Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

Journal of

Marine Science
and Engineering

Article
Numerical Simulation on Vortex Shedding from a
Hydrofoil in Steady Flow
Jian Hu, Zibin Wang, Wang Zhao, Shili Sun, Cong Sun and Chunyu Guo *
College of Shipbuilding Engineering, Harbin Engineering University, Harbin 150001, China;
hujian791018@163.com (J.H.); wangzibin@hrbeu.edu.cn (Z.W.); zhaowang54@163.com (W.Z.);
shili_sun@163.com (S.S.); suncong@hrbeu.edu.cn (C.S.)
* Correspondence: guochunyu_heu@outlook.com or guochunyu@hrbeu.edu.cn

Received: 19 February 2020; Accepted: 3 March 2020; Published: 12 March 2020 

Abstract: This paper presents a numerical modeling procedure for the idealization of vortex shedding
effects in the wake flow field of a NACA0009 hydrofoil. During the simulation, the lift and drag
acting on the hydrofoil were monitored, and the vortex-shedding frequency of the hydrofoil was
analyzed. The effects of inflow velocity, trailing-edge thickness, angle of attack, and maximum
hydrofoil thickness on vortex shedding were investigated. The results indicate that an increase in the
inflow velocity led to an increase in the vortex-shedding frequency and a negligible change in the
Strouhal number. Furthermore, as the thickness of the trailing edge increased, the vortex-shedding
frequency decreased gradually, whereas the Strouhal number first increased and then decreased.
Vortex shedding and lift curve oscillations ceased altogether after the angle of attack of the hydrofoil
increased beyond a certain threshold. When the maximum hydrofoil thickness was increased
while keeping the thickness and chord length of the trailing edge constant, the vortex-shedding
frequency decreased.

Keywords: vortex shedding; wake; flow-induced vibrations

1. Introduction
Vortex shedding from the trailing edge of a foil is a significant phenomenon in various engineering
disciplines. Flows separated by a foil will recombine at the upper and lower surfaces of the trailing
edge of the foil, forming periodic eddies in the wake field and inducing mechanical vibrations.
This phenomenon was first observed in aircraft airfoils; because airfoils typically exhibit very high
wake-vortex shedding frequencies, they are at a considerable risk of fatigue damage. The damage
caused by vortex shedding is particularly severe if the vortex-shedding frequency approaches a natural
frequency of the airfoil. A similar phenomenon is known to occur in propellers and is particularly
concerning in marine propellers because an intense degree of vortex shedding leads to periodic fatigue
stresses in the blades of the propeller that shorten their service life. Because vortex shedding can lead
to a series of failures in engineered structures, it is important to study the causes of this phenomenon
and elucidate the behavior of vortex shedding in a variety of operating conditions.
The mechanical vibrations induced by vortex shedding have been widely investigated. Vincenc
Strouhal discovered that the audible tone of a wire “singing” in the wind is proportional to the quotient
of the wind speed by the wire thickness (i.e., tone = proportionality coefficient × wind speed/wire
thickness) and that each cross-section has a certain proportionality coefficient in the noncritical range
of Reynolds numbers. This is known as the Strouhal number, and it is an important parameter to
evaluate the vortex-shedding frequency. Theodore von Kármán linked this phenomenon to the stable
and staggered vortices that form behind a cylindrical object, which was later termed the Kármán
vortex street effect. In addition, he proposed the theory of vortex street stability [1]. Perry et al. [2]

J. Mar. Sci. Eng. 2020, 8, 195; doi:10.3390/jmse8030195 www.mdpi.com/journal/jmse


J. Mar. Sci. Eng. 2020, 8, 195 2 of 20

obtained images of the vortex-shedding process from the wake field of a circular cylinder by tracing
the motion of aluminum particles in long-exposure photography. Griffin [3] and Williamson et al. [4]
investigated the alternating shedding of vortices from a cylinder and discovered that this process is
caused by interactions between vortices on the upper and lower surfaces of the cylinder; once a vortex
on a given surface reaches a sufficient strength, it draws the opposing shear layer across the near wake,
causing the vortices to shed alternatingly from the upper and lower surfaces. Gerrard [5] observed
vortex shedding from circular and bluff cylinders with different cross-sectional shapes and deduced
that the vortex generation is affected by the end conditions of the cylinders. Gerich [6] proved that the
shape of the ends of a cylinder affect the mechanism of vortex shedding from these ends. In addition,
Achenbach and Heinecke [7] conducted wind tunnel experiments and discovered that the wake of a
cylinder became more regular as the roughness of the cylinder’s surface increased. Furthermore, the
Strouhal number of a vortex shedding from a rough cylinder was estimated to be much smaller than
that of a vortex shedding from a smooth cylinder.
Through the various studies conducted on the vortex-shedding phenomenon, it was observed that
intense resonance can occur if the vortex-shedding frequency of the Kármán vortex street is equal to one
of the eigenfrequencies of a given object. This discovery shifted the focus of the study of vortex shedding
from simple cylindrical objects to objects of practical interest, such as hydrofoils. A series of analytical
studies were also conducted on the basis of experimental data. Ausoni [8] experimentally investigated
the effects of cavitation on vortex shedding from symmetrical hydrofoils with a blunted trailing edge
and analyzed the wake structure, vortex-shedding frequency, and hydrofoil resonance frequency
associated with each stage of cavitation development. He discovered that the vortex-shedding
frequency can “lock-in” to an eigenfrequency of a hydrofoil. Lotfy and Rockwell [9] investigated
vortex shedding in the near-wake of an oscillating trailing edge. Prasad and Williamson [10] analyzed
the vortex dynamics of hydrofoil wakes based on the instability of the shear layer separating the
water from a bluff body. Dwayne et al. [11] investigated vortex shedding at a high Reynolds number
from hydrofoils with different terminating bevel angles on their trailing edges and found that the
vortex-shedding intensity is higher in blunter and thicker trailing edges. Zobeiri et al. [12] conducted
an experimental study on the effects of the trailing-edge shape on vortex shedding and flow-induced
vibrations and found that oblique trailing edges significantly reduce flow-induced vibrations. This
was attributed to the collisions between upper and lower vortices in an oblique trailing edge, which
leads to vorticity redistribution.
Numerical modeling has been widely employed for the study of vortex shedding. Many previous
numerical studies focused on vortex shedding from circular objects. For example, Williamson [13]
numerically simulated the wake flows of a cylindrical object. William [14] used complex analysis to
formulate an analytical expression for the shedding frequency of Kármán vortex streets from the trailing
edge of an airfoil. However, the shedding frequencies predicted by this expression were typically
much lower than the experimentally observed ones. Huerre and Monkewitz [15] and Oertel et al. [16]
analyzed the effects of flow instabilities on vortex-shedding behaviors of blunt hydrofoils analytically.
Potential flow theory is often used to study the wake-vortex shedding of swinging or flapping objects.
However, current studies on vortex shedding focus on the formation of Kármán vortex streets behind
static objects, which is essentially a boundary layer dissociation problem; these problems are very
difficult to model with the potential flow theory. Another significant feature of this problem is
that the trailing edge of the hydrofoil is often very thin, which leads to very large local Reynolds
numbers. This results in unique physical phenomena that can only be modeled by precise numerical
simulations. Lee et al. [17] studied the vortex shedding of hydrofoils, numerically considering a
variety of trailing-edge shapes, and compared their findings with the experimental findings of Ausoni
and Zobeiri. In addition, they investigated the effects of periodically varying free-stream flows on
vortex shedding.
Many studies using numerical simulations of or experimental tests on vortex shedding have been
conducted, and the subsequent hydrodynamic force fluctuations of hydrofoils have been analyzed.
J. Mar. Sci. Eng. 2020, 8, 195 3 of 20

The main purpose of the abovementioned studies was to evaluate the amplitude and frequency of force
fluctuation, which have significant effects on marine structures. Ausoni [8] experimentally studied
the vortex-shedding frequency of hydrofoils with perpendicularly truncated trailing edges and found
that the increase in frequency was almost linear with the inflow velocity. Dwayne et al. [11] studied
the vortex-shedding frequency of hydrofoils with obliquely truncated trailing edges and found that
thicker hydrofoils or blunter trailing edges correspond to higher vortex-shedding frequencies. Another
experimental study was conducted by Zobeiri [12], wherein it was found that the obliquely truncated
trailing edges can significantly reduce the vortex-shedding strength, and subsequently, the amplitude
of the force fluctuation. Analytical studies of vortex shedding were performed by Blake [14] and
Oertel [16] on hydrofoils and cylinders, respectively. However, the force of analytical results is much
less than that by experiments. The vortex shedding of hydrofoils is strongly related to the formation of
viscous boundary layers. From this perspective, methods based on the potential flow assumptions can
hardly predict vortex shedding precisely. Instead, computational fluid dynamics (CFD) simulations can
effectively represent vortex shedding. Although Reynolds-averaged Navier–Stokes simulations [8] or
large Eddy simulations (LES) [17] cannot fully simulate turbulent flows, their numerical results generally
agree well with experimental results and can reflect the physical characteristics of vortex shedding.
The above discussion shows that vortex shedding from static hydrofoils is a problem of practical
significance and academic interest as it plays a significant role in flow-induced vibrations and noise
and fatigue damage. However, research on this problem is still insufficient, and further elucidations are
required. In this study, we conducted CFD simulations of two-dimensional flows around a NACA0009
hydrofoil in a variety of operating conditions to observe vortex shedding at the trailing edge of a
hydrofoil. We investigated the lift curve and trailing-edge vortex shedding of the hydrofoil with
different inflow velocities, angles of attack, truncation points, and maximum hydrofoil thicknesses
to reveal their correlations with the vortex-shedding frequency. The findings of this study will
provide general rules for the investigation of flow-induced vibrations in objects such as hydrofoils and
marine propellers.

2. Materials and Methods

2.1. Mathematical Modeling


In this study, the finite volume method, which satisfies the conservations of mass and momentum,
was used to solve the integral equations. The vertex-centered method was adopted to form the control
elements; the second-order upwind scheme was used for the discrete differentials; the semi-implicit
method for pressure linked equations scheme was used for the flow solution. It firstly assumes a
velocity distribution u∗ , v∗ , Φ∗ (u, v, Φ) in the flow field and then calculates the coefficients and constants
in the discrete-form momentum equations at the first iteration. Then, it guesses the pressure p∗ ; it
solves the discrete momentum equation from the pressure field and the pressure correction equation
from the velocity field. After correcting the calculated pressure and velocity, all other variables of the
control equation are solved. Finally, the results are verified and iterated until convergence.
Numerical simulations were performed using the Spalart–Allmaras (S-A) turbulence model,
which is suitable for high-precision boundary-layer computations, e.g., lift computations.
Spalart and Allmaras [18] argued that energy and information flow from large scales to smaller
scales in free shear flows and represented the turbulent eddy viscosity coefficient by two terms:
a production term and a diffusion term. They constructed a one-equation turbulence model by
utilizing the experience gained from other turbulence models and by amending the other turbulent
flow-field models.
In the S-A equation, the definition of Reynold stress is:
 
−ui u j = 2vt Sij , Sij ≡ ∂Ui /∂x j + ∂U j /∂xi /2 (1)
J. Mar. Sci. Eng. 2020, 8, 195 4 of 20

where, the turbulent eddy viscosity vt can be expressed as:

χ3 ṽ
vt = ṽ fv1 , fv1 = ,χ ≡ , (2)
χ3 + c3v1 v

where v is the kinematic molecular viscosity and ṽ is the working variable. From the above equations,
we obtain the transport equation:

Dṽ 1h i  c ṽ
 
= cb1 [1 − ft2 ]S̃ṽ + ∇((v + ṽ)∇ṽ) + cb2 (∇ṽ)2 − cw1 fw − b12 ft2 + ft1 ∆U2 , (3)
Dt σ κ d

where S(~) is the magnitude of vorticity; b is the distance to the nearest wall; ∆U is the velocity
difference between a point in the fluid and a point on the wall; ft1 , ft2 , and fw are functional equations,
and the rest of the symbols represent constants. The specific definitions of these symbols are as follows:

ṽ χ
S̃ ≡ S + fv2 , fv2 = 1 − (4)
κ2 d2 1 + χ fv1

w2t h 2
 
i
2 2
d + gt dt , gt ≡ min(0.1, ∆U/wt ∆x),

ft1 = ct1 gt exp−ct2 (5)
∆U2
 
ft2 = ct3 exp −ct4 χ2 , (6)
1/6
 1 + c6w3 

 , g = r + cw2 r6 − r , r ≡ ṽ ,
 
fw = g
6
(7)
6
g + cw3 S̃κ2 d2

The S-A turbulence model is more computationally efficient than two-equation turbulence models,
and it is also very stable and accurate. The S-A model accurately simulates the region of the wake
where strain rates are dominated by vorticity, and it predicts the turbulent eddy viscosity.

2.2. Test Case and Computational Mesh


In this study, a two-dimensional flow method was used to study vortex shedding. The model
used in this study was a NACA0009 hydrofoil with a chord length c of 100 mm and a trailing-edge
thickness b of 3.22 mm, which is identical to the model used by Ausoni et al. (2006). Figure 1 shows a
diagram of the hydrofoil model with the chord length (c), trailing-edge thickness (b), and angle of
attack (α).
J. Mar. Sci. Eng. 2020, 8, x FOR PEER REVIEW 5 of 20

Figure 1. Diagram of the hydrofoil model.


Figure 1. Diagram of the hydrofoil model.

To avoid the effects of the computational boundaries illustrated in Figure 2 on the flow simulation,
the domain was set to relatively large dimensions (13c × 6c). Typically, the outlet boundary has a
greater effect than the inlet boundary owing to the downstream vortex shedding. Thus, the hydrofoil
J. Mar. Sci. Eng. 2020, 8, 195 5 of 20

was further from the outlet than from the inlet. Neglecting the mutual effect between the inlet boundary
J. Mar. Sci. Eng. 2020, 8, x FOR PEER REVIEW 5 of 20
and the hydrofoil, the inlet was set as the velocity inlet. To achieve better conservation and more stable
simulations, the outlet was set as the pressure outlet rather than the velocity outlet. To further avoid
the effects of the boundaries, the upper and lower boundaries were set as symmetrical and slipping,
different from the nonslipping surfaces of the hydrofoil.
Figure 1. Diagram of the hydrofoil model.

Figure 1. Diagram of the hydrofoil model.


Figure 2.
Figure 2. Computational domain.
Computational domain.

The grid around the hydrofoil is shown in Figure 3. Two mesh types were used in the computational
domain, namely trimmed and prism layer meshes. The trimmed mesh was the primary type of mesh
used in the fluid region. This part of the mesh consisted of quadrilateral elements parallel to the
inflow; this configuration reduces truncation errors in the calculation. Prism layer meshes were used
to generate the boundary layers. A 3c × 2c refined region was set up around the hydrofoil to refine
the flows around it, and the mesh size in this region was 5% of that of the unrefined region. The
computational domain contained a total of 2,495,989 mesh elements.
Figure 2. Computational domain.

Figure 3. Computational domain mesh.

The grid around the hydrofoil is shown in Figure 3. Two mesh types were used in the
computational domain, namely trimmed and prism layer meshes. The trimmed mesh was the
primary type of mesh used in the fluid region. This part of the mesh consisted of quadrilateral
elements parallel to the inflow; this configuration reduces truncation errors in the calculation. Prism
layer meshes were used to generate the boundary layers. A 3c × 2c refined region was set up around
the hydrofoil to refine the flows around it, and the mesh size in this region was 5% of that of the
unrefined region. The computational domain contained a total of 2,495,989 mesh elements.
Figure 3. Computational domain mesh.
Figure 3. Computational domain mesh.

The grid
Figure around
4 shows the hydrofoil
the boundary layer is shown
grid of the in Figure In
hydrofoil. 3. these
Two computations,
mesh types were used
physical in the
variables
computational domain, namely trimmed and prism layer meshes. The trimmed
tend to vary most significantly around boundaries. Therefore, a refined boundary layer mesh was used mesh was the
primary
in type of mesh
the calculations. Theused in the
thickness of fluid region.
the first meshThislayerpart
wasofdetermined
the mesh consisted
by the Y+of quadrilateral
value. Because
the Reynolds number in the simulation was quite large (Re = 1,194,257.2), the fluid flow becamePrism
elements parallel to the inflow; this configuration reduces truncation errors in the calculation. fully
layer meshes
turbulent were used
at positions to generate
close the boundary
to the surface layers. A
of the hydrofoil. 3c × 2c refined
Therefore, the Y+region was set
value was set to
up1,around
which
the hydrofoil
yielded a first to refine
mesh thethickness
layer flows around
of 1.97it,×and
10−6them.mesh size in this
The growth rate region was 5%
of the prism of that
layer mesh ofwas
the
unrefined region. The computational domain contained a total of 2,495,989 mesh
set to 1.2, and the boundary layer mesh was configured with a total of 20 layers to ensure a smoothelements.
transition into the outer mesh. The total thickness of the boundary layer mesh was 3.67 × 10−4 m.
J. Mar. Sci. Eng. 2020, 8, 195 6 of 20

J. Mar. Sci. Eng. 2020, 8, x FOR PEER REVIEW 6 of 20

Figure 4. Boundary layer mesh of the hydrofoil’s trailing edge.


Figure 4. Boundary layer mesh of the hydrofoil’s trailing edge.

Figure 4 shows
2.3. Convergence Analysisthe boundary layer grid of the hydrofoil. In these computations, physical
variables tend to vary most significantly around boundaries. Therefore, a refined boundary layer
2.3.1.
mesh Mesh Convergence
was used in the calculations. The thickness of the first mesh layer was determined by the Y+
value. Because the
Using the meshing Reynolds number
strategy in the
reported simulation
by Ausoni et al.was
[8], aquite large
refined (Rewas
mesh = 1,194,257.2),
configured the fluid
to assess
flow became fully
the mesh convergence. turbulent at positions close to the surface of the hydrofoil. Therefore, the Y+ value
was set
Thetomedium
1, whichmesh
yielded
hadaafirst
totalmesh layer thickness
of 2,495,989 elements,ofand
1.97the
× 10 −6 m. The growth rate of the prism
thickness of the boundary layer was
layer mesh−4 was set to 1.2, and the boundary layer mesh was configured
3.67 × 10 m. The mesh size was further increased and reduced to verify the mesh with a total of 20 layers
convergence, to
and
ensure
the a smooth
resulting mesh transition
had mediuminto the
andouter
fine mesh. The total
grids with thickness
1,562,259 and of the boundary
4,027,882 layer
elements, mesh was
respectively.
3.67 × 10 −4 m.
When the grid scale was changed between coarse, medium, and fine while keeping the number of
boundary layers constant, the extension rates of the layers changed to 1.4, 1.3, and 1.2, respectively.
2.3. Convergence Analysis
Then, the time step was set to 1 × 10−6 s. Table 1 lists the drag coefficient Cd , maximum lift
coefficient Cl (max), and Strouhal number (St ) obtained with different grid scales.
2.3.1. Mesh Convergence
Using the meshing strategyTable 1. Mesh
reported byconvergence analysis
Ausoni et al. results. mesh was configured to assess
[8], a refined
the mesh convergence.
Mesh Cd Cl (max) St
The medium mesh had a total of 2,495,989 elements, and the thickness of the boundary layer
Coarse 0.0270 0.06018 0.190
was 3.67 × 10−4 m. The mesh size was further increased and reduced to verify the mesh convergence,
Medium 0.0258 0.05605 0.192
and the resulting mesh had medium 0.0253
Fine and fine grids
0.05521 with 1,562,259
0.197 and 4,027,882 elements,
respectively. When the grid scale was changed between coarse, medium, and fine while keeping the
number of boundary layers constant, the extension rates of the layers changed to 1.4, 1.3, and 1.2,
From Table 1, it can be observed that the calculation results with the three grid conditions are
respectively.
similar. The mesh size was further reduced −6 under the medium mesh condition, and the resulting Cd ,
Then, the time step was set to 1 × 10 s. Table 1 lists the drag coefficient 𝐶𝑑 , maximum lift
Cl , and St changed by 1.9%, 1.5%, and 2.6%, respectively. The degree of change in data is small, and
coefficient 𝐶𝑙 (max), and Strouhal number (𝑆𝑡 ) obtained with different grid scales.
thus, it can be considered grid-independent.
Table 1. Mesh convergence analysis results.
2.3.2. Time-Step Convergence
We searched for theMesh
time step that was𝑪𝒅 required 𝑪to
𝒍 (max) 𝑺𝒕
converge the simulation when the mesh
Coarse 0.0270 0.06018 0.190
contained 2,495,989 elements. The time step of the simulation was gradually decreased from 2 × 10−4
to 1 × 10−6 s. The Cd , ClMedium
(max), and St that0.0258
were obtained0.05605 0.192value (Dt) are presented
with each time step
in Table 2. Fine 0.0253 0.05521 0.197

From Table 1, it can be observed that the calculation results with the three grid conditions are
similar. The mesh size was further reduced under the medium mesh condition, and the resulting 𝐶𝑑 ,
𝐶𝑙 , and 𝑆𝑡 changed by 1.9%, 1.5%, and 2.6%, respectively. The degree of change in data is small, and
thus, it can be considered grid-independent.

2.3.2. Time-Step Convergence


J. Mar. Sci. Eng. 2020, 8, 195 7 of 20

Table 2. Time step convergence.

Dt (s) Cd Cl (max) St
2× 10−4 0.0145 0.000299 0.112
1 × 10−4 0.0145 8.98 × 10−5 0.150
5 × 10−5 0.0157 0.01663 0.0171
6 × 10−6 0.0239 0.05293 0.188
2 × 10−6 0.0252 0.05599 0.197
1 × 10−6 0.0258 0.05605 0.197

From Table 2, it can be observed that the results become increasingly similar as the time step
shortens. In particular, the results obtained at Dt = 2 × 10−6 and 1 × 10−6 s are very similar, indicating
that the results obtained with Dt = 1 × 10−6 s have converged.

2.3.3. Turbulence Model Selection


The mesh and time-step convergence analyses show that a converged result can be obtained from
the simulation with a total mesh number of 2,495,989 and a time step of 1 × 10−6 s. Therefore, these
two parameters were used to verify the convergence of the S-A, shear stress transport k-omega, and
K-epsilon turbulence models. The comparison results are shown in Table 3.

Table 3. Comparison of Cl and St obtained by different turbulence models.

Turbulence Model Cl (max) St


S-A 0.05605 0.192
Shear stress transport k-omega 0.06199 0.220
K-epsilon 0.05602 0.186

Ausoni [8] conducted a comparative test between a normal hydrofoil and a hydrofoil with rough
stripes at the leading edge. In the latter case, a strong turbulent boundary layer was developed.
With an inflow velocity of 12 m/s, the normal and rough-stripe hydrofoils had St of 0.245 and 0.174,
respectively, indicating that the vortex-shedding frequency is highly related to the occurrence of
turbulence. However, there is insufficient understanding of the turbulence around the hydrofoil. The
calculation results obtained with the three models are all within the experimental value range and
have no significant differences. The S-A model could accurately calculate the near-wall flow of the
hydrofoil, and thus, was used for the subsequent simulations.

3. Results

3.1. Simulation Results


Based on the results of the mesh and time-step convergence analyses, a base mesh size of 0.005 m
and time step of 1 × 10−6 s were used as the computational parameters of the simulations. The
inflow velocity Ure f was set to 12 m/s. The lift and drag of the hydrofoil were then obtained from
these simulations. A dimensionless time quantity, t = TU/c, was defined as the abscissa to plot the
relevant curves.
Figure 5 shows the changes in the lift curve from the beginning of the simulation until a steady
state was reached. Two distinct stages can be identified from these plots—at the beginning of the
simulation, the flow field is in an unsteady state, and the amplitude of the lift curve oscillations
increases; after some time, the amplitude of the lift curve oscillations stops changing and the flow
field becomes stable. Figure 6 shows the Fourier transform plot of the convergence-time history curve.
From this figure, the frequency (735.3 Hz) and amplitude (0.0689) of the lift curve oscillations under
the given set of operating conditions were obtained.
J. Mar. Sci. Eng. 2020, 8, x FOR PEER REVIEW 8 of 20
increases; after some time, the amplitude of the lift curve oscillations stops changing and the flow
field becomes stable. Figure 6 shows the Fourier transform plot of the convergence-time history
increases; after some time, the amplitude of the lift curve oscillations stops changing and the flow
curve. From this figure, the frequency (735.3 Hz) and amplitude (0.0689) of the lift curve oscillations
field becomes stable. Figure 6 shows the Fourier transform plot of the convergence-time history
under the given set of operating conditions were obtained.
curve. From this figure, the frequency (735.3 Hz) and amplitude (0.0689) of the lift curve oscillations
J. Mar. Sci. Eng. 2020, 8, 195 8 of 20
under the given set of operating conditions were obtained.
Fully developed
0.10
Fully developed
0.10
0.05

0.05

Cl Cl
0.00

0.00
-0.05 Unsteady

-0.05 Unsteady
-0.10
0 1 2 3 4 5
-0.10 t
0 1 2 3 4 5
t
Figure 5. Convergence-time history curve of 𝐶𝑙 .
Figure 5. Convergence-time history curve of Cl .
Figure 5. Convergence-time history curve of 𝐶𝑙 .
0.08
0.07 (735.30,0.0689)
0.08
0.06
0.07
Amplitude

(735.30,0.0689)
0.05
0.06
0.04
Amplitude

0.05
0.03
0.04
0.02
0.03
0.01
0.02
0.00
0.01
0 2,000 4,000 6,000 8,000 10,000
0.00
f [Hz]
0 2,000 4,000 6,000 8,000 10,000
Figure 6. Characteristics of the liftf curve
[Hz] in the frequency domain.
Figure 6. Characteristics of the lift curve in the frequency domain.
Figure 7 illustrates the changes in the drag curve of the hydrofoil over time. Because the flow field
Figure 6. Characteristics of the lift curve in the frequency domain.
Figure 7inillustrates
was unsteady the early the changes
stages of theinsimulation,
the drag curve
large of the hydrofoil
changes over time.
were initially Because
observed thedrag
in the flow
field was unsteady
coefficient. Once the in thefield
flow earlybecame
stages of the simulation,
stable, large changes
the drag coefficient were initially
still exhibited observed
periodic in the
oscillation,
Figure 7 illustrates the changes in the drag curve of the hydrofoil over time. Because the flow
drag coefficient.
similar Once the
to the lift curve. Theflow field became
steady-state stable,frequency
oscillation the drag ofcoefficient
the drag still exhibited
curve (1460.92 periodic
Hz) is
field was unsteady in the early stages of the simulation, large changes were initially observed in the
oscillation, similar
approximately twice thatto the liftliftcurve.
of the The steady-state
curve, indicating oscillation
that a pair of vortex frequency of the
shedding form one drag curve
oscillation
drag coefficient. Once the flow field became stable, the drag coefficient still exhibited periodic
of(1460.92Hz)
the lift while is approximately
a single vortex twice shedding that from
of thethe
liftupper
curve,orindicating that aforms
lower surface pair of
onevortex shedding
oscillation of
oscillation,
J. Mar. Sci. Eng.similar to the
2020, 8, x FOR PEER lift curve. The steady-state oscillation frequency of the drag curve
REVIEW 9 of 20
form
the one oscillation of the lift while a single vortex shedding from the upper or lower surface forms
drag.
(1460.92Hz) is approximately twice that of the lift curve, indicating that a pair of vortex shedding
one oscillation of the drag.
form one oscillation of the lift while a single vortex shedding from the upper or lower surface forms
0.32
one oscillation of the drag.
0.28
Cd

0.24 Fully developed

Unsteady
0.20

0.16

0 1 2 3 4 5
t

Figure 7. Drag curve of the hydrofoil under convergent conditions.


Figure 7. Drag curve of the hydrofoil under convergent conditions.
Here, we define the pressure at some arbitrary point as p, which can be nondimensionalized as
Cp = 12Here,
p−p∞ we define the pressure at some arbitrary point as p, which can be nondimensionalized as
2 , where p∞ = 0.
1 p  p
ρUre f

C pFigure 8 shows that the pressure distribution on the upper and lower surfaces of the hydrofoil
2 U 2
are similar at t = 2.68; because at this time, the difference in pressures between the upper and lower
ref
, where p∞ = 0.
Figure 8 shows that the pressure distribution on the upper and lower surfaces of the hydrofoil
are similar at t = 2.68; because at this time, the difference in pressures between the upper and lower
surfaces is small, the corresponding lift coefficient is also small (0.016). Figure 8b,c show the surface
pressure distributions of the hydrofoil when the lift coefficient was at its minimum (t = 2.88) and
maximum (t = 3.12), respectively. In these figures, it may be observed that a significant pressure
Here, we define the pressure at some arbitrary point as p, which can be nondimensionalized as
1 p  p
Cp 
2 U ref
2
, where p∞ = 0.
Figure 8 shows
J. Mar. Sci. Eng. 2020, 8, 195
that the pressure distribution on the upper and lower surfaces of the hydrofoil 9 of 20
are similar at t = 2.68; because at this time, the difference in pressures between the upper and lower
surfaces is small, the corresponding lift coefficient is also small (0.016). Figure 8b,c show the surface
surfaces
pressureis distributions
small, the corresponding lift coefficient
of the hydrofoil when theislift also small (0.016).
coefficient Figure
was at 8b,c show(tthe
its minimum surface
= 2.88) and
pressure distributions of the hydrofoil when the lift coefficient was at its minimum
maximum (t = 3.12), respectively. In these figures, it may be observed that a significant pressure (t = 2.88) and
maximum = 3.12),
difference (tarises afterrespectively. In these
x/c = 0.75, which figures,
induces a lift.it may be observed that a significant pressure
difference arises after x/c = 0.75, which induces a lift.

1.0
1
Upper
0.5
Upper
0

Cp
Cp

Cut
0.0
Cut
Lower
-1
-0.5
Lower

-1.0 -2
0.00 0.25 0.50 0.75 1.00 0.00 0.25 0.50 0.75 1.00
x/c x/c

(a) t = 2.68 (b) t = 2.88

Lower
Cp

0
Cut

Upper
-1

0.00 0.25 0.50 0.75 1.00


x/c

(c) t = 3.12
Figure8.8.Distribution
Figure Distributionofofthe
thepressure
pressureon
onthe
thesurfaces
surfacesofofthe
thehydrofoil
hydrofoilover
overtime.
time.

Twomonitoring
Two monitoringpoints
points(Point
(Point11and
andPoint
Point2)2)were
wereset
setup
upatatthe
theupper
upperand
andlower
lowerturning
turningpoints
points
ofofthe
thetrailing
trailingedge
edgeofofthe
thehydrofoil.
hydrofoil.The
Thevelocity
velocityand
andpressure
pressureatatthe
themonitoring
monitoringpoints
pointsare
aredenoted
denoted
as U and P, respectively, and we further define a dimensionless variable u = U/Ure f to simplify the
velocity term. The dimensionless velocity–time and velocity–pressure curves obtained at these points
are presented in Figure 9. In the figure, the velocity and pressure curves at both the monitoring points
exhibit similar oscillatory amplitudes and frequencies but in opposing phases. This indicates that the
vortices on the upper and lower surfaces of the trailing edge are symmetrical and that they are shed
alternatingly. At each monitoring point, the velocity and pressure curves have identical periodicities
but opposing amplitudes, i.e., the maximum of one curve always corresponds to the minimum of the
other curve.
Figure 10 presents a vorticity map of the shed vortices. As can be seen from the figure, vortices
are shed alternatingly from the upper and lower surfaces of the trailing edge. The vorticities at the
core of vortices 1, 2, 3, 4, 5, and 6 are 26,000, 20,000, 16,000, 13,500, 12,000, and 11,000, respectively.
Therefore, the vorticities of the shed vortices dissipate as they move away from the trailing edge, and
the vorticity dissipation is most pronounced when the vortices first begin to dissociate from the surface
of the hydrofoil.
are presented in Figure 9. In the figure, the velocity and pressure curves at both the monitoring points
exhibit similar oscillatory amplitudes and frequencies but in opposing phases. This indicates that the
vortices on the upper and lower surfaces of the trailing edge are symmetrical and that they are shed
alternatingly. At each monitoring point, the velocity and pressure curves have identical periodicities
but opposing amplitudes, i.e., the maximum of one curve always corresponds to the minimum of the
J. Mar. Sci. Eng. 2020, 8, 195 10 of 20
other curve.

1.00
Point1
Point2

0.75

u
0.50

0.25

0.00
0 1 2 3 4
t

(a) (b)

Point1
Point2
0.4 1.00 0.4

Cp
0.2 0.75 0.2
Cp

0.0
u

Cp
0.50 0.0

-0.2
0.25 -0.2

-0.4 u

0 1 2 3 4 0.00 -0.4
0 1 2 3 4
t t

(c) (d)
Figure
Figure 9.
9. Velocity
Velocity and
and pressure
pressure variations
variations at
at the
the trailing
trailing edge
edge of
of the hydrofoil. (a)
the hydrofoil. (a) Positions
Positions of
of the
the
monitoring points. (b) Velocity variations at the two monitoring points. (c) Pressure changes at
monitoring points. (b) Velocity variations at the two monitoring points. (c) Pressure changes at the two the
two monitoring
monitoring points.
points. (d) Relationship
(d) Relationship between
between velocity
velocity and and pressure
pressure at Point
at Point 1. 1.
J. Mar. Sci. Eng. 2020, 8, x FOR PEER REVIEW 11 of 20

Figure 10 presents a vorticity map of the shed vortices. As can be seen from the figure, vortices
are shed alternatingly from the upper and lower surfaces of the trailing edge. The vorticities at the
core of vortices 1, 2, 3, 4, 5, and 6 are 26,000, 20,000, 16,000, 13,500, 12,000, and 11,000, respectively.
Therefore, the vorticities of the shed vortices dissipate as they move away from the trailing edge, and
the vorticity dissipation is most pronounced when the vortices first begin to dissociate from the
surface of the hydrofoil.

Figure 10. Vortex shedding from the trailing edge (t = 3.12).


Figure 10. Vortex shedding from the trailing edge (t = 3.12).

Because
Because vortex
vortexshedding
sheddingoften
oftenleads
leadsto
to vibrations
vibrationsinin various
various structures
structures and
and objects
objects in
in practical
practical
applications,
applications, different operating conditions were set up to investigate the factors that influencevortex
different operating conditions were set up to investigate the factors that influence vortex
shedding
shedding and
and toto elucidate
elucidate the
the vortex-shedding
vortex-shedding behaviors.
behaviors.

3.2. Frequency of Vortex Shedding at Different Velocities


The value of 𝑈𝑟𝑒𝑓 was varied to examine the effects of the inflow velocity on vortex shedding.
The value of 𝑈𝑟𝑒𝑓 was changed by altering the velocity at the inlet. A time step of 1 × 10−6 s was used,
and the thickness of the boundary layer was recalculated considering Y+ = 1 and 𝑈𝑟𝑒𝑓 that was used
in the simulation. The wake-vortex shedding frequencies that were obtained with different inflow
velocities are presented in Table 4, where f1 is the frequency that was obtained in this simulation, f2
J. Mar. Sci. Eng. 2020, 8, 195 11 of 20

3.2. Frequency of Vortex Shedding at Different Velocities


The value of Ure f was varied to examine the effects of the inflow velocity on vortex shedding. The
value of Ure f was changed by altering the velocity at the inlet. A time step of 1 × 10−6 s was used, and
the thickness of the boundary layer was recalculated considering Y+ = 1 and Ure f that was used in the
simulation. The wake-vortex shedding frequencies that were obtained with different inflow velocities
are presented in Table 4, where f 1 is the frequency that was obtained in this simulation, f 2 and f 3 are the
experimental values obtained by Ausoni et al. [8] in their smooth and rough leading-edge experiments,
respectively, and f 4 is the result that they obtained through a CFD simulation.

Table 4. Vortex-shedding frequency at different inflow velocities; f 1 : frequency obtained in this study;
f 2 and f 3 : frequencies obtained from smooth and rough leading edge experiments, respectively [8]; f 4 :
frequency obtained through a CFD simulation in this study.

Uref (m/s) f 1 (Hz) f 2 (Hz) f 3 (Hz) f 4 (Hz) St pre.


10 601.78 795.24 532.04 599.51 0.194
11 672.02 912.15 590.42 0.197
12 735.30 912.15 649.31 0.197
13 780.69 912.15 717.00 0.193
14 861.62 999.18 775.09 0.198
15 902.63 1096.76 833.92 0.194
16 963.06 1174.99 883.14 957.87 0.194
17 1024.57 1223.34 921.82 0.194
18 1118.94 1291.91 989.51 0.200
19 1146.51 1378.94 1058.08 0.194
20 1207.48 1418.49 1116.10 1212.35 0.194
21 1268.45 1184.66 0.194

Figure 11 compares the vortex-shedding frequencies that were obtained in the present simulation
with those obtained by Ausoni et al. (2006). As can be seen from the figure, the CFD calculation results
do not differ significantly from each other, but they do deviate from the experimental results to some
extent. The simulated results are similar to those obtained in the rough leading-edge experiment both
qualitatively and quantitatively, and the results of the latter may be approximated by a linear function
of J.UMar. = 60.97U
(f Sci. Eng. 2020,−8, 11.925).
x FOR PEERAusoni
REVIEW et al. argued that the differences between the numerical and
12 of 20
experimental values were a result of the differences in flow field around the hydrofoil in the numerical
the numerical
simulation simulations, the
and experiments. flows around
Because the hydrofoil
a turbulence wereused
model was assumed tonumerical
in the be fully turbulent. In thethe
simulations,
model experiments, there was a distinct turning point, where the fluid flows changed from
flows around the hydrofoil were assumed to be fully turbulent. In the model experiments, there was a laminar
to turbulent.
distinct turning point, where the fluid flows changed from laminar to turbulent.

Smooth (Ausoni(2006))
2000 Rough leading edge (Ausoni(2006))
Present CFD

1500
f [Hz]

1000

500

0
0.0 5.0x105 1.0x106 1.5x106 2.0x106 2.5x106 3.0x106 3.5x106
Re
Figure11.
Figure 11. Oscillation
Oscillation frequency
frequencywith
withdifferent
differentinflow
inflowvelocities.
velocities.

Figure 12 presents the values of 𝑆𝑡 as a function of inflow velocity. These values in the “smooth”
experiment generally fluctuated around 0.24, and the changes in these values with respect to inflow
velocity were significant. The values of 𝑆𝑡 that were obtained in the “rough leading edge”
experiment are similar to those obtained in our CFD simulation (0.19 versus 0.18), and the changes
in 𝑆𝑡 with respect to inflow velocity were insignificant in both cases.
0
0.0 5.0x105 1.0x106 1.5x106 2.0x106 2.5x106 3.0x106 3.5x106
J. Mar. Sci. Eng. 2020, 8, 195
Re 12 of 20
Figure 11. Oscillation frequency with different inflow velocities.

Figure
Figure1212presents
presentsthe
thevalues ofS𝑆t 𝑡asasa afunction
valuesof functionofofinflow
inflowvelocity.
velocity.These
Thesevalues
valuesininthe
the“smooth”
“smooth”
experiment
experiment generally fluctuated around 0.24, and the changes in these values with respect toinflow
generally fluctuated around 0.24, and the changes in these values with respect to inflow
velocity
velocitywere
weresignificant. TheThe
significant. values of St that
values of 𝑆were obtained
𝑡 that in the “rough
were obtained leading
in the edge”
“rough experiment
leading edge”
are similar toare
experiment those obtained
similar in our
to those CFD simulation
obtained in our CFD(0.19 versus (0.19
simulation 0.18),versus
and the changes
0.18), in Schanges
and the t with
respect to inflow
in 𝑆𝑡 with velocity
respect werevelocity
to inflow insignificant in both cases.in both cases.
were insignificant

0.50
Smooth(Ausoni(2006))
Present CFD
St Rough leading edge(Ausoni(2006))

0.25

0.00
10 15 20
Uref [m/s]

Figure 12.Strouhal
Figure12. Strouhalnumber
numberatatdifferent
differentinflow
inflowvelocities.
velocities.
3.3. Effects of the Trailing-Edge Thickness
3.3. Effects of the Trailing-Edge Thickness
A sharp trailing edge is usually undesirable in marine propellers and hydrofoils. To avoid
A sharp trailing edge is usually undesirable in marine propellers and hydrofoils. To avoid
resonances in marine propellers, an appropriate trailing-edge thickness must be selected to shift the
resonances in marine propellers, an appropriate trailing-edge thickness must be selected to shift the
vortex-shedding frequency away from the natural frequencies of the propeller. Therefore, numerical
vortex-shedding frequency away from the natural frequencies of the propeller. Therefore, numerical
studies on the relationship between the vortex-shedding frequency and trailing-edge thickness are
studies on the relationship between the vortex-shedding frequency and trailing-edge thickness are
highly important. The NACA0009 hydrofoil was truncated at different positions to investigate the
highly important. The NACA0009 hydrofoil was truncated at different positions to investigate the
effects of the trailing-edge thickness (b) on vortex shedding. The inflow velocity Ure f was set to 12 m/s,
effects of the trailing-edge thickness (b) on vortex shedding. The inflow velocity 𝑈𝑟𝑒𝑓 was set to 12
and the thickness of the boundary-layer mesh was set according to the chord length of the hydrofoil.
m/s, and the thickness of the boundary-layer mesh was set according to the chord
Each set of vortex shedding data in Table 5 corresponds to a hydrofoil shape with a different length of the
truncation.
hydrofoil. Each set of vortex shedding data in Table 5 corresponds to a hydrofoil shape with a
different truncation.
Table 5. Vortex-shedding frequency corresponding to each trailing-edge thickness.

C (mm)
Table 5. Vortex-shedding B (mm)corresponding
frequency F (Hz)to each trailing-edge
St thickness.
29.66 9.00 239.92 0.180
65.45 6.14 464.25 0.237
79.39 4.04 666.22 0.224
88.00 2.69 845.18 0.190
93.30 1.54 1216.55 0.157

Figure 13 illustrates the lift curves that were obtained with different hydrofoil truncations. It can
be inferred from the figure that the lift curves always undergo a period of change before stabilization.
The lift curve has a rather irregular shape when b = 9 mm, as indicated by the different maximum and
minimum values in each cycle, even in the steady state. The changes in lift curve when b = 1.54 mm are
different from those observed with the other trailing-edge thicknesses; in addition to the initial growth
in oscillation amplitude, the oscillations of the b = 1.54 mm lift curve also decrease in amplitude before
they eventually stabilize. The trailing-edge thickness gradually decreases as the truncation approaches
the rear end of the hydrofoil. As this occurs, the lift-curve oscillations gradually decrease in amplitude
but increase in frequency.
mm are different from those observed with the other trailing-edge thicknesses; in addition to the
initial growth in oscillation amplitude, the oscillations of the b = 1.54 mm lift curve also decrease in
amplitude before they eventually stabilize. The trailing-edge thickness gradually decreases as the
truncation approaches the rear end of the hydrofoil. As this occurs, the lift-curve oscillations
J. Mar. Sci. Eng.decrease
gradually 2020, 8, 195in amplitude but increase in frequency. 13 of 20

1.5 0.2

1.0
0.1
0.5
Cl

0.0 0.0

Cl
-0.5
-0.1
-1.0

-1.5 -0.2
0 5 10 15 0 1 2 3 4 5 6 7
t t

(a) c = 29.66 mm, b = 9.00 mm (b) c = 65.45 mm, b = 6.14 mm

0.15 0.08

0.10
0.04

0.05

Cl
0.00
Cl

0.00

-0.05
-0.04

-0.10
-0.08
-0.15 0 1 2 3 4
0 1 2 3 4 t
t

(c) c = 79.39 mm, b = 3.04 mm (d) c = 88.00 mm, b = 2.69 mm

0.03

0.02

0.01
Cl

0.00

-0.01

-0.02

-0.03
0 1 2 3 4
t

(e) c = 93.30 mm, b = 1.54 mm

Figure 13. Variation of the lift coefficient Cl with respect to the trailing-edge thickness, b.

Figures 14–16 indicate that the oscillation frequency of the lift curve (f ) decreases with increase in
b. Figure 16 shows that St first increases and then decreases with increase in b; the value of St peaks at
b = 6.138 mm. Figure 14 indicates that the steady-state lift curve is rather irregular when b is large, as
indicated by the multiple frequency modes in the lift curve (the higher-order modes are significantly
stronger when b is large). Figure 17 illustrates the changes in vortex shedding that occurred as the
truncation shifted rearward (thus increasing chord length and decreasing b).
Figures
in 14–16
b. Figure 16indicate
shows thatthat 𝑆thefirst
oscillation frequency
increases and then of decreases
the lift curve
with(f)increase
decreases in with
b; theincrease
value of 𝑆𝑡
𝑡
in b. peaks
Figureat16 shows that 𝑆
b = 6.138 mm. Figure 14 indicates that the steady-state lift curve is rather irregular of
𝑡 first increases and then decreases with increase in b; the value 𝑆𝑡 b is
when
peaks at b = 6.138 mm. Figure 14 indicates that the steady-state lift curve is rather
large, as indicated by the multiple frequency modes in the lift curve (the higher-order modes irregular when b is are
large,significantly
as indicatedstronger
by the multiple frequency modes in the lift curve (the higher-order
when b is large). Figure 17 illustrates the changes in vortex shedding that modes are
significantly stronger when b is large). Figure 17 illustrates thechord
changes in vortex sheddingb).that
J.occurred as the
Mar. Sci. Eng. truncation
2020, 8, 195 shifted rearward (thus increasing length and decreasing 14 of 20
occurred as the truncation shifted rearward (thus increasing chord length and decreasing b).

b/c Ratio
b/c
1.5 Ratio 0.303
1.5 0.3030.094
0.0940.051
1.
0
0.0510.031
1. 0.0310.017
0

Amplitude
0.017

Amplitude
0.
5
0.
5

0.
0
00 0.
22 0
00 5
00 20 00
22 00 18 00 54
0
2 0 0 16 00
18 00 14 00
6 43
1 00 12 00
14 0 0f [H 1 0 0 0
12 0 8 3
f [ 1 0 0 z]0 60
0 2
Hz
] 80 0 40
0
60 00 2
0 2 1
40 0 0
20 10
0
- 20
00
-2
Figure 14. Vortex shedding characteristics of each model in the frequency domain.
Figure 14. Vortex
Figure shedding
14. Vortex characteristics
shedding of each
characteristics model
of each in theinfrequency
model domain.
the frequency domain.

1500
1500

1000
1000
[Hz]
[Hz]

500
f

500

0
0 0 3 6 9
0 3
b6[mm] 9
J. Mar. Sci. Eng. 2020, 8, x FOR PEER REVIEW 15 of 20
b [mm]
Figure 15.
Figure 15. Change
Change in
in ff with various values of b.
Figure 15. Change in f with various values of b.

0.24

0.21
St

0.18

0.15
0 3 6 9
b[mm]

Figure 16.
Figure 16. Change in 𝑆St with
Changein various values of b.
𝑡 with various values of b.

(a) c = 29.66 mm, b = 9.00 mm (b) c = 65.45 mm, b = 6.14 mm


0.18

0.15
0 3 6 9
b[mm]
J. Mar. Sci. Eng. 2020, 8, 195 15 of 20
Figure 16. Change in 𝑆𝑡 with various values of b.

(a) c = 29.66 mm, b = 9.00 mm (b) c = 65.45 mm, b = 6.14 mm

(c) c = 79.39 mm, b = 3.04 mm (d) c = 93.30 mm, b = 1.54 mm


Figure
Figure 17.
17. Vortex
Vortex shedding
shedding with
with different
different trailing-edge thicknesses.

3.4. Effects
3.4. Effects of
of Angle
Angle of
of Attack
Attack
In most
In most scenarios,
scenarios, the
the hydrofoil
hydrofoilhashasaacertain
certainangle attack,α,α,that
angleofofattack, thatinfluences
influences the drag
the dragandand liftlift
of
thethe
of hydrofoil. In this
hydrofoil. section,
In this thethe
section, value of αofofαthe
value of hydrofoil
the hydrofoilwas was
varied to examine
varied its effects
to examine on wake
its effects on
vortex shedding. The parameters of the hydrofoil model were c = 88 mm, b = 2.69
wake vortex shedding. The parameters of the hydrofoil model were c = 88 mm, b = 2.69 mm, and 𝑈𝑟𝑒𝑓 mm, and U re f = 12
m/s. The mesh parameters are the same as those in the previous simulations.
= 12 m/s. The mesh parameters are the same as those in the previous simulations.
As shown in Table 6, the shedding of wake vortices directly depends on α. At small values of α
(Figure 18), the vortex-shedding behavior
Table 6. Effectsisof
similar to that
α on wake at αshedding.
vortex = 0◦ . As α increases to a certain value
(Figure 19), the boundary layer at the frontal edge of the hydrofoil begins to dissociate from the surface,
Α (°) 𝑪𝒍 F (Hz) 𝑺𝒕
causing the boundary layer near the trailing edge to become thinner. In this scenario, the vortices
shed from the suction 0 surface are much 0weaker than those 845.18
shed from the pressure 0.190 surface. Therefore,
1 wake is dominated
the vortex street in the 0.086 by vortices from 865.05
the pressure surface.0.194 The vortices from
2
the suction surface move 0.230 from the pressure
along the vortices 856.60surface and rapidly 0.192 dissipate. When α
3 0.343 855.43
increases further (Figure 20), the boundary layer dissociates from the surface 0.192
of the hydrofoil at its
6 0.618 812.98
frontal edge, and the Kármán vortex street is no longer observed in the wake field. 0.183
9 1.042 757.99 0.170
12 1.304
Table 6. Effects of α on wake vortex shedding.

A ( shedding
As shown in Table 6, the ) ofCwake
l F (Hz)
vortices directly dependsSt on α. At small values of α
(Figure 18), the vortex-shedding
0 behavior 0is similar to 845.18
that at α = 0°. As α increases to a certain value
0.190
1 0.086 865.05 0.194
(Figure 19), the boundary layer at the frontal edge of the hydrofoil begins to dissociate from the
2 0.230 856.60 0.192
3 0.343 855.43 0.192
6 0.618 812.98 0.183
9 1.042 757.99 0.170
12 1.304

By comparing Figure 21a,b and Figure 22, it is found that there is always a stagnation point on the
pressure side at which Cp = 1.0 and that the stagnation point moves away from the leading edge as
the angle of attack increases. As shown in Figure 21a, at α = 3, the lift on the hydrofoil is relatively
small. At the trailing edge, the pressure on the pressure side is lower than that on the suction side.
As shown in Figure 21b, when the angle of attack increases to 6◦ , the flow becomes stronger, and the
J. Mar. Sci. Eng. 2020, 8, x FOR PEER REVIEW 16 of 20
J. Mar. Sci. Eng. 2020, 8, x FOR PEER REVIEW 16 of 20
J. Mar. Sci. Eng. 2020, 8, x FOR PEER REVIEW 16 of 20
surface, causing the boundary layer near the trailing edge to become thinner. In this scenario, the
surface,
J. Mar. Sci.causing
vortices Eng. 2020,
shed fromthe boundary
8, the
195 layer near
suction surface are the
muchtrailing
weakeredge to those
than become thinner.
shed from the In this scenario,
pressure the
16 of
surface. 20
surface,
vortices causing
shed the boundary layer near the trailing edge to become thinner. In this scenario, the
Therefore,
vortices shed thefrom
vortex
from
thestreet
suction
thestreet in surface
suction the wake
surface
areismuch
areismuch
weakerby
dominated
weakerby
than
than
those shed
vortices from
those from
from the pressure
the pressure
shed the
from
surface.
surface.
the pressure The
surface.
Therefore,
vortices the
from vortex
the suction in the
surface wake
move alongdominated
the vortices vortices
from the pressure pressure
surface surface.
and The
rapidly
pressure
Therefore, difference
thethevortex between
street the pressure
in the and
wakealong suction sides
is dominated by increases,
vortices which theeffectively increases the
vortices
dissipate. from
When α suction
increases surface
further move
(Figure 20),the
thevortices
boundaryfrom thefrom
layer pressure
dissociates
pressure
surface
from the
surface.
and The
rapidly
surface of
lift on thefrom
vortices hydrofoil.
the suction surface move along the vortices from the pressure surface and rapidly
dissipate.
the hydrofoilWhen α increases
at its frontal further
edge, (Figure
and the Kármán 20), the boundary
vortex layer
street is no dissociates
longer observed from the
in the surface
wake of
field.
dissipate.
the hydrofoil When
at itsαfrontal
increases further
edge, (Figure
and the Kármán 20),vortex
the boundary
street is nolayer dissociates
longer observed from thewake
in the surface of
field.
the hydrofoil at its frontal edge, and the Kármán vortex street is no longer observed in the wake field.

Figure 18. Vortex shedding when α = 1°.◦


Figure 18.
Figure Vortex shedding
18. Vortex when αα ==1°.
shedding when 1 .
Figure 18. Vortex shedding when α = 1°.

Figure 19. Vortex shedding when α = 9°.


Figure 19. Vortex shedding when α = 9°.◦
Figure
Figure 19. Vortex shedding
19. Vortex when αα =
shedding when 9 .
= 9°.

J. Mar. Sci. Eng. 2020, 8, x FOR PEER REVIEW


Figure 20. Vortex shedding when α = 12°. 17 of 20
Figure 20. Vortex shedding when α = 12°.
Figure 20. Vortex shedding when α = 12°.
Figure
By comparing Figures 21a,b 22,Vortex
and 20. shedding
it is found thatwhen α=
there is12
◦.
always a stagnation point on the
By comparing
pressure side at which Figures
Cp = 21a,b
1.0 andand 22,the
that it is found that there is always
awayaafrom stagnation pointedge
on the
pressureBy comparing
1.5
side at whichFigures
C 21a,b and 22, it stagnation
is found that
p = 1.0 and that the stagnation point
point moves
there
2 moves is always
away from
the leading
stagnation
the pointedge
leading
as
on the
as
the angle
pressure of
side attack increases.
at which As
Cp = 1.0 shown
and in
thatin
theFigure 21a,
stagnation at α = 3, the lift on the hydrofoil is relatively
the
small.angle of attack
At1.0the increases.
trailing edge, the Aspressure
shown on Figure 21a, atpoint
the pressure α = 3,
side
movesliftaway
is the
lower on the
than
from the leading
hydrofoil
that
edge as
is relatively
on the suction side.
the
small. angle
At theof attack increases. As shown in Figure 21a, at α = 3, the lift on the hydrofoil
trailing edge, the pressure on the pressure side is lower than that on the suction side. is relatively
As shown
small. 0.5in Figure 21b, when the angle of attack increases to 6°, the flow becomes stronger, and the
At the trailing edge, thethe
pressure onattack
the pressure side is lower than that onstronger,
the suction
As shown
pressure in Figure
difference 21b, when
between the angle of
pressure and increases
suction sides
0
to 6°, the flow
increases, which becomes
effectively andside.
increases the
the
As shown 0.0 in Figure 21b, when the angle of attack increases to 6°, the flow becomes stronger, and the
Cp

pressure difference between the pressure and suction sides increases, which effectively increases the
Cp

lift on
pressurethe hydrofoil.
difference
lift hydrofoil.between the pressure and suction sides increases, which effectively increases the
on the-0.5
lift on the hydrofoil. -2
-1.0

-1.5
0.00 0.25 0.50 0.75 1.00 -4
x/c 0.00 0.25 0.50 0.75 1.00
x/c

(a) α = 3° (b) α = 6°
Figure21.
Figure 21.Pressure
Pressuredistribution
distributionon
onthe
thesurfaces
surfacesof
ofthe
thehydrofoil.
hydrofoil.

Figure 22 shows the pressure distribution on the surfaces of the hydrofoil with a stable lift
coefficient. As shown in Figure 23, where α is large, the lift curve of the hydrofoil stabilizes after a
short period of growth. It can be inferred from the figure that the pressure on the suction and pressure
surfaces of the trailing edge gradually converge to the same value. Then, because these pressures are
Figure 22 shows the pressure distribution on the surfaces of the hydrofoil x/c with a stable lift
coefficient. As shown (a)in
α =Figure
3° 23, where α is large, the lift curve of (b) the αhydrofoil
= 6° stabilizes after a
short period of growth. It can be inferred from the figure that the pressure on the suction and pressure
surfaces of the trailing Figure
edge21. Pressure converge
gradually distribution
toon
thethe surfaces
same of the
value. hydrofoil.
Then, because these pressures are
J. the
Mar.same,
Sci. Eng. 2020, 8, 195
the lift can no longer change. 17 of 20
Figure 22 shows the pressure distribution on the surfaces of the hydrofoil with a stable lift
coefficient. As shown in Figure 23, where α is large, the lift curve of the hydrofoil stabilizes after a
short period of growth. It can be inferred2.5 from the figure that the pressure on the suction and pressure
surfaces of the trailing edge gradually 0.0 converge to the same value. Then, because these pressures are
the same, the lift can no longer change.
-2.5

Cp
-5.0
2.5
-7.5
0.0
-10.0
-2.5
-12.5

Cp
-5.00.00 0.25 0.50 0.75 1.00
x/c
-7.5
Figure 22.Surface
Figure22. Surface pressure distribution of the hydrofoil when α = 12◦ .
-10.0 pressure distribution of the hydrofoil when α = 12°.

Figure 22 shows the pressure -12.5 distribution on the surfaces of the hydrofoil with a stable lift

2.00where α is large, the lift curve of the hydrofoil stabilizes after a


coefficient. As shown in Figure 23, 0.00 0.25 0.50 0.75 1.00
x/c
short period of growth. It can be inferred from the figure that the pressure on the suction and pressure
1.75
surfaces of the trailing edge
Figure 22.gradually converge
Surface pressure to the same
distribution of thevalue. Then,
hydrofoil because
when these pressures are
α = 12°.
the same, the lift can no longer change.
1.50
Cl

1.25
2.00
1.00
1.75
0.75
1.50
0.50
Cl

1.25 0 5 10 15
t
1.00

Figure 23. Lift curve of the hydrofoil when α = 12°.


0.75

3.5. Effects of Hydrofoil Thickness 0.50


0 5 10 15
t
Based on the NACA00 series of hydrofoil models, we designed a set of hydrofoil models with a
Figure
trailing-edge thickness and chord Lift curve
23.length of bof=the hydrofoil
2.69 mm and c = α88=mm,
when 12◦ . respectively. These models
Figure 23. Lift curve of the hydrofoil when α = 12°.
3.5. Effects of Hydrofoil Thickness
3.5. Effects of Hydrofoil Thickness
Based on the NACA00 series of hydrofoil models, we designed a set of hydrofoil models with a
Based on
trailing-edge the NACA00
thickness serieslength
and chord of hydrofoil
of b = models,
2.69 mmwe designed
and a setrespectively.
c = 88 mm, of hydrofoilThese
models with a
models
trailing-edge thickness and chord length of b = 2.69 mm and c = 88 mm, respectively. These
were placed in a Ure f = 12 m/s flow field to calculate their flow and drag coefficients. The mesh models
parameters used in this simulation are the same as those used in the previous simulations. Vortex
shedding data for different hydrofoils are shown in Table 7.

Table 7. Results obtained after changing the thickness of the hydrofoil.

Hydrofoil Type f (Hz) Cl (max) St


0009 845.18 0.0561 0.190
0012 794.18 0.0497 0.178
0014 743.03 0.0353 0.167
0016 697.51 0.0287 0.157
0018 654.81 0.0215 0.147

The test yielded a time-averaged Cl value of the symmetrical hydrofoil of 0 at α = 0◦ . Therefore,


Cd is used to compare the experimental and simulated values. Owing to the lack of test results on
hydrofoils with truncated trailing edges, we only compared the results of the original hydrofoil test
with the simulation results, as shown in Table 8. As shown in the table, the S-A model predicted the
drag fairly precisely.
0009 845.18 0.0561 0.190
0009
0012 845.18
794.18 0.0561
0.0497 0.190
0.178
0012
0014 794.18
743.03 0.0497
0.0353 0.178
0.167
0014
0016 743.03
697.51 0.0353
0.0287 0.167
0.157
0016
0018 697.51
654.81 0.0287
0.0215 0.157
0.147
J. Mar. Sci. Eng. 2020, 8, 195 18 of 20
0018 654.81 0.0215 0.147
Table 8. Results obtained after changing the thickness of the hydrofoil (Re = 3.4 × 10−6).
Table8.8. Results
Table Results obtained
obtained after
after changing
changingthe
thethickness
thicknessof
ofthe
thehydrofoil
hydrofoil(Re
(Re== 3.4
3.4 ×
× 10
10−6).).
−6

Hydrofoil
Hydrofoil 𝑪𝒅 (exp.) [19] 𝑪𝒅 (pre.)
TypeType
Hydrofoil 𝑪𝒅 C(exp.)
d (exp.) [19]
[19] Cd (pre.)
𝑪𝒅 (pre.)
Type
NACA0009 0.00654 0.00649
NACA0009 0.00654 0.00649
NACA0009
NACA0012
NACA0012 0.00654
0.00657
0.00657 0.00649
0.00652
0.00652
NACA0012
NACA0018
NACA0018 0.00657
0.00725
0.00725 0.00652
0.00720
0.00720
NACA0018 0.00725 0.00720
By
By observing
observing thethe vortex-shedding
vortex-shedding states
states in
in Figures
Figures 24
24 and
and 25,
25, itit is
is clear
clear that
that vorticity
vorticity ofof the
the shed
shed
By observing
vortices induced thethicker
by vortex-shedding
hydrofoils states
are in Figures 24lower.
significantly and 25,Table
it is clear6 that vorticity
indicates that of the
the shed
vortex-
vortices induced by thicker hydrofoils are significantly lower. Table 6 indicates that the vortex-shedding
vortices induced
shedding by thicker hydrofoils are in
significantly lower. Table 6 indicates that theofvortex-
frequencyfrequency
decreasesdecreases withinincrease
with increase hydrofoil hydrofoil
thickness.thickness.
Furthermore,Furthermore,
the effectsthe effects
of vortex vortex
shedding
shedding
shedding frequency
on the lift decreases
also with
decrease increase
with in
increasehydrofoil
in thickness.
hydrofoil Furthermore,
thickness. From the effects
Figure 26, of can
it vortex
be
on the lift also decrease with increase in hydrofoil thickness. From Figure 26, it can be observed that
shedding
observed on
that the lift
increasesalso
in decrease
the with
hydrofoil increase
thickness in hydrofoil
lead to thickness.
larger From
position-dependent Figure 26,
changes it can
in be
the
increases in the hydrofoil thickness lead to larger position-dependent changes in the surface pressure;
observed
surface that increases
pressure; in thethat
the position hydrofoil
producesthickness leaddifferential
a pressure to larger position-dependent
also changes in the
the position that produces a pressure differential also becomes closer tobecomes closer
the trailing to the trailing
edge.
surface
edge. pressure; the position that produces a pressure differential also becomes closer to the trailing
edge.

Figure 24. Vortex shedding of the 0012 model.


Figure 24.
Figure Vortex shedding
24. Vortex shedding of
of the
the 0012
0012 model.
model.

J. Mar. Sci. Eng. 2020, 8, x FOR PEER REVIEW 19 of 20


Figure 25. Vortex shedding of the 0018 model.
Figure 25. Vortex shedding of the 0018 model.
Figure 25. Vortex shedding of the 0018 model.

1.0 1.0

0.5 0.5
Cp

Cp

0.0
0.0

-0.5
-0.5

-1.0
0.00 0.25 0.50 0.75 1.00
0.00 0.25 0.50 0.75 1.00
x/c x/c

(a) (b)
Figure26.26.Surface
Figure Surface pressure
pressure distribution
distribution on different
on different hydrofoilhydrofoil models.
models. (a) Surface (a) Surface
pressure pressure
distribution
ofdistribution
Model 0012of(C Model 0012 (𝐶𝑙 =Surface
l = 0.0282). (b)
0.0282).pressure
(b) Surface pressure distribution
distribution of Model 0018of Model 0018 (𝐶𝑙 = −0.0232).
(Cl = −0.0232).

4. Conclusions
We simulated vortex shedding in the wake flow of an NACA0009 hydrofoil using the S-A
turbulence model. The main purpose of the simulation was to investigate the factors that affect the
vortex-shedding frequency, including the inflow velocity, angle of attack, and trailing-edge
thickness. Firstly, it was found that the increase in vortex-shedding frequency was almost linear with
the inflow velocity. However, the nondimensional Strouhal number remained almost constant at
0.19. Secondly, the oscillation of vortex shedding could be reduced by increasing the angle of attack.
J. Mar. Sci. Eng. 2020, 8, 195 19 of 20

4. Conclusions
We simulated vortex shedding in the wake flow of an NACA0009 hydrofoil using the S-A
turbulence model. The main purpose of the simulation was to investigate the factors that affect the
vortex-shedding frequency, including the inflow velocity, angle of attack, and trailing-edge thickness.
Firstly, it was found that the increase in vortex-shedding frequency was almost linear with the inflow
velocity. However, the nondimensional Strouhal number remained almost constant at 0.19. Secondly,
the oscillation of vortex shedding could be reduced by increasing the angle of attack. When the angle
of attack was sufficiently large, the oscillation was negligible. Thirdly, increases in the trailing-edge
breadth led to increases in the vortex-shedding frequency. Importantly, there was a specific breadth
corresponding to the maximum Strouhal number. Finally, it was found that a larger maximum thickness
of the hydrofoil led to lower frequency and more condense vorticity of vortex shedding.

Author Contributions: Conceptualization, J.H. and C.G.; Formal analysis, S.S. and C.S.; Investigation, Z.W. and
W.Z.; Methodology, J.H. and C.G.; Software, Z.W. and W.Z.; Writing—original draft, Z.W.; Writing—review &
editing, J.H. and Z.W. All authors have read and agreed to the published version of the manuscript.
Funding: This research received no external funding.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Milne-Thomson, L.M. Theoretical Hydrodynamics; Macmillan: London, UK, 1972; p. 375.
2. Perry, A.E.; Chong, M.S.; Lim, T.T. The vortex-shedding process behind two-dimensional bluff-bodies. J.
Fluid Mech. 1982, 116, 77–90. [CrossRef]
3. Griffin, O.M. A note on bluff body vortex formation. J. Fluid Mech. 1995, 284, 217–224. [CrossRef]
4. Williamson, C.H.K.; Roshko, A. Vortex formation in the wake of an oscillating cylinder. J. Fluids Struct. 1988,
2, 355–381. [CrossRef]
5. Gerrard, J.H. Wakes of cylindrical bluff bodies at low Reynolds-number. Philos. Trans. R. Soc. 1978, 288, 351.
6. Gerich, D.; Eckelmann, H. Influence of end plates and free ends on the shedding frequency of circular
cylinders. J. Fluid Mech. 1982, 122, 109–122. [CrossRef]
7. Achenbach, E.; Heinecke, E. On vortex shedding from smooth and rough cylinders in the range of Reynolds
numbers 6 × 103 to 5 × 106 . J. Fluid Mech. 1981, 109, 239–251. [CrossRef]
8. Ausoni, P.; Farhat, M.; Ait Bouziad, Y.; Kueny, J.L.; Avellan, F. Kármán vortex shedding in the wake of a 2D
hydrofoil: Measurement and numerical simulation. In Proceedings of the IAHR International Meeting of
WG on Cavitation and Dynamic Problems in Hydraulic Machinery and Systems, Barcelona, Spain, 28–30
June 2006.
9. Lotfy, A.; Rockwell, D. The near-wake of an oscillating trailing edge: Mechanisms of periodic and aperiodic
response. J. Fluid Mech. 1993, 251, 173–201. [CrossRef]
10. Prasad, A.; Williamson, C.H. The instability of the shear layer separating from a bluff body. J. Fluid Mech.
1997, 333, 375–402. [CrossRef]
11. Dwayne, A. Vortex shedding from a hydrofoil at high Reynolds number. J. Fluid Mech. 2005, 531, 293–324.
12. Zobeiri. A. How oblique trailing edge of a hydrofoil reduces the vortex-induced vibration. J. Fluids Struct.
2012, 32, 78–89. [CrossRef]
13. Williamson, C.H.K. Vortex dynamics in the cylinder wake. Annu. Rev. Fluid Mech. 1996, 28, 477–539.
[CrossRef]
14. William, K.B. A near-wake model for the aerodynamic pressures exerted on singing trailing edges. J. Acoust.
Soc. Am. 1976, 60, 594–599.
15. Huerre, P.; Monkewitz, P.A. Local and global instabilities in spatially developing flows. Annu. Rev. Fluid
Mech. 1990, 22, 473–537. [CrossRef]
16. Oertel, H., Jr. Wakes behind blunt bodies. Annu. Rev. Fluid Mech. 1990, 22, 539–562. [CrossRef]
17. Lee, S.J.; Lee, J.H.; Suh, J.C. Numerical investigation on vortex shedding from a hydrofoil with a beveled
trailing edge. Model. Simul. Eng. 2015, 2015, 565417. [CrossRef]
J. Mar. Sci. Eng. 2020, 8, 195 20 of 20

18. Spalart, P.R.; Allmaras, S.R. A one-equation turbulence model for aerodynamic flows. AIAA 1992, 92, 0439.
19. Harry, J.; Goett, W.; Kenneth, B. Tests of NACA 0009, 0012, and 0018 airfoils in the full-scale tunnel. NACA
Tech. Rep. 1938, 647.

© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

You might also like