cl
Cl

0.0 0.0

Cl
0.0

Cl
0.0

-0.6 -0.6 -0.8 -0.8


1.70 2.05 2.40 1.65 1.82 1.99 1.60 1.68 1.76 1.58 1.64 1.70
Cd Cd Cd Cd

pro
1.5 2 0.6 0.3

U* = 4
U* = 3

U* = 3.5

U* = 4.5
Cl
0.0

Cl
Cl

0 0.0
Cl

0.0

Downstream cylinder
-2 -0.6 -0.3
-1.5
0.425 0.445 0.465 0.1 0.7 1.3 -0.01 0.03 0.07 0.14 0.34 0.54
Cd Cd Cd Cd

1.5 1.5
U* = 6.5

U* = 7

1.5 1.5

U* = 10

U* = 15
re-
Cl

0.0
Cl

0.0
Cl

Cl
0.0 0.0

-1.5 -1.5 -1.5 -1.5


0.3 0.8 1.3 0.6 1.5 2.4 0.3 0.8 1.3 0.36 0.44 0.52
Cd Cd Cd Cd

Figure 15: (Color online) Phase portraits of Cd − Cl for upstream and downstream diamond cylinders at
representative U ∗ values (= 3, 3.5, 4, 4.5, 6.5, 7, 10, and 15) across multiple response branches. The
horizontal zero line (green) in each plot represents zero lift. The top two rows are for the upstream, and the
lP
bottom two are for the downstream cylinder.

periodicity. For most periodic cases, the drag frequency is twice that of the lift, implying
that the drag signals go through two cycles by the time the lift signals go through one.
For such cases, the Cd − Cl phase portraits are symmetric, two-looped, and assume a figure
rna

‘eight.’ The shape and size of the loop correspond to the frequency and amplitude of the
drag and lift coefficients. The phase portraits are asymmetric (due to asymmetric lift) on
the upper and some portions of the lower branches (detailed discussion made in Section 4.3).
Notably, some of these plots are multi-looped due to the presence of multiple frequencies. On
the other hand, for the downstream cylinder exposed to the wake of the upstream cylinder,
the Cd − Cl phase portraits are expected to be irregular for most of the values of U ∗ . The
disturbances in the flow caused by the vortices of the upstream cylinder significantly impact
the flow downstream, making it quasi-periodic. As seen from Figure 15, the flow for the
Jou

downstream cylinder can be considered periodic only in the desynchronization regime. The
presence of multiple bands in the Cd − Cl phase portraits on other response branches ensures
quasi-periodic flow. These plots are irregular, with time-varying closed loops.
Single-looped Cd −Cl phase portraits are resolved for upstream and downstream cylinders,
confirming the equivalence of drag and lift frequencies. In coherence with the discussions in
the previous section, the multiple frequencies are analyzed at U ∗ = 4.7 and 5 using the Fast
Fourier Transform (FFT) on sufficiently large data sets for drag and lift coefficient signals.
Figure 16 plots the frequencies of drag and lift coefficients for U ∗ = 4.7 and 5. Figures in
the first (16a, 16c) and second (16b, 16d) columns correspond to FFTs of lift and drag

28
Journal Pre-proof

UC DC UC DC
1.0 1.0 1.0 1.0
FC FC FC FC
l l l l
FCd FCd FCd FCd
d)

d)

d)

d)
P(C ), P(C

P(C ), P(C

P(C ), P(C

P(C ), P(C
0.5 U* = 4.7 0.5 U* = 4.7 0.5 U* = 5 0.5 U* = 5
l

of
0.0 0.0 0.0 0.0
0.0 0.5 1.0 0.0 0.5 1.0 0.0 0.5 1.0 0.0 0.5 1.0
F F F F
(a) (b) (c) (d)
Figure 16: (Color online) Fast Fourier Transforms (FFTs) of drag and lift coefficient signals for upstream

pro
(U C) and downstream (DC) cylinders at U ∗ = 4.7 and 5. In these plots, F refers to the non-dimensional
frequency, and P(Cd ) and P(Cl ) are the power spectral density of drag and lift coefficients. Similarly, FCd
and FCl are the corresponding non-dimensional fluid loading frequencies.

coefficients acting on upstream and downstream cylinders, respectively. These FFT plots
clearly show that the drag frequency is twice that of lift for the upstream cylinder and that
these frequencies are equivalent for the downstream cylinder. Notably, the various frequencies
for the fluid forces acting on the upstream cylinder are harmonics of the dominant frequency.
re-
However, weak incommensurate frequencies appear near the downstream cylinder’s dominant
frequency, resulting in banded Cd − Cl phase plots in Figures 12f and 12h.

4.5. The flow


FIV is a two-way coupled phenomenon where the fluid flow affects the oscillation, and
oscillations affect the fluid flow. So far, we have discussed the effect of fluid on oscillation
lP
characteristics. Here, we are discussing the wake of the tandem diamond cylinder.

4.5.1. Mode transition in the gap region


The interdependence of the vortex-shedding and oscillation amplitude for isolated oscil-
lating cylinders has been well-established in the literature (Navrose et al., 2014; Kumar et
al., 2018c). However, this relationship is not as well-defined for tandem cylinders as for a
single cylinder. The interaction between the cylinders can significantly affect the oscillation
rna

amplitude of the cylinders. In some cases, the wake of one cylinder can reduce the amplitude
of the oscillations of the other cylinder. In other cases, the wake can enhance the oscillation
amplitude and lead to large-amplitude vibrations and potential structural failure. It depends
on the specific geometry and flow conditions.
The vortex-shedding frequency (FCl ) of the upstream cylinder is depicted in 17(a). A
dashed line represents the vortex-shedding frequency of stationary tandem diamond cylinders
(= 0.1671). The figure illustrates that FCl in the desynchronization regimes (DS I and DS
Jou

II) is very close to that of stationary tandem cylinders. It should be noted that this does not
entail that the vortex-shedding modes of stationary and vibrating cylinders will be the same
or similar in the desynchronization regimes. A value of FCl < 0.1671 in DS II suggests that
cylinder vibrations interfere with the shedding of vortices in the cylinder’s wake, slowing the
rate of vortex-shedding. A lower FCl value indicates a decrease in fluid forces acting on the
cylinder. This is demonstrated in Figures 10(a) and (c). However, in the lock-in zone, where
the oscillation amplitudes are high, the vortex-shedding frequency differs significantly from
that of a stationary counterpart, resulting in different vortex-shedding patterns.
Three modes (Mode I, Mode II, and Mode III) of vortex-shedding in the gap region are
observed depending on the arrangement of vortices in the gap region between the cylinders.

29
Journal Pre-proof

(c) 2
U* = 3.8 (a) (d) 2
U* = 4.7
0.27
Y/D

Y/D
0 Mode II 0

Mode III
0.24
Mode II Mode III
-2 -2
0 5 0 5

of
l
FC
X/D 0.21 X/D
(b) U* = 3 (e) U* = 8.5
2 Mode I 2
Mode I
0.18 0.1671
Y/D

Y/D
0 0

pro
0.15
-2 Mode I -2
Mode I
0 5 0 4 8 12 16 20 0 5
*
X/D U X/D

Figure 17: (Color online) Figure (a) demonstrates the correlation between U ∗ and the vortex-shedding
frequency (FCl ) for the diamond cylinder located upstream. The dashed line on the graph represents the
vortex-shedding frequency of a stationary tandem diamond cylinder. Figures (b−e) illustrate three distinct
vortex modes (Mode I, II, and III) observed in the gap region between the cylinders, where the vorticity
values are scaled between -1 and 1. In Modes I (b, e) and III (d), two and three counter-rotating vortices

re-
are shed in the gap region, respectively, while the vortices coalesce in Mode II.

The distinct nature of these modes is evident in FCl . These have a significant impact on
the oscillation amplitude of the downstream cylinder. While the upstream cylinder sheds
vortices in Mode I in the DS I and LB & DS II regimes, the downstream cylinder experiences
Modes II and III vortices during the upstream cylinder’s lock-in. The range of U ∗ for which
lP
these modes are observed is shaded in Figure 17(a). Figure 17(b−e) shows the representation
of Modes I, II, and III at one representative value U ∗ in each zone. In Mode I, a vortex is
shed, and other alternating vortices form in the gap region. The vorticity plots also show
that the vortices in Mode I are more diffusive than those in the other two modes, implying
that the strength of the vortices in Mode I is lower than in the other modes. Figures
17(b) and 17(e) show the similarity of the vortex-shedding pattern for U ∗ = 3 and 8.5 in
rna

desynchronization regimes DS I and DS II, respectively. The vortices shed by the upstream
cylinder coalesce in the gap region in mode II, suppressing the oscillations of the downstream
cylinder. A reduction in oscillation amplitude of the downstream cylinder can be seen in its
initial branch in Figures 4 (b) and 6(b). When the oscillations for the upstream cylinder
grow further in the upper branch (UB), it sheds more vortices in the gap region, increasing
the oscillation amplitude for the downstream cylinder (see Figures 4b and 6b).
In summary, two single vortices (2S) from top and bottom are shed in Mode I, while
three (P+S) are shed in Mode III. The vortices coalesce between the cylinders in Mode II.
Jou

4.5.2. Downstream of the trailing cylinder


The instantaneous vortex-shedding plots on various response branches of the downstream
cylinder are depicted in Figure 18. The vortices in the wake of the downstream cylinder
coalesce near the cylinder in the desynchronization branches DS I (Figure 18a) and DS II
(Figure 18h). The extent of the region of coalescence of the vortices in DS I increases with U ∗
(Figure 19a-c) before the cylinder jumps to the initial branch at U ∗ = 3.5. Due to the sudden
increase in the oscillation amplitude at U ∗ = 3.5 (onset of lock-in), the flow structure in the
wake changes, and the coalescence of vortices near the cylinder surface is no longer observed

30
Journal Pre-proof

of
pro
re-
Figure 18: (Color online) Instantaneous vorticity contours captured at different U ∗ values (= 2, 3.5, 4, 4.5,
6.5, 7.3, 14, and 18) across multiple response branches, with vorticity measurements ranging from -1 to 1.
Near-wake coalescence of vortices and deflection of wake are evident in Figures (a, c, d, e, g, and h), while
far-wake coalescence is observed in Figure (h).
lP
(Figure 18b). However, in the far field, they travel as a pair. The lost coalescence is found to
take over again when the initial branch closes at U ∗ = 4 (Figure 18c). This is because the flow
in the gap region is in Mode II (see Figure 17c), where the vortices coalesce and suppress the
oscillations (Figure 4b) of the downstream cylinder. The evolution of coalescence is depicted
in Figure (Figure 19d-h). With a shift on the next response branch EIB, the oscillation
rna

amplitude increases, interrupting the vortices’ coalescence near the downstream cylinder. It
is evident in the vorticity plot (Figure 18d) at U ∗ = 4.5 where the width of the wake zone
near the cylinder has increased, compared to the width observed at the closure of IB (Figure
18c). The increased width of the wake is due to the asymmetric wake mode (Mode III) in the
gap region between the cylinders. The arrangement of vortices at U ∗ = 6.5 in Figure 18(e)
is towards the closure of EIB. Note that the flow is no more asymmetric in the gap region,
and the cylinder displacement is close to its maxima (at U ∗ = 6.8). This makes the flow
Jou

vibrant in the wake of the downstream cylinder. As oscillations shift on the lower branch
LB, the flow is more systematic near the cylinder at U ∗ = 7.3 and coalesces in the far field
(Figure 18f). As the system approaches DS II from LB, the oscillation amplitude decreases,
and coalescence of vortices prevails (see Figures 18g and 18h).

4.6. Phase between transverse displacement and lift


VIV involves a closed loop feedback between the oscillator motion and the fluid forces
exerted on the surface of the cylinders. The coupling between the in-line and transverse
response and the corresponding drag and lift coefficients can be depicted from Equations 6
and 7. It can be understood from Equations 6 and 7 that the amplitude of oscillations will

31
Journal Pre-proof

of
pro
Figure 19: (Color online) Illustration of the enhancement in merging distant wake vortices in DS I (at
U ∗ = 2, 3, 3.1 in Figures a−c), and the progression of vortex merging in the vicinity of the cylinder in IB (at
U ∗ = 3.5, 3.6, 3.7, 3.8, and 4 in Figures d−h). Vorticity measurements are scaled from -1 to 1.

largely depend upon both magnitude and alignment of the fluid forces with the direction
of the oscillator’s motion. It is a fact that the phase between transverse response and lift
coefficient change by 180◦ well within the lock-in zone ( Sen and Mittal, 2011). The ‘in-
re-
phase’ to ‘anti-phase’ zones largely decides the ‘Bell-shape’ of the response curve. Figure 20
plots the variation of Y /D − Cl for upstream and downstream diamond cylinders. The in-
phase relation between Y /D and Cl can be depicted by a first-third quadrant orientation of
Y /D − Cl phase portraits. Similarly, a second-fourth quadrant orientation of Y /D − Cl phase
portraits depicts the anti-phase relation. Generally, the phase change between Y /D and Cl
is observed as abrupt. Thus, the Ymax − U ∗ variation (response curve) can be divided into
lP
two Zones. The in-phase zone (from U ∗ = 1 to the U ∗ corresponding to phase change) and
anti-phase zone (after phase change to the end of U ∗ ). Here, the upstream and downstream
rna

Figure 20: (Color online) Phase portraits of Y /D − Cl at different U ∗ values are shown for the upstream
Jou

cylinder (at U ∗ = 4.5, 5.5, 6, 6.5, 7, and 12 in Figures a−f) and downstream cylinders (at U ∗ = 4.5, 5.5,
6, 6.5, 14, and 15 in Figures g−l). The horizontal zero line (green) in each plot represents zero lift. The
relationship between Y and Cl can be identified as in-phase or anti-phase based on the orientation of the
Y − Cl plots in the first-third and second-fourth quadrants.

cylinders bear an in-phase relation between Y /D and Cl till U ∗ = 5.5 (Figure 20a,b and
20g,h)and go out-of-phase at U ∗ = 6.5 (Figure 20d) and continues to be out-of-phase for
the upstream cylinder. For both cylinders, at U ∗ = 6, the inclination of the plot is nearly
parallel to the zero-lift line indicating that the transverse displacement and lift force bear
an angle close to π2 . Since the flow interacting with the downstream cylinder is non-uniform,

32
Journal Pre-proof

it is tedious to determine the exact phase values. However, for U ∗ ≥ U ∗ corresponding to


closure of lock-in, Y /D and Cl are out-of-phase (Figure 20l).
Although the Y /D and Cl are in phase, the low value of U ∗ offers a high spring stiffness
value. As a result, with an increase in U ∗ values, the response increases gradually in DS
I. The oscillation amplitude of the oscillators undergoes a sharp jump with the onset of
lock-in. The transverse response should follow an ever-increasing trend without the phase

of
change from in-phase to anti-phase. However, as the U ∗ enters the anti-phase regime, the
lift forces oppose the transverse response amplitude. Therefore, even when the increase in
U ∗ values further lowers the spring stiffness, the anti-phase relationship between Y /D and
Cl lowers cylinder response. Eventually, the oscillation amplitude stabilizes as the cylinders

pro
go out of lock-in. This explains the bell shape of the response curve.

4.7. Future prospects


Non-circular multi-structure systems are becoming increasingly popular and require sys-
tematically comprehending the intricate interactions between structural oscillation and the
surrounding fluid. To achieve this goal, a comprehensive range of investigations must be
conducted to determine the impact of various factors, including the presence of multiple
structures, mass ratio, damping, and spacing between structures, on the coupled dynamics
re-
of fluid and structural motion in both laminar and turbulent flow regimes. Additionally,
investigating the effects of flow and structural non-linearity on oscillation and flow dynamics
could yield valuable insights.

5. Conclusions
This numerical study investigates the flow-induced vibrations of tandem diamond cylin-
lP
ders, focusing on the distinct response characteristics between the upstream and downstream
cylinders. The results showed that the upstream diamond cylinder displayed an ‘upper
branch’ similar to an isolated diamond cylinder, while the downstream cylinder exhibited an
‘extended initial branch’ and a wider lower branch, resulting in an ‘extended lock-in’ with a
low oscillation frequency. Importantly, this unique behavior is exclusive to tandem diamond
cylinders and is not observed in single diamond or tandem square cylinders. Furthermore,
rna

both diamond cylinders exhibited high amplitude transverse oscillations within a narrow
range of reduced velocities with identical oscillation frequencies along and across the flow,
resulting in single-looped ‘raindrop’ shaped motion trajectories.
The vortex dynamics in the wake of the downstream cylinder were significantly influ-
enced by the impingement of vortices from the upstream cylinder, resulting in distinctive
loading behavior for both cylinders. The downstream cylinder experienced less drag and
approximately twice the lift compared to the upstream cylinder, except for a small range of
Jou

reduced velocities. The region between the cylinders exhibited three distinct vortex modes,
namely Modes I, II, and III. Mode I was observed when the vortex-shedding frequency was
close to or less than the vortex-shedding frequency of a stationary isolated diamond cylinder,
while Mode II resulted from the coalescence of vortices between the cylinders. Mode III was
characterized by the arrangement of three vortices in the gap for a specific range of reduced
velocities.
The study revealed that diamond cylinders exhibited unique response characteristics
compared to square cylinders in tandem configuration, including differences in oscillation
frequencies, amplitudes, and vortex dynamics. The impingement of vortices from the up-
stream cylinder significantly affected the downstream cylinder’s loading behavior, resulting

33
Journal Pre-proof

in distinctive drag and lift characteristics. Further investigations are warranted to better
understand the observed vortical arrangements and their impact on the wake dynamics of
the downstream diamond cylinder.
Data Availability
The data supporting this study’s findings are available from the corresponding author

of
upon reasonable request.
Declaration of competing interest
The authors declare that they have no known competing financial interests or personal
relationships that could have appeared to influence the work reported in this paper.

pro
Appendix A. Derivation of the non-dimensional form of the structural equation
Dimensionless forms of Equations 4 and 5 are obtained by using the following defini-
tions: X ∗ = X
D
, Y∗ = DY
and t∗ = tU
D
. The natural structural frequency (fN ) of the oscillator
is normalized by D and U and is expressed as FN = fNUD . The other basic quantities of
q
k Fx (t) Fy (t)
interest are defined as: ω = 2πfN = m , Cd (t) = 0.5ρU 2 D , Cl (t) = 0.5ρU 2 D and m

= mmd .

re-
Thus, using the above definitions, non-dimensionalization of Equation 4 can be done as:

d2 X
dt2
+ ω2X =
Fx
m
(A.1)
d2 X 0.5ρU 2
DC d
+ (2πfN )2 X = . (A.2)
dt2 m∗ md
lP
∗ d2 X U 2 d2 X ∗
Normalizing X by D and t by U and D, we get dX dt
= U dX
dt∗
and dt2
= D dt∗2
. Substi-
tuting these in Equation A.2 and then rearranging, we obtain:

U 2 d2 X ∗ 2 ∗ 0.5ρU 2 DCd
+ (2πf N ) DX = (A.3)
D dt∗2 m∗ md
rna

2 ∗
dX fN D 2 ∗ ρD2 Cd
=⇒ + (2π ) X = (A.4)
dt∗2 U 2m∗ md
fN D
Substituting FN = U
, we get;

d2 X ∗ 2 ∗ Cd ρD2
+ (2πF N ) X = ( ) (A.5)
dt∗2 2m∗ md
d2 X ∗ Cd
+ (2πFN )2 X ∗ =
Jou

∗2
q. (A.6)
dt 2m∗
2
Here, q = ρD md
attains different values for different geometric cross-sections. For the
2 2
present study, a diamond cylinder is used for which A = D2 and md = ρA = ρ D2 . Thus,
ρD2
q = 0.5ρD 2 = 2. The non-dimensionalization of Equation 5 can also be achieved similarly.

Thus, the non-dimensional form of Equations 4 and 5 (in terms of U ∗ ) for cylinder motion
along and across the flow is given by:

34
Journal Pre-proof

 2
d2 X ∗ 2π Cd
+ X∗ = ∗ , (A.7)
dt∗2 U ∗ m
2 ∗
 2
dY 2π Cl
∗2
+ ∗
Y ∗ = ∗. (A.8)
dt U m

of
In Equations A.7 and A.8, X ∗ and Y ∗ are the non-dimensional displacements of the
cylinder center along and across the flow, respectively. The second-order derivatives of these
instantaneous dimensionless displacements are the instantaneous dimensionless accelerations

pro
of the cylinder. U ∗ is the reduced velocity and m∗ is the mass ratio. The drag and lift
coefficients couple the fluid flow equations with Equations A.7 and A.8.

Appendix B. Convergence and validation studies


B.1 Grid independence analysis
Presented is the mesh convergence of 2-DOF vortex-induced vibration simulations at
a Reynolds number of 100, using a diamond cylinder with a mass ratio of 10 and a fixed

re-
reduced velocity of 5. The mesh employed in the investigation comprises 24149 nodes and
23780 elements, with 60 nodes on each face of the cylinder, and is referred to as M2. To
assess the accuracy of the results, coarse (M1) and fine (M3) meshes with (nodes, elements)
= (14769, 14480) and (42949, 42460), respectively, were compared. Meshes M1 and M3 have
40 and 80 nodes on each face of the cylinder, respectively. Table B.3 compares the flow and
vibration characteristics of the three meshes.
lP
Table B.3: Mesh convergence results: comparison of characteristic flow and vibration quantities obtained on
meshes M1, M2, and M3 for a vibrating diamond cylinder of m∗ = 10 free to oscillate in 2-DOF in uniform
flow at Re = 100 and U ∗ = 5. The percentage deviation of the estimated values for meshes M1 and M3 from
the corresponding values for mesh M2 is shown, denoted by the values in parentheses.

Mesh Nodes Elements NF Ymax /D Clrms Cdavg Fy FCl


M1 14769 14480 40 0.5613 (2.50%) 0.5085 (2.71%) 2.3094 (0.08%) 0.1919(0%) 0.1919 (0%)
rna

M2 24149 23780 60 0.5757 0.4951 2.2902 0.1919 0.1919


M3 42949 42460 80 0.5841 (1.46%) 0.4916 (0.71%) 2.3053 (0.70%) 0.1920 (0.05%) 0.1921 (0.10%)

B.2 Time-step size convergence studies


To examine the adequacy of the time-step size, computations were performed at a
Reynolds number of 100, a mass ratio of 10, and a reduced velocity of 5 using time-steps of
Jou

∆tU
D
= 0.01, 0.02, and 0.03 on mesh M2, which comprises 24149 nodes and 23780 elements.
The results are summarized in Table B.4, which indicates that there is no significant devi-
ation in the properties of flow-induced vibrations (FIV) and flow when ∆tU D
≤ 0.02. This
establishes that ∆tU
D
= 0.02 is sufficient for the present computations on mesh M2. Therefore,
all subsequent computations were conducted using a non-dimensional time-step size of 0.02.
B.3 Validation studies
The validation studies are performed for the following cases: (a) VIV of an isolated
diamond cylinder, (b) flow around a pair of stationary circular cylinders in tandem, (c) flow
around a pair of stationary square cylinders in tandem, and (d) VIV of a pair of square
cylinders in tandem.

35
Journal Pre-proof

Table B.4: Time-step size convergence results: effect of the normalized time-step size ∆tU
D on characteristic
flow and vibration quantities of a freely oscillating isolated diamond cylinder (m∗ = 10) along and across
the flow at Re = 100 and U ∗ = 5.

∆tU
D
Ymax /D Clrms Cdavg Fy FCl
0.01 0.5778 0.4898 2.2920 0.1920 0.1920

of
0.02 0.5757 0.4951 2.2902 0.1919 0.1919
0.03 0.5792 0.5003 2.2904 0.1919 0.1919

pro
(a) VIV of an isolated diamond cylinder
The accuracy of the VIV computations for an isolated cylinder is done for a diamond
cylinder. The computed results are compared with those by Zhao et al. (2013) at Re =
100 and U ∗ = 5. The results are tabulated in Table B.5. It is apparent from the table that
our results are well within the range of values predicted by Zhao et al. (2013). Thus, there
is a satisfactory agreement of the projected response data for a diamond oscillator. This
comparison is also shown in Sourav et al. (2022).

Authors
re-
Table B.5: Comparison of Ymax /D at Re = 100 and U ∗ = 5 with those reported by Zhao et al. (2013) for
an isolated diamond cylinder of mass ratio = 3 and executing oscillations along and across the flow.

Blockage Ymax /D
Zhao et al. (2013) 0.0354 0.6183 (extracted from graph)
Present 0.025 0.6453
lP
Next, the solver is validated for tandem stationery and vibrating circular and square
cross-section cylinders.
(b) Flow around a pair of stationary circular cylinders in tandem
The accuracy of the computed results for stationary tandem circular cylinders sep-
rna

S
arated by a distance ( D ) of 4 is compared with those of Sharman et al. (2005). The
computations were performed at Re = 100 and blockage (B) = 0.02. The table below (Ta-
ble B.6) compares the time-averaged drag coefficient and r.m.s. of the lift coefficient. It is
evident from Table B.6 that there is an excellent match between the computed results and
the results by Sharman et al. (2005).

Table B.6: Unsteady flow past a pair of stationary in-line circular cylinders of identical size at Re = 100:
S
comparison of predicted integral parameters for D = 4 with those reported by Sharman et al. (2005).
Jou

Study S/D B Re
Upstream cylinder Downstream cylinder
Cd Clrms Cd Clrms
Sharman et al. (2005) 4 0.02 100 1.2752 0.3020 0.7035 0.9892
Present 4 0.02 100 1.2592 0.3020 0.6992 0.9705

(c) Flow around a pair of stationary square cylinders in tandem


S
At D = 5 and Re = 100, the validity of the solver is checked for stationary tandem square
cylinders. The results from Bao et al. (2012) compare the mean drag, r.m.s. lift and the

36
Journal Pre-proof

Table B.7: Unsteady flow past a pair of stationary in-line square cylinders of identical size at Re = 100:
S
comparison of predicted integral parameters for D = 5 with those reported by Bao et al. (2012).

Study B
Upstream cylinder Downstream cylinder
Cd Clrms St Cd Clrms St
Bao et al. (2012) 0.02 1.426 0.289 0.130 1.099 1.211 0.130

of
Present 0.02 1.3910 0.2901 0.128 1.0956 1.206 0.128

vortex-shedding frequency (Strouhal number St). Table B.7 establishes the excellent match

pro
between the two sets of results.
(d) VIV of a pair of square cylinders in tandem
For an elastically mounted square cylinder in the wake of a stationary downstream square
cylinder separated by a distance of 5D, Figure B.21 compares the transverse oscillation
amplitudes for the present and Han et al. (2018). For both sets of simulations, the Re was
fixed at 80, and U ∗ was varied between 3 and 18. It can be observed that there is an excellent
match of the results with that of Han et al. (2018). This figure is also shown by Kumar and
Sen (2021).

0.8
re-
0.6
lP
Ymax/D

0.4

0.2
Han et al. (2018)
rna

Present
0.0
2 4 6 8 10 12 14 16 18
U*
Figure B.21: Comparison of normalized transverse response in 2−DOF wake-induced vibrations of a square
cylinder obtained by Han et al. (2018) with the present results. The parameters used for the computations
S
are: m∗ = 2; Re = 100; B = 0.0166; and D = 5.
Jou

Thus, we establish here that the solver is well-validated for many problems. It can
efficiently handle the problems of both isolated and tandem stationary/ vibrating cylinders
of different shapes.

References
[1] Aswathy, M.S. and Sarkar, S., 2019. Effect of stochastic parametric noise on vortex
induced vibrations. International Journal of Mechanical Sciences, 153, 103–118.

37
Journal Pre-proof

[2] Bao, Y., Wu, Q. and Zhou, D., 2012. Numerical investigation of flow around an inline
square cylinder array with different spacing ratios. Computers & Fluids, 55, 118–131.

[3] Bearman, P.W., 1984. Vortex shedding from oscillating bluff bodies. Annual Review of
Fluid Mechanics, 16, 195–212.

of
[4] Bernitsas, M.M., Raghavan, K., Ben-Simon, Y. and Garcia, E.M.H., 2008. VIVACE (Vor-
tex Induced Vibration Aquatic Clean Energy): A new concept in generation of clean and
renewable energy from fluid flow. Journal of Offshore Mechanics and Arctic Engineering,
130, 041101.

pro
[5] Bhatt, R. and Alam, M.M., 2018. Vibrations of a square cylinder submerged in a wake.
Journal of Fluid Mechanics, 853, 301–332.

[6] Borazjani, I. and Sotiropoulos, F., 2009. Vortex-induced vibrations of two cylinders in
tandem arrangement in the proximity–wake interference region. Journal of Fluid Mechan-
ics, 621, 321–364.

[7] Carmo, B.S., Sherwin, S.J., Bearman, P.W. and Willden, R.H.J., 2011. Flow-induced
vibration of a circular cylinder subjected to wake interference at low Reynolds number.
re-
Journal of Fluids and Structures, 27, 503–522.

[8] Chatterjee, D. and Mondal, B., 2012. Forced convection heat transfer from tandem square
cylinders for various spacing ratios. Numerical Heat Transfer, Part A: Applications, 61,
381–400.

[9] Ding, L., Zhang, L., Bernitsas, M.M. and Chang, C.C., 2016. Numerical simulation and
lP
experimental validation for energy harvesting of single-cylinder VIVACE converter with
passive turbulence control. Renewable Energy, 85, 1246–1259.

[10] Dorogi, D. and Baranyi, L., 2020. Identification of upper branch for vortex-induced
vibration of a circular cylinder at Re= 300. Journal of Fluids and Structures, 98, 103135.
rna

[11] Etminan, A., Moosavi, M. and Ghaedsharafi, N., 2011. Determination of flow configu-
rations and fluid forces acting on two tandem square cylinders in cross-flow and its wake
patterns. International journal of mechanics, 5(2), 63–74.

[12] Gu, W., Xu, X., Wei, N., Yao, W., Yu, G., Lian, X., Gao, J. and Zhou, J., 2022. The
space effect on WIV interference between a fixed and oscillating diamond cylinder at a
low Reynolds number of 100. Ocean Engineering, 264, 112428.

[13] Gu, W., Xu, X., Jiang, M. and Yao, W., 2023. Effects of spacing ratio on vortex-induced
Jou

vibration of twin tandem diamond cylinders in a steady flow. Physics of Fluids, 35, 043604.

[14] Haider, B.A. and Sohn, C.H., 2018. Effect of spacing on a pair of naturally oscillating
circular cylinders in tandem arrangements employing IB-LB methods: Crossflow-induced
vibrations. International Journal of Mechanical Sciences, 142, 74–85.

[15] Han, Z., Zhou, D., Malla, A., Nepali, R., Kushwaha, V., Li, Z., Kwok, K.C., Tu, J. and
Bao, Y., 2018. Wake-induced vibration interference between a fixed square cylinder and
a 2-DOF downstream square cylinder at low Reynolds numbers. Ocean Engineering, 164,
698–711.

38
Journal Pre-proof

[16] Inoue, O., Mori, M. and Hatakeyama, N. 2006. Aeolian tones radiated from flow past
two square cylinders in tandem. Physics of Fluids, 18, 046101.

[17] Jaiman, R.K., Pillalamarri, N.R. and Guan, M.Z., 2016. A stable second-order par-
titioned iterative scheme for freely vibrating low-mass bluff bodies in a uniform flow.
Computer Methods in Applied Mechanics and Engineering, 301, 187–215.

of
[18] Jauvtis, N. and Williamson, C. H. K., 2004. The effect of two degrees of freedom on
vortex-induced vibration at low mass and damping. Journal of Fluid Mechanics, 509,
23–62.

pro
[19] Kang, Z., Ni, W. and Sun, L., 2016. An experimental investigation of two-degrees-of-
freedom VIV trajectories of a cylinder at different scales and natural frequency ratios.
Ocean engineering, 126, 187–202.

[20] Khalak, A. and Williamson, C.H.K., 1996. Dynamics of hydroelastic cylinder with very
low mass and damping. Journal of Fluids and Structures, 10, 455–472.

[21] Kumar, D., Singh, A.K. and Sen, S., 2018a. Identification of response branches for

616–627. re-
oscillators with curved and straight contours executing VIV. Ocean Engineering, 164,

[22] Kumar, D., Sourav, K., Sen, S. and Yadav, P.K., 2018b. Steady separation of flow from
an inclined square cylinder with sharp and rounded base. Computers & Fluids, 171, 29–40.

[23] Kumar, D., Mittal, M. and Sen, S., 2018c. Modification of response and suppression of
lP
vortex-shedding in vortex-induced vibrations of an elliptic cylinder. International Journal
of Heat and Fluid Flow, 71, 406–419.

[24] Kumar, D., Sourav, K., Yadav, P.K. and Sen, S., 2019. Understanding the secondary
separation from an inclined square cylinder with sharp and rounded trailing edges. Physics
of Fluids, 31, 073607.
rna

[25] Kumar, D., Sourav, K. and Sen, S., 2019a. Steady separated flow around a pair of
identical square cylinders in tandem array at low Reynolds numbers. Computers & Fluids,
191, 104244.

[26] Kumar, D. and Sen, S., 2021. Flow-induced vibrations of a pair of in-line square cylin-
ders. Physics of Fluids, 33, 043602.

[27] Li, X., Lyu, Z., Kou, J. and Zhang, W., 2019. Mode competition in galloping of a square
Jou

cylinder at low Reynolds number. Journal of Fluid Mechanics, 867, 516–555.

[28] Meneghini, J. R., Saltara, F., Siqueira, C. L. R. and Ferrari, J. A., 2001. Numerical
simulation of flow interference between two circular cylinders in tandem and side-by-side
arrangements. Journal of Fluids and Structures, 15, 327–350.

[29] Mishra, S.K., Yadav, P.K., Sarkar, H. and Sen, S., 2022. Correspondence between the
number of no-slip critical points and nature of rear stagnation point of a symmetric object.
Physics of Fluids, 34, 111702.

39
Journal Pre-proof

[30] Mittal, S. and Tezduyar, T.E., 1992. Notes on the stabilized space-time finite-element
formulation of unsteady incompressible flows. Computer physics communications, 73, 93–
112.

[31] Mittal, S. and Tezduyar, T.E., 1994. Massively parallel finite element computation
incompressible flows involving fluid-body interactions. Computer Methods in Applied Me-

of
chanics and Engineering, 112, 253–282.

[32] Mittal, S., Kumar, V. and Raghuvanshi, A., 1997. Unsteady incompressible flows past
two cylinders in tandem and staggered arrangements. International Journal for Numerical
Methods in Fluids 25, 1315–1344.

pro
[33] Mysa, R.C., Kaboudian, A. and Jaiman, R.K., 2016. On the origin of wake-induced
vibration in two tandem circular cylinders at low Reynolds number. Journal of Fluids and
Structures 61, 76–98.

[34] Navrose, Yogeswaran, V., Sen, S. and Mittal, S., 2014. Free vibrations of an elliptic
cylinder at low Reynolds numbers. Journal of Fluids and Structures, 51, 55–67.

re-
[35] Nemes, A., Zhao, J., Jacono, D.L. and Sheridan, J., 2012. The interaction between flow-
induced vibration mechanisms of a square cylinder with varying angles of attack. Journal
of Fluid Mechanics, 710, 102–130.

[36] Nepali, R., Ping, H., Han, Z., Zhou, D., Yang. H., Tu, J., Zhao, Y. and Bao , Y.,
2020. Two-degree-of-freedom vortex-induced vibrations of two square cylinders in tandem
arrangement at low Reynolds numbers. Journal of Fluids and Structures, 97, 102991.
lP
[37] Prasanth, T.K. and Mittal, S., 2008. Vortex-induced vibrations of a circular cylinder at
low Reynolds numbers. Journal of Fluid Mechanics, 594, 463–491.

[38] Prasanth, T.K. and Mittal, S., 2009a. Flow-induced oscillation of two circular cylinders
in tandem arrangement at low Re. Journal of Fluids and Structures, 25(6), 1029–1048.
rna

[39] Prasanth, T.K. and Mittal, S., 2009b. Vortex-induced vibration of two circular cylinders
at low Reynolds number. Journal of Fluids and Structures, 25(4), 731–741.

[40] Qiu, T., Lin, W., Du, X. and Zhao, Y., 2021b. Mass ratio effect on vortex-induced
vibration for two tandem square cylinders at a low Reynolds number. Physics of Fluids,
33, 123604.

[41] Qiu, T., Zhao, Y., Du, X. and Lin, W., 2021a. Spacing effect on the two-degree-of-
Jou

freedom VIV of two tandem square cylinders. Ocean Engineering, 236, 109519.

[42] Saad, Y. and Schultz, M.H., 1986. GMRES: A generalized minimal residual algorithm
for solving nonsymmetric linear systems. SIAM Journal on Scientific and Statistical Com-
puting, 7, 856–869.

[43] Sarpkaya, T., 2004. A critical review of the intrinsic nature of vortex-induced vibrations.
Journal of Fluids and Structures, 19, 389–447.

[44] Sen, S. and Mittal, S., 2011. Free vibration of a square cylinder at low Reynolds numbers.
Journal of Fluids and Structures, 27, 875–884.

40
Journal Pre-proof

[45] Sen, S. and Mittal, S., 2015. Effect of mass ratio on free vibrations of a square cylinder
at low Reynolds numbers. Journal of Fluids and Structures, 54, 661–678.

[46] Seyed-Aghazadeh, B., Carlson, D.W. and Modarres-Sadeghi, Y., 2017. Vortex-induced
vibration and galloping of prisms with triangular cross-sections. Journal of Fluid Mechan-
ics, 817, 590–618.

of
[47] Sharman, B., Lien, F.S., Davidson, L. and Norberg, C., 2005. Numerical predictions of
low Reynolds number flows over two tandem circular cylinders. International Journal for
Numerical Methods in Fluids, 47(5), 423–447.

pro
[48] Sohankar, A. and Etminan, A., 2009. Forced-convection heat transfer from tandem
square cylinders in cross flow at low Reynolds numbers. International Journal for Numer-
ical Methods in Fluids, 60, 733–751.

[49] Sohankar, A., 2012. A numerical investigation of the flow over a pair of identical square
cylinders in a tandem arrangement. International Journal for Numerical Methods in Fluids,
70, 1244–1257.

re-
[50] Sourav, K. and Sen, S., 2017. On the response of a freely vibrating thick elliptic cylinder
of low mass ratio. Journal of Applied Fluid Mechanics, 10, 899–913.

[51] Sourav, K. and Sen, S., 2019. Transition of VIV-only motion of a square cylinder to
combined VIV and galloping at low Reynolds numbers. Ocean Engineering, 187, 106208.

[52] Sourav, K. and Sen, S., 2020. Determination of the transition mass ratio for onset of
lP
galloping of a square cylinder at the least permissible Reynolds number of 150. Physics of
Fluids, 32, 063601.

[53] Sourav, K., Kumar, D. and Sen, S., 2020. Undamped transverse-only VIV of a diamond
cylinder at low Reynolds numbers. Ocean Engineering, 197, 106867.

[54] Sourav, K., Yadav, P.K., Tallapragada, P. and Kumar, D., 2022. Simultaneous stream-
rna

wise and cross-stream oscillations of a diamond oscillator at low Reynolds numbers. Physics
of Fluids, 34, 063601.

[55] Sumner, D., 2010. Two circular cylinders in cross-flow: A review. Journal of Fluids and
Structures, 26, 849-899.

[56] Tamimi, V., Esfehani, M.J., Zeinoddini, M., Naeeni, S.T.O., Wu, J. and Shahvaghar-
Asl, S., 2020. Marine hydrokinetic energy harvesting performance of diamond and square
Jou

oscillators in tandem arrangements. Energy, 202, 117749.

[57] Tang, B., Fan, X., Wang, J. and Tan, W., 2022. Energy harvesting from flow-induced
vibrations enhanced by meta-surface structure under elastic interference. International
Journal of Mechanical Sciences, 236, 107749.

[58] Tatsutani, K., Devarakonda, R. and Humphrey, J.A.C., 1993. Unsteady flow and heat
transfer for cylinder pairs in a channel. International Journal of Heat and Mass Transfer,
36, 3311-3328.

41
Journal Pre-proof

[59] Tezduyar, T.E., Behr, M. and Liou, J., 1992a. A new strategy for finite element compu-
tations involving moving boundaries and interfaces—the deforming-spatial-domain/space-
time procedure: I. The concept and the preliminary numerical tests, 94, 339–351.

[60] Tezduyar, T.E., Behr, M., Mittal, S. and Liou, J., 1992b. A new strategy for finite
element computations involving moving boundaries and interfaces—the deforming-spatial-

of
domain/space-time procedure: II. Computation of free-surface flows, two-liquid flows, and
flows with drifting cylinders. Computer methods in applied mechanics and engineering,
94, 353–371.

[61] Tezduyar, T.E., Mittal, S., Ray, S.E. and Shih, R., 1992c. Incompressible flow com-

pro
putations with stabilized bilinear and linear equal-order-interpolation velocity-pressure
elements. Computer Methods in Applied Mechanics and Engineering, 95, 221–242.

[62] Wang, J.S., Fan, D. and Lin, K., 2020. A review on flow-induced vibration of offshore
circular cylinders. Journal of Hydrodynamics, 32, 415–440.

[63] Wu, X., Ge, F. and Hong, Y., 2012. A review of recent studies on vortex-induced
vibrations of long slender cylinders. Journal of Fluids and structures, 28, 292–308.
re-
[64] Xu, X., Gu, W. and Yao, W., 2021. Numerical simulation of the VIV of twin tandem
diamond cylinders at low Reynolds numbers. Ocean Engineering, 238, 109745.

[65] Yadav, P.K., Sourav, K., Kumar, D. and Sen, S., 2021. Flow around a diamond-section
cylinder at low Reynolds numbers. Physics of Fluids, 33, 053611.
lP
[66] Yadav, P.K., Sharma, S. and Sen, S., 2023. Damped flow-induced vibrations of a square
cylinder at low Reynolds numbers. Journal of Flow Visualization and Image Processing,
30, 87–113.

[67] Yang, G. and Wu, J., 2013. Effect of side ratio and aiding/opposing buoyancy on
the aerodynamic and heat transfer characteristics around a rectangular cylinder at low
Reynolds numbers. Numerical Heat Transfer, Part A: Applications, 64(12), 1016–1037.
rna

[68] Zaki, T.G., Sen, M. and Gad-el-hak, M., 1994. Numerical and experimental investigation
of flow past a freely rotatable square cylinder. Journal of Fluids and Structures, 8, 555–582.

[69] Zdravkovich, M.M., 1977. Review of flow interference between two circular cylinders in
various arrangements. Journal of Fluids Engineering, 99, 618–633.

[70] Zdravkovich, M.M., 1985. Flow induced oscillations of two interfering circular cylinders.
Jou

Journal of Sound and Vibration, 101, 511–521.

[71] Zhang, B., Mao, Z., Song, B., Ding, W. and Tian, W., 2018. Numerical investigation on
effect of damping-ratio and mass-ratio on energy harnessing of a square cylinder in FIM.
Energy, 144, 218–231.

[72] Zhang, H., Zhou, L. and Tim, K.T., 2022. Mode-based energy transfer analysis of
flow-induced vibration of two rigidly coupled tandem cylinders. International Journal of
Mechanical Sciences, 228, 107468.

42
Journal Pre-proof

[73] Zhao, G.M, Pillalamarri, N.R., Mysa, R.C. and Jaiman, R.K., 2015, May. Flow-induced
vibrations of single and tandem square columns. In International Conference on Offshore
Mechanics and Arctic Engineering (56482, V002T08A053). American Society of Mechan-
ical Engineers.

[74] Zhao, M., Cheng, L. and Zhou, T., 2013. Numerical simulation of vortex-induced vibra-

of
tion of a square cylinder at a low Reynolds number. Physics of Fluids, 25, 023603.

[75] Zhao, M., 2013. Flow induced vibration of two rigidly coupled circular cylinders in
tandem and side-by-side arrangements at a low Reynolds number of 150. Physics of Fluids,
25, 123601.

pro
[76] Zhao, X., Cheng, D., Zhang, D. and Hu, Z., 2016. Numerical study of low-Reynolds-
number flow past two tandem square cylinders with varying incident angles of the down-
stream one using a CIP-based model. Ocean Engineering, 121, 414–421.

[77] Zhou, L., Li, H., Tim, K.T., He, X., Maceda, G.Y.C. and Zhang, H., 2023. Sensitivity-
aided active control of flow past twin cylinders. International Journal of Mechanical Sci-
ences, 242, 108013.
re-
[78] Zhou, Y. and Alam, M.M., 2016. Wake of two interacting circular cylinders: A review.
International Journal of Heat and Fluid Flow, 62, 510–537.
lP
rna
Jou

43
Journal Pre-proof

Highlights

of
• Extended lock-in range for the downstream diamond cylinder.

pro
• Raindrop-shaped cylinder trajectory for both cylinders.
• Asymmetric mean lift for the upstream cylinder.
• Three modes of vortex arrangement in the gap region.

re-
lP
rna
Jou
Journal Pre-proof

of
pro
Vortex−Induced Vibrations of Tandem Diamond Cylinders: A Novel Lock−in Behavior

Problem Setup Branching Behavior and Extended lock−in Asymmetric fluid loading

Upstream cylinder U* = 4.7 Upstream cylinder


σ xx = 0,

IB LB
σ xy = 0

0.30 0.64 3.5


Ymax/D 3.0 Cl Cd

Fy
FN 1.5

Cl, Cd
0.26 0.48
10 D 10 D
re- 0.0

Ymax/D
-1.5
Fy

0.22 0.32 360 380 400 420 430


tU/D

DSI U* = 5
100 D

3.5
0.18 DSII 0.16 Cl Cd

UB

Cl, Cd
1.5

0.14 0.00
0.0
σ xy = 0, v = 0
σ xy = 0, v = 0

0 4 8 12 16 20
lP
-1.0
*
k

U 600 620 640 650


tU/D
.

Downstream cylinder
u = X, v = Y

0.28 IB 1.2 U* = 4.7 Downstream cylinder


0.5
k

Ymax/D Cl Cd
5D
.

0.3
Fy Cl, Cd

0.22 0.9
x

FN 0.0
Re = 100, m* = 10

Ymax/D
O

Fy

0.16 0.6 -0.3


rna

360 380 400 420


k

tU/D
DSI EIB LB DSII U* = 5
25 D

0.8
0.10 0.3 Cl Cd
0.6
Cl, Cd

0.3
0.04 0.0
0.0
0 4 8 12 16 20
u=U

* -0.3
v=0

U 600 620 640 660


tU/D

Deepak Kumar and Kumar Sourav


Jou
Journal Pre-proof

CRediT authorship contribution statement

Deepak Kumar: Conceptualization (equal); Data curation (equal), Formal analysis (equal),

of
Investigation (equal), Methodology (equal), Validation (equal), Visualization (equal), Writing – original
draft (equal), Writing – review & editing (equal). Kumar Sourav: Conceptualization (equal); Data
curation (equal), Formal analysis (equal); Investigation (equal), Methodology (equal), Validation
(equal), Visualization (equal), Writing – original draft (equal), Writing – review & editing (equal)

pro
re-
lP
rna
Jou
Journal Pre-proof

Declaration of interests

☒ The authors declare that they have no known competing financial interests or
personal relationships that could have appeared to influence the work reported in
this paper.

of
☐ The authors declare the following financial interests/personal relationships which
may be considered as potential competing interests:

pro
re-
lP
rna
Jou

View publication stats

You might also like