You are on page 1of 10

International Journal of Naval Architecture and Ocean Engineering 15 (2023) 100521

Contents lists available at ScienceDirect

International Journal of Naval Architecture and Ocean Engineering


journal homepage: http://www.journals.elsevier.com/
international-journal-of-naval-architecture-and-ocean-engineering/

Development of a hydrofoil wake oscillator model based on a near-


vortex strength for predicting vortex-induced vibration on a hydrofoil
Hyun-Gyu Choi a, b, Suk-Yoon Hong a, b, **, Jee-Hun Song c, *, Won-Seok Jang a, b,
Woen-Sug Choi d
a
Department of Naval Architecture and Ocean Engineering, Seoul National University, Seoul, Republic of Korea
b
Institute of Engineering Research, Seoul National University, Seoul, Republic of Korea
c
Department of Naval Architecture and Ocean Engineering, Chonnam National University, Yeosu, Republic of Korea
d
Department of Ocean Engineering, Korea Maritime and Ocean University, Busan, Republic of Korea

a r t i c l e i n f o a b s t r a c t

Article history: Recently, the evaluation of vortex-induced vibration has emerged as a significantly important issue
Received 9 November 2022 owing to the development of high-speed and lightweight ships and submarines. To derive an accurate
Received in revised form vortex-induced vibration response, it is essential to consider the fluid-structure interaction. Moreover, it
19 January 2023
is necessary to evaluate the generation of the fluid-structure interaction to effectively prevent cata-
Accepted 19 February 2023
strophic failure in the structures. In this study, a hydrofoil wake oscillator model was developed based on
Available online 28 February 2023
a near-vortex strength that considers the fluid-structure interaction. The near-vortex strength was
calculated from the boundary layer on a trailing edge to overcome the empirical parameter of lift fluc-
Keywords:
Fluid-Structure Interaction (FSI)
tuation in conventional wake oscillator models. To predict the vortex-induced vibration on a hydrofoil,
Vortex-Induced Vibration (VIV) procedures for calculating the near-vortex strength and coupling the structural equations and fluid
Wake oscillator equation were introduced. The vortex-induced vibration derived using the developed hydrofoil wake
Vortex strength oscillator model was verified by comparison it against the experimental results. The results reveal that
Hydrofoil the derived amplitude and lock-in range of the vortex-induced vibration were consistent with the
experimental results. In addition, the extent of occurrence of the fluid-structure interaction and its
contribution to vortex-induced vibration were evaluated using a non-dimensional wake parameter.
© 2023 Society of Naval Architects of Korea. Production and hosting by Elsevier B.V. This is an open
access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).

1. Introduction conditions tends to be stronger than that under low-speed sailing


conditions. Furthermore, compared to conventional structures
The sailing of a ship or submarine is accompanied by the gen- composed of steel, the recent application of composite materials to
eration of a vortex behind the rudder or sail owing to the fluid achieve a lightweight design result in large deformations on
viscous effect. Consequently, the shedding of the vortex from these structures (Mouritz et al., 2001). This indicates that the fluid-
structures induces vibration on the hull. This phenomenon is structure interaction methodology should be applied as the large
defined as vortex-induced vibration, and it continuously loads the deformations caused by the fluid force constantly affect the flow
structures and increases the risk of fatigue failure. Recently, the field around the structures (Vandiver et al., 2006). Therefore, it is
development of high-speed ships and submarines has emerged as a essential to consider the fluid-structure interaction methodology to
critical issue for various reasons, such as performance enhance- derive accurate vibration response on structures.
ment and sailing time and energy saving. However, the lift force Vortex-induced vibration has special characteristics of lock-in
fluctuation caused by vortex shedding under high-speed sailing and lock-off phenomena. In the lock-in range, fluid force induces
a significantly higher vibration as the vortex shedding frequency is
synchronized to the wet natural frequency of the hydrofoil. This
indicates that even if the frequency is slightly lower or higher than
* Corresponding author.
the wet natural frequency, a higher vibration might be induced at
** Corresponding author. Department of Naval Architecture and Ocean Engineer-
ing, Seoul National University, Seoul, Republic of Korea. the same frequency as the wet natural frequency. In contrast, in the
E-mail addresses: syh@snu.ac.kr (S.-Y. Hong), jhs@jnu.ac.kr (J.-H. Song). lock-off range, the frequency of the vortex-induced vibration is not
Peer review under responsibility of The Society of Naval Architects of Korea.

https://doi.org/10.1016/j.ijnaoe.2023.100521
2092-6782/© 2023 Society of Naval Architects of Korea. Production and hosting by Elsevier B.V. This is an open access article under the CC BY-NC-ND license (http://
creativecommons.org/licenses/by-nc-nd/4.0/).
H.-G. Choi, S.-Y. Hong, J.-H. Song et al. International Journal of Naval Architecture and Ocean Engineering 15 (2023) 100521

related to the wet natural frequency of the hydrofoil. Thus, only a In this study, a hydrofoil wake oscillator model based on a near-
very small vibration is induced at the same frequency as the vortex vortex strength was developed to predict the vortex-induced vi-
shedding frequency. It is important to determine the lock-in and bration on a hydrofoil. To consider the fluid-structure interaction,
lock-off ranges as the lock-in range can exist under the sailing speed wake equation and structural equations were coupled with the
condition of ships or submarines. If ships or submarines continu- acceleration of the bending and twisting motions and wake
ously sail in the lock-in range, the risk of fatigue failure rapidly parameter. The structural equations were composed of the bending
increases, thus resulting in the propagation of cracks in the struc- and twisting dynamic motion equations, which consider its mode
tures. Therefore, it is essential to predict the vortex-induced vi- shape functions as a hydrofoil predominantly moves with bending
bration, as well as the lock-in and lock-off range, at the initial design and twisting motions. In addition, a new methodology was apply to
stage to prevent catastrophic failure. the hydrofoil wake oscillator model to overcome the empirical
To consider the fluid-structure interaction, wake oscillator parameter using the function of near-vortex strength. The results
models have been introduced to derive vortex-induced vibration reveal that the amplitude of vortex-induced vibration can be pre-
on structures (Kurushina et al., 2022). The wake oscillator model dicted, and the lock-in and lock-off ranges can be determined.
exhibits an advantage in terms of the fluid-structure interaction
methodology as it enables the calculation of the vortex-induced
vibration by representing near wake force of vortex shedding 2. Modelling of a hydrofoil wake oscillator model based on a
with non-linear characteristics and periodically loading of the near-vortex strength
structures. Previous studies have derived the lift force on a cylinder
and the vortex-induced vibration on a bluff body by coupling the Hydrofoil dynamic equations of motions and wake equation are
oscillator model with a structural equation (Bishop and Hassan, defined under the several assumptions. The elastic axis is not
1964; Parkinson, 1989). Facchinetti et al. (2004) proposed the coincided with the center of gravity to couple the bending and
coupling of a classical van der Pol wake oscillator model with a twisting motion and the elastic axis is the same point along the
transverse structural equation to derive the transverse amplitude spanwise, thus, the deflection of the elastic axis is restrained to the
and lock-in range on a cylinder. The spanwise characteristics had vertical direction and the rotation is generated on elastic axis only.
been considered by distributing along the span from Thus, the mode shape can be derived from a beam model as the foil
Balasubramanian and Skop (1996); Skop and Balasubramanian behaves like a beam.
(1997); Skop and Luo (2001). Jauvtis and Williamson (2004);
Sanchis et al. (2008) studied the in-line and transverse vibration (2-
2.1. Structural equations of a hydrofoil with bending and twisting
DoF) with different mass ratio. Subsequently, the model proposed
motions
by Fachhinetti was modified to a frequency-dependent model for
an elastically mounted cylinder to predict the free and forced vi-
To derive the displacement from the bending and twisting
bration (Ogink and Metrikine, 2010). Thereafter, Bai and Qin (2014)
motion, each mode shape should be considered. Bending and
proposed a wake oscillator model that considers near-vortex effect.
twisting mode shape function of a hydrofoil with the fixed-free
Srinivasan et al. (2018) developed the reduced order wake oscillator
boundary condition can be represented as follows:
for predicting the in-line and transverse motions of cylinder which
was good agreement with the experimental results (Table 1). coshð~sÞ  cosð~sÞ  0:734½sinhð~sÞ  sinð~sÞ
Thereafter, the hydrodynamic coefficient of the model was derived bðsÞ ¼
2
from the wake oscillator model and this method was extended to 3-
D domain to predict in-line and transverse vibration on a riser ps
system in two-degree-of-freedom (Qin et al., 2021; Feng et al., tðsÞ ¼ sin (1)
2
2022). where, s is a non-dimensional span length (s ¼ s=L) and ~s is 1.875s.
The inability of previously proposed wake oscillator models to By integrating the modal shape function over the non-
set fluctuating vortex force as the empirical value during the dimensional span, the displacement on the hydrofoil can be
coupling procedure between the wake and structural equations has derived to consider the bending and twisting motions (Chae et al.,
limited its application for the prediction of the vortex-induced vi- 2016).
bration on a hydrofoil. In case of a hydrofoil, it is difficult to define
the fluctuating vortex force as the fluctuating responses change ð1
significantly according to the sectional shapes. Hence, a new Sbb ¼ b2 ðsÞds
methodology was applied to the developed hydrofoil wake oscil- 0
lator model to overcome the empirical parameter for vortex force
ð1
by calculating the theoretical vortex strength based on the
Sbt ¼ bðsÞtðsÞds (2)
boundary layer on the trailing edge. Thus, the fluctuating lift force
on a hydrofoil could be determined as the function of a near-vortex 0
strength(G). ð1
Stt ¼ t 2 ðsÞds
0
Table 1
Comparisons of wake oscillator models for FSI. The governing equation of a hydrofoil can be derived using
Lagrange's approach, which can be represented as Eq. (3). Using Eq.
2D/3D DOF Structure CL0
(3), the governing equation for the bending and twisting motions of
Facchinetti (2004) 2D 1DOF cylinder 0.3 a hydrofoil can be expressed as Eq. (5).
Ogink (2010) 2D 1DOF cylinder 0.3842
Bai (2014) 2D 2DOF cylinder 0.3
d vT vT vU
Qin (2021) 2D 2DOF cylinder 0.3  þ ¼Q (3)
Feng (2022) 3D 2DOF cylinder 0.2967 dt vz_ vz vz
In this study 2D 2DOF hydrofoil based on G

2
H.-G. Choi, S.-Y. Hong, J.-H. Song et al. International Journal of Naval Architecture and Ocean Engineering 15 (2023) 100521

 
h
z¼ (4)
q

where, z is a matrix for bending and twisting displacement on a


hydrofoil and Q is a matrix related to generalized forces on a
hydrofoil.

ðMS þ MF Þz€ þ CS z_ þ KS z ¼ FS (5)


Fig. 1. Hydrofoil dynamic model for bending and twisting motions in flow condition.
Eq. (6) is the matrix for the structural mass, damping and
stiffness characteristics. The added mass for the bending and
twisting motions was considered using Theodorsen's potential flow
approximation (Theodorsen, 1979) shown in Eq. (7). The fluid force L ¼ rU G (10)
acting on the thin plate from the potential flow field was assumed
Fidler (1974) introduced an approximation methodology for
to be the fluid force acting on the hydrofoil section. Eq. (8) shows
calculating the vortex strength from a slender body. Bourgoyne
the lift and moment forces caused by trailing edge vortex fluctua-
(2005) proposed a modified method for determining the vortex
tions that consider the mode shape function.
strength from the trailing edge. In addition, the vortex strength of
  diverse structures, such as foils and vortex generators, has been
ms Sbb ms bsc Sbt
MS ¼ calculated to optimize their performances (Prangemeier et al.,
ms bsc Sbt Iq Stt
  2010; Ahmad and Proctor, 2014; Gutierrez et al., 2019; Hansen
Ch Sbb 0 et al., 2019). The vortex strength is expressed as a function of the
CS ¼ (6)
0 Cq Stt circulation which is defined as the surface integral of vorticity. The
  vorticity can be derived using the curl of the velocity vector in the
Kh Sbb 0 fluid domain. In 2D conditions, it is expressed in the terms of the
KS ¼
0 Kq Stt gradient of velocities and represents the vorticity of Z-component,
as expressed in Eq. (11).
2 3
2
6 prb Sbb prsc b3 Sbt u ¼V!
! u
MF ¼ 4   7 5 (7)
prsc b3 Sbt prab4 0:125 þ s2c Stt  
vv vu ^
uk ¼  k in 2D (11)
  vx vy
LSbb
FS ¼ (8) As shown in Figs. 2 and 3, vortex shedding occurs when the
MStt
boundary layer near the trailing edge is separated from the wall. To
calculate the vortex strength, several assumptions were applied in
where, ms is the structural mass per span and r is a fluid density. Ch,
this study. First, the near trailing edge vortex (G1) was assumed to
Cq are structural viscous damping coefficients and Kh, Kq are stiff-
dominantly induces the lift force on the hydrofoil. Second, the near-
ness coefficients for bending and twisting motions, respectively. As
vortex was assumed to be fully generated during period Dt, which is
shown from Fig. 1, b is a half of chord length and sc is a variable
half vortex shedding frequency. Lastly, the thickness of the shear
related to the static unbalance of hydrofoil.
layer was represented as the thickness of the displacement
boundary layer as the vortex shedding is mainly caused by fluid
viscosity effect. Hence, the vortex strength can be calculated using
2.2. Wake equation for expressing lift fluctuation from vortex
Eq. (12).
shedding
ðð
G¼ !
u dS
To couple the structural equations in section 2.1, the wake
equation should be modelled. Lift force fluctuation on a bluff body S
is caused by trailing edge vortex shedding even under a uniform
flow condition (Gerrard, 1966). This time-varying fluctuation can be ð
dshear
modelled using a van der Pol wake equation which is a second- vu umax
G¼ dy (12)
order differential equation. It is widely used to derive the vortex- vy fvortex
0
induced vibration as it exhibits a non-linear characteristics
composing of a non-linear damping term of q2 variable. By substituting Eq. (12) to Eq. (10), the lift force fluctuation can
be calculated. From Eq. (12), it is expected that the amplitude of lift
€ þ εuv ðq2  1Þq_ þ u2v q ¼ FW
q (9) fluctuation might increase in turbulent flow rather than in laminar
flow since the vortex strength is proportional to the gradient of
where, q is a non-dimensional variable related to vortex fluctuation. stream flow velocity.
ε is a non-linear damping coefficient. uv and FW are an angular
vortex shedding frequency and fluid-structure coupling term,
respectively. 2.3. Coupling of the structural equations and wake equation to
The lift force fluctuation can be calculated using the vortex consider the FSI effect
strength from the near-vortex shedding. Wu (1981) and Obasaju
et al. (1988) derived the lift fluctuation using Joukowski To couple the structural equations and wake equation, a non-
circulation-lift theorem, and their results exhibited a similar ten- dimensionalization process is required for each equation. In case
dency as the experimental measurement. of the structural equations, the process was divided into space
3
H.-G. Choi, S.-Y. Hong, J.-H. Song et al. International Journal of Naval Architecture and Ocean Engineering 15 (2023) 100521

2 3
L
" # 6 S 7
~LS 6 ms bu2 bb 7
bb 6 q 7
NFS ¼ ¼6 7 (16)
~
MS 6 M 7
bb 4 Stt
5
ms b uq
2 2

where, Iq is the moment of inertial about the elastic axis (EA) and rq
is the non-dimensional radius of gyration of the hydrofoil section.
zh, zq are the structural damping coefficient for the bending and
twisting motions. uh is the angular bending natural frequency and
uq is the angular twisting natural frequency.
In terms of the wake equation, non-dimensionalization is
Fig. 2. Hydrofoil bending and twisting motions model for coupled structure and wake needed as well to couple with bending and twisting motion
oscillator. equations. Likewise, t is replaced with T by multiplying uq to t.

T ¼ uq t (17)
Thereafter, the non-dimensional van der Pol wake equation was
derived as Eq. (18). To couple the structural equations, the coupling
force term (NFW) can be substituted for the acceleration, velocity
and displacement motions of the hydrofoil. Facchinetti et al. (2004)
compared three different coupling models with cylinder experi-
mental results, and found that the highest accuracy for the pre-
diction of the vortex-induced vibration response was obtained
when the acceleration coupling term is used. Thus, the coupling
force term was set as the acceleration of the bending and twisting
motions, as expressed in Eq. (19).

€ þ εuq ðq2  1Þq_ þ u2q q ¼ NFW


q (18)

€þA €
NFW ¼ A1 h (19)
Fig. 3. Schematics of trailing edge vortex dynamic from the boundary layer. 2q

where, uq is the non-dimensional vortex shedding angular fre-


domain and time domain. First, the spatial variable, h, was non- quency (uq ¼ uv/uq). A1 and A2 are the bending and twisting
dimensionalized by dividing it into b which is a half of a chord coupling effect coefficients, respectively.
length. Additionally, the time variable, t, was replaced with the Lastly, the structural equations, Ep. (13), and wake equation, Ep.
non-dimensional time variable, T, by multiplying uq by t. Subse- (18) can be coupled using the relationship between the lift coeffi-
quently, the non-dimensional matrix for the mass, damping and cient and non-dimensional wake variable. As shown in Eq. (20), the
stiffness (NMS, NCS, NKS) was derived as Eqs. (13)e(16). lift coefficient can be expressed as a function of the non-
dimensional wake variable (Facchinetti et al., 2004). Hence, the
vortex-induced vibration can be calculated while considering the
fluid-structure interaction.
NMS €
~z þ NCS ~z_ þ NKS ~z ¼ NFS (13)

1
CL ¼ CL0 q (20)
2
" #
h=b where, CL0 is the reference lift coefficient representing the fluctu-
~z ¼ (14)
q ation of the lift force when the structure was subjected to vortex-
induced vibration. In a previous study, CL0 was defined as an
empirical value of 0.3 (Blevins, 2001). However, there is a limit to
defining the value of CL0 as a constant as the fluctuating lift force
2 3 can be significantly changed owing to the diverse hydrofoil
Sbb sc Sbt
NMS ¼4 5 sectional shapes. Therefore, a new methodology was developed to
sc Sbt rq2 Stt calculate the reference lift coefficient from the function of the
2 3 vortex strength by substituting Eq. (10) and Eq. (12) to Eq. (20).
2zh ðuh =uq ÞSbb 0
NCS ¼ 4 5 (15)
0 2zq rq2 Stt ð
dshear
2 3 L0 2 vu umax
ðuh =uq Þ Sbb 2
0 CL0 ¼ ¼ dy (21)
NKS ¼ 4 5 0:5rU 2 c Uc vy fvortex
0
0 rq2 Stt
Finally, the coupled matrix for non-dimensional structural and
wake equations can be derived as Eq. (22e25).
4
H.-G. Choi, S.-Y. Hong, J.-H. Song et al. International Journal of Naval Architecture and Ocean Engineering 15 (2023) 100521

trailing edge shape was cut as blunt type with a 3.22 mm thickness

~þC Z ~_ ~ (22)
MFSI Z FSI þ KFSI Z ¼ FFSI to form periodic vortices and to observe the vortex-induced vi-
bration clearly. The vibration measurement point was located at
23 0.01m from the trailing edge to the upstream direction at mid-span.
h=b
6 7 The boundary condition was fixed-free condition as shown in Fig. 4.
~¼6 q 7
Z (23)
4 5 The procedure for predicting the vortex-induced vibration is illus-
q trated in Fig. 5. The acceleration of bending and twisting motions
and vortex parameter were continuously updated at each time step
2 3 considering the fluid-structure interaction .
Sbb sc Sbt 0
6 7
6 7
MFSI ¼ 6 sc Sbt rq2 Stt 0 7 3.1. Calculation of the near-vortex strength based on boundary
4 5
A1 A2 1 layer
2   3
uh The vortex strength should be calculated to derive the lift fluc-
6 2zh u Sbb 0 0 7
6 q 7 tuation caused by vortex shedding. Based on Eq. (12), the vortex
6 7
6 7 strength of a near-vortex can be derived from the boundary layer of
CFSI ¼ 6 0 2zq rq2 Stt 0 7
6 7 (24) a trailing edge. The boundary layer was derived using computa-
6   7
4 5 tional fluid dynamic (CFD) based on Reynolds Average Navier-
0 0 εuq q2  1 Stokes equation (RANS) in a steady condition. The simulation
2  2 3 condition was presented in Table 2, and the set-up for simulation
h u domain is shown in Fig. 6.
6 u Sbb 0 0 7
6 q 7 For a more accurate extraction of the boundary layer, the tur-
6 7
KFSI ¼6
6
7 bulence model utilized the k-u(SST) model, and the transition
6 0 rq2 Stt 0 7
7
4 5 model utilized a two-equations g-Req model. As shown in Fig. 6, the
0 0 u2q fluid domain was set as the form of the O-gird type and the radius
of the domain was set to 20 times the chord length to prevent the
2 3
6 rU 2
7
6 CL0 q 7
6 2m s uq l
2 7
6 7
6  7
FFSI ¼6
6 1 rU 2 7
7 (25)
6 sc þ C q 7
6 2 2ms u2 l L0 7
6 q 7
4 5
0
Fig. 4. Blunt trailing edge hydrofoil with NACA0009 sectional shape and boundary
condition (Zobeiri, 2012).
where, MFSI, CFSI, KFSI are the coupled non-dimensional matrixes for
mass, damping and stiffness. Z~ is the coupled matrix for vortex
variable, bending and twisting motions, and l is the ratio of the
total mass to structural mass.
In addition, the non-dimensional coupled lift force and moment
induced from the near-vortex can be expressed as Eq. (25) as the lift
force and moment on a hydrofoil can be calculated using Eq. (26)
and Eq. (27)(Hodges and Pierce, 2011). As shown in the coupled
matrix, the bending and twisting motions were subjected to the
fluid force, q, and the hydrofoil motions, h;€ €
q, exerted an influence
on the wake parameter. Hence, the fluid-structure interaction can
be considered.

L ¼ rU 2 bCL (26)

 
1
M¼ sc þ rU 2 b2 CL (27)
2

where, U is the free stream velocity and CL is the lift coefficient


which is caused by near wake force.

3. Prediction of vortex-induced vibration on a hydrofoil

The predicted vortex-induced vibration was compared to the


experimental result (Zobeiri, 2012), which was obtained in an EPFL
high-speed cavitation tunnel with a test section of
150  150  750 mm. The hydrofoil was based on the section shape Fig. 5. Procedure for predicting vortex-induced vibration on a hydrofoil considering
of NACA0009 with 0.1m chord length and 0.15m span length. The fluid-structure interaction.

5
H.-G. Choi, S.-Y. Hong, J.-H. Song et al. International Journal of Naval Architecture and Ocean Engineering 15 (2023) 100521

Table 2
Conditions for CFD analysis.

Values

Governing Eq. Reynolds Averaged NeS Eq. (RANS)


Turbulence model k-u (SST)
Transition model g Req
Flow velocity 5 ~ 20 m/s
Angle of attack 0
number of grid points 71,000

Fig. 8. Non-dimensional boundary layer of a wide range of flow speed conditions (5


~20 m/s) at the trailing edge (boundary layer thickness:; fitting curve:).

Fig. 6. Computational grid and boundary condition of CFD domain with 20 times
chord length, C.

backflow effect induced at the outer boundary. In addition, a mesh


Fig. 9. Verification of the calculated vortex strength based on the boundary layer
was generated to satisfy the condition of CFL < 1 and y þ ¼ 1. compared with the experimental results.
The boundary layer was derived by probing the stream velocity
at the measuring points as shown in Fig. 7. The measuring points
consisted of 40 points with a spacing of 1.8  105 m to Furthermore, the vortex strength was derived to calculate the
5.4  105 m. The spaces of each point near the trailing edge were fluctuating lift force. By substituting the velocity distribution to Eq.
densely set as the gradient of the steam velocity was significantly (12), the vortex strength was derived, as shown in Fig. 9. As the flow
changed. velocity increased, the vortex strength tends to increase linearly
Fig. 8 shows the boundary layers at a wide range of flow speed (Ausoni, 2009). This linear increase can be expected as the vortex
conditions (5 ~20 m/s). The blue point represents the boundary strength is proportional to the gradient of flow velocity. Conse-
layer thickness of each flow velocity condition and the blue dashed quently, the calculated vortex strength was in a good agreement
line is the fitting curve of each boundary layer thickness. The with the experimental results.
boundary layer thickness was observed to gradually decrease as the
flow velocity increased. In addition, the gradient of flow velocity
3.2. Prediction of the vortex-induced vibration using a hydrofoil
increased owing to the increase in the Reynolds number. This val-
wake oscillator model
idates the velocity distribution as the boundary layer thickness
decreased as the Reynolds number increased.
In this section, the vortex-induced vibration on a hydrofoil was
derived considering the fluid-structure interaction. The parameters
for the hydrofoil wake oscillator are expressed in Table 3. The
reference lift coefficients were calculated using the near-vortex
strength value, Eq. (21), which were derived to be approximately
0.038e0.046, according to the flow velocity. The non-linear
damping coefficient, ε was set as 0.3, which was proposed by
Facchinetti et al. (2004). The vortex shedding frequency, fv, was
calculated using StU/tTE where St is the Strouhal number (0.23) and
tTE is the trailing edge thickness (0.0322 mm).
Particularly, the van der Pol wake equation exhibits a charac-
teristics in which the amplitude of the non-dimensional wake value
converges to 2 fluctuating with vortex shedding frequency when
the fluid-structure interaction is barely noticeable (Nayfeh, 2011).
In contrast, the amplitude of the non-dimensional wake value
tends to increase above 2 if the fluid-structure interaction highly
Fig. 7. The measuring points for the boundary layer at the trailing edge. increases. Thus, it can be expected that the extent of fluid-structure
6
H.-G. Choi, S.-Y. Hong, J.-H. Song et al. International Journal of Naval Architecture and Ocean Engineering 15 (2023) 100521

Table 3 effect on the structural displacement. The results reveal that the
Parameters for the hydrofoil wake oscillator model. non-dimensional wake variable fluctuating with a peak value of 2
Parameters Values and an appropriate vortex shedding frequency was observed in all
Structural density, rs 8000 kg/m3
cases.
Mass per span, ms 6.4096 kg/m In contrast, Fig. 11 shows that the effect of the fluid-structure
Young's modulus, E 1.9e11 Pa interaction was considered. The results reveal that the non-
Poisson's ratio, nu 0.305 dimensional wake exhibited a value of 2 before and after the
Elastic axis, EA 0.0509 m
lock-in range, which is similar to the tendency shown in Fig. 10,
Static unbalance, sc 0.0431
Radius of gyration, rq 0.026 m whereas the non-dimensional wake value was higher than 2 in the
Inertial moment, Iq 0.0046 kg ▪ m lock-in range. This is because the vortex shedding frequency was
Structural damping, zh, zq 0.02 synchronized with the wet twisting frequency of the hydrofoil at
Natural frequency for bending motion, fh 309Hz
12 m/s. Therefore, the hydrofoil wake oscillator model can evaluate
Natural frequency for twisting motion, fq 1084Hz
Non-linear damping coefficient, ε 0.3
the extent of the fluid-structure interaction and its contribution to
Fluid-structure interaction coefficient, A1, A2 500 the vortex-induced vibration using non-dimensional wake
Time step, △t 2E  5 parameter.
The fluid-structure interaction coefficient determines the de-
gree of fluid-structure interaction. As shown in Fig. 12, the peak of
interaction can be predicted using the non-dimensional wake non-dimensional wake were derived according to the flow velocity
value. and fluid-structure interaction coefficients. The higher non-
It is important to evaluate the extent of fluid-structure inter- dimensional wake values were derived near 12 m/s owing to the
action effect to prevent the catastrophic failure of structures. lock-in. At a velocity of 5 ~9 m/s, which is pre lock in condition,
Basically, the degree of fluid-structure interaction is associated there was no change at 2 even with an increase in the fluid-
with the amplitude of vortex-induced vibration. Hence, an increase structure coefficient. However, at the velocity (z12 m/s) around
in the fluid-structure interaction causes severe vibration, and it the wet twisting natural frequency, a gradual extension of the lock-
indicates that the vortex shedding frequency was synchronized to in range with an increase the fluid-structure interaction coefficient
the natural frequency of the structure. This phenomenon is was observed. Especially, when the fluid-structure interaction co-
commonly defined as “lock-in”. Thus, it is essential to reduce the efficient is zero, the peak values of q converged to 2 regardless of
degree of fluid-structure interaction in the initial design stage. the flow velocity conditions. Otherwise, the lock-in range was
Figs. 10 and 11 show the non-dimensional wake values in the extended to the 11 ~14 m/s range as the fluid-structure interaction
absence and presence of fluid-structure interaction. The simulation coefficient increased.
cases were composed of three cases: pre lock-in range, lock-in Fig. 13 shows the effect of the fluid-structure coefficient on the
range and post lock-in range to distinguish the non-dimensional amplitude of the vortex-induced vibration. The gradient of ampli-
wake characteristics for each case. Fig. 10 shows the fluid- tude is highly small order approximately 107 m except the lock-in
structure coefficient is zero. Thus, the displacement of the hydro- point. This indicates that the fluid-structure coefficient exerts a
foil exerts no effect on the fluid domain and the fluid force exerts no small effect on the amplitude of the vortex-induced vibration as it is

Fig. 10. Non-dimensional wake values(q) in time domain with A1,2 ¼ 0: (a) pre lock-in range (7 m/s); (b) lock-in range (12 m/s); (c) post lock-in range (15 m/s).

Fig. 11. Non-dimensional wake values(q) in time domain with A1,2 ¼ 500: (a) pre lock-in range (7 m/s); (b) lock-in range (12 m/s); (c) post lock-in range (15 m/s).

7
H.-G. Choi, S.-Y. Hong, J.-H. Song et al. International Journal of Naval Architecture and Ocean Engineering 15 (2023) 100521

Fig. 14. Vortex-induced vibration derived from a hydrofoil oscillator model based on a
near-vortex strength(G).

Fig. 12. Non-dimensional wake values(q) according to the flow velocity(U) and fluid-
structure interaction coefficients (A1,2): (a) x-y-z plane; (b) x-y plane.

Fig. 15. Modal analysis of the hydrofoil for wet bending and twisting modes.

with the vortex shedding frequency in the lock-off range, whereas


it is correlated with the wet twisting frequency in the lock-in range.
The lock-in range was generated from 11 to 13 m/s with a signifi-
cantly high vibration and strong fluid-structure interaction. The
amplitude in the lock-in range was found to be approximately 25
times higher than that in lock-off condition.
The predicted vortex-induced vibration response was verified
using the experimental results shown in Figs. 16 and 17. Fig. 16
shows the peak frequency of vortex-induced vibration for each
flow velocity. In the lock-off range, the peak frequency of each
velocity exhibited a similar tendency to the experimental results for
deriving the vortex shedding frequency. Furthermore, the hydrofoil
Fig. 13. Gradient of vortex-induced vibration amplitude according to the flow veloc-
ity(U) and fluid-structure interaction coefficients (A1,2). wake oscillator predicted the lock-in range from 11 to 13 m/s fixing
at the wet twisting frequency.
Fig. 17 shows the comparison of the amplitude of the vortex-
approximately 1/100 of the amplitude of the vortex-induce vibra- induced vibration against the experimental results. Only small
tion. The extension of the lock-in range can be determined using vortex-induced vibration was observed in the lock-off range.
the peak of the gradient. However, a very large amplitude was induced in the lock-in range
Based on the hydrofoil wake oscillator model, the vortex- (11 ~13 m/s), which was almost 100 times higher than that under
induced vibration was derived as shown in Fig. 14. The blue line the lock-off condition. Thus, the hydrofoil wake oscillator derives
indicates the vibration response and the black line is the peak the relevant vortex-induced vibration with a similar tendency and
frequency of vibration. The wet bending and twisting frequency are it was consistent with the experimental results.
represented by the green line and red line, respectively. The wet The degree of fluid-interaction was determined, as shown in
bending and twisting mode shape and frequency are determined in Fig. 18. The fluid-structure interaction exhibited no effect in the
Fig. 15. From the black line, the peak frequency was not correlated lock-off range as the amplitude of the non-dimensional wake

8
H.-G. Choi, S.-Y. Hong, J.-H. Song et al. International Journal of Naval Architecture and Ocean Engineering 15 (2023) 100521

dimensional wake were determined with a value of above 2 in lock-


in range. This indicates that the fluid-structure interaction was
actively generated in the lock-in range. Thus, the accurate vibration
can be predicted compared to a 1-way FSI system. Particularly, the
largest vibration occurred at 13 m/s as the degree of fluid-structure
interaction was at the highest value. The result revealed that the
developed hydrofoil wake oscillator model can predict the ampli-
tude of vortex-induced vibration, as well as the lock-in range that
determines the contribution of the fluid-structure interaction
effect.

4. Conclusion

In this study, a hydrofoil wake oscillator model was developed


by coupling hydrofoil dynamic motion equations and wake equa-
tion. During the coupling process, a near-vortex strength function
was developed to overcome the empirical parameter used in the
conventional wake oscillator model. The near-vortex strength was
Fig. 16. Peak frequency of vortex-induced vibration derived from hydrofoil oscillator
calculated using the boundary layer on the trailing edge, and it was
model based on a near-vortex strength(G) comparison against experimental results. verified by comparison the predicted value against the experi-
mental results.
Subsequently, the vortex-induced vibration was derived based
on the hydrofoil wake oscillator model. The results revealed that
the peak frequency and amplitude of vortex-induced vibration
were consistent with the experimental results. The vortex-induced
vibration frequency was locked on the wet twisting frequency from
11 to 13 m/s, which accompanied a very large vibration. In addition,
the extent of the effect of the fluidestructure interaction on the
vortex-induced vibration could be evaluated by determining the
non-dimensional wake value. The fluid-structure interaction was
weak in the lock-off range, whereas strong interaction was actively
generated in the lock-in range with a value of above 2. The largest
vibration occurred at 13 m/s due to the highest fluid-structure
interaction effect. This result confirmed the usability of the devel-
oped hydrofoil wake oscillator model on the initial design of a
rudder or sail for the effective prevention of severe vortex-induced
vibration.
In future studies, the structural equations should be improved to
consider hydrofoils with sweep-shape rather than the same shape
with spanwise direction. Additionally, the boundary condition
should be improved to depict the rudder characteristics including a
Fig. 17. Amplitude of vortex-induced vibration derived from hydrofoil oscillator model stock. From a wake oscillator perspective, the wake equation might
based on a near-vortex strength(G) comparison against experimental results. be modified to reflect the vortex characteristics from the oblique
trailing edge which generates the asymmetric vortices. Subse-
quently, studies should be conducted to obtain the vibration
response according to the trailing edge shapes.

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.

Acknowledgements

This research was funded by the Institute of Engineering


Research at Seoul National University, the BK21 program(Graduate
research and education program for environ-ment-friendly and
Fig. 18. Amplitude of non-dimensional wake value(q) from hydrofoil oscillator model
digitalized naval architecture and ocean engineering) funded by the
based on a near-vortex strength(G).
Ministry of Education, Republic of Korea and Basic Science Research
Program through the National Research Foundation of Korea (NRF)
converged to 2. Hence, the fluid-structure interaction effect was funded by the Ministry of Education, Science and Technology
negligible in the lock-off range. In contrast, the amplitudes of non- (2021R1F1A1059914).
9
H.-G. Choi, S.-Y. Hong, J.-H. Song et al. International Journal of Naval Architecture and Ocean Engineering 15 (2023) 100521

References Kurushina, V., Postnikov, A., Franzini, G.R., Pavlovskaia, E., 2022. Optimization of the
wake oscillator for transversal viv. J. Mar. Sci. Eng. 10, 293.
Mouritz, A.P., Gellert, E., Burchill, P., Challis, K., 2001. Review of advanced composite
Ahmad, N.N., Proctor, F., 2014. Review of idealized aircraft wake vortex models. In:
structures for naval ships and submarines. Compos. Struct. 53, 21e42.
52nd Aerospace Sciences Meeting, p. 927.
Nayfeh, A.H., 2011. Introduction to Perturbation Techniques. John Wiley & Sons.
Ausoni, P., 2009. Turbulent Vortex Shedding from a Blunt Trailing Edge Hydrofoil.
Obasaju, E., Bearman, P., Graham, J., 1988. A study of forces, circulation and vortex
EPFL Phd. Thesis.
patterns around a circular cylinder in oscillating flow. J. Fluid Mech. 196,
Bai, X., Qin, W., 2014. Using vortex strength wake oscillator in modelling of vortex
467e494.
induced vibrations in two degrees of freedom. Eur. J. Mech. B Fluid 48, 165e173.
Ogink, R.H.M., Metrikine, A.V., 2010. A wake oscillator with frequency dependent
Balasubramanian, S., Skop, R., 1996. A nonlinear oscillator model for vortex shed-
coupling for the modeling of vortex-induced vibration. J. Sound Vib. 329,
ding from cylinders and cones in uniform and shear flows. J. Fluid Struct. 10,
5452e5473.
197e214.
Parkinson, G., 1989. Phenomena and modelling of flow-induced vibrations of bluff
Bishop, R.E.D., Hassan, A.Y., 1964. The lift and drag forces on a circular cylinder
bodies. Prog. Aero. Sci. 26, 169e224.
oscillating in a flowing fluid. Proc. Roy. Soc. Lond. Series A. Math. Phys. Sci. 277,
Prangemeier, T., Rival, D., Tropea, C., 2010. The manipulation of trailing-edge
51e75.
vortices for an airfoil in plunging motion. J. Fluid Struct. 26, 193e204.
Blevins, R.D., 2001. Flow-induced Vibration, second ed. Krieger publishing company.
Qin, X., Bai, W., Bai, Y., 2021. Analysis and prediction for structural responses and
Bourgoyne, D.A., 2005. Vortex shedding from a hydrofoil at high Reynolds number.
hydrodynamic properties of two-degree-of-freedom vortex induced vibration.
J. Fluid Mech. 531, 293e324.
Mech. Adv. Mater. Struct. 28, 16e26.
Chae, E.J., Akcabay, D.T., Lelong, A., Astolfi, J.A., Young, Y.L., 2016. Numerical and
Sanchis, A., Sælevik, G., Grue, J., 2008. Two-degree-of-freedom vortex-induced vi-
experimental investigation of natural flow-induced vibrations of flexible hy-
brations of a spring-mounted rigid cylinder with low mass ratio. J. Fluid Struct.
drofoils. Phys. Fluids 28, 075102.
24, 907e919.
Facchinetti, M.L., De Langre, E., Biolley, F., 2004. Coupling of structure and wake
Skop, R.A., Balasubramanian, S., 1997. A new twist on an old model for vortex-
oscillators in vortex-induced vibrations. J. Fluid Struct. 19, 123e140.
excited vibrations. J. Fluid Struct. 11, 395e412.
Feng, Y., Chen, D., Li, S., Xiao, Q., Li, W., 2022. Vortex-induced vibrations of flexible
Skop, R., Luo, G., 2001. An inverse-direct method for predicting the vortex-induced
cylinders predicted by wake oscillator model with random components of
vibrations of cylinders in uniform and nonuniform flows. J. Fluid Struct. 15,
mean drag coefficient and lift coefficient. Ocean Eng. 251, 110960.
867e884.
Fidler, J.E., 1974. Approximate method for estimating wake vortex strength. AIAA J.
Srinivasan, S., Narasimhamurthy, V., Patnaik, B., 2018. Reduced order modeling of
12, 633e635.
two degree-of-freedom vortex induced vibrations of a circular cylinder. J. Wind
Gerrard, J., 1966. The mechanics of the formation region of vortices behind bluff
Eng. Ind. Aerod. 175, 342e351.
bodies. J. Fluid Mech. 25, 401e413.
Theodorsen, T., 1979. General Theory of Aerodynamic Instability and the Mecha-
Gutierrez, A.R., Fernandez, G.U., Errasti, I., Zulueta, E., 2019. Analysis and prediction
nism of Flutter. National Advisory Committee for Aeronautics, USA. Report 496.
for structural responses and hydrodynamic properties of two-degree-of-
Vandiver, J.K., Swithenbank, S.B., Jaiswal, V., Jhingran, V., 2006. Fatigue damage from
freedom vortex induced vibration. Energies 11, 3107.
high mode number vortex-induced vibration. Int. Conf. Offshore Mech. Arctic
Hansen, M.O., Charalampous, A., Foucaut, J.M., Cuvier, C., Velte, C.M., 2019. Vali-
Eng. 47497, 803e811.
dation of a model for estimating the strength of a vortex created from the
Wu, J.C., 1981. Theory for aerodynamic force and moment in viscous flows. AIAA J.
bound circulation of a vortex generator. Energies 12, 2781.
19, 432e441.
Hodges, D.H., Pierce, G.A., 2011. Introduction to Structural Dynamics and Aero-
Zobeiri, A., 2012. Effect of Hydrofoil Trailing Edge Geometry on the Wake Dynamics.
elasticity, vol. 15. cambridge university press.
EPFL Phd. Thesis.
Jauvtis, N.a., Williamson, C., 2004. The effect of two degrees of freedom on vortex-
induced vibration at low mass and damping. J. Fluid Mech. 509, 23e62.

10

You might also like