Download as pdf or txt
Download as pdf or txt
You are on page 1of 36

J. Fluid Mech. (2022), vol. 952, A26, doi:10.1017/jfm.2022.

893

Vortex-induced vibration of a two


degree-of-freedom flexibly mounted circular
cylinder in the crossflow direction

Seyedmohammad Mousavisani1, ‡, Naumi Noshin Chowdhury1, ‡,


Hadi Samsam-Khayani1 , Hamed Samandari1 and
Banafsheh Seyed-Aghazadeh1, †
1 Department of Mechanical Engineering, University of Massachusetts, Dartmouth, MA 02747, USA

(Received 26 April 2022; revised 30 August 2022; accepted 9 October 2022)

Vortex-induced vibration (VIV) of a two degree-of-freedom (DOF) circular cylinder,


placed in the test section of a recirculating water tunnel and free to oscillate in its
first two vibrational modes in the crossflow direction, is studied experimentally. The
dynamic response of the cylinder is studied for a reduced velocity range of U ∗ = 4–30
for eigenfrequency ratios in the range of 1.3–3.0. For the two DOF system, while the onset
of the VIV response followed a similar lock-in region as those observed for a classical
VIV response of a single DOF system, by increasing the reduced velocity a secondary
lock-in region was observed over which the oscillations of the cylinder were locked into
the system’s second mode. In addition, there existed an intermediate range of reduced
velocity over which the VIV response consisted of oscillations at a combination of the first
https://doi.org/10.1017/jfm.2022.893 Published online by Cambridge University Press

two natural modes of the system. As the eigenfrequency ratio between the first two modes
increased, the secondary lock-in range was extended to higher reduced velocities and the
reduced velocity range over which multi-modal oscillations were observed was decreased.
A full map of vortex dynamics in the wake of the cylinder was developed qualitatively
and quantitatively using hydrogen bubble flow visualization and time-resolved volumetric
particle tracking velocimetry techniques, respectively. A Q-criterion analysis revealed
the existence of highly three-dimensional vortex structures in the wake of the cylinder.
The spatiotemporal mode analysis using the proper orthogonal decomposition technique
revealed strong coupling between the vortex shedding modes in the wake of the cylinder
and the structural vibration modes.
Key words: vortex dynamics, vortex shedding

† Email address for correspondence: b.aghazadeh@umassd.edu


‡ These authors contributed equally to this work.
© The Author(s), 2022. Published by Cambridge University Press 952 A26-1
S. Mousavisani and others
1. Introduction
The vortex-induced vibration (VIV) response of various systems have been investigated
in early studies by modelling the system as a flexibly mounted rigid cylinder placed in
a flow, free to oscillate with one degree-of-freedom (DOF) in the transverse (crossflow)
direction (Bearman 1984; Khalak & Williamson 1999; Govardhan & Williamson 2002;
Sarpkaya 2004; Raghavan & Bernitsas 2011). The complex dynamic response of the
structures coupled with their surrounding flow has been investigated to characterize the
fundamentals of the fluid–structure interaction response for such systems. In many real
world applications, such as mooring lines used to stabilize the offshore oil platforms
and the conceptual designs for offshore wind turbines, VIV is observed in long flexible
cylinders, where oscillations are possible in both the crossflow and streamwise (inline)
directions. Later studies have investigated the influence of the second DOF in the inline
direction on the overall VIV response of a circular cylinder (Sarpkaya 1995; Jauvtis &
Williamson 2004; Dahl, Hover & Triantafyllou 2006; Dahl et al. 2007). These studies
have demonstrated that giving the structure a second DOF in the inline direction can
change the dynamic response of the system, significantly. In an experimental study on
a two DOF VIV response of a circular cylinder (Dahl et al. 2007), the ratio between
the crossflow to inline natural frequency of the system was altered systematically to
study the impact of the system’s frequency ratios on the overall VIV response of the
system. It was concluded that under perfect vortex shedding resonance, in which the inline
natural frequency of the system is twice the crossflow natural frequency, a fundamental
change in both the orbital motion of the structure and the frequency content of the fluid
forces occurs. The VIV response of the system was observed to be dominated by a
complex vortex wake, resulting in high frequency lift forces. In VIV applications such
large contributions of higher harmonics of the lift force could consequently contribute to
the fatigue life of structures undergoing large amplitude VIV. Additionally, flow-induced
vibration of a circular cylinder, placed in the wake of a stationary upstream cylinder, with
two DOF in both the inline and crossflow directions has been studied (Chaplin & Batten
2014). The results suggested that the presence of both the inline and crossflow DOF, and
additionally the presence of higher modes can significantly influence some features of the
fluid–structure interactions when compared with those of an isolated cylinder with one
DOF.
https://doi.org/10.1017/jfm.2022.893 Published online by Cambridge University Press

While VIV studies of elastically mounted rigid cylinders have facilitated our
understanding of VIV occurring in a more complex case of flexible cylinders, the VIV
response of a flexible cylinder can undergo complex large amplitude oscillations at higher
vibrational modes, due to the additional variables associated with dynamics of flexible
systems. Experimental (Chaplin et al. 2005; Trim et al. 2005; Vandiver et al. 2005; Lie
& Kaasen 2006; Huera-Huarte & Bearman 2009; Seyed-Aghazadeh & Modarres-Sadeghi
2016; Seyed-Aghazadeh, Edraki & Modarres-Sadeghi 2019), and numerical (Evangelinos,
Lucor & Karniadakis 2000; Bourguet, Karniadakis & Triantafyllou 2011a; Bourguet et al.
2011b; Bourguet, Karniadakis & Triantafyllou 2013; Zanganeh & Srinil 2016) studies have
well documented the VIV response of a flexible cylinder, placed perpendicular to the flow.
Excitation of mono- and multi-frequency modes and transition from lower modes to higher
modes excitation have been observed in the VIV response of such flexible cylinders. In
addition, a travelling wave observed in the VIV response of flexible cylinders (Marcollo
et al. 2011; Bourguet, Lucor & Triantafyllou 2012; Seyed-Aghazadeh et al. 2021) adds to
the complexity of the response that cannot be characterized in simpler models of a flexible
cylinder, i.e. flexibly mounted cylinder studies with one or two DOF.

952 A26-2
VIV of a two DOF circular cylinder in the crossflow direction
Most VIV studies on flexible cylinders have considered the flexible cylinder to be a
tension-dominated one. If the flexible tension-dominated cylinder can be modelled as a
string under oscillations in fluid flow, the ratio between the system’s natural frequencies
in both the crossflow and inline directions is approximately multiple integers, i.e. the
second natural frequency will have values twice the fundamental frequency, the third
natural frequency will be thrice the fundamental frequency and so on. In a recent study, the
VIV response of a bending-dominated cylinder has been studied experimentally (Gedikli,
Chelidze & Dahl 2018), in which the ratio between the system’s natural frequencies was
designed to take different values in each directions. It was concluded that independent
from the natural frequency ratios in the crossflow and inline directions, the inline response
is a forced response that depends heavily on the crossflow response of the system. The
VIV response of the cylinder in the inline direction was found to be unlikely to oscillate
with an even mode shape due to symmetric drag loading, even when the system was
tuned to have an even mode to be excited at the expected frequency of vortex shedding.
This observation has been different from those reported for a tension-dominated flexible
cylinder for which both odd and even modes have been excited in the crossflow and inline
directions (Seyed-Aghazadeh & Modarres-Sadeghi 2016; Seyed-Aghazadeh et al. 2019).
The question that arises here is what if the flexible cylinder cannot be modelled as either
a tension-dominated or bending-dominated one and the frequency ratios in each direction
have non-integer values as one would expect to see for a beam under tension? How
the VIV response of the system will be compared with those of a tension-dominated or
bending-dominated one? Our understanding of the complex multi-modal VIV behaviour
in such flexible cylinders can be well improved if a study can be designed to accommodate
for a systematic and controlled change in the system’s natural frequencies and structural
excitation at a single or multiple modes higher than one, while some complex features of
the response, such as travelling wave behaviour, can be limited or entirely eliminated. This
is the goal in our current study.
Recent experimental studies have investigated the VIV in a dual mass system, where
the oscillating cylinder subjected to the incoming flow has two DOF to oscillate in the
crossflow direction (Nishi 2013; Nishi, Fukuda & Shinohara 2017). These studies have
investigated the advantages of using such a dual mass system in expanding the flow
velocity range at which large amplitude oscillations could occur. These studies have shown
the selective advantages of such systems in energy harvesting over a wider excitation range
https://doi.org/10.1017/jfm.2022.893 Published online by Cambridge University Press

in the VIV response of the system; the excitation was first observed for the flow velocity
range at which the oscillation frequency locked into the first eigenfrequency of the system,
and later shifted to a limited range of velocities over which lock-in occurred at frequencies
close to the second eigenfrequency of the system. However, the experimental apparatus
design in these studies limited the upper ranges of flow velocities at which the system could
be tested and, therefore, the results were limited to VIV oscillations observed up to only the
onset of excitation at the second eigenfrequencies. Similarly, in a recent theoretical study
(Vicente-Ludlam, Barrero-Gil & Velazquez 2015), the galloping response of a dual mass
system has been studied using the quasi-steady assumption. Improvements in the energy
extraction efficiencies have been studied by tuning the mechanical system properties of
the system.
While the recent studies on the VIV of a dual mass system, free to oscillate with
two DOF in the crossflow direction, have shed light on their potentials to be used in
fluidic energy harvester designs to extend their operating range, there is still a gap in
our understanding of the fundamentals of VIV in multiple DOF systems, in particular
when (a) there is no limitation in the experimental set-up to prevent the VIV response
952 A26-3
S. Mousavisani and others
to cover excitation at both the first and second eigenfrequencies of the system, and
(b) the eigenfrequency ratios can be systematically varied to study a wide range of
VIV response occurrences, from mono-frequency excitation at each eigenfrequency to a
multi-modal excitation over overlapping flow velocities, over which interaction between
the excited modes can be studied. These goals will be achieved in our current study
where VIV of a two DOF elastically mounted rigid cylinder, free to oscillate in the
crossflow direction at its first two eigenfrequencies, is studied experimentally. We have
developed a unique experimental set-up that would realize the study of the VIV response
of a two DOF elastically mounted circular cylinder at both its eigenmodes. The two
eigenfrequencies of the system include one at the lower first mode and one at the higher
second mode. The cylinder in our study was mounted on an air-bearing set-up, allowing
frictionless motion of the cylinder in the crossflow direction, and placed vertically in the
test section of a recirculating water tunnel. The unique air-bearing set-up design allows
for testing a variation of eigenfrequency ratios of the system, which in return can facilitate
understanding of the physical mechanisms responsible for the VIV of multi-DOF cylinders
when excited at higher vibrational modes. The dynamic response of the cylinder is studied
for a range of flow velocities in the range of U = 0.04–0.31 m s−1 , corresponding to a
Reynolds number range of Re = 1033–7478.
In what follows, § 2 describes the experimental set-up and methodology used to capture
the VIV response of the system. Section 3 presents theoretical modal analysis of the
forced vibration of a system with two DOF and describes the system’s eigenfrequencies
and eigenmodes. The results presented in § 4 describe the overall VIV response of the
system over the entire range of reduced velocities tested for a range of eigenfrequency
ratios of the system. Section 4.1 describes the amplitudes and frequencies of oscillations as
well as response phase evolution over a wide range of flow velocities and eigenfrequency
ratios. Section 4.2 presents the flow forces acting on the cylinder. Section 4.3 presents
the qualitative and quantitative wake structure at a sample eigenfrequency ratio for the
entire range of flow velocities previously tested. A full map of vortex shedding patterns
is presented at a sample eigenfrequency ratio of the system. Finally, § 4.4 presents the
prevalent shedding modes in the wake of the cylinder undergoing VIV using the technique
of proper orthogonal decomposition (POD).
https://doi.org/10.1017/jfm.2022.893 Published online by Cambridge University Press

2. Experimental set-up and data collection


2.1. Structural response measurements
The experiments were conducted in a recirculating water tunnel, with a test section of
0.45 m × 0.45 m × 1.5 m and a turbulence intensity of less than 1 % for up to a velocity of
1 m s−1 . A uniform circular cylinder with a diameter of D = 21.6 mm, immersed length
of L = 250 mm and an aspect ratio of L/D = 11.57 was tested. The mass ratio of the
cylinder was calculated to be m∗ = 4 m/πρD2 L = 21.1, where m is total moving mass
of the system and ρ is the flow density. The cylinder was attached to an air-bearing
set-up, allowed to oscillate in the transverse (crossflow) direction. A schematic of the
set-up is shown in figure 1. The air-bearing set-up had eight air bearings, each pair
mounted on ceramic coated aluminium shafts running in parallel, to reduce the structural
damping in the system. Two moving carts were mounted on aluminium shafts, each free
to oscillate smoothly, constraining the motion to two DOF in the crossflow direction only.
The elasticity in the system was provided by two identical springs with stiffness k1 that
were attached from the moving carts to a fixed housing, together with a central spring with
952 A26-4
VIV of a two DOF circular cylinder in the crossflow direction

Cart 2 – 2nd DOF


Cart 1 – 1st DOF

Air bearing

Springs

Flow

Cylinder

MiniShaker & LED


(Time-resolved volumetric PTV)
Figure 1. Schematic of the experimental apparatus.

stiffness k2 that connected the two moving carts together. The cylinder was attached to one
of the carts (cart 1) and placed in the test section of the water tunnel, being subjected to
fluid flow. The motion of cart 1 holding the cylinder in the flow was transmitted to cart 2
through the central spring (k2 ).
Two laser displacement sensors (Panasonic HL-G1 series) were synchronously used to
record displacement data of the oscillating carts in the crossflow direction. The oscillations
of the carts were recorded for 120 s, sampled at a rate of 500 Hz, capturing at least 50 cycles
of oscillations of the cylinder. Decay tests were conducted in still water and air to obtain
the two natural frequencies and damping ratios of the system at each mode by giving
the system an initial displacement and releasing it to oscillate freely while recording the
amplitude of oscillations. Figure 2(a) shows a sample decay test response in still water in
terms of time series of amplitude of oscillations normalized by the diameter of the cylinder,
y∗ . The first natural frequency of the system is used to make the time dimensionless, t∗ .
https://doi.org/10.1017/jfm.2022.893 Published online by Cambridge University Press

Figure 2(b) shows the frequency content of the free decay response analysed through fast
Fourier transform (FFT). This air-bearing set-up enabled us to change the ratio between
the first two natural frequencies by replacing the central spring in the set-up and varying
its stiffness, while keeping the side springs untouched. This would yield to a constant
first natural frequency of the system, while changing the second natural frequency of the
system. The first natural frequency of the system in still water was measured to be fN1 =
0.48 Hz, following the free decay tests. Table 1 summarizes the first and second natural
frequencies and damping ratios of the system, obtained from the free decay tests. The
natural frequencies for first and second modes were used to calculate first and second
reduced velocities, respectively, as Ui∗ = U/fNi D for i = 1 and 2, where U is the flow
velocity, and fN is the natural frequency of the system at each mode.
A six-axis force sensor (ATI-Nano17 IP65) was used to measure the total flow forces
of the cylinder in the crossflow and inline directions. The displacement data measured
through the laser displacement sensors was synchronised with the measured flow forces
obtained through the force sensor. During these experiments, the water level was held
constant and the flow velocity was increased from zero in small steps of 0.003 m s−1 .
952 A26-5
S. Mousavisani and others
(a) (b)
0.5 0.2
f1∗ f2∗
y∗ 0 |y ∗|

–0.5
0 10 20 30 40 50 0 1 2 3
t∗ f∗
Figure 2. Free decay response sample for the eigenfrequency ratio of R = 1.5. (a) Time history of normalized
amplitude of oscillations (y∗ ) versus normalized time (t∗ ), and (b) normalized frequency content of the
response. The in-water damping ratio is obtained as 0.008 from free decay tests.

Cylinder’s diameter D = 21.6 mm


Mass of cart-1 m1 = 0.95 kg
Mass of cart-2 m2 = 0.95 kg
First natural frequency fN = 0.48 Hz
Frequency ratio 1.3, 1.5, 1.7, 2.1, 2.5, 3
In air damping ratios ζ1 = ζ2 = 0.003
Aspect ratio L/D = 11.57
Mass ratio m∗ = 21.1
Reynolds number Re = 1033–7478
Reduced velocity U ∗ = 4–30
Table 1. Experimental parameters and the test cylinder’s structural characteristics.

The system’s response in the crossflow direction was measured for the Reynolds number
range of Re = 1033–7478, corresponding to a reduced velocity range of U ∗ = 4–30,
calculated based on the system’s first natural frequency. Details of the test matrix and
governing dimensionless system parameters are given in table 1.

2.2. Qualitative and quantitative flow visualization


https://doi.org/10.1017/jfm.2022.893 Published online by Cambridge University Press

Both qualitative and quantitative flow visualization tests were conducted to study vortex
dynamics in the wake of the cylinder. For qualitative flow visualization purposes, the
hydrogen bubble visualization technique was used. The hydrogen bubble generation
technique works based on the electrolysis of water. A platinum wire with a diameter of
0.0508 mm was strung across the test section perpendicular to the flow and the spanwise
length of the cylinder. The platinum wire served as a cathode, while a 50–100 V, 2 A power
supply positively charged a graphite plate to serve as an anode. The potential between the
two caused a build-up of hydrogen bubbles on the platinum wire. The bubbles separated
from the wire once their diameter exceeded the wire diameter, creating a bubble film that
was used to view the vortex structures in the wake of the cylinder. A high-speed camera
(Victorem 32B216MCX), placed underneath the water tunnel’s test section, was employed
to record the instantaneous wake structure from the bottom view capturing images at
200 frames per second at a spanwise location of z = 0.5 along the length of the cylinder
measured from the water free surface, where z is the dimensionless spanwise location
normalized by the length of the cylinder. To illuminate the bubble sheet, light-emitting
diode (LED) lightings were mounted on each side of the test section at an angle to ensure
uniform lighting conditions.
952 A26-6
VIV of a two DOF circular cylinder in the crossflow direction
For quantitative flow measurement purposes, a time-resolved volumetric particle
tracking velocimetry (PTV) technique was used to measure the flow field dynamics in the
wake of the cylinder at sample reduced velocities. The flow field dynamics were measured
in a time-resolved fashion using the state-of-the-art three-dimensional (3-D) Lagrangian
PTV system (Shake-the-Box, LaVision Inc., Ypsilanti, MI, USA). The shake-the-box
system uses a ‘MiniShaker’ integrated camera system and a flashlight LED array for
recording and illumination of the seeding particles, respectively. Seeding particles used
were polyamide high-quality hollow glass spheres of 60 μm nominal diameter. The
particle concentration used was approximately 0.02 g l−1 . Particles were illuminated by
a 300 × 100 mm2 LED (FLASHLIGHT 300 array, LaVision Inc.) with 72 high-power
LEDs that were operated above the nominal LED current to generate short pulses at very
high light intensities. The recording system including the four pre-aligned digital cameras
in the MiniShaker box were equipped with 16 mm lenses to capture images at 120 Hz
trigger rate at full resolution (1984 × 1264). Multiple separate subsets of 1200 images
corresponding to 20 s of measurement (which cover the minimum of eight cycles of the
cylinder’s oscillation) were acquired at each sample reduced velocity. The flashlight and
set of cameras were triggered simultaneously using a LaVision programmable timing unit
driven by DaVis 10 acquisition software. Cameras were calibrated using a 3-D calibration
plate with two planes with calibration marks separated by the pre-defined distance between
its two planes. A polynomial fit that uses a third-order polynomial was integrated to model
the perspective and optical image distortion for the mapping from a world plane to the
CCD camera sensor. The shake-the-box algorithm that uses the 3-D Lagrangian PTV
method allowed for measurements of densely seeded flows at the highest spatial resolution.
The particle-based tracking method built into this method uses an iterative particle
reconstruction technique combined with an advanced four-dimensional PTV algorithm
using the time information for the reconstruction process of particle trajectories. In the
shake-the-box algorithm, the available velocity information is used to estimate particle
distribution compared with previous methods in which first the particle distribution is
determined followed by a deduction of the velocity (Schanz, Gesemann & Schröder 2016).
This algorithm can achieve a higher reconstruction accuracy at much faster processing
speed compared with the tomographic particle imaging velocimetry (Chen, Wu & Cheng
2019).
https://doi.org/10.1017/jfm.2022.893 Published online by Cambridge University Press

Figure 1 shows the schematic of the PTV measurement set-up, and locations of the LED
flashlight and MiniShaker camera box used in this system. The platinum wire used in the
hydrogen bubble generation process was placed horizontally in the test section of the water
tunnel located at a distance of one cylinder diameter upstream of the cylinder (not shown
in the schematics).

3. Theoretical modal analysis


The experimental set-up shown in figure 1 can be modelled as a two DOF system. The
equations of motion of this two DOF system can be written as

m1 ÿ1 + (c1 + c2 ) ẏ1 − c2 ẏ2 + (k1 + k2 ) y1 − k2 y2 = Fy (t), (3.1)


m2 ÿ2 + (c1 + c2 ) ẏ2 − c2 ẏ1 + (k1 + k2 ) y2 − k2 y1 = 0, (3.2)

where mi , ki and ci denote the mass, stiffness and damping of the system, respectively. The
displacement in the crossflow direction with respect to the zero equilibrium position of
the system is denoted by yi . The subscript i = 1, 2 stands for the crossflow DOF for cart
952 A26-7
S. Mousavisani and others

(a)
Magnitude (b)

0
100 R = 3.0

φ(deg.)
R = 2.5 R = 3.0
–90 R = 2.5
R = 2.1
R = 1.7 R = 2.1
R = 1.5 –180 R = 1.7
0 1 R = 1.3 R = 1.5
2 3 0 1 R = 1.3
r = f /f 4 2 3
ext 1 r = f /f 4
ext 1

Figure 3. Frequency-response function curves of the two DOF system for different frequency ratios.

1 and cart 2, respectively. The dot symbol indicates the derivative with respect to time.
Here Fy (t) is the instantaneous fluid force on the submerged cylinder, connected to cart
1, which can be used to calculate the crossflow force coefficient, Cy (t) = 2Fy (t)/ρDU 2 L;
ρ is the fluid density, U is the flow velocity and D is the diameter of the cylinder. The
numerical values of the parameters used for the two DOF system are provided in table 1.
The system has two fundamental eigenfrequencies of f1 and f2 with the corresponding
mode shapes of 𝞥1 = [1, 1]T and 𝞥2 = [1, −1]T , respectively. For a harmonic excitation
of Fy (t) = F0 ei(2πfext )t , figure 3(a) shows the magnitude of frequency-response function
(FRF) curves of cart 1 for the different values of k2 . The plot shows that the first
eigenfrequency of the system remains constant, independent from changes in the value
of k2 , whereas the second eigenfrequency increases with any increase in the value of
k2 , resulting in eigenfrequency ratios (R = f2 /f1 ) between 1.3 and 3.0 for the system.
Figure 3(b) shows the phase of FRF curves of cart 1 for the different values of k2 . It
shows that the displacement remains in phase with forcing for frequencies smaller than
the first eigenfrequency of the system, and experiences a sudden transition to out-of-phase
at the frequencies very close to the resonance (r = 1). This is a typical behaviour observed
in systems with a very low damping ratio (ζ  1). The experimental set-up in our current
study is designed to uniquely achieve such a low damping ratio. A similar behaviour is
observable in frequency regions near the second eigenfrequency of the system.
The two DOF system’s response to fluid forces in the crossflow direction, Fy (t),
https://doi.org/10.1017/jfm.2022.893 Published online by Cambridge University Press

can include single or multiple harmonics. These harmonics may or may not match the
eigenfrequencies of the system in general. The FFT plot can be used to analyse the
frequency content of the oscillation, but a FFT plot does not provide any information on
which mode shapes of the system are excited. In a nonlinear system, for example, a FFT
plot with multiple frequency peaks can be either associated with a multi-modal oscillation
or a multi-harmonic oscillation. In a multi-modal oscillation, multiple modes of the
system are excited at their corresponding eigenfrequencies, whereas in a multi-harmonic
oscillation only one mode is excited at multiple different frequencies. To distinguish
between these two different dynamic responses, periodic orbits in the 3-D space of
( y1 , y2 , ẏ1 ), shown in figure 4, can be used to visually differentiate between a single
harmonic oscillation, a multi-modal oscillation and a multi-harmonic oscillation for a two
DOF system. The projection in the x–y plane provides information on mode shapes that
have been excited, whereas the projection in the x–z or y–z planes provides information on
frequency contents of the oscillation. The periodic orbits are commonly used in nonlinear
dynamic studies to visualize the steady-state response of a nonlinear system (Ardeh &
Allen 2013).
952 A26-8
VIV of a two DOF circular cylinder in the crossflow direction
(a) (b) (c)

2 2 2
∙ ∙ ∙
ŷ1 0 ŷ1 0 ŷ1 0
–2 –2 –2 –2 –2 –2
0 0 0 ŷ
–2 ŷ1 –2 ŷ1 –2
0 1
0 2 0 2 2
ŷ2 2 ŷ2 2 ŷ2 2

(d ) (e) (f )

2 2 2
∙ ∙ ∙
ŷ1 0 ŷ1 0 ŷ1 0
–2 –2 –2 –2 –2 –2
0 0 0 ŷ
–2 ŷ1 –2 ŷ1 –2
0 1
0 2 0 2 2
ŷ2 2 ŷ2 2 ŷ2 2

(g) (h) (i)

1 1 1
∙ ∙
ŷ1 0 ŷ1 0 y∙1 0
–1 –1 –1 –1
–1 –1
0 0 0
–1 ŷ1 –1 ŷ1 –1 y1
0 1 0 1 0 1 1
z ŷ2 1 ŷ2 1 y2
y
x
Figure 4. Periodic orbits of a two DOF system at (a–c) single harmonic motions, (d–f ) higher harmonic
motions and (g–i) multi-modal motions.

The first row plots in figure 4(a–c) show samples for the periodic orbits of a two DOF
system for a single harmonic oscillation, where the excitation frequencies are selected as
fext = f1 , ( f1 + f2 )/2 and f2 , respectively. Here y1 and y2 are normalized displacements
https://doi.org/10.1017/jfm.2022.893 Published online by Cambridge University Press

with respect to the maximum oscillation amplitude, ymax , obtained for y1 and y2 in our
simulations; ẏ1 is normalized with respect to 2πf1 ymax . The hat symbol in these figures
denotes the normalized values. The sample eigenfrequency ratio is set to R = 1.7. In
the first row plots, the projection of periodic orbits in the x–y plane forms an ellipse
(approximately a line here due to very small damping ratios) with +45◦ and −45◦
orientations for fext = f1 and f2 , representing harmonic oscillation around the 1st mode
shape and the 2nd mode shape of the system, respectively. The projection takes a form
between the projection for the 1st mode shape and 2nd mode shape if the system is
excited at a frequency that does not match the fundamental eigenfrequencies of the
system, similar to the one shown in figure 4(b). The second and third row plots in
figure 4(d–i) show the difference between periodic orbits of a multi-harmonic oscillation
and a multi-modal oscillation for the two DOF system, respectively. In a multiharmonic
oscillation (second row), the projection of periodic orbits in the x–y plane shows that
only the 1st mode or 2nd mode, or a linear combination of them are excited, even though
the projection in the x–z or y–z planes shows existence of two harmonics (two closed
circles). The third row plots in figure 4(g–i) show the case of a multi-modal oscillation
952 A26-9
S. Mousavisani and others
where the 1st and 2nd mode shapes of the system are excited by a harmonic excitation
such as Fy (t) = F1 e(i(2πf1 )t) + F2 e(i(2πf2 )t) , where F2 = F1 , F2 = 4F1 and F2 = 8F1 ,
respectively. The projection of periodic orbits in the x–y plane shows a transition from
the 1st mode being dominantly excited to the 2nd mode being dominantly excited. To
study the modal transition in our experimental VIV study, we have obtained similar plots
for our experiments that will be discussed in our results. Such an analysis provides us
with a powerful tool to better characterize the dynamic VIV response of the system
and distinguish between a multi-modal oscillation or a multi-harmonic oscillation in the
system.

4. Results
4.1. Structural response of the system
In this section an overall view of the VIV response is given for the experimental campaign
discussed in the present work, where the cylinder is allowed to oscillate with two DOF in
the crossflow direction only. The response is discussed for a range of eigenfrequency ratios
of R = 1.3, 1.5, 1.7, 2.1, 2.5 and 3.0 over a wide range of flow velocities. The first row
plots in figure 5 show the dimensionless amplitudes of oscillation (defined as y∗ = Y/D,
where Y is the amplitude of oscillation calculated based on the root-mean-square (r.m.s.)
amplitude of the cylinder’s displacement time series that is calculated around the mean
value of the response, y), versus reduced velocities of U ∗ and U2∗ , that are reduced
velocities calculated based on the first (fN1 ) and the second (fN2 ) natural frequencies of
the system, respectively. The second row plots are r.m.s. values of modal contributions,
normalized by the diameter of the cylinder plotted versus reduced velocity for each
eigenfrequency ratio. The term ’modal contribution’ is used to determine how each
vibrational mode is contributing to the dynamic response of the system over the range
of reduced velocities tested. This will better facilitate our understanding of the observed
VIV response of the cylinder and any possible multi-modal oscillations of the system. The
third row plots show frequency contents of the oscillations plotted as contour plots for the
crossflow oscillations as a function of the reduced velocity, U ∗ . These contours are formed
by taking the FFTs of displacement time series at every given value of U ∗ , normalizing
them by the first natural frequency of the system in water, and placing them next to each
https://doi.org/10.1017/jfm.2022.893 Published online by Cambridge University Press

other. The dashed lines represent the Strouhal frequencies for a fixed cylinder at similar
Reynolds number ranges, also measured in our experimental campaign by analysing the lift
force frequency content of a stationary rigid cylinder using the six-axis force sensor set-up
discussed in § 2. The response of the cylinder is plotted for the range of eigenfrequency
ratios listed in table 1. The response is also compared with those of a single DOF cylinder
by modifying the experimental apparatus, such that a rigid link replaced the centre spring
connecting the two carts, allowing oscillation of both carts to be limited only to one DOF
in the crossflow direction. That being said, the amplitude of oscillations of the circular
cylinder undergoing classical VIV in a single DOF are also plotted for comparison in the
first row plots of figure 5 in a grey colour.
At a frequency ratio of R = 1.3 (figure 5a), oscillations start at a reduced velocity of
U ∗ ≈ 4.8, with frequencies of oscillation around the first eigenfrequency of the system
(figure 5c), reach a maximum amplitude of y∗ ≈ 0.48 at the reduced velocity of U ∗ ≈ 6.4.
The amplitude of oscillations versus reduced velocity response up to this point follows
a very similar trend as those observed for a classical one DOF VIV response of a
circular cylinder, as also plotted on top of the amplitude plots, shown in grey markers.
952 A26-10
VIV of a two DOF circular cylinder in the crossflow direction

U∗2 U∗2
(a) 0 5 10 15 20 25 0 5 10 15 20
0.6 (d) 0.6
y∗ 0.4 0.4
0.2 0.2

0 5 10 15 20 25 30 0 5 10 15 20 25 30
(b) 0.6 (e) 0.6
Mode 1 Mode 1
Mode 2
A∗
0.4 0.4 Mode 2
0.2 0.2
0 0
0 5 10 15 20 25 30 0 5 10 15 20 25 30
(c) 2.0 0.6 ( f ) 2.0 0.6
0.4 0.4
f ∗ 1.5 1.5
1.0 0.2 1.0 0.2
0.5 0 0.5 0
0 5 10 15 20 25 30 0 5 10 15 20 25 30

(g) 0 3 6 9 12 15 18 0 3 6 9 12 15
0.6 ( j) 0.6
y∗ 0.4 0.4
0.2 0.2

0 5 10 15 20 25 30 0 5 10 15 20 25 30
(h) 0.6 Mode 1
(k) 0.6 Mode 1
A∗ 0.4 Mode 2 0.4 Mode 2
0.2 0.2

0 5 10 15 20 25 30 0 5 10 15 20 25 30
(i) 2.5 0.6 (l) 2.5 0.6
2.0 0.4 2.0 0.4
f∗ 1.5 1.5
1.0 0.2 0.2
1.0
0.5 0 0.5 0
0 5 10 15 20 25 30 0 5 10 15 20 25 30

(m) 0 3 6 9 12 (p) 0 2 4 6 8 10
0.6 0.6
y∗ 0.4 0.4
0.2 0.2

0 5 10 15 20 25 30
https://doi.org/10.1017/jfm.2022.893 Published online by Cambridge University Press

0 5 10 15 20 25 30
(n) 0.6 (q) 0.6
Mode 1 Mode 1
A∗ 0.4 Mode 2 0.4 Mode 2
0.2 0.2

0 5 10 15 20 25 30 0 5 10 15 20 25 30
(o) 0.6 (r) 0.6
3 3
0.4 0.4
f∗ 2 2
0.2
0.2
1 1
0 0
0 5 10 15 20 25 30 0 5 10 15 20 25 30
U∗ U∗
Figure 5. Amplitudes of oscillation (first row), modal contributions (second row) and frequency content of
oscillation (third row) versus the reduced velocities of U ∗ and U2∗ for all eigenfrequency ratios. Plots are shown
for the eigenfrequency ratio of R = 1.3 (a–c), R = 1.5 (d–f ), R = 1.7 (g–i), R = 2.1 (j–l), R = 2.5 (m–o)
and R = 3.0 (p–r). The dark grey dashed line in the frequency contour plots represents the Strouhal line of
St = 0.17. The grey solid symbols in the amplitude of oscillations plots (first row) represent the VIV response
of a single DOF circular cylinder.

952 A26-11
S. Mousavisani and others
Modal contributions at this range of reduced velocity exhibit contributions from pure mode
one of the oscillations (figure 5b). At a reduced velocity range of U ∗ ≈ 7.63–7.95, the
amplitude of oscillations suddenly drop to values around y∗ ≈ 0.35. Modal contributions
in this reduced velocity range exhibit a decrease in the contribution of the first structural
mode and an increase in the second structural mode contributions. A multi-modal response
is observed at this small reduced velocity range with comparable contributions from both
the eigenfrequencies of the system. Beyond this reduced velocity range, the amplitude of
oscillations start to increase again, followed by a jump in the frequency of oscillations
that take values close to the second eigenfrequency of the system (figure 5c). Beyond the
second peak in the amplitude of oscillations, occurring at the reduced velocity of U ∗ ≈
8.3, by increasing the reduced velocity, the amplitude of oscillations decreases and drops
to zero at the reduced velocity of U ∗ ≈ 15.8, that indicates the end of the lock-in region.
The modal contributions at this reduced velocity range exhibit pure contributions from the
second structural mode. The second peak in the amplitude of oscillations observed at this
reduced velocity range has a maximum value around A∗ ≈ 0.46, that is relatively the same
as those observed at the first peak in the lock-in region.
As the frequency ratio between the first two modes increases, the VIV response of the
system follows relatively similar trends as those discussed for the ratio of R = 1.3. For
lower reduced velocity ranges, the VIV response is dominated by the first structural mode
(first lock-in region), then switches to a multi-modal VIV response, where both the first
and second modes contribute to the VIV response of the system. As the reduced velocity
further increases, the response exits the overlapping multi-modal oscillations range and
enters the second lock-in region corresponding to the oscillations at the second mode of
vibration. The distinct trend observed here is that as the frequency ratio between the first
two modes increases and the system’s eigenfrequencies get values apart from each other,
the two lock-in regions at each mode get separated with a subtle and sudden decrease in the
amplitude of oscillation that separates the two lock-in regions observed for each mode. The
two lock-in regions become completely separated at high frequency ratios of R = 2.5 and
R = 3.0 (figure 5m,p), at which the VIV response of the system at the first mode follows
a very similar trend as observed for the single DOF system, where the lock-in region is
extended in the reduced velocity range of U ∗ ≈ 4.8–12, with a maximum amplitude of
y∗ ≈ 0.47. For the second lock-in region, at which the oscillations occur at frequencies
close to the second eigenfrequency of the system, the lock-in region is extended in the
https://doi.org/10.1017/jfm.2022.893 Published online by Cambridge University Press

reduced velocity range of U ∗ ≈ 12.6–27.7 for the eigenfrequency ratio of R = 2.5, and
reduced velocity range of U ∗ ≈ 13–30 for the eigenfrequency ratio of R = 3.0. For all
eigenfrequency ratios, although the second lock-in region is wider than those observed for
mode one, they are extended over relatively similar reduced velocity ranges of ≈5–11 for
R = 2.5 and ≈4.5–10 for R = 3.0, if the reduced velocity is defined as the flow velocity
normalized by the second eigenfrequency of the system (U2∗ ), shown as the upper x-axis
ranges in the amplitude plots of figure 5.
Figure 6 shows three sample displacement time histories (first row), and their
corresponding frequency content (FFT plots at the second row) at three reduced velocities
of U ∗ = 6.05, 9.21 and 12.05 for the eigenfrequency ratio of R = 1.7. The time and
frequency values in the x axes are plotted as dimensionless values (t∗ and f ∗ ), normalized
by the first natural frequency of the system. The third row plots represent the periodic
orbits in the 3-D space of ( y∗1 , y∗2 , ẏ∗1 ), as those studied earlier in figure 4. These sample
reduced velocities are selected such that they represent three different regions of modal
contributions in the VIV response of the system; at the low reduced velocity of U ∗ = 6.05,
the VIV response of the system consists of oscillations occurring at pure mode one,

952 A26-12
VIV of a two DOF circular cylinder in the crossflow direction
(a) (b) (c)
0.5
0.5 0.5
y∗ 0 0 0
–0.5 –0.5
–0.5
0 10 20 30 0 10 20 30 0 10 20 30
(d) t∗ (e) t∗ (f) t∗
0.50 0.50 0.75
|y∗| 0.25 0.25
0.50
0.25
0 0 0
1 2 3 1 2 3 1 2 3
f∗ f∗ f∗
(g) (h) (i)

1 2 2
y·1∗ 0 y·1∗ 0 y·1∗ 0
–1 –2 –2
–1 –2 –2
0 –2 0 –2 0
y1∗ y1∗
–1
0 y1∗ 0
2 2
0
2 2
z y∗2
1 1
y∗2 y∗2
y
x

Figure 6. Displacement time histories (a–c) and the corresponding FFTs (d–f ) and periodic orbits (g–i) at the
reduced velocity of U ∗ = 6.05 representing mode 1 (a,d,g), U ∗ = 9.20 representing transition from mode 1 to
mode 2 (b,e,h), and U ∗ = 12.05 representing mode-2 oscillations (c, f,i).

represented with a single sharp peak at its frequency content (FFT) plot (figure 6d). At
a higher reduced velocity of U ∗ = 9.21, the VIV response of the system is a multi-modal
response selected at the transition region from mode one to mode two, at which oscillations
have contributions from both modes with two sharp peaks at values close to the first
and second eigenfrequencies of the system (figure 6e). Finally, at the reduced velocity
of U ∗ = 12.05, the VIV response of the system consists of oscillations at the second
structural mode, with frequency contents represented by a sharp peak at the frequency
close to the second eigenfrequency of the system (figure 6f ). The projections of periodic
orbits (third row plots) on the x–y plane confirm that indeed modes one and two are excited
at a reduced velocity of U ∗ = 6.05 and U ∗ = 12.05, as they form an ellipse with +45◦
https://doi.org/10.1017/jfm.2022.893 Published online by Cambridge University Press

and −45◦ orientations, respectively. More importantly, the observed periodic orbits with
a stretched rectangular form in the x–y plane at the reduced velocity of U ∗ = 9.20 assure
that the observed VIV response is a multi-modal response where both modes one and two
are excited.

4.2. Flow forces


An additional series of water tunnel experiments were conducted during which the force
sensor was attached to the upper end of the cylinder, measuring the total flow forces
acting on the cylinder in both the crossflow and inline directions as the cylinder undergoes
VIV. The measured forces in the crossflow direction comprised both the hydrodynamic
flow forces and the inertia force due to the motion of the force sensor. Therefore, the
inertia force was removed from the total measured transverse force by post-processing the
experimental data. When the force sensor was added at the top of the submerged cylinder,
the mass ratio and the submerged length of the cylinder were kept constant compared with
the experiments done without the force sensor attached, the results of which were reported
952 A26-13
S. Mousavisani and others
in the previous section. This was achieved by adjusting the length and mass of the bracket
connecting the cylinder to the force sensor, which was kept above the water level. However,
the mechanical coupling between the sensor and the bracket holding the cylinder added to
the flexibility of the system, such that at high reduced velocities beyond U ∗ ≈ 16.0, the
cylinder in our experiments experienced shaking at its free end, causing the flow forces to
take unrealistic values at such high reduced velocities. As a solution, an alternative method
was used here to obtain the flow forces in the crossflow direction. This technique works
based on using the time series of the cylinder’s crossflow displacement to calculate and
reconstruct the crossflow forces as
       
Fy ÿ1 2ζ1 ωn1 0 ẏ1
= 𝞥𝞥 T
+𝞥 𝞥 T
0 ÿ2 0 2ζ2 ωn2 ẏ2
 2   
ωn1 0 y1
+𝞥 𝞥T , (4.1)
0 ωn22 y2
where Fy is the crossflow force acting on the cylinder, 𝞥 is the mass-normalized mode
shape of the system (obtained only for the structure, with no added mass contribution), ωni
is the natural frequency of the system in air. Following the vibration models for discrete
parameter systems (Inman & Singh 1994), we have used a proportional viscous damping
model to capture damping effects for the two DOF system. Assuming a proportional
viscous model, we have measured the structural damping ratios at each mode of the
system, ζi , using free decay tests in air and have obtained similar damping ratios of
ζ1 = ζ2 = 0.003. Here, yi is the displacement data recorded using the non-contacting laser
displacement sensor. This force reconstruction method was previously verified and used
to measure the flow forces acting on a tapered cylinder and a triangular prism undergoing
large amplitude flow-induced vibration (Seyed-Aghazadeh, Carlson & Modarres-Sadeghi
2015, 2017). Following this flow reconstruction technique, the reconstructed flow forces
were shown to match very well with the results of the direct measurements using a
force sensor during experiments (Seyed-Aghazadeh et al. 2015). Here, sample cases are
presented to show how this reconstruction method works for VIV response measurements
of a two DOF circular cylinder.
Figure 7(a,b) shows the comparison between the dimensionless crossflow force
coefficients (Cy ) and their corresponding frequency content measured directly by the force
https://doi.org/10.1017/jfm.2022.893 Published online by Cambridge University Press

sensor and those reconstructed using (4.1), plotted over the range of reduced velocities
tested for a two DOF VIV case at the eigenfrequency ratio of R = 1.7. The crossflow
force coefficients are calculated as Cy = 2Fy /ρDU 2 L. Additionally, three sample cases of
the crossflow force coefficient time histories and their corresponding frequency contents
are plotted in figure 7(c–h) at sample reduced velocities of U ∗ = 6.05, 9.21 and 12.05
(also shown with dashed lines in figures 7a,b). It is observed that the force reconstruction
method not only is capable of reproducing the general force distribution over the range of
reduced velocities tested, but also it can reproduce very fine details of the flow forces,
as shown both in the time history and in the FFT plots of all three cases. Following
this validation, the flow forces in the crossflow direction reported in this work are the
reconstructed values, while the inline flow forces are directly measured via the force
sensor, reported up to the reduced velocity of U ∗ ≈ 16, at which the cylinder started to
experience shaking. In addition, the force sensor measurements, as well as reconstructed
crossflow forces, were synchronized with the displacement data recordings from the
non-contacting displacement sensors, making it possible to directly obtain the phase
difference between the flow forces and the displacement of the cylinder in the crossflow
direction.
952 A26-14
VIV of a two DOF circular cylinder in the crossflow direction
(a) (b) 2.5 1.5
P1 P2 P3 P1 P2 P3
1.5
2.0
1.0
Cy 1.0 f ∗C 1.5
y
0.5
0.5 1.0

0 0.5 0
4 6 8 10 12 14 16 4 6 8 10 12 14 16
U∗ U∗
(c) (d)
1 1.0
Cy 0 |Cy| 0.5
–1
0
0 5 10 15 20 25 30 0.5 1.0 1.5 2.0 2.5 3.0 3.5
(e) (f)
2
1.0
Cy 0 |Cy| 0.5
–2 0
0 5 10 15 20 25 30 0.5 1.0 1.5 2.0 2.5 3.0 3.5
(g) 2 (h)
1.0
Cy 0 |Cy| 0.5
–2 0
0 5 10 15 20 25 30 0.5 1.0 1.5 2.0 2.5 3.0 3.5
t∗ f∗
Figure 7. (a) Comparison between the crossflow force coefficients measured directly by the force sensor
(grey dashed line) to those reconstructed using displacement data (blue solid line), (b) frequency content
of the measured crossflow force coefficient, (c–g) time histories and (d–h) their corresponding frequency
contents of the dimensionless crossflow forces at sample reduced velocities of U ∗ = 6.05, 9.21 and 12.05
at the eigenfrequency ratio of R = 1.7.

Figure 8 shows the crossflow force coefficients together with their frequency contents
and phase differences between the flow forces and the displacement of the cylinder in
the crossflow direction plotted against the reduced velocity for the VIV response of the
https://doi.org/10.1017/jfm.2022.893 Published online by Cambridge University Press

cylinder with two DOF with varying eigenfrequency ratios. Shown with grey markers
are the crossflow force coefficients for the VIV response of the cylinder with one DOF
measured by the force sensor, plotted for comparison. The phase values are calculated
based on filtering the displacement and force signals around the dominant frequency of
oscillations. At the reduced velocity range at which multi-modal oscillations occur, the
phase values are calculated by filtering the vibration response around each eigenfrequency
of the system.
At the eigenfrequency ratio of R = 1.3, the crossflow force coefficients increase sharply
from zero at the beginning of the first VIV region to the maximum value of Cy ≈ 1.6 at
the reduced velocity of U ∗ = 6.4. This trend, reduced velocity ranges and the maximum
Cy values stay the same for all the eigenfrequency ratios, following a very similar trend
compared with those observed for the VIV response of a one DOF cylinder, shown in grey
dashed lines. In this reduced velocity range that the crossflow force coefficients increase
to their maximum values, the frequency content of the forces exhibits contributions from
the first eigenfrequency of the system. As the reduced velocity is further increased,
contributions from the second eigenfrequency of the system are observed in the frequency
content plots of the crossflow forces. Once the multi-modal flow forces dominate the
952 A26-15
S. Mousavisani and others
U2∗ U2∗
(a) 0 5 10 15 20 25 (d) 0 5 10 15 20
2 2
Cy 1 1

0 5 10 15 20 25 30 0 5 10 15 20 25 30
(b) 2.0 2.5 (e) 2.0 2.5
2.0 fair,2 2.0

f ∗C 1.5 fair,2 1.5 1.5 1.5


y fair,1 1.0 fair,1 1.0
1.0 1.0
0.5 0.5
0.5 0 0.5 0
5 10 15 20 25 30 5 10 15 20 25 30
(c)
90 (f) 90
φ(deg.)

45 45

0 5 10 15 20 25 30 0 5 10 15 20 25 30

(g) 0 3 6 9 12 15 18 ( j) 0 3 6 9 12 15

2 2
Cy
1 1

0 5 10 15 20 25 30 0 5 10 15 20 25 30
(h) 2.0 fair,2
2.5 (k) 2.5 2.5
fair,2
2.0 2.0
2.0
f ∗C 1.5 1.5
fair,1 1.0 1.5
1.5
y fair,1 1.0
1.0 1.0
0.5 0.5
0.5 0 0.5 0
5 10 15 20 25 30 5 10 15 20 25 30
(i) 90 (l) 90
φ(deg.)

45 45

0 5 10 15 20 25 30 0 5 10 15 20 25 30

(m) 0 2 4 6 8 10
0 2 4 6 8 10 12 (p)
2 2
Cy 1
https://doi.org/10.1017/jfm.2022.893 Published online by Cambridge University Press

0 5 10 15 20 25 30 0 5 10 15 20 25 30
3
2.5 (q) fair,2 2.5
(n) fair,2
2.0 3 2.0
f ∗C 2 1.5
2
1.5
y fair,1
1.0 1.0
fair,1
1 0.5 1 0.5
0 0
0 5 10 15 20 25 30 0 5 10 15 20 25 30
(o) 90 (r) 90
φ(deg.)

45 45

0 5 10 15 20 25 30 0 5 10 15 20 25 30
U∗ U∗
Figure 8. The crossflow force coefficients (first row), their corresponding frequency contents (second row),
and the phase difference between the cylinder’s displacement and the crossflow forces acting on the cylinder
(third row) versus the reduced velocities of U ∗ and U2∗ for all the eigenfrequency ratios tested. Plots are shown
for the eigenfrequency ratio of R = 1.3 (a–c), R = 1.5 (d–f ), R = 1.7 (g–i), R = 2.1 (j–l), R = 2.5 (m–o) and
R = 3.0 (p–r). The grey markers show the crossflow force coefficients for the VIV response of a single DOF
system.
952 A26-16
VIV of a two DOF circular cylinder in the crossflow direction
VIV response of the system, by increasing the reduced velocity the crossflow force
coefficient decreases. The force coefficient values then exhibit a plateau, at which the force
coefficients stay around the value of Cy ≈ 1.1. As the reduced velocity further increases,
the multi-modal response gives way to oscillations at the second mode. For this range of
reduced velocity at the second lock-in region, the force coefficient values pick up again and
reach to a second maxima. As the reduced velocity further increases, the force coefficients
decrease and reach zero once the response exits the second lock-in region. The crossflow
force coefficients have a relatively similar trend for all eigenfrequency ratios. As the
eigenfrequencies take values further away from each other, the two peaks in the crossflow
force coefficient plots grow separately from each other. For example, at the eigenfrequency
ratio of R = 2.1 and beyond, the two maxima peaks of the force coefficients are completely
separated, where the first region with the corresponding maximum peak of the force
coefficient occurs at the reduced velocity range of U ∗ = 4.8–10.8 over which the VIV
response is dominated by oscillations at the first mode. The second region is followed
over the reduced velocity range of U ∗ = 10.8–23.8, at which the second mode dominates
the VIV response of the system. The overall force coefficient values in the first peak
region, dominated by the first structural mode, closely match the force coefficients of those
observed for the VIV response of a single DOF circular cylinder.
The phase difference between the crossflow displacements and forces plotted in the
third row of figure 8 shows a change with reduced velocity for each eigenfrequency
ratio. For eigenfrequency ratios below R = 2.1, for which the VIV response represents
a multi-modal oscillation for an intermediate reduced velocity range during the transition
from mode one to mode two, the phase values stay below ≈12◦ for the entire range of
reduced velocities. This behaviour is different from what is observed in the VIV response
of a circular cylinder, for which there is a phase jump from 0◦ to 180◦ between flow forces
and the cylinder displacement (Khalak & Williamson 1999). This is due to the fact that
with a non-zero structural damping, the phase jump to 180◦ occurs only if the frequency
ratio between the oscillation frequency and the natural frequency of the system measured
in air (fosc /fna ) gets values far from unity (Sarpkaya 2004). In the current experiments,
however, the frequency ratio stays close to unity, resulting in a phase difference of less
than 180◦ . The dashed lines in the frequency contents of the crossflow forces represent
the first and second natural frequencies of the system in air. For eigenfrequency ratios of
R = 2.1 and R = 2.5, there is a phase change from small values to values close to 45◦
https://doi.org/10.1017/jfm.2022.893 Published online by Cambridge University Press

for the circular cylinder tested here, which means that the flow forces start opposing the
displacement at a certain reduced velocity close to the end of the first lock-in region. Once
the second mode oscillation starts to ramp up, the phase goes back to zero and it stays
around values smaller than 5◦ for the entire reduced velocity range over the second lock-in
range.
Figure 9 shows the overall VIV response of the circular cylinder, free to oscillate in two
DOF in the crossflow direction, compared with those of a single DOF VIV response of a
circular cylinder. The overall amplitude of oscillation versus reduced velocity is plotted in
figure 9(a) for all the eigenfrequency ratios tested. As discussed earlier, by increasing
the eigenfrequency ratio, the lock-in range gets extended to higher reduced velocities.
Beyond a critical eigenfrequency ratio of R = 2.1, the lock-in region is split into two
very separate regions, at which each eigenmode is excited, separately. The corresponding
modal contributions versus the reduced velocities are plotted in figure 9(b) for the range
of eigenfrequencies tested. The modal contributions exhibit oscillations that shift from
pure mode one at low reduced velocities to a range of intermediate reduced velocities
at which multi-modal oscillations occur. By increasing the reduced velocity values, the

952 A26-17
S. Mousavisani and others
contributions from mode one decrease to zero and the VIV response gets dominated by
pure mode two oscillations. While for the eigenfrequency ratios below R = 2.1, the VIV
response of the system follows this behaviour, for relatively larger eigenfrequency ratios of
R = 2.5 and 3.0, the two lock-in regions are well separated from each other such that there
does not exist any intermediate reduced velocity range at which multi-modal oscillations
occur. Time-averaged and fluctuating inline force coefficients are shown in figures 9(c)
and 9(d), respectively. The inline force coefficients here are directly measured through
the force sensor attached atop of the cylinder. The force coefficients are plotted for the
reduced velocities up to which the cylinder does not experience any shaking at its tip.
The time-averaged inline force coefficients evolution follows similar trends as those of the
amplitude of oscillations at each eigenfrequency ratio, where they increase to a maximum
value at the reduced velocity at which the maximum oscillation occurs, and decrease to
lower values as the amplitude of oscillation decreases. As for fluctuating components of the
inline force coefficients, while their maximum values appear to be relatively the same for
a wide range of eigenfrequency ratios, multiple peaks are observed as the eigenfrequency
ratios increase. The crossflow force coefficients shown in figure 9(e) increase from zero
to their maximum values and decrease back to zero for a single DOF system, while they
exhibit two maxima for a two DOF system. The two maxima get further away from each
other as the eigenfrequency ratio gets larger. Figure 9( f,g) shows the crossflow force
coefficients in phase with velocity (Cyv ) and the crossflow force coefficients in phase with
acceleration, known as the added mass coefficient (Cma ), versus the reduced velocity for
all the eigenfrequency ratios, respectively. These values have been calculated as (Wang,
Fan & Triantafyllou 2021)
     
2 Ts Cy (t)ẏ(t) dt 2U 2 Ts Cy (t)ÿ(t) dt
Cyv =
Ts
·    , and Cma = − πd2 ·  ÿ2 (t) dt , (4.2)
2
Ts Ts ẏ2 (t) dt Ts

where Cy is the instantaneous crossflow force coefficient, Ts is one vibration cycle and the
dots represent derivatives with respect to time. The calculated crossflow force coefficients
in phase with velocity have all positive values for all the reduced velocities tested,
confirming that the flow forces contribute to the excitation of the cylinder for all the cases
studied here. The crossflow force coefficients in phase with velocity values also follow
https://doi.org/10.1017/jfm.2022.893 Published online by Cambridge University Press

similar trends as those observed for the crossflow force coefficients, where two peaks
are observed for the two DOF system. The distance between two maxima increases with
the increase in eigenfrequency ratio. Figure 9(g) shows the crossflow force coefficients in
phase with acceleration, known as the added mass coefficient (Cma ), versus the reduced
velocity for all the eigenfrequency ratios tested. These added mass coefficients have all
positive values for all the reduced velocities tested. As shown previously (Wang et al.
2021), the value of the added mass coefficient reflects the relative motion between the
oscillating cylinder and the shed vortex in the near wake of the cylinder. This relation
will be discussed further by studying the wake dynamics of the cylinder in the following
section.
While in figure 9(d) we discussed the values of the inline force coefficients versus
reduced velocity for the range of eigenfrequencies tested, it is also worth discussing
their frequency contents here. Figure 10 represents the frequency contents of the inline
force coefficients for the sample cases of a single DOF cylinder (figure 10a) and three
eigenfrequencies of R = 1.3, 1.7 and 2.1 (figure 10b–d). For the single DOF system, the
inline force coefficients follow a trend that is in agreement with previous VIV studies, for
which the second harmonic frequencies, with values twice the natural frequency of the
952 A26-18
VIV of a two DOF circular cylinder in the crossflow direction

(a) (b)

0.6 0.6
0.4
y∗ 0.4 A∗ 0.2
R = 3.0 R = 3.0
R = 2.5 R = 2.5
0.2 R = 2.1 R = 2.1
R = 1.7 0 R = 1.7
0 R = 1.5 R = 1.5
5 R = 1.3 5 10 R = 1.3
10 15 15 20 1DoF
20 25
1DoF 25 30
30

(c) (d)

2
Cx 0.2 R = 3.0
– R = 3.0 R = 2.5
Cx 1 R = 2.5 R = 2.1
R = 2.1 R = 1.7
0 R = 1.7 0 R = 1.5
R = 1.5 5 R = 1.3
5 R = 1.3 10 15
10 15 20 25 30
1DoF
20 25 30
1DoF

(e) (f)
0.2
2 Cyv R = 3.0
R = 3.0 0.1 R = 2.5
Cy 1 R = 2.5 R = 2.1
R = 2.1 0 R = 1.7
R = 1.7 R = 1.5
0 R = 1.5 5 10 R = 1.3
5 15 20 1DoF
R = 1.3 25 30
10 15 20 25 30
1DoF U∗
U∗

(g)
4
R = 3.0
Cma 2 R = 2.5
R = 2.1
0 R = 1.7
R = 1.5
5 10 R = 1.3
15 20 1DoF
25 30
U∗

Figure 9. An overall comparison of the VIV response of the system in terms of its (a) dimensionless amplitude
of oscillations, (b) modal contributions, (c) time-averaged inline force coefficients, (d) inline force coefficients,
(e) crossflow force coefficients, ( f ) crossflow force coefficients in phase with velocity and (g) crossflow force
https://doi.org/10.1017/jfm.2022.893 Published online by Cambridge University Press

coefficients in phase with acceleration versus the reduced velocity for all eigenfrequency ratios.

system, are observed as the dominant frequency, with small contributions from the first
harmonic frequencies (Carlson, Currier & Modarres-Sadeghi 2021). When the cylinder is
given the second DOF, the frequency content still exhibits the first and second harmonic
contributions of each eigenfrequency. In addition to these frequencies, with two DOF
oscillations additional frequencies are also present in the frequency contents shown in
figure 10(b–d). These components are associated with the relative velocity of the cylinder
moving in the crossflow direction that makes an angle with respect to the oncoming flow.
Figure 11 shows the flow forces acting on an oscillating cylinder. As the cylinder oscillates,
due to the relative velocity of the body with respect to the oncoming flow, the lift and drag
forces make an angle with respect to the crossflow and inline directions. Previous VIV
studies have shown the existence of second and third harmonic frequency components in
the flow forces, when the cylinder was free to oscillate in the crossflow direction (Du, Jing
& Sun 2014; Seyed-Aghazadeh et al. 2015). Of particular interest in the current study is the
frequency contents of the inline flow forces at the eigenfrequency ratios for which the VIV
952 A26-19
S. Mousavisani and others
response of the system exhibits multi-modal oscillations over a range of reduced velocities.
Assuming sinusoidal multi-modal oscillations for the cylinder in the inline direction as
Y(t) = Y1 sin (ω1 t) + Y2 sin (ω2 t) , (4.3)
in which ω1 and ω2 are the two frequencies of cylinder oscillation and Y(t) is the
dimensionless amplitude of oscillation with respect to the diameter of the cylinder, D.
The flow forces acting in the inline direction can be expressed as
Fx (t) = FD (t) cos θ − FL (t) sin θ, (4.4)
where FL and FD are the lift and drag forces with combined frequencies of ω1 and ω2 and
the values twice those, respectively, which can be expressed as
FL (t) = FL1 sin (ω1 t + ψ1 ) + FL2 sin (ω2 t + ψ2 ) , (4.5)
FD (t) = F̄D + FD1 sin (ω1 t + ψ3 ) + FD2 sin (ω2 t + ψ4 ) + FD3 sin (2ω1 t + ψ5 )
+FD4 sin (2ω2 t + ψ6 ) , (4.6)
where ψ values are the phase angles between the lift and drag forces with respect to
the cylinder’s displacement. It is worth noting that ideally, for a symmetric wake of the
cylinder, one would expect the drag force frequency content to include only frequencies
twice those observed in the lift direction. However, in general, any asymmetry of the
wake can lead to frequency contents at similar values of those in the lift forces and their
corresponding second harmonics, as those considered here in (4.6) (Carlson et al. 2021).
For a small angle between the flow velocity and the relative velocity, θ,
sin θ ∼ tan θ ∼ Ẏ, cos θ ∼ 1, (4.7)
where Ẏ is the dimensionless velocity of the cylinder in the crossflow direction. Following
the assumption of a small angle θ , (4.4) can be written as
Fx (t) = FD (t) − FL Ẏ(t). (4.8)
Substituting the lift and drag forces from (4.5) and (4.6), as well as Ẏ in the form of Ẏ =
Y1 ω1 cos(ω1 t) + Y2 ω2 cos(ω2 t) in (4.8), the flow force coefficient, Cx (t) = 2Fx /ρDU 2 L,
in the inline direction can be obtained as
https://doi.org/10.1017/jfm.2022.893 Published online by Cambridge University Press

       
Cx (t) = C1 sin ω1 t + 𝟇1 + C2 sin ω2 t + 𝟇2 + C3 sin 2ω1 t + 𝟇3 + C4 sin 2ω2 t + 𝟇4
   
+ C5 sin (ω1 + ω2 ) t + 𝟇5 + C6 sin (ω1 − ω2 ) t + 𝟇6 , (4.9)
in which C1 –C6 are functions of the lift and drag coefficients, frequencies of oscillations
and the phase difference between the flow force and the cylinder displacement. As
expressed in (4.9), the flow force in the inline direction consists of the first and second
frequencies and their second harmonics (terms with ω1 and ω2 and values twice those), as
well as other harmonics with frequencies of (ω1 + ω2 ) and (ω1 − ω2 ). These frequency
contents are shown for the sample cases in figure 10(b,c) for the two cases at which
the oscillations exhibit multi-modal behaviour over a range of reduced velocities. For
the single DOF system as well as the two DOF at the eigenfrequency of R = 2.1, the
oscillations of the cylinder exhibit a mono-frequency oscillation and, therefore, only the
dominant frequencies and their second harmonics at values twice the system’s natural
frequencies are observed. An advantage of having a two DOF set-up to study the VIV
response of the cylinder shows itself here as it has enabled us to observe, identify and
discuss the existence and the source of multiple frequency components in the VIV system
952 A26-20
VIV of a two DOF circular cylinder in the crossflow direction
(a) (b)
0.20 0.20
4 4
0.15 2 f2∗ 0.15
3 3 f1∗ + f2∗
f ∗C 2 f1∗ 0.10 2 f1∗ 0.10
x 2 2
f2∗
f1∗ f1∗
1 0.05 1 0.05
f2∗ – f1∗

0 0
0 2 4 6 8 10 12 14 16 0 2 4 6 8 10 12 14 16

(c) (d )
0.20 0.20
4 2 f2∗
4 2 f2∗
0.15 0.15
3 f1∗ + f2∗ 3
f ∗C 2 f1∗ 0.10 f2∗ 0.10
x 2 2 2 f1∗
f∗
f1∗ 2 0.05 f1∗ 0.05
1 1
f2∗ – f1∗
0 0
0 2 4 6 8 10 12 14 16 0 2 4 6 8 10 12 14 16
U∗ U∗
Figure 10. Frequency contents of the inline force coefficients for (a) the sample case of a single DOF
cylinder, and (b–d) three eigenfrequencies of R = 1.3, 1.7 and 2.1.

Fy

FL
U

Vrel FD

θ

Fx x
https://doi.org/10.1017/jfm.2022.893 Published online by Cambridge University Press

Figure 11. Flow forces acting on a cylinder undergoing VIV.

of the circular cylinder in our study in a straightforward fashion. These frequencies often
remain unnoticed when we study the complex multi-modal VIV response of a flexible
cylinder, for which the complexity of the response makes it very challenging and difficult
to separately study the frequency contents of the response as we did here and they mostly
remain unidentified or not discussed.

4.3. Wake structure


An overall map of shedding patterns observed in the wake of the circular cylinder
undergoing VIV in two DOF at the sample eigenfrequency ratio of R = 1.7 is given
in figure 12. This sample eigenfrequency ratio has been selected such that the cylinder
952 A26-21
S. Mousavisani and others

(a) U2∗ (b)


0 5 10 15
0.6

(viii)
(vii)
(iii)
(iv)

(vi)
(ii)

(v)
(i)
i 2S v 2S
0.5

0.4 ii 2S∗ vi 2S∗


y∗ 0.3

0.2 iii 2P vii 2P

0.1 iv SST viii 2S

0 5 10 15 20 25 30
U∗
Figure 12. (a) A map of the vortex patterns in the wake of the cylinder at the eigenfrequency ratio of R = 1.7
at different reduced velocities, together with (b) schematics of the shedding patterns.

exhibits mono-frequency oscillations at each of its first two eigenfrequencies over a


range of reduced velocities, and also experiences multi-modal oscillations over a range of
reduced velocity over which the VIV response of the system shifts from oscillations at the
dominant mode one to oscillations dominated by the second eigenfrequency of the system.
The predominant shedding patterns at lower reduced velocities, for which the VIV enters
the first lock-in region, are symmetric two single vortices shed per one cycle of oscillations,
known as 2S shedding (region i), that later transition to an asymmetric 2S shedding as the
amplitude of oscillation reaches the maximum amplitude (region ii). The asymmetry into
the shedding is introduced as a shear layer from one side of the cylinder gets stretched
and entrained into the wake of an already shed single vortex from the opposite side of
the cylinder. The asymmetric 2S shedding later develops into two pairs of vortices being
shed per cycle of oscillation, known as 2P vortex shedding. The 2S and its transition to 2P
shedding have been observed previously for the VIV response of a circular cylinder with a
single DOF to oscillate in the crossflow direction (Govardhan & Williamson 2000). As the
reduced velocity is further increased, the oscillations enter an overlapping range at which
both eigenfrequencies dominate the VIV response of the system (region iv). At this range
https://doi.org/10.1017/jfm.2022.893 Published online by Cambridge University Press

of reduced velocity, the shedding pattern per one cycle of oscillations exhibits two single
vortices being shed followed by a triplet of vortices, named here as SST shedding. Once
the VIV response enters the second lock-in region at which the oscillations are dominated
by the second eigenfrequency of the system, similar vortex shedding patterns as those
observed in the first lock-in region repeat themselves, i.e. 2S vortex shedding patterns at
lower reduced velocity (region v), that transition to 2P patterns at region (vii) through an
asymmetric 2S shedding pattern observed in region (vi). At high reduced velocity the VIV
response of the system drops to lower amplitudes and the response exits the second lock-in
region (region viii), the shedding pattern exhibits a 2S vortex shedding.
Figure 13 shows the vortex shedding patterns in the wake of the cylinder using PTV
and hydrogen bubble flow visualization, together with the time histories and frequency
content (FFT plots) of the VIV response of the cylinder at sample reduced velocities of
U ∗ = 5.4, 6.7 and 7.9. The snapshots on the left (the first column) are the instantaneous
out-of-plane vorticity plots obtained through the time-resolved 3-D PTV measurements,
as discussed in § 2 and the snapshots at the right are directly from hydrogen bubble
flow visualization (the second column). The vorticity plots are taken at a normalized
spanwise submerged length of z = 0.5, measured from the free bottom end of the cylinder.
952 A26-22
VIV of a two DOF circular cylinder in the crossflow direction
The time histories and their corresponding frequency contents shown on the third column
are plotted against dimensionless time and frequency, where the first eigenfrequency of
the system is used to obtain those values as f ∗ = fosc /fN and t∗ = t × fN . The sample
reduced velocities are selected such that they represent all the possible shedding patterns
that could occur over the first lock-in region, as discussed earlier in figure 12. The
hydrogen bubble flow visualization videos at these reduced velocities are also provided as
electronic supplementary movies 1–3 available at https://doi.org/10.1017/jfm.2022.893. At
the reduced velocity of U ∗ = 5.4, two single vortices are shed per one cycle of oscillations
(2S pattern). The vortex shedding frequency at this reduced velocity is equal to the
frequency of oscillations (fvs = fosc = 0.46 Hz at U ∗ = 5.4), which is a feature of a VIV
response. The frequency content of the response exhibits a single peak at the dimensionless
frequency close to unity, that represents a mono-frequency oscillation at values close to
the first eigenfrequency of the system. At the reduced velocity of U ∗ = 6.7, the 2S vortex
shedding pattern persists, but the shedding loses its symmetry, such that once a vortex is
shed from one side of the cylinder a shear layer from the other side of the cylinder gets
wound around the already shed vortex. As the cylinder moves to the opposite side, a second
vortex is fully shed. The extended shear layer from the bottom of the cylinder gets wound
around the shed vortex from the top, as also presented in the vorticity field measurements
shown in figure 13(e). The time history and the frequency content of the VIV response
at this reduced velocity still exhibit a mono-frequency oscillation at a frequency close to
the first eigenfrequency of the system. At the reduced velocity of U ∗ = 7.9, the response
is still in the first lock-in region and the shedding pattern observed in the wake of the
cylinder exhibits a pair of vortices being shed per one cycle of oscillation (2P shedding).
The pairs are both observed in the hydrogen bubble flow visualization and the vorticity
plots of figure 13. The time histories and the frequency content at this reduced velocity
exhibits a multi-modal oscillation; while the response is dominated by a frequency of
oscillation at the first eigenfrequency of the system, small contributions from the second
eigenfrequency of the system are observed at the dimensionless frequency of f ∗ = 1.7.
The time histories also exhibit their deviation from a single harmonic oscillatory response
to a periodic oscillation with more than one frequency.
Figure 14 shows the shedding patterns at three sample time stamps in one cycle of
oscillation of the cylinder, followed by the oscillation time histories and frequency content
at the reduced velocity of U ∗ = 9.8. This is a sample reduced velocity from region
https://doi.org/10.1017/jfm.2022.893 Published online by Cambridge University Press

(iv) of figure 12 at which the VIV response of the system is dominated by comparable
contributions from both eigenfrequencies of the system. The vortex shedding at this region
exhibits a unique pattern; in one full cycle of oscillation the cylinder sheds two single
vortices followed by a triplet of vortices (SST). This shedding pattern and the time stamps
at which each shedding occurs are presented in figure 14. The hydrogen bubble flow
visualization video is also provided as electronic supplementary movie 4.
Figure 15 shows the shedding patterns and the oscillation time histories and frequency
content at the reduced velocities of U ∗ = 10.1, 11.1, 13.0 and 17.7. At the reduced
velocity of U ∗ = 10.1, that is, a sample reduced velocity from region (v) from figure 12,
the VIV response of the system is dominated by comparable contributions from both
eigenfrequencies of the system. This region coincides with the onset of the second
lock-in region and the wake exhibits two single vortex shedding patterns per one cycle
of oscillation (2S). The 2S shedding pattern resembles those previously seen at the onset
on the first lock-in range in region (i). The vortex shedding frequency at this point is
equal to the frequency of oscillations in the second eigenfrequency (fvs = fosc = 0.82 Hz
at U ∗ = 10.1), which is a feature of a VIV response. At the reduced velocity of U ∗ = 11.1,

952 A26-23
S. Mousavisani and others
(a) (b) (c)
0.5

Velocity (m s–1) Vorticity ωz(S)


40 0.025
Single vortices (S) y∗ 0
0
20 –0.5
–0.025
(d)
0 5 10
Y (mm)

0 t∗
0.050
0.50
–20
0 |y∗| 0.25

–40 0
–0.050 1 2 3
–100 –80 –60 –40 –20 0 20 40 f∗
(e) (f) (g)

Velocity (m s–1) Vorticity ωz(S)


0.05 0.5
40
0 y∗ 0

20 –0.5
0 5 10
Y (mm)

Single vortex (S)


t∗
0
–0.05
(h)
Entrained shear layer 0.04 0.50
–20 |y∗| 0.25
0

–40 –0.04 0
1 2 3
f∗
–100 –80 –60 –40 –20 0 20 40
(i) ( j) (k)
Velocity (m s–1) Vorticity ωz(S)

0.08 0.5
40 y∗ 0
0 –0.5
20
0 5 10
Y (mm)

t∗
0 –0.08 (l)
A vortex pair (P) 0.50
0.05
–20 |y∗|
0 0.25

–40 –0.05 0
1 2 3
f∗
–100 –80 –60 –40 –20 0 20 40
X (mm)

Figure 13. Vortex shedding patterns in the wake of the cylinder using out-of-plane vorticity plots obtained
through the time-resolved 3-D PTV measurements (a,e,i), and hydrogen bubble flow visualization (b, f,j),
together with the time histories and frequency content (FFT plots) of the VIV response of the cylinder
(c,d,g,h,k,l) at sample reduced velocities of U ∗ = 5.4 (a–d), 6.7 (e–h) and 7.9 (i–l). The videos are provided as
supplementary movies 1–3.
https://doi.org/10.1017/jfm.2022.893 Published online by Cambridge University Press

that is, a sample reduced velocity from region (vi) from figure 12, mono-frequency
oscillations are observed at which the VIV response of the system exhibits frequency
contents at values close to the second eigenfrequency of the system. The wake of the
cylinder exhibits an asymmetric 2S vortex shedding pattern, where similar to region (ii) in
the first lock-in region, the symmetry of the two single vortices are broken. At the reduced
velocity of U ∗ = 13, that is, a sample reduced velocity from region (vii) of figure 12, the
vortex shedding patterns follow the same trend as those observed in the first lock-in region
where the asymmetric 2S shedding transitions to the 2P shedding pattern. The frequency
content of the dynamic response of the system exhibits mono-frequency oscillations at a
value close to the second eigenfrequency of the system. Finally, at the reduced velocity of
U ∗ = 17.7, that is, a sample reduced velocity from region (viii) of figure 12, the amplitude
of oscillations decreases to smaller values that mark the end of the second lock-in region.
At this reduced velocity the wake of the cylinder exhibits a very turbulent 2S vortex
shedding pattern. The PTV measurements represent two elongated shear layers being shed
from each side of the cylinder.
We would like to expand the wake structure discussion by pointing out the 3-D nature
of the wake of the cylinder undergoing VIV in our study. While the wake snapshots shown
952 A26-24
VIV of a two DOF circular cylinder in the crossflow direction
(a) (b)

Velocity (m s–1) Vorticity ωz(S)


40 0.05
0
20
–0.05
Y (mm)

0 Single vortex (S)


0.10
–20 0

–40
–0.10
–100 –80 –60 –40 –20 0 20 40
(c) (d ) (e)
t2

Velocity (m s–1) Vorticity ωz(S)


0.5
40 0.04
Single vortex (S) y∗ 0
0
20 –0.5 t1 t3
–0.04
Y (mm)

0 5 10
0 (f) t∗
0.10
0.50
–20
0 |y∗|
0.25
–40
–0.10 0
1 2 3
–100 –80 –60 –40 –20 0 20 40 f∗
(g) (h)
Velocity (m s–1) Vorticity ωz(S)

40 0.05
A triplet (T)
0
20
–0.05
Y (mm)

0
0.10
–20
0
–40
–0.10
–100 –80 –60 –40 –20 0 20 40
X (mm)

Figure 14. Vortex shedding patterns in the wake of the cylinder using out-of-plane vorticity plots obtained
through the time-resolved 3-D PTV measurements (a,c,g), and hydrogen bubble flow visualization (b,d,h),
together with the time histories and frequency content (FFT plots) of the VIV response of the cylinder (e, f ) at
sample reduced velocity of U ∗ = 9.8 at three time stamps during one cycle of oscillation of t1 (a,b), t2 (c–f )
and t3 (g,h). The video is provided as supplementary movie 4.

in figures 13–15 have been taken at one spanwise location (z = 0.5) along the length of
https://doi.org/10.1017/jfm.2022.893 Published online by Cambridge University Press

the cylinder, our time-resolved volumetric PTV measurements can provide insights into
the 3-D wake of the cylinder. To acquire further details of the vortex structures, the Q
criterion was implemented using both the time-averaged and instantaneous vector fields.
The Q criterion is defined as (Hunt, Wray & Moin 1988)

(Ω2 − S2 )
, Q= (4.10)
2
where Ω and S represent vorticity and strain rate in the flow field, respectively. Positive
Q values in the flow field indicate places where vorticity reigns supreme. Samples of the
normalized volumetric and two-dimensional (2-D) Q criterion are represented in figure 16
at three sample reduced velocities of U ∗ = 5.4, 9.8 and 13 for the eigenfrequency ratio of
R = 1.7. Positive values of Q distinguish dominant vortices in the wake of the cylinder.
The time-averaged volumetric Q-criterion plots (first column plots) show spanwise change
in the location of the shed vortex along the length of the cylinder, indicating that the
vortical structures in the wake of the cylinder undergoing VIV have a 3-D nature. The
three-dimensionality effects get more dominant at higher reduced velocities, as those
952 A26-25
S. Mousavisani and others
(a) (b) (c)
0.1 0.5

Velocity (m s–1) Vorticity ωz(S)


40
y∗ 0
Single vortices (S)
20 0 –0.5
Y (mm)

0 5 10
0 –0.1
(d) t∗
0.1 0.50
–20
0 |y∗| 0.25
–40
–0.1 0
–100 –80 –60 –40 –20 0 20 40 1 2 3
f∗
(e) (f) (g)
0.1

Velocity (m s–1) Vorticity ωz(S)


40 0.5
y∗ 0
0
20 –0.5
Y (mm)

–0.1 0 5 10
t∗
0 Single vortex (S)
0.1 (h)
Entrained shear layer 0.50
–20
0 |y∗|
0.25
–40 –0.1
0
–100 –80 –60 –40 –20 0 20 40 1 2 3
f∗
(i) (j) (k)
0.10
Velocity (m s–1) Vorticity ωz(S)

40 0.5
0 y∗ 0
20 –0.5
Y (mm)

–0.10 0 5 10
0 A vortex pair (P) 0.15 (l) t∗

–20 0.50
0 |y∗|
0.25
–40
–0.15 0
–100 –80 –60 –40 –20 0 20 40 1 2 3
f∗
(m) (n) (o)
0.15
Velocity (m s–1) Vorticity ωz(S)

40 0.5
0 y∗ 0
20
–0.5
Y (mm)

–0.15
0 5 10
0 0.20 (p) t∗
–20 0.50
0
|y∗| 0.25
–40
Single vortices (S)
–0.20
–100 –80 –60 –40 –20 0 20 40 0
https://doi.org/10.1017/jfm.2022.893 Published online by Cambridge University Press

1 2 3
X (mm) f∗

Figure 15. Vortex shedding patterns in the wake of the cylinder using out-of-plane vorticity plots obtained
through the time-resolved 3-D PTV measurements (a,e,i,m), and hydrogen bubble flow visualization (b, f,j,n),
together with the time histories and frequency content (FFT plots) of the VIV response of the cylinder
(a,d,g,h,k,l,o,p) at sample reduced velocities of U ∗ = 10.1 (a–d), 11.1 (e–h), 13.0 (i–l) and 17.7 (m–p).

shown in figure 16(e,h). While the wake exhibits a 3-D behaviour, it is worth noting that the
very same vortex shedding patterns discussed earlier in figure 12 still dominate the wake
of the cylinder. Figure 16(e, f ) shows instantaneous Q for a reduced velocity of U ∗ = 9.8,
for which the vortex shedding patterns were observed to be of SST-type (figure 14). A
triplet and two single vortices are observed in the wake of the cylinder at the instantaneous
Q contour at this reduced velocity. However, the triplet vortices are not observed in the
time-averaged Q-criterion contour (figure 16d). This can be explained by the fact that the
trailing triplet vortices in the wake of the cylinder are weaker than the leading two single
vortices in the repeated shedding pattern, resulting in dominant 2S patterns observed in the
time-averaged contour. At the reduced velocity of U ∗ = 13, the 2-D and 3-D instantaneous
952 A26-26
VIV of a two DOF circular cylinder in the crossflow direction
y

(a) z
(b) z (c) Q/(U/D)2 0 0.8 1.6 z x
Q/(U/D)2 0 0.1 0.2 Q/(U/D)2 0 0.4 0.8
x y
) x y
(S 50
es
r tic
vo

Y (mm)
le 0
ng
Si

–100
Z (mm)

20 20 –100 –50 Single vortices (S)


0 –50 0 –50
–20 –20
0 0

)
–100 –50 0 50 100
m
50 (m

)
–50 50

m
–50 X (mm)

(m
0
X

Y (mm 50 100 0 100

X
) Y (mm 50
)
z y
(d ) z
(e) (f)
Q/(U/D)2 0 0.05 0.10 Q/(U/D)2 0 0.45 0.90 Q/(U/D)2 0 1 2 z x
y x y
x )
(T
let 50 A triplet (T)
rt ip
A

Y (mm)
0
Z (mm)

20 –100 –100
20
0 –50 0 –50 –50
–20 –20
0 0
)
m

50
(m

)
m
–50 –50 50 –100 –50 0 50 100

(m
X

Y (mm0 50 100 0 100 X (mm)


X
) Y (mm 50
)
y
(g) z
(h) z
(i) Q/(U/D)2 0 1 2 z x
Q/(U/D)2 0 0.05 Q/(U/D)2 0 0.5 1.0
x y y
) x
(P
p air 50
tex
or
Y (mm)

v A vortex pair (P)


A 0
Z (mm)

20 –100
20 –100
0 –50 –50
0 –50
–20 0 –20
0
)
m

–50 50 0
(m

50 –100 –50 50 100


m

0 –50
(m

100
X

Y (mm 50 0 100 X (mm)


X

) Y (mm 50
)

Figure 16. Time-averaged 3-D (a,d,g), instantaneous 3-D (b,e,h) and 2-D (c, f,i) normalized Q-criterion
contours at reduced velocities of U ∗ = 5.4 (a–c), U ∗ = 9.8 (d–f ) and U ∗ = 13 (g–i).

Q-criterion contours show the dominant 2P vortex shedding patterns in the wake of the
https://doi.org/10.1017/jfm.2022.893 Published online by Cambridge University Press

cylinder, as also expected from previous observations in figure 12. Since the shed vortices
are advected downstream, the 2P vortex shedding pattern is no longer present in the
time-averaged Q-criterion contour (figure 16g).

4.4. Proper orthogonal decomposition


Proper orthogonal decomposition is one of the most robust techniques to investigate the
turbulent flow’s large (energetic) coherent structures (Berkooz, Holmes & Lumley 1993;
Lumley 2007; Sieber, Paschereit & Oberleithner 2016; Taira et al. 2017). The unsteady
flow field, u, could be represented as a long-time mean field, Ū, and fluctuation flow, u ,
via the general Reynolds decomposition as

u = Ū + u . (4.11)

In the current study, the method of snapshots POD introduced in Sirovich (1987) is used
to present the fluctuation of a vector field using a linear combination of orthonormal modes
to solve an eigenvalue problem. The snapshots approach decomposes the time-dependent
952 A26-27
S. Mousavisani and others
(a) (b) (c)
30 30 30
25 25 25
Energy %

20 20 20
15 15 15
10 10 10
5 5 5

0 10 20 30 40 50 60 0 10 20 30 40 50 60 0 10 20 30 40 50 60
POD mode POD mode POD mode
Figure 17. Energy distribution for the first 60 POD modes at the reduced velocities of (a) U ∗ = 5.4, (b)
U ∗ = 9.8 and (c) U ∗ = 13.

fluctuation velocity field, ui (x, t), into the unsteady spatial basis functions of 𝟇k (x) and
their corresponding temporal modal coefficients of ak (t) as
N
ui (x, t) = ak (t)𝟇k (x), (4.12)
k=1
with N being the number of snapshots employed in the calculation. In the current study, the
POD method is applied to the resulting 3-D velocity field obtained via the time-resolved
3-D PTV measurements. To obtain spatial (𝟇k (x)) and temporal (ak (t)) POD modes, single
value decomposition (SVD) was applied to the fluctuation velocity field (u − Ū)) as (Horn
& Johnson 1985; Trefethen & Bau 1997; Saad 2011):

(𝟇, λ, a) = SVD(u − Ū), (4.13)
where the eigenvalues λ represent the collected energy from POD modes 𝟇(x). A zeroth
mode is assumed for the time-averaged flow field, which is then subtracted from each
flow field signal. Figure 17 shows each mode’s relative kinetic energy contribution as a
percentage of the total energy at the three sample reduced velocities studied before. In the
current work, to identify unstable wake flow by POD, the criteria to decide the number of
contributing modes is set at 5 % of the total energy of modes. It is shown that the first few
modes contribute to the majority of energy of the system in the samples reduced velocities
selected, whereas the energy proportion of higher modes steadily drops to zero. In this
https://doi.org/10.1017/jfm.2022.893 Published online by Cambridge University Press

study, the first four POD modes, containing higher than 5 % of the total energy levels,
have been selected to be investigated in the following section.
The three-dimensional spatial structure and corresponding temporal coefficients of the
first four POD modes in terms of the streamwise and transverse velocity components at
the three selected sample reduced velocities of U ∗ = 5.4, 9.8 and 13 at the eigenfrequency
ratio of R = 1.7 are shown in figures 18–20, respectively. The sample reduced velocities
are selected such that three distinct regions of the modal contributions to the VIV response
of the system, i.e. pure modes one and two and the multi-modal VIV response, are
covered. As figure 17 shows, at the reduced velocity of U ∗ = 5.4, where the 2S shedding
pattern was identified, the first two POD modes contain a dominant percentage of the
total energy (55 %), and the third and fourth had less than 3 % energy. However, at the
reduced velocities of U ∗ = 9.8 and 13, at which the SST and 2P shedding patterns were
identified, while the first two POD modes contain approximately 36 % and 48 % of the
total energy of the system, respectively, the third and fourth POD modes are still energetic
with contributions of around 14 % to the total energy of the system. Figure 20 shows the
corresponding temporal evolution of the first four POD modes at the same sample reduced
velocities versus dimensionless time (figures 20a1–a4, b1–b4 and c1–c4), together with
952 A26-28
VIV of a two DOF circular cylinder in the crossflow direction
(a) z (b) z (c) z (d ) z
u (m s–1) –0.5 u (m s–1) –0.4
0 0.5 0 0.4 u (m s–1) –0.12 0 0.12 u (m s–1) –0.1 0 0.1
y
x y x y y x
x

–100 20 –100 –100 20 –100


Z (mm)

20 20
0 –50 0 –50 0 –50 0 –50
–20 0 –20 0 –20 –20 0
0

)
m

m
m
50 (m 50

(m
50

(m
50

(m
–50 –50 –50 –50
X

X
0 0 0

X
Y (m 100 Y (m 100 0 100 Y (m 100
m) 50 m) 50 Y (m 50 m) 50
m)

(e) u (m s–1) –0.8 0


z
(f) z (g) z (h) z

0.8 u (m s–1) –0.8 0 0.8 u (m s–1) –0.8 0 0.8 u (m s–1) –0.6 0 0.6
y y y x y
x x x

20 –100 20 –100 –100 20 –100


Z (mm)

20
0 –50 0 –50 0 –50 0 –50
–20 0 –20 –20 0 –20 0
0

)
)

m
m

m
50

(m
50

(m
50 50
(m

(m
–50 –50 –50 –50

X
X

0 0 X 0 100 0 100
Y (m 100 Y (m 100 Y (m 50 Y (m 50
m) 50 m) 50 m) m)

(i) u (m s–1) –1.5 0 1.5


z
( j) u (m s –1)
–1 0 1
z
(k) u (m s–1) –0.7 0 0.7
z (l) u (m s–1) –0.6 0 0.6
z

x y y
x x y x y

20 –100 20 –100 20 –100 20 –100


Z (mm)

0 –50 0 –50 0 –50 0 –50


–20 0 –20 0 –20 0 –20 0

)
)

m
m

50

(m
50 50

(m
50
(m

(m

–50 –50 –50 –50

X
X

0 0 0 100 0 100
Y (m 100 Y (m 100 Y (m 50 Y (m 50
m) 50 m) 50 m) m)

Figure 18. Spatial structures of the streamwise velocity for the first four POD modes (shown in each column
from mode one to four) at the reduced velocities of U ∗ = 5.4 (a–d), U ∗ = 9.8 (e–h) and U ∗ = 13 (i–l).

their frequency content shown through FFT plots (figures 20a5–a8, b5–b8 and c5–c8). The
time and frequency in these plots are normalized by the system’s first natural frequency.
Scalogram plots (figures 20a9–a12, b9–b12 and c9–c12), obtained using the wavelet
transform of the associated temporal coefficients time series, show the time evolution of
the frequency contents normalized by the system’s first natural frequency, that is plotted
against the dimensionless time, t∗ .
The first sample reduced velocity of U ∗ = 5.4 is selected at the region where the
https://doi.org/10.1017/jfm.2022.893 Published online by Cambridge University Press

VIV response of the system experiences oscillations at only mode one (figures 18 and
19a–d). The first and second POD modes show the analogous spatial structures and
shift to the downstream in the streamwise direction. Furthermore, the first and second
POD modes illustrate the travelling wave’s spatiotemporal structure, which is created
as a result of vortices being shed from each side of the cylinder to the downstream.
Asymmetric and symmetric patterns about the centreline of the spatial structure are
observed in the first two modes in the streamwise and spanwise directions, respectively.
Asymmetric velocity field distributions around the centreline indicate symmetric vorticity
contour distributions, whereas symmetric patterns represent the mirrored symmetric ones
(Konstantinidis, Balabani & Yianneskis 2007). Therefore, the first two POD modes at this
reduced velocity represent a symmetric 2S vortex shedding pattern. Figure 20(a1,a2,a5,a6)
also shows the time history and FFT analysis of temporal coefficients for the first two
modes at this reduced velocity. These coefficients fluctuate with a single frequency content
with a sharp peak at the dominant frequency (f ∗ = 1), that is the same as the cylinder’s
first natural frequency, as shown in the frequency contents of FFT plots. This frequency
corresponds to the vortex shedding frequency at this reduced velocity, that is also the same
as the cylinder’s oscillation frequency (shown with dotted lines). In contrast, the third and
952 A26-29
S. Mousavisani and others
(a) v (m s–1) –1 0 1
z
(b) v (m s–1) –0.6 0 0.6
z
(c) v (m s –1)
–0.1 0
z
(d ) v (m s –1) –0.1 0 0.1
z
0.1
x y x y x y x y

20 –100 20 –100 –100 20 –100


Z (mm)

20
0 –50 0 –50 0 –50 0 –50
–20 0 –20 0 –20 –20
0 0

)
m

m
50 50

(m

(m
50 50

(m

(m
–50 –50 –50 –50
0

X
0 0

X
Y (m 100 100 0
m) 50 Y (m 50 Y (m 50 100 Y (m 50 100
m) m) m)

(e) z (f) z (g) z (h) z


v (m s–1) –1.5 0 1.5 v (m s–1) –1 0 1 v (m s–1) –0.6 0 0.6 v (m s–1) –0.5 0 0.5
y x y x y
x x y
Z (mm)

20 –100 20 –100 20 –100 20 –100


0 –50 0 –50 0 0
–50 –50
–20 –20 –20 –20
0 0 0 0
)

)
m

m
50 50
(m

(m
50 50

(m

(m
–50 –50 –50 –50
0
X

X
Y (m 100 0 100 0 0
m) 50 Y (m 50 Y (m 50 100 Y (m 50 100
m) m) m)

(i) z
( j) z (k) z (l) z
v (m s–1) –2 0 2 v (m s–1) –1 0 1 v (m s–1) –0.7 0 0.7 v (m s–1) –0.6 0 0.6
x y y y
x y x
x

20 20 20 –100 20 –100
Z (mm)

–100 –100
0 0 0 –50 0 –50
–50 –50
–20 –20 –20 –20
0 0 0 0

)
)

m
)

m
m

50 50

(m
(m
50 50 –50 –50
(m

(m

–50 –50
0 0

X
X
0 0
X

Y (m 100 Y (m 100 Y (m 50 100 Y (m 50 100


m) 50 m) 50 m) m)

Figure 19. Spatial structures of the transverse velocity for the first four POD modes (shown in each column
from mode one to four) at the reduced velocities of U ∗ = 5.4 (a–d), U ∗ = 9.8 (e–h) and U ∗ = 13 (i–l).

fourth POD modes show irregular spatial (figures 18 and 19c,d) and temporal patterns
(figure 20a3,a4,a7,a8) associated with small-scale turbulent structures that exhibit several
low-frequency peaks in the frequency content of the temporal coefficient modes. The
scalograms of the associated temporal coefficient modes also show that, for the first and
second modes, the vortex shedding frequency stays close to the first natural frequency
of the cylinder during the oscillation (figure 20a9,a10). On the other hand, for the third
and fourth temporal coefficients with lower levels of energy, the frequency contents are
smaller than the cylinder’s first natural frequency and they show temporal fluctuations in
https://doi.org/10.1017/jfm.2022.893 Published online by Cambridge University Press

their frequency contents (figure 20a11,a12).


The second sample reduced velocity of U ∗ = 9.8 is selected such that the VIV response
of the system exhibits multi-modal oscillations that transition from oscillations at mode
one to mode two. The two sharp peaks in the frequency content of the structure’s
oscillation response shown by the dotted lines in the FFT plots (figure 20b5–b8) represent
such a multi-modal response with contributions from both modes with oscillation
frequencies at the values close to the first and second natural frequencies of the system. The
first two POD modes (figures 18 and 19e–h) represent almost identical structures of vortex
shedding. As previously shown in figure 14, at this reduced velocity the wake exhibited
the SST shedding pattern. Analysing the POD modes here, it is observed that the first two
POD modes behave similar to those at the reduced velocity of U ∗ = 5.4, at which the
2S vortex shedding pattern was observed. On the other hand, the third and fourth POD
modes are significantly different. The third spatial structure in the streamwise direction
shows nearly two parallel vortical structures, whereas the fourth POD mode exhibits
uneven small-scale turbulence structures as the energy level at this mode is relatively
low (figure 17). Additionally, the symmetric spatial pattern, as observed for the first two
POD modes, is replaced with an erratic pattern in the transverse direction of the third and
952 A26-30
VIV of a two DOF circular cylinder in the crossflow direction

(a1) (a2) (a3) (a4)


0.1 1 0.1 1 0.1 1 0.1 1
a1 0 0 y∗ a2 0 0 y ∗ a3 0 0 y∗ a4 0 0 y∗
–0.1 –1 –0.1 –1 –0.1 –1 –0.1 –1
0 2 4 6 8 0 2 4 6 8 0 2 4 6 8 0 2 4 6 8
t∗ t∗ t∗ t∗
(a5) (a6) (a7) (a8)
0.05 0.7 0.05 0.7 0.05 0.7 0.05 0.7
|a1| |y∗| |a2| |y∗| |a3| |y∗| |a4| |y∗|
0 0 0 0
0 1 2 3 0 1 2 3 0 1 2 3 0 1 2 3
f∗ f∗ f∗ f∗
(a9) (a10) (a11) (a12)
3 0.025 3 0.025 3 0.025 3 0.025
f∗ 2 f∗ 2 f∗ 2 f∗ 2
1 0.005 1 0.005 1 0.005 1 0.005
0 2 4 6 8 0 2 4 6 8 0 2 4 6 8 0 2 4 6 8
t∗ t∗ t∗ t∗
(b1) (b2) (b3) (b4)
0.1 1 0.1 1 0.1 1 0.1 1
a1 0 0 y∗ a2 0 0 y ∗ a3 0 0 y∗ a4 0 0 y∗
–0.1 –1 –0.1 –1 –0.1 –1 –0.1 –1
0 2 4 6 8 0 2 4 6 8 0 2 4 6 8 0 2 4 6 8
t∗ t∗ t∗ t∗
(b5) (b6) (b7) (b8)
0.05 0.7 0.05 0.7 0.05 0.7 0.05 0.7
|a1| |y∗| |a2| |y∗| |a3| |y∗| |a4| |y∗|
0 0 0 0
0 1 2 3 0 1 2 3 0 1 2 3 0 1 2 3
f∗ f∗ f∗ f∗
(b9) (b10) (b11) (b12)
3 0.025 3 0.025 3 0.025 3 0.025
f∗ 2 f∗ 2 f∗ 2 f∗ 2
1 0.005 1 0.005 1 0.005 1 0.005
0 2 4 6 8 0 2 4 6 8 0 2 4 6 8 0 2 4 6 8
t∗ t∗ t∗ t∗

(c1) (c2) (c3) (c4)


0.1 1 0.1 1 0.1 1 0.1 1
a1 0 0 y∗ a2 0 0 y ∗ a3 0 0 y∗ a4 0 0 y∗
https://doi.org/10.1017/jfm.2022.893 Published online by Cambridge University Press

–0.1 –1 –0.1 –1 –0.1 –1 –0.1 –1


0 2 4 6 8 0 2 4 6 8 0 2 4 6 8 0 2 4 6 8
t∗ t∗ t∗ t∗
(c5) (c6) (c7) (c8)
0.05 0.7 0.05 0.7 0.05 0.7 0.05 0.7
|a1| |y∗| |a2| |y∗| |a3| |y∗| |a4| |y∗|
0 0 0 0
0 1 2 3 4 5 0 1 2 3 4 5 0 1 2 3 4 5 0 1 2 3 4 5
f∗ f∗ f∗ f∗
(c9) (c10) (c11) (c12)
5 0.025 5 0.025 5 0.025 5 0.025
f∗ 3 f∗3 f∗3 f∗3
1 0.005 1 0.005 1 0.005 1 0.005
0 2 4 6 8 0 2 4 6 8 0 2 4 6 8 0 2 4 6 8
t∗ t∗ t∗ t∗
Figure 20. Time evolution of the first four temporal coefficients (a1 –a4 ), their corresponding frequency
contents (FFT plots) and scalogram plots at the reduced velocities of U ∗ = 5.4 (box 1 – rows 1–3), U ∗ = 9.8
(box 2 – rows 4–6) and U ∗ = 13 (box 3 – rows 7–9). Dotted lines in the time histories and FFT plots
represent the cylinder’s amplitude of oscillations (y∗ ) and their frequency contents (|y∗ |), respectively, at the
corresponding reduced velocities.
952 A26-31
S. Mousavisani and others
fourth POD modes. Similarly, corresponding temporal coefficients show the multi-modal
behaviour in which the dominant frequency for the first and second temporal modes is
close to the cylinder’s second natural frequency (figure 20b1,b2,b5,b6). In contrast, the
time history and frequency contents of the third and fourth temporal modes show that
the dominant vortex shedding frequency is close to the cylinder’s first natural frequency
(figure 20b3,b4,b7,b8). Such a behaviour confirms that the first two modes with the
highest energy levels (figure 17b) and high shedding frequency can be associated with
the shedding of single vortices while the third mode with the moderate level of energy and
low frequency (close to system’s first natural frequency) corresponds to the shedding of
triplet vortices. The fourth mode with a relatively lower energy level and low frequency
can be associated with the small-scale turbulence structures in the wake of the cylinder.
The scalograms, representing the time evolution of frequency content for the temporal
coefficients, confirm the multi-modal oscillations of the first four temporal coefficient
at the reduced velocity of U ∗ = 9.8 with the shedding frequency values close to the
cylinder’s first and second natural frequencies (figure 20b9–b12).
The third sample reduced velocity of U ∗ = 13 is selected such that the VIV response
of the system exhibits oscillations at the second structural mode, with the frequency
content represented by a sharp peak near the system’s second natural frequency (shown
by dotted lines in figure 20c5–c8). The third row figures (figures 18–19i–l) show spatial
patterns of the first four POD modes at this sample reduced velocity. The first two POD
modes in the streamwise and transverse directions show similar asymmetric and symmetric
vortical structures, respectively. The spatial structure of the third and fourth POD modes
exhibits two small-scale parallel structures with opposing signs that are asymmetric
in both the streamwise and transverse directions. The temporal POD modes at this
reduced velocity show a mono-frequency behaviour in which frequency contents and the
corresponding time evolution for the first two temporal coefficients (figure 20c1,c2,c5,c6)
show a strong peak with its value close to the system’s second natural frequency. The
dominant frequency is doubled for the third and fourth POD modes compared with the
first and second temporal modes (figure 20c3,c4,c7,c8). As suggested by previous studies
(Konstantinidis et al. 2007), these POD modes are attributed to the 2P symmetrical vortex
structures in the wake of the cylinder. Scalogram plots at this reduced velocity indicate
that the frequency of the first and second temporal mode coefficients stays close to the
system’s second natural frequency over time (figure 20c9,c10), and doubles for the third
https://doi.org/10.1017/jfm.2022.893 Published online by Cambridge University Press

and fourth coefficients (figure 20c11,c12).


In summary, the aforementioned spatiotemporal mode analysis using the POD technique
shows that vortex shedding modes in the wake of the cylinder are strongly coupled with
the structural vibration modes, as discussed earlier in § 4.1. Both approaches of wavelet
transform and the POD produced a strikingly similar frequency content for the structural
oscillations response and the vortex shedding in the wake of the cylinder, in which the
peak frequencies for the 3-D POD modes are well in agreement with those reported in
the frequency analysis of the structural VIV response. Here, we demonstrated that the 3-D
POD technique can well capture the proper vortex shedding modes and their corresponding
frequency contents for a circular cylinder undergoing VIV. The POD technique used in
conjunction with the structural response analysis techniques in our study provided us
with a unique tool to better understand the VIV response of a circular cylinder and the
associated complex and 3-D vortex dynamics in the wake of the cylinder. These combined
techniques have potentials to be extended to studies of complex fluid–structure interaction
systems, for which a strong coupling between the structural response and the surrounding
complex and 3-D fluid flow exists.

952 A26-32
VIV of a two DOF circular cylinder in the crossflow direction
5. Conclusion
An experimental study of VIV of a flexibly mounted rigid cylinder, free to oscillate in the
crossflow direction at its first two modes, was conducted.
Dynamic response of the cylinder, in terms of amplitude of oscillations, flow
forces and wake structures were studied for a range of reduced velocities of U ∗ =
4–30, corresponding to a Reynolds number range of Re = 1033–7478 for a range of
eigenfrequency ratios of the system. The VIV response of the system was shown to be
heavily dependant on the ratio between the first two natural frequencies of the system
and how far away they are from each other; for frequency ratios as small as R = 1.3,
the onset of the VIV response of the system was very similar to that observed for a
classical VIV response of a single DOF circular cylinder. However, the lock-in width
was extended to higher reduced velocities as the frequency ratios increased. The extended
lock-in range was split between two regions where, for the first region at low reduced
velocities, mono-frequency oscillations at the first mode were observed. As the reduced
velocity was increased, the VIV response of the system transitioned to a range at which
multi-modal oscillations dominated the response of the system. By further increasing the
reduced velocity, the second lock-in region was observed, over which the VIV response
of the system was dominated by the mono-frequency excitation at the second mode of
the system. The modal contributions also exhibited oscillations that shifted from pure
mode one at low reduced velocities to a range of intermediate reduced velocities at
which multi-modal oscillations occurred. By increasing the reduced velocity values, the
contributions from mode one decreased to zero and the VIV response got dominated by
pure mode two oscillations. While, for the eigenfrequency ratios below R = 2.1, the VIV
response of the system followed this behaviour, for relatively larger eigenfrequency ratios
of R = 2.5 and 3.0, the two lock-in regions were well separated from each other such
that there did not exist any intermediate reduced velocity range at which multi-modal
oscillations occurred.
The crossflow and inline flow force coefficients were also studied for the range of
reduced velocities and eigenfrequency ratios tested here. The crossflow force coefficients
followed similar trends to those observed for the amplitudes of oscillation where their
values increased to a maximum at reduced velocities where maximum oscillations
occurred. By further increasing the reduced velocity, the flow force coefficients were
https://doi.org/10.1017/jfm.2022.893 Published online by Cambridge University Press

dropped in their values at the transition region between mode one and mode two. At the
reduced velocity range over which the second mode started to dominate the response of the
system, the crossflow force coefficients picked up in their magnitudes and reached a second
maximum value. By further increasing the reduced velocity, when the VIV response of the
system dropped to lower amplitudes and exited the lock-in region, the crossflow force
coefficient followed a similar decrease in value and got to zero when the oscillations
stopped. In the inline direction, flow forces consisted of frequencies at the values close to
the system’s natural frequencies and their corresponding second harmonics, as expected
from the classical VIV response of a single DOF system. However, for the range of reduced
velocities and eigenfrequencies that the VIV response of the system exhibited multi-modal
oscillations, other frequency components were also present in the frequency content of the
inline forces. Existence of such harmonics were studied and related to the relative velocity
of the cylinder as it underwent a multi-frequency VIV response.
Both qualitative and quantitative flow visualization tests were conducted to study vortex
dynamics in the wake of the cylinder. For qualitative flow visualization purposes, the
hydrogen bubble visualization technique was used. For quantitative flow measurement

952 A26-33
S. Mousavisani and others
purposes, the time-resolved volumetric PTV technique was used to measure the flow field
dynamics in the wake of the cylinder at sample reduced velocities. The vortex shedding
patterns observed in the wake of the cylinder at each lock-in region followed those
expected for a classical single DOF VIV response, where at the low reduced velocities
two single vortices were shed per one cycle of oscillation (2S shedding pattern). This
pattern was followed by a pair of vortices shed per one cycle of oscillation (2P shedding
pattern). For the intermediate reduced velocity range over which the response of the
system transitioned from mode one oscillation to mode two, a new shedding pattern was
observed where two single vortices and a triplet of vortices were shed in one cycle of
oscillation (SST shedding pattern). These shedding patterns were also confirmed from
the instantaneous vorticity field information obtained through the time-resolved PTV
measurements.
The Q criterion was used to demonstrate the 3-D characteristics of the cylinder’s wake.
The 2-D and 3-D visualizations showed 2S and 2P patterns in the wake of the cylinder at
the two selected reduced velocities of U ∗ = 5.4 and 13, confirming our qualitative wake
visualizations using the hydrogen bubble flow visualizations, as well as those reported in
instantaneous PTV vorticity field. At the sample reduced velocity of U ∗ = 9.8, the 3-D
time-averaged Q criterion revealed a 2S structure; however, the instantaneous 3-D and
2-D contours revealed a triplet vortex pattern that were not presented in the time-averaged
contour. This was attributed to the fact that the trailing triplet vortices in the wake of the
cylinder were weaker than the leading two single vortices in the repeated shedding pattern,
resulting in dominant 2S patterns observed in the time-averaged contour.
Additionally, the POD technique was implemented to extract the prevalent shedding
modes in the wake of the cylinder undergoing VIV. The POD analysis illustrates the
differences in the energy of each mode, temporal coefficient and, most importantly, the
dominant frequency of the POD modes. At the sample reduced velocity of U ∗ = 5.4,
the first two POD modes exhibited an asymmetric vortex mode, similar to those of a
classical Kármán vortex street. The wavelet analysis of the temporal POD coefficients
at this reduced velocity revealed a synchronization of vortex shedding frequency and
the cylinder’s oscillation frequency. At the reduced velocity of U ∗ = 9.8, the first two
POD modes were similar in their shedding frequency having values at the second natural
frequency of the system, while the third and fourth POD modes were linked to the first
natural frequency of the system. At the third sample reduced velocity of U ∗ = 13, the
https://doi.org/10.1017/jfm.2022.893 Published online by Cambridge University Press

doubling in the frequency of the third and fourth POD modes illustrated the 2P vortex
structure in the wake of the cylinder that were well in agreement with the qualitative
hydrogen bubble flow visualizations discussed in this study.
Supplementary movies. Supplementary movies are available at https://doi.org/10.1017/jfm.2022.893.

Funding. This research work is supported by University of Massachusetts Dartmouth’s Marine and Undersea
Technology (MUST) Research Program funded by the Office of Naval Research (ONR) under Grant No.
N00014-20-1-2170.

Declaration of interests. The authors report no conflict of interest.

Author ORCIDs.
Hadi Samsam-Khayani https://orcid.org/0000-0001-6416-1135;
Hamed Samandari https://orcid.org/0000-0001-8312-8509;
Banafsheh Seyed-Aghazadeh https://orcid.org/0000-0002-4260-7084.

Author contributions. S.M. and N.N.C. contributed equally to this work.

952 A26-34
VIV of a two DOF circular cylinder in the crossflow direction
R EFERENCES
A RDEH , H.A. & A LLEN , M.S. 2013 Investigating cases of jump phenomenon in a nonlinear oscillatory
system. In Topics in Nonlinear Dynamics (ed. G. Kerschen, D. Adams & A. Carrella), vol. 1, pp. 299–318.
Springer.
B EARMAN , P.W. 1984 Vortex shedding from oscillating bluff bodies. Annu. Rev. Fluid Mech. 16, 195–222.
B ERKOOZ , G., H OLMES , P. & LUMLEY, J.L. 1993 The proper orthogonal decomposition in the analysis of
turbulent flows. Annu. Rev. Fluid Mech. 25 (1), 539–575.
BOURGUET, R., K ARNIADAKIS , G.E. & T RIANTAFYLLOU, M.S. 2011a Vortex-induced vibrations of a long
flexible cylinder in shear flow. J. Fluid Mech. 677, 342–382.
BOURGUET, R., K ARNIADAKIS , G.E. & T RIANTAFYLLOU, M.S. 2013 Distributed lock-in drives broadband
vortex-induced vibrations of a long flexible cylinder in shear flow. J. Fluid Mech. 717, 361–375.
BOURGUET, R., LUCOR , D. & T RIANTAFYLLOU, M.S. 2012 Mono-and multi-frequency vortex-induced
vibrations of a long tensioned beam in shear flow. J. Fluids Struct. 32, 52–64.
BOURGUET, R., M ODARRES -SADEGHI , Y., K ARNIADAKIS , G.E. & T RIANTAFYLLOU, M.S. 2011b
Wake-body resonance of long flexible structures is dominated by counterclockwise orbits. Phys. Rev. Lett.
107 (13), 134502.
C ARLSON , D.W., C URRIER , T.M. & M ODARRES -SADEGHI , Y. 2021 Flow-induced vibrations of a square
prism free to oscillate in the cross-flow and inline directions. J. Fluid Mech. 919, A2.
C HAPLIN , J.R. & BATTEN , W.M.J. 2014 Simultaneous wake- and vortex-induced vibrations of a cylinder
with two degrees of freedom in each direction. Trans. ASME J. Offshore Mech. Arctic Engng 136 (3),
031101.
C HAPLIN , J.R., B EARMAN , P.W., H UARTE , F.J.H. & PATTENDEN , R.J. 2005 Laboratory measurements of
vortex-induced vibrations of a vertical tension riser in a stepped current. J. Fluids Struct. 21 (1), 3–24.
C HEN , L., W U, J. & C HENG , B. 2019 Volumetric measurement and vorticity dynamics of leading-edge vortex
formation on a revolving wing. Exp. Fluids 60 (1), 12.
DAHL , J.M., H OVER , F.S. & T RIANTAFYLLOU, M.S. 2006 Two-degree-of-freedom vortex-induced
vibrations using a force assisted apparatus. J. Fluids Struct. 22 (6–7), 807–818.
DAHL , J.M., H OVER , F.S., T RIANTAFYLLOU, M.S., D ONG , S. & K ARNIADAKIS , G.E.M. 2007 Resonant
vibrations of bluff bodies cause multivortex shedding and high frequency forces. Phys. Rev. Lett. 99 (14),
144503.
D U, L., J ING , X. & S UN , X. 2014 Modes of vortex formation and transition to three-dimensionality in the
wake of a freely vibrating cylinder. J. Fluids Struct. 49, 554–573.
E VANGELINOS , C., LUCOR , D. & K ARNIADAKIS , G.E. 2000 DNS-derived force distribution on flexible
cylinders subject to vortex-induced vibration. J. Fluids Struct. 14 (3), 429–440.
G EDIKLI , E.D., C HELIDZE , D. & DAHL , J.M. 2018 Observed mode shape effects on the vortex-induced
vibration of bending dominated flexible cylinders simply supported at both ends. J. Fluids Struct.
81, 399–417.
G OVARDHAN , R. & W ILLIAMSON , C.H.K. 2000 Modes of vortex formation and frequency response of a
freely vibrating cylinder. J. Fluid Mech. 420, 85–130.
https://doi.org/10.1017/jfm.2022.893 Published online by Cambridge University Press

G OVARDHAN , R. & W ILLIAMSON , C.H.K. 2002 Resonance forever: existence of a critical mass and an
infinite regime of resonance in vortex-induced vibration. J. Fluid Mech. 473, 147–166.
H ORN , R.A. & J OHNSON , C.R. 1985 Matrix Analysis. Cambridge University Press.
H UERA-H UARTE , F.J. & B EARMAN , P.W. 2009 Wake structures and vortex-induced vibrations of a long
flexible cylinder–Part 1: Dynamic response. J. Fluids Struct. 25 (6), 969–990.
H UNT, J.C.R., W RAY, A.A. & M OIN , P. 1988 Eddies, streams, and convergence zones in turbulent flows. In
Studying turbulence using numerical simulation databases, 2. Proceedings of the 1988 summer program.
I NMAN , D.J. & S INGH , R.C. 1994 Engineering Vibration, vol. 3. Prentice Hall.
JAUVTIS , N.A. & W ILLIAMSON , C.H.K. 2004 The effect of two degrees of freedom on vortex-induced
vibration at low mass and damping. J. Fluid Mech. 509, 23–62.
K HALAK , A. & W ILLIAMSON , C.H.K. 1999 Motions, forces and mode transitions in vortex-induced
vibrations at low mass-damping. J. Fluids Struct. 13 (7), 813–851.
KONSTANTINIDIS , E., BALABANI , S. & Y IANNESKIS , M. 2007 Bimodal vortex shedding in a perturbed
cylinder wake. Phys. Fluids 19 (1), 011701.
L IE , H. & K AASEN , K.E. 2006 Modal analysis of measurements from a large-scale VIV model test of a riser
in linearly sheared flow. J. Fluids Struct. 22 (4), 557–575.
LUMLEY, J.L. 2007 Stochastic Tools in Turbulence. Courier Corporation.
M ARCOLLO , H., E ASSOM , A., F ONTAINE , E., T OGNARELLI , M., B EYNET, P., C ONSTANTINIDES , Y. &
OAKLEY, O.H. J R . 2011 Traveling wave response in full-scale drilling riser VIV measurements. In 30th
International Conference on Offshore Mechanics and Arctic Engineering, vol. 44397, pp. 523–537.

952 A26-35
S. Mousavisani and others
N ISHI , Y. 2013 Power extraction from vortex-induced vibration of dual mass system. J. Sound Vib. 332 (1),
199–212.
N ISHI , Y., F UKUDA , K. & S HINOHARA , W. 2017 Experimental energy harvesting from fluid flow by using
two vibrating masses. J. Sound Vib. 394, 321–332.
R AGHAVAN , K. & B ERNITSAS , M.M. 2011 Experimental investigation of Reynolds number effect on vortex
induced vibration of rigid circular cylinder on elastic supports. Ocean Engng 38 (5), 719–731.
SAAD , Y. 2011 Numerical Methods for Large Eigenvalue Problems. Society for Industrial and Applied
Mathematics.
SARPKAYA , T. 1995 Hydrodynamic damping, flow-induced oscillations, and biharmonic response. Trans.
ASME J. Offshore Mech. Arctic Engng 117 (4), 232–238.
SARPKAYA , T. 2004 A critical review of the intrinsic nature of vortex-induced vibrations. J. Fluids Struct. 19
(4), 389–447.
S CHANZ , D., G ESEMANN , S. & S CHRÖDER , A. 2016 Shake-the-box: Lagrangian particle tracking at high
particle image densities. Exp. Fluids 57 (5), 70.
S EYED -AGHAZADEH , B., B ENNER , B., G JOKOLLARI , X. & M ODARRES -SADEGHI , Y. 2021 An
experimental investigation of vortex-induced vibration of a curved flexible cylinder. J. Fluid Mech. 927,
A21.
S EYED -AGHAZADEH , B., C ARLSON , D.W. & M ODARRES -SADEGHI , Y. 2015 The influence of taper ratio
on vortex-induced vibration of tapered cylinders in the crossflow direction. J. Fluids Struct. 53, 84–95.
S EYED -AGHAZADEH , B., C ARLSON , D.W. & M ODARRES -SADEGHI , Y. 2017 Vortex-induced vibration and
galloping of prisms with triangular cross-sections. J. Fluid Mech. 817, 590–618.
S EYED -AGHAZADEH , B., E DRAKI , M. & M ODARRES -SADEGHI , Y. 2019 Effects of boundary conditions on
vortex-induced vibration of a fully submerged flexible cylinder. Exp. Fluids 60 (3), 38.
S EYED -AGHAZADEH , B. & M ODARRES -SADEGHI , Y. 2016 Reconstructing the vortex-induced-vibration
response of flexible cylinders using limited localized measurement points. J. Fluids Struct. 65, 433–446.
S IEBER , M., PASCHEREIT, C.O. & O BERLEITHNER , K. 2016 Spectral proper orthogonal decomposition.
J. Fluid Mech. 792, 798–828.
S IROVICH , L. 1987 Turbulence and the dynamics of coherent structures, I. Coherent structures. Q. Appl. Maths,
45 (3), 561–590.
TAIRA , K., B RUNTON , S.L., DAWSON , S.T.M., ROWLEY, C.W., C OLONIUS , T., M C K EON , B.J.,
S CHMIDT, O.T., G ORDEYEV, S., T HEOFILIS , V. & U KEILEY, L.S. 2017 Modal analysis of fluid flows:
an overview. AIAA J. 55 (12), 4013–4041.
T REFETHEN , L.N. & BAU, D. III 1997 Numerical Linear Algebra, vol. 50. SIAM.
T RIM , A.D., B RAATEN , H., L IE , H. & T OGNARELLI , M.A. 2005 Experimental investigation of
vortex-induced vibration of long marine risers. J. Fluids Struct. 21 (3), 335–361.
VANDIVER , J.K., M ARCOLLO , H., S WITHENBANK , S. & J HINGRAN , V. 2005 High mode number
vortex-induced vibration field experiments. In Offshore Technology Conference. OnePetro.
V ICENTE -LUDLAM , D., BARRERO -G IL , A. & V ELAZQUEZ , A. 2015 Enhanced mechanical energy
extraction from transverse galloping using a dual mass system. J. Sound Vib. 339, 290–303.
WANG , Z., FAN , D. & T RIANTAFYLLOU, M.S. 2021 Illuminating the complex role of the added mass during
https://doi.org/10.1017/jfm.2022.893 Published online by Cambridge University Press

vortex induced vibration. Phys. Fluids 33 (8), 085120.


Z ANGANEH , H. & S RINIL , N. 2016 Three-dimensional VIV prediction model for a long flexible cylinder with
axial dynamics and mean drag magnifications. J. Fluids Struct. 66, 127–146.

952 A26-36

You might also like