Download as pdf or txt
Download as pdf or txt
You are on page 1of 24

Marine Structures 84 (2022) 103247

Contents lists available at ScienceDirect

Marine Structures
journal homepage: www.elsevier.com/locate/marstruc

An experimental investigation on interfering VIVs of double and


triple unequal-diameter flexible cylinders in tandem
Bing Zhao a, b, Shixiao Fu a, b, *, Mengmeng Zhang a, b, **, Haojie Ren a, b, Liangbin Xu c,
Wenhui Xie d, Lusheng Jia d
a
State Key Laboratory of Ocean Engineering, Shanghai Jiao Tong University, Shanghai, 200240, China
b
Institute of Polar and Ocean Technology, Institute of Marine Equipment, Shanghai Jiao Tong University, Shanghai, 200240, China
c
Sun Yat-Sen University, Zhuhai, 528478, China
d
China National Offshore Oil Corporation (CNOOC) Research Institute, Chaoyang District, Beijing, 10010, China

A R T I C L E I N F O A B S T R A C T

Keywords: Many unequal-diameter risers have been widely applied to satisfy different operational re­
Wake interference responses quirements in ocean oil and gas drilling engineering, and adjacent tandem risers can undergo
Tandem configuration complex vibration responses and even collisions under ocean flow conditions. To investigate the
Unequal-diameter flexible cylinders
wake interference responses for multiple risers in tandem configuration, an experiment of double
Vortex-induced vibrations
Wake-induced vibrations
and triple unequal-diameter flexible cylinders was conducted under uniform flow with Reynolds
numbers Re ranging from 2832 to 11,328 for the larger cylinder and from 1805 to 7220 for the
smaller cylinder. The two cylinders were 8 m in length with aspect ratios equal to 282.49 and
443.21, respectively, and the corresponding diameter ratio was 0.64. Two different tandem
configurations were arranged with a large cylinder upstream and downstream, and the wall
spacing between adjacent cylinders was fixed at 4D, where D is the diameter of small cylinders.
Fiber Bragg grating strain sensors were used to measure both in-line (IL) and cross-flow (CF)
strain responses. The modal superposition method was applied to reconstruct displacement re­
sponses based on measured strains, and drag forces were further inversely identified from the
mean IL displacements. Equivalent wake velocities were calculated from drag force reductions.
The results show that the downstream large cylinder in a triple-cylinder system endures a stronger
wake interference effect than that in a double-cylinder system. Wake interference can decrease
the dominant mode of downstream cylinders and increase the vibration amplitudes due to the
reduction in the local wake velocity. In addition, a downstream cylinder undergoing wake-
induced vibrations has the same dominant frequency in the CF and IL directions, and its CF
frequency is identical to that of the upstream cylinder.

1. Introduction

In ocean engineering, marine riser systems usually comprise a variety of risers with unequal diameters because of different
operational requirements, such as, water injection risers, mineral lifting risers, and oil and gas production risers. As opposed to the

* Corresponding author. State Key Laboratory of Ocean Engineering, Shanghai Jiao Tong University, Shanghai, 200240, China.
** Corresponding author. State Key Laboratory of Ocean Engineering, Shanghai Jiao Tong University, Shanghai, 200240, China.
E-mail addresses: shixiao.fu@sjtu.edu.cn (S. Fu), claire_zhang@sjtu.edu.cn (M. Zhang).

https://doi.org/10.1016/j.marstruc.2022.103247
Received 29 December 2021; Received in revised form 3 May 2022; Accepted 8 May 2022
Available online 19 May 2022
0951-8339/© 2022 Published by Elsevier Ltd.
B. Zhao et al. Marine Structures 84 (2022) 103247

vortex-induced vibrations (VIV) of an isolated cylinder, multiple-cylinder systems experience more complex interference responses,
such as, wake-induced vibrations (WIV) and even collisions under the action of ocean flow, which may threaten structural safety in
service. Considerable research has made significant contributions to understanding the phenomenon and mechanism of VIV for an
isolated rigid or flexible cylinder through experimental and numerical methods, while fewer investigations have been conducted on
two or more cylinders.
Multiple-cylinder systems with two identical rigid cylinders operating in tandem have been studied thoroughly. With respect to
cylinder center-to-center spacing ratio S/D and the corresponding flow pattern around two identical stationary cylinders, Zdravkovich
classified wake interference into three regimes: a single bluff body region (1 < S/D < 1.2–1.8), a reattachment region (1.2–1.8 < S/D <
3.4–3.8) and a binary vortex street region (S/D > 3.4–3.8) [1]. Relevant studies on dual fixed rigid cylinders in tandem have been
conducted widely and have offered significant advancements [2–7]. However, in practical engineering applications, risers can vibrate
freely in cross-flow (CF) and in-line (IL) directions under flow or wake conditions.
Researchers often fixed the upstream cylinder and released the downstream cylinder in the CF or both CF and IL directions. Hover
and Triantafyllou conducted model experiments for two identical cylinders in tandem with the spacing between two model centers at
4.75D and the Reynolds number Re at 3 × 104 [8]. The study found the upstream cylinder to be unaffected by the downstream cylinder,
while the downstream cylinder’s vibration frequencies were controlled by the upstream vortex-shedding frequency. Assi et al.
experimentally investigated two tandem circular rigid cylinders with center-to-center spacing ratios S/D ranging from 4 to 20 [9] with
the upstream cylinder restrained to vibrate in both the CF and IL directions, while the downstream cylinder remained free in the CF
direction. This work demonstrated that the vortex excitation mechanism of unsteady wake-structure interactions provides positive
energy for the downstream cylinder to undergo WIV. Furthermore, they proposed the concept of wake stiffness to predict the
amplitude and frequency response characteristics of WIV [10,11]. As a fluid dynamic force, wake stiffness can be regarded as a spring
for the downstream cylinder to balance the flow excitation and produce vibration responses. In such a scenario, the WIV response
would be strengthened due to the wake stiffness effect depending on the Reynolds number Re instead of the reduced velocity Ur. In
addition to experiments in water, Hu et al. conducted a wind tunnel experiment and found that the downstream cylinder presents three
vibration responses: pure upstream shedding vortex resonance (VR), separated VR and wake-induced galloping (WIG), and coexisting
VR and WIG [12]. However, Assi et al. proposed that WIV may be different from the classic galloping [9].
When the upstream cylinder is also free to oscillate, the interference responses will be more complex. Kim et al. divided the spacing
ratio S/D into five regimes according to the vibration responses for two identical cylinders. Especially in regimes 1.1 < S/D < 1.2 and 3
< S/D < 3.7, the two cylinders had no vibration responses, while violent vibration occurred in regime 1.1 < S/D < 1.2 at reduced
velocity Ur > 6 [13]. Huang and Herfjord conducted an experiment on two elastically supported rigid cylinders subjected to steady
flows and concluded that the interference response for the downstream cylinder depended on the reduced velocity of the upstream
cylinder and its own reduced velocity based upon the actual mean wake velocity [14]. Armin et al. studied a pair of identical rigid
cylinders at subcritical Re and found that the downstream cylinder had multiple frequency components derived from upstream
shedding vortices and its own natural frequency and vortices [15]. In addition, they noted that WIV exerted considerable influence on
IL responses. Qin et al. presented an experimental investigation on wake interference for two tandem cylinders with different natural
frequencies by adjusting the spring stiffness [16]. Considering six natural frequency ratios ranging from 0.6 to 1.6, they found that the
frequency ratio affects the critical reduced velocity such that the vibration amplitude for the downstream cylinder increases, and that
for the upstream cylinder may decrease. Other investigation results by computational fluid dynamics (CFD) methods have also been
published widely [17–20].
For two tandem cylinders of different diameters, Qin et al. fixed a more slender upstream cylinder whose diameter ratios d/D to the
downstream free cylinder are 0.2, 0.4, 0.6, 0.8 and 1.0 in a wind tunnel [21], demonstrating that the downstream cylinder was excited
to vibrate violently due to the lift force induced by a switching gap shear layer from side to side in the upstream narrow wake at a small
d/D ratio. Subsequently, Wang et al. investigated the influence of the upstream cylinder on wake dynamics based on flow visualization
measurements and found that it was related to the flow structure, Strouhal number St, wake width, vortex formation length, and lateral
width between the two gap shear layers [22]. Huera-Huarte and Jiménez-González also presented an experimental investigation on the
interference characteristics of two rigidly coupled circular cylinders in tandem, and the results showed that vibration amplitudes for
d/D > 0.2 were smaller than those of an isolated cylinder [23].
However, slender flexible cylinder structures whose responses are more complicated than those of rigid cylinders under ocean flow
are mainly used in ocean engineering, especially for multi-riser systems. Limited to computational resources, numerical investigations
on multiple flexible cylinders are relatively rare [24,25]. Therefore, most current investigations concentrate on the studies involving
two identical flexible cylinders.
Huera-Huarte and Bearman experimentally investigated the near wake interference responses of a downstream flexible cylinder
exposed to the wake of an upstream similar cylinder [26]. The two 1.5-m-long cylinders were vertically attached to the water flume
under different center-to-center spacing ratios (S/D = 2 to 4), exposing only the lower 40% of the cylinders to water. Reynolds number
Re could reach 12,000. The experiment showed that the upstream cylinder experiences larger vibrations than the downstream cylinder
under small gap spacing, but for large gap spacing, the downstream cylinder undergoes nonclassic VIV or WIV with large amplitudes at
a high Ur. In addition, they found that WIV was influenced by reduced velocity, gap distance and natural frequency differences be­
tween two cylinders. Sequentially, Huera-Huarte and Gharib further investigated the far wake interference responses with S/D ratios
between 4 and 8 [27]. This work confirmed that the upstream cylinder presents a classic VIV response, while the downstream cylinder
experiences VIV in lock-in reduced velocity region, or WIV beyond the region. However, due to the shallow immersion depth, only the
first mode was excited in their experiments. Subsequently, Huera-Huarte et al. placed a cylinder 3 m in length downstream of a
stationary cylinder with the lower 1.59 m submerged in the water under Reynolds number Re up to 22,000 [28]. This experiment

2
B. Zhao et al. Marine Structures 84 (2022) 103247

Fig. 1. Experimental facility.

identified multi-mode overlapping behaviors. In addition, separation distance between the cylinders seemingly had little influence on
the amplitude and frequency responses. However, the incoming wake magnified the amplitudes remarkably for the downstream
cylinder. Liu et al. also investigated the dynamic responses of flexible cylinders vertically placed in tandem. Their overall lengths were
6.2 m with an aspect ratio of 310 [29]. In contrast to other work, the top 1.2 m was exposed to uniform flow, while the bottom part was
in still water. The experiment showed that the upstream cylinder experienced classic VIV, while the downstream underwent WIV in a
larger lower mode region compared to the isolated model.
Multiple flexible risers of unequal diameters are widely applied in practical engineering, but the corresponding investigations are
relatively fewer. Wang et al. conducted experiments on three cylinders under different incoming flow attack angles ranging from 0◦ to
90◦ [30]. For the tandem configuration, the upstream wake shielding effect resulted in the gradual reduction of the Strouhal number St
and the sudden reduction of the dominant frequencies for the midstream and downstream cylinders. Usually, the wake shielding effect
can be enhanced as the diameter of the upstream cylinder increases. Other similar investigations on double or more unequal-diameter
cylinders in tandem can be also referred to Refs. [31,32].
Multiple-cylinder investigations typically have focused on several identical cylinders rather than those of unequal-diameter.
Although previous studies have achieved results on wake interference responses, an investigation on the response of varying flex­
ible cylinders in water is needed.
In present work, we focused on investigating double and triple unequal-diameter cylinders towed vertically in tandem by exper­
iment. We employed the modal superposition method to reconstruct the displacement response based on tested strain signals, and drag
forces were inversely identified from the structural responses [33,34]. Detailed analyses and discussions were performed to investigate
the corresponding interference responses.

2. Experimental conditions

The experiments were performed in the Ocean Engineering Basin at Shanghai Jiao Tong University in 2017. A photograph of the
experimental facility is shown in Fig. 1. The max adjustable depth of the false bottom is 10 m. Before the test, all the devices were
installed on the false bottom above the water. Then the false bottom gradually descended until the entire cylinder models were
submerged in the water and the other top devices were above the water line. The top hinged point was below the water line, and there
was a flange connecting the top end of the cylinder and the force sensor. The flange across water line was designed to be fluid-deflector-
shaped to make flow disturbance much smaller. Then a vertical shaft connected the force sensor, the spacing-adjusting device and the
pretensioner sequentially.
In our experiments, the entire flexible cylinder models were immersed vertically in water, different from most experiments with the
models towed horizontally in basins. In a horizontal arrangement, gravity will cause the flexible cylinder to generate an initial bending
deformation, even under a large axial tension, and it is oriented in the CF direction, which may affect the CF vibration responses. To
remove the influences of gravity, the cylinders were arranged vertically in this experiment. In addition, the vertical arrangement is
very common in ocean engineering, such as top tensioned risers (TTRs), hybrid standing risers (HSRs) and drilling risers. Currently,
few experimental techniques can satisfy the vertical models to be completely submerged in water [26,27,35]. Our experiments with

3
B. Zhao et al. Marine Structures 84 (2022) 103247

Table 1
Physical parameters of cylinder models.
Model Parameter LC SC

Length l (m) 8.0 8.0


Diameter D (mm) 28.32 18.05
Slenderness ratio l/D 282.49 443.21
Dry mass per unit length m (kg/m) 1.684 0.511
Mass ratio m* 2.73 2.01
Bending stiffness EI (N m2) 115.5 14.4
Tensile stiffness EA (N) 4.30E6 1.65E6

Fig. 2. Measured position configuration along cylinders.

entire models vertically immersed in water can enrich current investigations. To ensure the cylinder oriented in a vertical direction, the
model was adjusted well to meet the vertical requirements in air and descended into the water. During the tests, two encoders were
installed at the top and bottom driving motors to record the displacements at the two ends of the cylinder. Their results were
completely coincident with no delay, which indicated that the cylinder kept a vertical orientation steadily.
Two kinds of 8-m-long cylinders with unequal hydrodynamic diameters were installed vertically in two parallel horizontal tracks,
with their top end pinned and bottom end fixed to the connection devices. The two cylinders were 18.05 mm and 28.32 mm in
diameter, respectively, so the corresponding small to large diameter ratio is 0.64. The slenderness ratios are 282.49 and 443.21 for the
large and small cylinders (LC and SC). Also, the bending stiffness of the large cylinder was more than 8 times that of the small cylinder.
More specific physical parameters of riser models are listed in Table 1.
To obtain the strain responses, Fiber Bragg Grating (FBG) strain sensors were attached to the inner surface of the main cylinder
body in both the CF and IL planes and protected by a soft outer sheath that had little bending and tensile stiffness. As shown in Fig. 2,
two pairs of strain sensors were located on ortho-symmetric points of every cross section, denoted as CF1, CF2, IL1 and IL2. Meanwhile,
the sensors were uniformly installed at 15 cross sections of the LC model and 12 cross sections of the SC model, with a spacing of 0.50 m
and 0.62 m, respectively. All strain signals were collected synchronously at a frequency of 250 Hz to ensure no phase lag between the

4
B. Zhao et al. Marine Structures 84 (2022) 103247

Fig. 3. LC and SC models in tandem configurations.

Table 2
Pretension and natural frequency in water of riser models.
Cylinder model LC SC

Pretension T0 (N) 144.0 70.0


Tension to wet weight ratio T* 1.75 3.50
1st natural frequency in still water (Ca = 1.0) fn1 (Hz) 0.524 0.607
2nd natural frequency in still water (Ca = 1.0) fn2 (Hz) 1.209 1.269

Fig. 4. Global coordinate definitions.

strain signals at different measured locations along the model axis.


In the experiments, the LC and SC models are in a tandem configuration with a 4D wall spacing (D refers to the SC diameter), as
shown in Fig. 3. Both the double- and triple-cylinder systems with LC upstream and downstream were evaluated. In the triple-cylinder
experiment, SC1 was always placed in the middle of the tandem configuration. The motion control system towed the models at a steady
velocity ranging from 0.1 m/s to 0.4 m/s to simulate uniform flows, and the corresponding Reynolds numbers Re ranged from 2832 to
11,328 for the LC and from 1805 to 7220 for the SC. Simultaneously, constant axial pretensions were exerted on the cylinders through
the compressing springs at the top end. To make the tension to wet weight ratio of the small cylinder twice that of the large cylinder,
the pretension was set as 144.0 N/70.0 N for LC/SC, respectively, corresponding to tension ratios of 1.75/3.50. Accordingly, the first

5
B. Zhao et al. Marine Structures 84 (2022) 103247

two order natural frequencies in still water are listed in Table 2 based on Eq. (1) with the added mass coefficient equal to 1.0 constantly
[34]:
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
( )̅
( )2 √
1 iπ √ EI Tl 2
fni = √ / 1+ (1)
2π l m + Ca πρw D2 4 (iπ)2 EI

where i is the order of natural frequency, Ca is the added mass coefficient in water, ρw is the water density, and T is the pre-tension; the
other parameters are referenced in Table 1.

3. Data processing

3.1. Coordinate definitions

As shown in Fig. 4, the origin of the coordinates is at the top end of the riser model, and the z-axis is along the central axis from top
to bottom. The positive direction of the x-axis is the same as the uniform flow direction, so the xoz and yoz planes are the IL and CF
planes, respectively.

3.2. Displacement reconstruction

Due to the structural vibration responses under flows or wakes, the initial constant pretension acting on the top end of the models
varies periodically. Therefore, the measured strain includes two parts: the bending strain caused by hydrodynamic force and the
tension strain caused by top tension. Compared to the CF direction, the riser models sustain a large mean drag force in the IL direction,
so the IL bending strain can be divided into the mean bending strain and vibrating bending strain. In addition, considering the
symmetry of two measured locations in the CF or IL plane, every pair of strains contains the same tension strain and opposite bending
strain. Thus, the strains εCF1 , εCF2 , εIL1 and εIL2 at the four points CF1, CF2, IL1 and IL2 can be written as
εCF1 = εT + εV− CF (2)

εCF2 = εT − εV− CF (3)

εIL1 = εT + εV− IL + εM (4)

εIL2 = εT − εV− IL − εM (5)

where εT is the tensile strain caused by the axial tension, εV− CF and εV− IL are the vibrating bending strains in the CF and IL directions,
and εM is the mean bending strains in the IL direction.
From Eqs. (2)~(5), we can obtain the bending strains in the CF and IL directions as follows:
εCF1 − εCF2
εCF = εV− CF = (6)
2
εIL1 − εIL2
εV− IL + εM = (7)
2
Considering that the average bending strains εV− IL and εM are close to zero and constant in steady experimental phases, respectively,
Eq. (8) can be derived further as
εIL1 − εIL2
εV− IL + εM = = εV− IL + εM = εM = εM (8)
2
From Eqs. (6) and (7), the strain caused by VIV in the IL direction can be written as
εIL1 − εIL2 εIL1 − εIL2
εIL = εV− IL = − (9)
2 2
The modal superposition method can be employed to reconstruct the vibration displacement response from the above strain results
as follows:

n
w(z, t) = pi (t)φi (z), z ∈ [0, l] (10)
i=1

where z is the coordinate of the cylinder axis, t refers to time, pi (t) presents the mode weight, and the i-th mode shape φi (z) of the
pinned-fixed beam can be approximated as
( ) ( )
4i + 1 4i + 1
φi (z) = sin π z + ai sinh πz ​ (i = 1, 2, ⋯) (11)
4l 4l

6
B. Zhao et al. Marine Structures 84 (2022) 103247

Fig. 5. The first five mode shapes of a pinned-fixed beam.

where ai is the mode shape coefficient. The first five order mode shapes are shown in Fig. 5.
According to the small deformation assumption, the curvatures of the riser models can be expressed as the second derivative of the
displacement with respect to the spatial variable z:

∂2 w(z, t) ∑ n ∑n
κ(z, t) = = pi (t)φ′′i (z) = pi (t)θi (z) (12)
∂z 2
i=1 i=1

where κ(z, t) is the bending curvature, and the second derivative of the spatial coordinate of φi (z) is the mode shape of curvature θi (z).
Simultaneously, based on the geometric relationship between curvature κ(z, t) and bending strain ε(z, t) of the beam, the curvature
can also be written as
ε(z, t)
κ(z, t) = (13)
R

where R is the cylinder radius.


Therefore, the mode weights of displacement or curvature can be obtained from the measured bending strain caused by vibration in
light of Eqs. (1) and (10) Combining the mode weights and the mode shapes, the vibration displacement response, y(z,t) and x(z,t), in
the CF and IL directions can be reconstructed from Eq. (12).

3.3. Frequency analysis

To obtain the response dominant frequencies, a fast Fourier transform (FFT) was applied to the original strain signals in the time
domain. We applied FFT to the strain signals of every measurement point, and the sum of PSD corresponding to the frequency of all the
measurement points was calculated to determine the dominant frequencies according to the following formula:
∫ +∞
Fj (ω) = εj (t)e− iωt dt (14)
− ∞

7
B. Zhao et al. Marine Structures 84 (2022) 103247

Fig. 6. Comparison of the mean drag displacement for the LC in the different tandem configurations with LC upstream and downstream.

where εj (t) is the time history of the original strain at the j-th measured point.

3.4. Drag force coefficients identification

To obtain the drag force distribution along the riser axis, an inverse identifying method can be applied to structural IL responses
based on the structural vibration control equation as follows [33,36]:

∂4 [x(z, t) + X(z)] ∂2 T(z)[x(z, t) + X(z)] ∂x(z, t) ∂2 x(z, t)


fIL (z, t) + fM,IL (z) = EI − +c +m (15)
∂z 4 ∂z 2 ∂t ∂t 2

where is EI is the bending stiffness, T(z,t) is the axial tension, c is the structural damping, m is the mass per unit length, x(z,t) and y(z,t)
are the vibration displacements obtained from Eq. (10), and X(z) is the mean drag displacement. The IL hydrodynamic force can be
divided into time-varying and constant parts:

∂4 x(z, t) ∂2 T(z)x(z, t) ∂x(z, t) ∂2 x(z, t)


fIL (z, t) = EI − +c +m (16)
∂z4 ∂z2 ∂t ∂t2

∂4 X(z) ∂2 T(z)X(z)
fM,IL (z) = EI − (17)
∂z4 ∂z2

where fIL (z, t) are the distributed hydrodynamic forces for vibration in the IL directions, fM,IL (z, t) is the mean drag force distribution
along the axis, and the wet weight of the cylinder has been considered in axial tension T(z) calculation:
T(z) = T0 − mw gz (18)

where T0 is the pretension at the top end and mw is the wet mass of the cylinder.
Submitting the Eqs. (18) into the (17), the mean drag force is derived as

8
B. Zhao et al. Marine Structures 84 (2022) 103247

Fig. 7. Comparison of the mean drag displacement for the SC in the different tandem configurations with LC upstream and downstream.

Fig. 8. Comparison of the mean drag force coefficients for the LCs and SCs in different tandem configurations.

∂4 X(z) ∂2 X(z) ∂X(z)


fM,IL (z) = EI − T(z) + 2mw g (19)
∂z 4 ∂z2 ∂z

9
B. Zhao et al. Marine Structures 84 (2022) 103247

Fig. 9. Equivalent wake velocities for the midstream and downstream cylinders in different tandem configurations.

Fig. 10. Max dimensionless CF and IL vibration amplitudes for the upstream LCs and the downstream SCs in the single-, double- and triple-cylinder
systems versus the reduced velocity Ur.

10
B. Zhao et al. Marine Structures 84 (2022) 103247

Fig. 11. Max dimensionless CF and IL vibration amplitudes for the upstream SCs and the downstream LCs in the single-, double- and triple-cylinder
systems versus the reduced velocity Ur.

Furthermore, the drag force coefficients can be acquired by normalizing the above values with characteristic drag force:
fM,IL (z)
Cd (z) = 1 (20)
ρ U2 D
2 w

Furthermore, the mean drag force coefficient Cd,m is defined by



1 L
Cd,m = Cd (z)dz (21)
L 0

4. Results and discussions

4.1. Mean drag displacements and force coefficients

4.1.1. Mean drag displacements


Fig. 6 shows the distribution of the mean drag displacements for LC in the IL direction along the cylinder axes under different
tandem arrangements with LC upstream and downstream, respectively. In both the double- and triple-cylinder systems, when arranged
upstream, the LCs behave like the single LC in Fig. 6(a), which reflects the reliability of the experiment. Due to the wake shielding
effect, a conspicuous displacement reduction for the downstream LCs can be found in Fig. 6(b), compared to the single LC displace­
ments [28]. However, the displacements for the downstream LC in the triple-cylinder system is smaller than those in the
double-cylinder systems, because more SCs interfere with the downstream wake around the LC.
Fig. 7 illustrates the SC mean drag displacement distributions. When the LC is upstream, the midstream and downstream SC
displacements present an obvious decrease, and the downstream SC in the double-cylinders and the midstream SC1 in the triple-
cylinder system have similar responses, as shown in Fig. 7(a). If the LC is placed downstream, the upstream SCs in the double- and
triple-cylinder systems show approximate displacements to the single SC, while the midstream SC1 in the triple-cylinder system
generates smaller displacements in Fig. 7(b). Furthermore, the midstream SC1 following the upstream LC has slightly smaller dis­
placements than those following SC2 in the triple-cylinder system. Thus, the larger the upstream cylinder diameter is, the stronger the

11
B. Zhao et al. Marine Structures 84 (2022) 103247

Fig. 12. Distributions of the CF vibration amplitude for the downstream SCs and LCs at U = 0.3 m/s.

Fig. 13. Explanation of the envelope of the displacement responses.

wake shielding effect.

4.1.2. Drag force coefficients


To further elucidate the wake shielding effect of the mean drag displacement, Fig. 8 gives the mean drag force coefficient for the LCs
and SCs in different tandem configurations. The coefficients for the upstream LCs in the double- and triple-systems are close to the
single LCs. For the downstream LCs, their mean drag force coefficients decrease to approximately 80% and 50% of the single LCs for the
double- and triple-cylinder systems, respectively. Because the downstream LC in the triple-cylinder system is farther away from the
most upstream SC and more cylinders stand upstream, the wake shielding effect on the downstream LC is more significant, leading a

12
B. Zhao et al. Marine Structures 84 (2022) 103247

Fig. 14. Envelope of displacement response for models in the double-cylinder system (dashed black line: initial position; solid blue line: defor­
mation position).

smaller drag force. When placed upstream, the SCs in the double- and triple-cylinder systems also have drag force coefficients like
those of the single SCs. However, the coefficients for the midstream and downstream SCs can drop to approximately one-third of the
single SC. Because the upstream LC has a larger hydrodynamic diameter and can generate strong enough wake shielding effect, the
midstream and downstream SCs in both the double-and triple-cylinder systems are subjected to nearly the same drag forces. Thus,
smaller drag coefficients can be obtained if a cylinder is exposed to the wake of upstream cylinders [22,37], which will result in the
reduction of mean drag displacements for downstream cylinders.

4.1.3. Equivalent wake velocity


As mentioned above, the downstream cylinders are exposed to the upstream wake bearing lower drag forces. Because the true
incoming wake velocities to downstream cylinders are lower than the free flow velocities U, the drag reduction possibly results from
decreased wake velocities rather than drag force coefficients. To obtain the equivalent wake velocity Uw, we assume that the drag force
per unit length can be expressed as
1 1
fd,d = Cd,d ρw U 2 D = Cd,s ρw Uw2 D (22)
2 2

where Cd,d and Cd,s are the drag force coefficients for the downstream cylinder and its corresponding single cylinder, respectively, and
Uw is the equivalent wake velocity for the downstream cylinder and can be derived as

13
B. Zhao et al. Marine Structures 84 (2022) 103247

Fig. 15. Envelope of displacement response for models in triple-cylinder systems (dashed black line: initial position; solid blue line: deforma­
tion position).

√̅̅̅̅̅̅̅̅
Cd,d
Uw = U = αU U (23)
Cd,s

where αU is defined as velocity reduction factor. Especially, the method is based on current experiments in a wall spacing of 4D.
However, if the spacing were closer, the downstream SC could end up in a recirculating region of the wake, where the mean wake
velocity might become negative. Thus, by inference, the two equations are suitable for the case where the spacing is larger than 4D.
Fig. 9 gives the equivalent wake velocity for the midstream and downstream cylinders in different tandem configurations. The
dashed line for the case where U = Uw is also plotted to provide a good preference. The mean wake velocities of all the downstream
cylinders are smaller than the free flow velocities. When the LC is upstream, the wake velocities at the downstream SC in the double-
cylinder system and at the midstream SC1 in the triple-cylinder systems are identical basically, as shown in Fig. 9(a). The downstream
SC2 in the triple-cylinder system presents slightly larger wake velocities at high flow velocities, because it is farther away from the
downstream LC and endures a weaker wake shielding effect. As illustrated in Fig. 9(b), the local wake velocities for the downstream LC
in the double-cylinder system are larger than those for the LC in the triple-cylinder systems, because two upstream SCs have a stronger
wake shielding effect. However, the midstream SC1 in the triple-cylinder systems presents the lowest local velocities in the complex
flow field due to the interaction of the upstream SC2 and the downstream LC.

4.2. Vibration response analysis

In this section, two reduced velocities, corresponding to the nominal free flow velocity U and the local wake velocity Uw, are used
and defined by

14
B. Zhao et al. Marine Structures 84 (2022) 103247

Fig. 16. Frequency spectra of the upstream LC and the downstream SC in the double-cylinder system.

U
Ur = (24)
(fn1 D)LC/SC

Uw
Uw,r = (25)
(fn1 D)LC/SC

where fn1 is the fundamental natural frequency, and D is the cylinder diameter.

4.2.1. Vibration amplitude characteristics


Fig. 10 and Fig. 11 illustrate the maximal RMS of dimensionless vibration amplitudes for the cylinders in the tandem configurations
with the LC upstream and downstream, respectively. It is important to note that our tested flow velocity conditions are relatively few to
obtain a clear and detailed trend of the vibration amplitude changes, but some phenomena can still be observed.
In Fig. 10, the upstream LCs in the double- and triple-cylinder systems are under the direct action of incoming flow, so their vi­
bration amplitudes in both the CF and IL directions generally display similar changing tendency to the single LC [15]. With regard to
the downstream SCs, the vibration amplitudes are basically amplified at most flow velocities, which also was seen by Huera-Huarte
et al. [28]. At U = 0.3 m/s, the single SC presents the third dominant mode, while the other SCs exhibit the second mode with
larger amplitudes in Fig. 12(a). In addition, in the triple-cylinder systems, the midstream SC1 presents larger CF amplitudes than
downstream SC2 because midstream SC1 is close enough to the upstream LC to bear a stronger wake interference effect. Compared to
the downstream SC in the double-cylinder system, the midstream SC1 in the triple-cylinder systems has smaller vibration amplitudes,
because the downstream SC2 possibly disturbs its surrounding flow.
As presented in Fig. 11, the vibration amplitudes of upstream SCs are similar to those of the single model in both the CF and IL
directions [15], while the downstream LC models show a relatively greater disparity. Especially when the LCs are arranged down­
stream at U = 0.30 m/s, their CF dimensionless amplitudes are 0.74 and 0.62 in the double- and triple-cylinder systems, respectively,
which are clearly larger than the value of 0.37 for the single LC. Similarly, in Fig. 12(b), their CF dominant mode orders are also
decreased to the second, but the amplitudes are increased accordingly. In the IL direction, when two tandem SC models are located
upstream, they will have a stronger influence on the downstream LC model to produce a larger amplitude as the flow velocity increases
[16]. However, at low flow velocities, the downstream LCs basically behave like a single LC because the wake of the upstream SC has a

15
B. Zhao et al. Marine Structures 84 (2022) 103247

Fig. 17. Frequency spectra of the upstream SC and the downstream LC in the double-cylinder system.

weak influence effect. Clearer and more accurate phenomena can be obtained by increasing the tested flow velocity conditions.

4.2.2. Potential collision analysis


To judge whether the two adjacent models collide, in practical engineering design, we conservatively assume that the max mean
drag displacement and the max vibration amplitude occur at the same cross section of the cylinder. Then, the displacement envelope is
determined by the maximum vibration amplitudes in the CF and IL directions and the maximum mean drag displacement. As explained
√̅̅̅
in Fig. 13, the blue ellipse has a long axis equal to the sum of the CF maximum amplitude 2ACF,rms and the cylinder radius R, and its
√̅̅̅
short axis is equal to the sum of the IL maximum amplitude 2AIL,rms and the cylinder radius R. In addition, XIL,drag refers to the
maximum mean drag displacement.
Fig. 14 and Fig. 15 show the dimensionless displacement envelopes for the cylinders in the double- and triple-cylinder systems
respectively. The wall gap between two adjacent models decreases as the flow velocity increases, especially when the LC model is
located upstream. In both the double- and triple-cylinder systems, the upstream LC and the adjacent SC can collide at U = 0.4 m/s. The
strong wake velocity reduction causes a small drag force on the downstream SC. Then, the downstream SC models produce smaller
drag displacements under the shielding effects shown in Figs. 14(a) and Fig. 15(a). If the SC is placed upstream, the result will not be as
severe. Compared to the double-cylinder system, the upstream two SCs in the triple-cylinder system are near colliding if the flow
velocity continues to increase. Therefore, in the actual ocean environment, it is necessary to avoid arranging a larger riser upstream in
the dominant current direction, and to choose the wall spacing of four times the diameter of the small riser.

4.2.3. Vibration frequency characteristics

(1) Dominant frequencies

Fig. 16 to Fig. 19 show the CF and IL frequency spectra for the cylinders in different arrangements at different flow velocities. The
FFT was applied to the CF or IL vibration strain time history of each measured point, and the results of all the measured points were
summed to obtain the spectrum in the figures. In the double-cylinder system, the cylinders show single-frequency-dominant responses
in the CF directions basically, except at U = 0.35 m/s where the 2nd and 3rd modal responses participated together for the LC in

16
B. Zhao et al. Marine Structures 84 (2022) 103247

Fig. 18. Frequency spectra of the upstream LC, the midstream SC1 and the downstream SC2 in the triple-cylinder system.

Figs. 16 and 17. In the IL direction, multiple-frequency responses can be seen clearly in most cases, but the upstream SC presents a
single-frequency-dominant response in Fig. 17. Compared to the double-cylinder system, the LC and SC2 show similar phenomena in
the triple cylinder system in Fig. 18 and Fig. 19. For the midstream SC1, its frequency responses are more complex when the SC2 is
upstream, and remarkable multiple-frequency responses in the CF and IL directions can be observed in Fig. 19.
Fig. 20 to Fig. 23 give the dimensionless dominant frequencies in both the CF and IL directions for the LC and SC models,
respectively, in the single-, double- and triple-cylinder systems versus the reduced velocity. The dominant frequencies are obtained
from the maximum spectral value by applying an FFT to the strain response time history, and normalized by dividing the first
theoretical natural frequency in Eq. (1).
Fig. 20 presents the dimensionless dominant frequencies versus the nominal reduced velocity for the LCs and SCs in the tandem
configuration with LC upstream. It is clear that the upstream LC models in both the double- and triple-cylinder systems vibrate at a
similar frequency to that of the single model in the CF and IL directions. The Strouhal number is defined according to the vortex-
shedding frequency. In VIV prediction of flexible risers in practical engineering, researchers introduce the concept of Strouhal
number from rigid cylinders, and based on the syntonization phenomenon between vortex-shedding and structural vibration, the
vortex-shedding frequency is treated to be identical to the structural dominant vibration frequency basically [38,39]. Then the
Strouhal number can be obtained by the slope of a linear best fit straight line based on the data of flow velocities and structural
dominant frequencies [40,41].
In the CF direction, the number St can reach approximately 0.17 and the Strouhal numbers in the IL direction are approximately
0.34. Downstream SCs, including midstream SC1, in the triple-cylinder system, present similar frequency characteristics but are lower
than those of a single SC. In addition, the Strouhal number for the downstream SCs is approximately 0.1 in both the CF and IL di­
rections. In the bottom right figure of Fig. 20, for the midstream and downstream SCs in the triple-cylinder system, the points at Ur
around 9 are off the line. It is noted that the line is fitted by the other points at higher flow velocities where the IL dominant frequencies
are the same to the CF frequencies, different from the result at Ur around 9. Similar phenomena can also be found in Fig. 22. In fact, the
downstream SCs experiencing different dominant responses under different velocities, which will be discussed in next part.
For the downstream SCs, if the dimensionless dominant frequencies are scattered versus the reduced wake velocity as shown in
Fig. 21, the Strouhal numbers are close to 0.17 for the single SC in the CF and IL directions generally at high flow velocities. Thus, when
the LC is arranged upstream, the downstream CF frequencies correspond to the local wake velocity [14].

17
B. Zhao et al. Marine Structures 84 (2022) 103247

Fig. 19. Frequency spectra of the upstream SC2, the midstream SC1 and the downstream LC in the triple-cylinder system.

When the arrangement is reversed with LC downstream, the upstream SCs still behave like the single SC in Fig. 22. In addition, due
to the smaller upstream wake reduction and the larger bending stiffness, the downstream LCs present CF frequencies like the single LC.
However, because the wake interference has a stronger influence in the IL direction [8], the downstream LCs do not behave like the
single LC. Their IL dominant frequencies decrease at U = 0.4 m/s and U = 0.3 m/s in the double- and triple-cylinder systems,
respectively. The downstream LC begin to be dominated by WIV rather than VIV. Therefore, two branches are found corresponding to
the Strouhal numbers equal to 0.34 and 0.14.
When the dimensionless frequencies are scattered versus the reduced wake velocity in Fig. 23, the corresponding Strouhal numbers
become slightly larger, because the local wake velocities may be slightly underestimated. However, for the midstream SC1 in the triple-
cylinder system, the Strouhal number strikingly reaches 0.39, because the midstream SC1 is under the mutual influence of the up­
stream wake effect and the downstream choking effect. Thus, the flow filed around midstream SC1 becomes complex and chaotic, and
the SC1 may not just excited by upstream wake velocity.

(2) IL to CF dominant frequency ratio

The above dimensionless dominant frequency shows that downstream cylinders have an approximate Strouhal number in the CF
and IL directions. Table 3 summarizes the IL to CF dominant frequency ratio. At the lower flow velocity 0.1 m/s, the measured results
are influenced by the noises, so the results at higher flow velocities are presented. It can be observed that the ratios for the upstream
cylinders remain at approximately 2.0, while the values for most of the downstream cylinders are close to 1.0. Therefore, the upstream
cylinders undergo classic VIV responses under free incoming flow. However, the downstream cylinders experience complex WIV
responses because the free shear layers reattach onto their surface after shedding from the upstream cylinders. For the downstream
LCs, the ratios change from 2.0 to 1.0 as the flow velocity increases. Compared to the double-cylinder system, the downstream LC
begins to undergo WIV at U = 0.3 m/s instead of 0.4 m/s because more SCs are upstream to increase the wake interference. Thus, it is
the strong wake interference that controls the downstream cylinders’ IL dominant frequency to be the same as the CF. Thus, WIV exerts
a stronger influence on IL responses [8].

18
B. Zhao et al. Marine Structures 84 (2022) 103247

Fig. 20. Dimensionless CF and IL dominant frequencies for the upstream LC and downstream SC models in the single-, double- and triple-cylinder
systems versus the nominal reduced velocity Ur.

Fig. 21. Dimensionless CF and IL dominant frequencies for the downstream SC models in the double- and triple-cylinder systems with LC upstream
versus the reduced wake velocity Uw,r.

(3) Comparison of the upstream and downstream CF frequencies

The CF dominant frequencies for the different cylinders in the double- and triple-cylinder systems are compared as illustrated in
Fig. 24. The CF dominant frequencies of slenderer downstream SCs, including midstream SC1 in the triple-cylinder system, are the

19
B. Zhao et al. Marine Structures 84 (2022) 103247

Fig. 22. Dimensionless CF and IL dominant frequencies for the upstream SC and downstream LC models in single-, double- and triple-cylinder
systems versus the reduced velocity Ur.

same as those of larger upstream LC in the tested conditions. However, when arranged downstream, the LC will not vibrate at the same
frequency as the upstream SC models in both the double- and triple-cylinder systems, because the smaller flow velocity reduction leads
to the larger spacing and weaker wake interference on the downstream LC. For midstream SC1 in the triple-cylinder system, its CF
dominant frequencies can maintain a value similar to those of the upstream SC2. From the above discussions, the dominant frequency
for the downstream cylinders experiencing WIV response can be identical to that of the upstream cylinders, which can be defined as the
“frequency-controlling effect”. In addition, the effect results from strong vortexes shedding from the upstream cylinders reattaching
onto the surface of the downstream cylinders and controlling their vibration frequencies.

4.2.4. Dominant vibration mode order


From Figs. 16–19, significant multiple-mode responses can be observed, especially in the IL direction. In the CF directions, the
structural responses are dominated by a single mode in most cases. Corresponding to the interesting characteristics of dominant
frequencies discussed above, the dominant mode order for the downstream cylinders experiencing WIV will be further investigated in
this section. The dominant vibration mode orders for the LC and SC models in the single-, double-, and triple-cylinder systems are
presented to better understand the interference characteristics in Table 4 and Table 5. In the multiple cylinder systems, the upstream
cylinders basically vibrate at identical mode order to their corresponding single cylinders under the same tested conditions. For the
downstream SCs in the double- and triple-cylinder systems, the dominant mode orders are similar but lower than those of the single
cylinder in both the CF and IL directions, as shown in Table 4. When the arrangement is reversed in Table 5, the downstream LC in the
triple-cylinder system vibrates at a lower mode due to more upstream SCs, compared to the triple-cylinder systems. Thus, the number
of upstream cylinders will reduce the downstream mode order. In addition, for midstream SC1 in the triple-cylinder system, its mode
orders in the cases with LC upstream are lower than those with SC2 upstream, because the upstream LC can generate a stronger wake
interference and reduce the downstream local wake velocity. The jump from one dominant mode to the other adjacent higher mode is
severely encumbered, resulting in the mode overlapping between the two neighboring orders. This phenomenon was also observed by
Huera-Huarte et al. [28] and Lin et al. [25] in experimental or numerical investigations on multiple-cylinder systems. In the present
interpretation, the mode order reduction mainly results from the decreased local wake velocity.

5. Conclusions

This paper describes an investigation on the interfering VIV responses for LCs and SCs in the double- and triple-cylinder tandem

20
B. Zhao et al. Marine Structures 84 (2022) 103247

Fig. 23. Dimensionless CF and IL dominant frequencies for the downstream models in the double- and triple-cylinder systems with LC downstream
versus the reduced wake velocity Uw,r.

Table 3
The IL to CF dominant frequency ratio.

systems. The results are based on measured strains, mean drag displacements, mean drag force coefficients, equivalent wake velocities,
and vibration amplitude and frequency characteristics obtained during our experimental study. It is noted that the results are limited to
the case of 4D wall spacing. Detailed analysis and discussion of the results in the present work allow us to draw the following
conclusions:

(1) Compared to a double-cylinder system, the downstream LC in a triple-cylinder system endures a stronger wake interference
effect from the two upstream SCs, making the flow velocity where the dominant response transforms from VIV to WIV becomes
lower.

21
B. Zhao et al. Marine Structures 84 (2022) 103247

Fig. 24. Comparisons of CF dominant frequency for different riser models in the double- and triple-cylinder systems.

Table 4
Dominant vibration mode order Nd for different cylinders in tandem with LC upstream.

22
B. Zhao et al. Marine Structures 84 (2022) 103247

Table 5
Dominant mode order Nd for different cylinders in tandem with LC downstream.

(2) Wake interference can decrease the downstream cylinders’ dominant mode orders and increase their vibration amplitudes
compared to a single cylinder, which results from the reduction of local wake velocity.
(3) Upstream cylinders undergo classic VIV responses with a frequency ratio fIL/fCF of approximately 2.0, while the downstream
cylinders will experience WIV responses with fIL/fCF equal to 1.0 when the reduced wake velocity Uw,r increases to around 20. In
addition, the strong vortex shedding from an upstream cylinder make downstream SCs to vibrate at the same CF frequency as
the upstream cylinder.

Declaration of competing interest

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

Acknowledgement

The authors gratefully acknowledge the National Natural Science Foundation of China (Grant No. 52088102), National Science
Fund for Distinguished Young Scholars (Grant No. 51825903), Shanghai Science and Technology Program (Grant No. 19JC1412801,
21ZR1434500), Joint Funds of the National Natural Science Foundation of China (Grant No. U19B2013), Key Projects for Intergov­
ernmental Cooperation in International Science, Technology and Innovation (Grant No. 2018YFE0125100), the Project of the Research
on Lingshui Semi-submersible Production Platform (Grant No. LSZX-2020-HN-05-0406), National Natural Science Foundation of
China (Grant No. 52111530135, 52001208), State Key Laboratory of Ocean Engineering (Shanghai Jiao Tong University) (Grant No.
GKZD010081) and Shenlan Project (Grant No. SL2021MS018SL2020PT102).

References

[1] Zdravkovich MM. The effects of interference between circular cylinders in cross flow. J Fluid Struct 1987;1:239–61.
[2] Igarashi T. Characteristics of the flow around two circular cylinders in tandem:1st report. Bull. JSME 1981;24:323–31.
[3] Igarashi T. Characteristics of the flow around two circular cylinders in tandem:2nd report. Bull. JSME 1984;27:2380–7.
[4] Lin JC, Yang Y, Rockwell D. Flow past two cylinders in tandem: instantaneous and averaged flow structure. J Fluid Struct 2002;16:1059–71.
[5] Zhou Y, Yiu MW. Flow structure, momentum and heat transport in a two-tandem-cylinder wake. J Fluid Mech 2006;548:17–48.
[6] Alam MM. Lift forces induced by phase lag between the vortex sheddings from two tandem bluff bodies. J Fluid Struct 2016;65:217–37.
[7] Carmo BS, Meneghini JR, Sherwin SJ. Secondary instabilities in the flow around two circular cylinders in tandem. J Fluid Mech 2010;644:395–431.
[8] Hover FS, Triantafyllou MS. Galloping response of a cylinder with upstream wake interference. J Fluid Struct 2001;15:503–12.
[9] Assi GRS, Bearman PW, Meneghini JR. On the wake-induced vibration of tandem circular cylinders: the vortex interaction excitation mechanism. J Fluid Mech
2010;661:365–401.
[10] Assi GRS, Bearman PW, Carmo BS, Meneghini JR, Sherwin SJ, Willden RHJ. The role of wake stiffness on the wake-induced vibration of the downstream
cylinder of a tandem pair. J Fluid Mech 2013;718:210–45.

23
B. Zhao et al. Marine Structures 84 (2022) 103247

[11] Assi GRS. Wake-induced vibration of tandem and staggered cylinders with two degrees of freedom. J Fluid Struct 2014;50:340–57.
[12] Hu Z, Wang J, Sun Y. Flow-induced vibration of one-fixed-one-free tandem arrangement cylinders with different mass-damping ratios using wind tunnel
experiment. J Fluid Struct 2020;96:103019.
[13] Kim S, Alam MM, Sakamoto H, Zhou Y. Flow-induced vibrations of two circular cylinders in tandem arrangement. Part 1: characteristics of vibration. J Wind
Eng Ind Aerod 2009;97:304–11.
[14] Huang S, Herfjord K. Experimental investigation of the forces and motion responses of two interfering VIV circular cylinders at various tandem and staggered
positions. Appl Ocean Res 2013;43:264–73.
[15] Armin M, Khorasanchi M, Day S. Wake interference of two identical oscillating cylinders in tandem: an experimental study. Ocean Eng 2018;166:311–23.
[16] Qin B, Alam MM, Ji C, Liu Y, Xu S. Flow-induced vibrations of two cylinders of different natural frequencies. Ocean Eng 2018;155:189–200.
[17] Papaioannou GV, Yue DKP, Triantafyllou MS, Karniadakis GE. On the effect of spacing on the vortex-induced vibrations of two tandem cylinders. J Fluid Struct
2008;24:833–54.
[18] Borazjani I, Sotiropoulos F. Vortex-induced vibrations of two cylinders in tandem arrangement in the proximity-wake interference region. J Fluid Mech 2009;
621:321–64.
[19] Prasanth TK, Mittal S. Flow-induced oscillation of two circular cylinders in tandem arrangement at low Re. J Fluid Struct 2009;25:1029–48.
[20] Bao Y, Huang C, Zhou D, Tu J, Han Z. Two-degree-of-freedom flow-induced vibrations on isolated and tandem cylinders with varying natural frequency ratios.
J Fluid Struct 2012;35:50–75.
[21] Qin B, Alam MM, Zhou Y. Two tandem cylinders of different diameters in cross-flow: flow-induced vibration. J Fluid Mech 2017;829:621–58.
[22] Wang L, Alam MM, Zhou Y. Two tandem cylinders of different diameters in cross-flow: effect of an upstream cylinder on wake dynamics. J Fluid Mech 2018;
836:5–42.
[23] Huera-Huarte FJ, Jiménez-González JI. Effect of diameter ratio on the flow-induced vibrations of two rigidly coupled circular cylinders in tandem. J Fluid Struct
2019;89:96–107.
[24] Lin K, Wang J, Zheng H, Sun Y. Numerical investigation of flow-induced vibrations of two cylinders in tandem arrangement with full wake interference. Phys
Fluids 2020;32:015112.
[25] Lin K, Wang J, Fan D, Triantafyllou MS. Flow-induced cross-flow vibrations of long flexible cylinder with an upstream wake interference. Phys Fluids 2021;33.
[26] Huera-Huarte FJ, Bearman PW. Vortex and wake-induced vibrations of a tandem arrangement of two flexible circular cylinders with near wake interference.
J Fluid Struct 2011;27:193–211.
[27] Huera-Huarte FJ, Gharib M. Vortex- and wake-induced vibrations of a tandem arrangement of two flexible circular cylinders with far wake interference. J Fluid
Struct 2011;27:824–8.
[28] Huera-Huarte FJ, Bangash ZA, González LM. Multi-mode vortex and wake-induced vibrations of a flexible cylinder in tandem arrangement. J Fluid Struct 2016;
66:571–88.
[29] Liu H, Wang F, Jiang G, Guo H, Li X. Laboratory measurements of vortex- and wake-induced vibrations of a tandem arrangement of two flexible risers. China
Ocean Eng 2016;30:47–56.
[30] Wang Y, Li P, Liu Y, Guo H, Lou M. Experimental investigation on the vortex-induced vibration of a three-riser group coupling interference effect. J Sound Vib
2021:491.
[31] Xu W, Zhang Q, Yu Y, Lai J, Chang Y. Fluid-structure interactions (FSI) behaviour of two unequal-diameter flexible cylinders in tandem configuration. Ocean
Eng 2020:218.
[32] Wang E, Xu W, Yu Y, Zhou L, Incecik A. Flow-induced vibrations of three and four long flexible cylinders in tandem arrangement: an experimental study. Ocean
Eng 2019;178:170–84.
[33] Song L, Fu S, Cao J, Ma L, Wu J. An investigation into the hydrodynamics of a flexible riser undergoing vortex-induced vibration. J Fluid Struct 2016;63:325–50.
[34] Ren H, Zhang M, Wang Y, Xu Y, Fu S, Fu X, et al. Drag and added mass coefficients of a flexible pipe undergoing vortex-induced vibration in an oscillatory flow.
Ocean Eng 2020:210.
[35] Baarholm R, Kristiansen T, Lie H. Interaction and clashing between bare or straked risers, analyses and experimental data. In: Proceedings of the 26th
international conference on offshore mechanics and arctic engineering, OMAE2007. San Diego: California; 2007. USA.
[36] Song L, Fu S, Dai S, Zhang M, Chen Y. Distribution of drag force coefficient along a flexible riser undergoing VIV in sheared flow. Ocean Eng 2016;126:1–11.
[37] Huang S, Sworn A. Some observations of two interfering VIV circular cylinders of unequal diameters in tandem. J Hydrodyn 2011;23:535–43.
[38] Gao Y, Fu S, Wang J, Song L, Chen Y. Experimental study of the effects of surface roughness on the vortex-induced vibration response of a flexible cylinder.
Ocean Eng 2015;103:40–54.
[39] Rao Z, Vandiver JK, Jhingran V. Vortex induced vibration excitation competition between bare and buoyant segments of flexible cylinders. Ocean Eng 2015;94:
186–98.
[40] Vandiver JK, Peoples WW. The effect of staggered buoyancy modules on flow-induces vibration of marine risers. In: The offshore Technology conference; 2003.
Houston, Texas, U.S.A.
[41] Potts AE, Potts DA, Marcollo H, Jayasinghe K. Strouhal number for VIV excitation of long slender structures. In: Proceedings of the 37th international conference
on ocean, offshore and arctic engineering, OMAE2018. Madrid, Spain; 2018. June 17–22.

24

You might also like