Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

water

Article
Experimental and Computational Fluid Dynamic Study of
Water Flow and Submerged Depth Effects on a Tidal
Turbine Performance
Erfan Ghamati 1 , Hamed Kariman 2 and Siamak Hoseinzadeh 3, *

1 Department of Mechanical and Energy Engineering, Shahid Beheshti University, Tehran 19839 69411, Iran
2 School of Engineering, Edith Cowan University, 270 Joondalup Drive, Joondalup, WA 6027, Australia
3 Department of Planning, Design, and Technology of Architecture, Sapienza University of Rome,
00185 Rome, Italy
* Correspondence: siamak.hosseinzadeh@uniroma1.it

Abstract: This study involves an experimental and numerical analysis of the Hunter turbine, a
vertical axis turbine utilized for tidal energy. A laboratory model of the Hunter turbine, featuring
an aspect ratio of 1.2, was designed and tested. Numerical equations, including the Reynolds-
averaged Navier–Stokes (RANS) constant, were analyzed through computational fluid dynamics
(CFD) software using the k-ω turbulence model to forecast turbine performance and other related flow
specifications, such as pressure lines, stream velocity, and pressure. This simulation was conducted
on the surface of the turbine blade, and the results were obtained accordingly. The experimental data
were utilized to verify the numerical results, and the difference between the two was reasonably
acceptable. The turbine was studied in six different flow coefficients and four different vertical
positions. The results indicated that the power coefficient increased as the submerged depth from a
water-free surface increased, and after a specific depth, the output power remained constant. It was
also observed that the minimum depth from a water-free surface for maximum power coefficient was
three times the diameter of the turbine drum (3D).

Keywords: computational fluid dynamic; tidal turbine; Hunter turbine; submerged depth effect;
vertical axis turbine
Citation: Ghamati, E.; Kariman, H.;
Hoseinzadeh, S. Experimental and
Computational Fluid Dynamic Study
of Water Flow and Submerged Depth 1. Introduction
Effects on a Tidal Turbine Climate change and energy security are two factors that have contributed to a global
Performance. Water 2023, 15, 2312. energy crisis. To address these issues, reducing fossil energy consumption and greenhouse
https://doi.org/10.3390/w15132312 gas emissions is one of the most critical factors. Rising energy efficiency and utilizing
Academic Editor: Roberto Gaudio renewable energy (RE) sources, such as wind, solar, waves, and geothermal, can help de-
crease greenhouse gases [1,2]. One of the key ways to encourage industries to use REs is by
Received: 12 May 2023 determining environmental limits, such as greenhouse gases and global warming problems.
Revised: 12 June 2023 Over the years, researchers have studied ocean waves, which are one of the most important
Accepted: 16 June 2023
sources of RE [3–5]. Compared to conventional energy sources, renewable energies are
Published: 21 June 2023
typically not economical. As a result, researchers and engineers are continuously designing
low-cost machines with higher efficiencies and lower costs [5,6]. Hydro-kinetic conversion
systems can be suitable for harnessing energy from tidal currents or artificially constructed
Copyright: © 2023 by the authors.
water channels. Several studies have demonstrated that on-land water resources can be
Licensee MDPI, Basel, Switzerland. utilized for energy harvesting [6,7]. One of the advantages of tidal energy is that it can
This article is an open access article be produced by placing turbines in the path of waves [8,9]. Although typical maximum
distributed under the terms and tidal currents are around 3 m/s, wind speeds may reach 12 m/s. Comparing dynamic
conditions of the Creative Commons pressures, they are still higher in water, which means that the dimensions of tidal turbines
Attribution (CC BY) license (https:// can be smaller than those of wind turbines for the same power output. The development
creativecommons.org/licenses/by/ of turbines, support structures, and cabling to land are the most critical factors for the
4.0/). development of tidal energy [10–13].

Water 2023, 15, 2312. https://doi.org/10.3390/w15132312 https://www.mdpi.com/journal/water


Water 2023, 15, 2312 2 of 18

Tidal turbines can be classified into two types: horizontal axis tidal turbines (HATT)
and vertical axis tidal turbines (VATT). So far, almost all tidal energy projects have primarily
used HATT due to its advancements in wind energy technology. However, VATTs have
various benefits for achieving tidal energy. Firstly, they operate independently of the
flow direction. Secondly, their rotation speed is lower compared to HATTs, making them
more efficient at low tidal speeds when the speed is less. Thirdly, they can be used
better in shallow tidal currents because they can maximize the use of the available cross-
sectional area [14,15]. However, VATTs also have some weaknesses, such as lower efficiency
compared to HATTs and more difficulty in starting the work [15].
A new type of HATT is the Hunter turbine, which was invented by John Hunter to
extract energy from a river or ocean stream [14]. The turbine consists of various flapping
blades attached to a rotating drum. The turbine operates by opening the flapping blades
when the impact flow hits the side of the drum, and the blades on the other side are allowed
to flow by the current as long as they are controlled by the stops in the fully open situation.
Yang and Yawn conducted a study to investigate the performance of the Hunter
turbine in a standalone type compared with fixed-blade turbines. They concluded that
the optimal value of the flow coefficient was obtained in the range of 0.44 to 0.47, and the
value of the optimal power coefficient in this range was equal to 0.19 by using a numerical
simulation [16].
Chen et al. analyzed numerical simulations to analyze the impacts of a booster channel
on the performance of VATT. They simulated the flow field with FLUENT software, used
the finite volume method to solve the Reynolds-averaged Navier–Stokes equations and the
SST k-ω turbulence model, and simulated the rotation of the rotor by solving the rotational
motion equation. The results showed that for a single-channel turbine, the fluctuations of
hydrodynamic torques and rotational speeds are significantly decreased, and the output
power is 30% more than an independent turbine [17].
Sun et al. simulated the aerodynamic performance of a vertical tidal turbine with
a suitable bionic super-hydrophobic surface on airfoils. They investigated the effect of
different inlet velocities, the number of blades, and vertical tidal axis turbine solids with
suitable bionic hydrophobic surfaces. The results showed that high energy efficiency could
be enhanced by about 16.5% at the input speed U = 3 m/s. The increase in power factor is
largely due to the blade number, input velocity, and strength at small tip ratios. However,
the increase in energy efficiency could be independent of these aerodynamic factors at high
tip ratios [18].
Jing et al. conducted various experiments on VATT and provided a wealth of exper-
imental information to analyze the hydrodynamic function of turbines. They discussed
the effect of different parameters on the hydrodynamic function of turbines and proposed
control strategies for the blades of various turbines [19].
Derakhshan et al. conducted a numerical study on a horizontal axis tidal turbine
(HATT) to investigate the impact of neighbor turbine distance, duct, and turbine layout on
the power factor. The study found that the power factor increased for ducts with larger
area ratios [20].
Maduka et al. performed flume tests on flanged duct turbines in tidal currents and
found that these turbines produced 40% more power than bare turbines. However, accurate
economic estimates for these turbines are not currently available [21].
De Arcos et al. used computational fluid dynamics (CFD) software to simulate the
hydrodynamic effects of isolated flapping and torsional deformations on tidal turbine
blades. The study showed that a pressure drop on the suction side of the blade increases
the internal load [22].
Moreau et al. analyzed the environmental conditions and the effect of waves on the
operation of vertical axis tidal turbines (VATTs) in France. The study analyzed the number
of loads, frequency, and speed of waves on this type of turbine and found that VATTs are
suitable for electricity generation in these areas [23].
Water 2023, 15, 2312 3 of 18

Sun et al. used CFD software to investigate the effects of the number of blades on the
performance of VATTs. The study showed that increasing the number of blades leads to a
more homogeneous flow field and can improve efficiency at the start of operation [24].
Ma et al. investigated the hydrodynamic performance of a moving-state VATT and
found that the twin-rotor in these systems can increase the power output efficiency. The
study also found that the frequency and amplitude of the waves have less effect on the
average output power of the turbine [25].
Satrio and Utama tested a VATT in Indonesia and presented a modified example of
a VATT that has inclined vanes. The study found that this modified system has a better
self-starting capability, making it work better in low tidal flow mode [26]. Chen et al.
investigated and simulated the spatial evolution mechanism of a tidal current turbine
experimentally. They concluded after this simulation that the kinetic energy in the wake
region improves significantly in the range x = (2D, 3D) [27].
Xie et al. designed an innovative water turbine whose blades automatically open and
close under the force of ocean currents, absorbing energy from various directions. The
blades were shaped like a fish based on fluid dynamics principles, resulting in increased
efficiency in electricity generation [28].
Han et al. investigated the macroscopic characteristics of tidal turbines and optimized
their energy production farms using OpenTidalFarm software. The study concluded that
adding more turbines to the set enhances power generation by 50% [29].
Manolesos et al. explored the application of vortex generators (VGs) to improve the
performance of tidal turbines. The research results showed that VGs shorter than the
local boundary layer can enhance blade profile performance and successfully reduce flow
separation [30].
Wang et al. conducted an analysis and simulation of a new method to test the perfor-
mance of a 300 W counter-rotating horizontal axis tidal turbine (HATT) using computational
fluid dynamics (CFD). The findings were consistent between CFD predictions and labo-
ratory data, and the turbine performed optimally at a blade tip speed ratio of 6–10 [31].
Samadi et al. investigated a hybrid turbine consisting of a new turbine with semi-cylindrical
deflectors and a Savino’s turbine in the Qeshm channel between the Persian Gulf and the
Oman Sea to extract energy from a low-speed tidal current. Fluid dynamic analysis revealed
that combining the proposed system with Savion’s system increased power from 0.38 to
0.68 [32].
Khanjanpour and Javadi utilized the Taguchi technique to simulate vertical axis tidal
turbines (VAT). Their analysis showed that the torsion angle is the most crucial factor
affecting the hydrodynamic efficiency of the turbine, while the ratio of the chord to radius
had the least impact [33].
Yang and Lavin studied four different designs of Hunter turbines with different
ratios using numerical simulation. They concluded that 3D models produced significantly
different results from 2D models and that 3D effects cannot be ignored for accurate turbine
performance calculations. As the aspect ratio increases, the hydrodynamic parameters of the
turbine become more uniform along the entire span, and with higher flow coefficients, the
four designs exhibit similar performances and approach two-dimensional calculations [34].
Current turbines are usually installed over water streams and are exposed to various waves;
different investigations have studied the performance of a floating structure [35–37].
According to predictions, a significant number of tidal turbines will be installed at
a certain height above the water surface to allow for easy maintenance and repair of the
system. Most turbines currently available are primarily designed for shallow water currents.
However, when considering the installation of a turbine in a shallow river, placing it directly
in the riverbed can lead to potential damage from debris carried by the flowing water. Thus,
the optimal solution lies in installing the turbine at a specific elevation above the water
surface. The objective of this research study is to investigate the impact of the turbine’s
height above the water surface on its performance. By studying the relationship between
turbine height and performance, researchers can identify the optimal height that maximizes
maximizes the turbine’s efficiency. Furthermore, understanding the eff
height can help prevent potential damage caused by debris or other mate
the water. By determining the appropriate elevation, engineers can minim
Water 2023, 15, 2312
turbine damage and prolong its operational lifespan. In this study, a Hun
4 of 18
six movable blades hinged on a revolving drum was manufactured and te
atory. The researchers studied the effect of the turbine’s submerged dept
the turbine’s efficiency. Furthermore, understanding the effect of turbine height can help
mance. During
prevent potentialexperiments,
damage caused by the turbine
debris was
or other submerged
materials carried by at
thefour
water.different
By
1D, determining
2D, 3D, and 4D (where
the appropriate D represents
elevation, engineers canthe diameter
minimize the riskof the turbine).
of turbine damage Thes
sentand
theprolong its operational lifespan. In this study, a Hunter turbine with six movable
distance between the upper surface of the turbine and the water s
blades hinged on a revolving drum was manufactured and tested in a laboratory. The
researchers studied the effect of the turbine’s submerged depth on its performance. During
2. experiments, the turbine was submerged at four different depths, namely 1D, 2D, 3D, and
Methodology
4D (where D represents the diameter of the turbine). These depths represent the distance
To provide
between the upper a detailed
surface description
of the turbine of surface.
and the water the Hunter turbine, Figure 1
portant factors. R in Figure 1 denotes the blade drum’s diameter, and R
2. Methodology
distanceTobetween the center
provide a detailed of ofthe
description the blade chordFigure
Hunter turbine, and1the blade
highlights drum when
important
factors. R in Figure 1 denotes the blade drum’s diameter, and
opened or closed. Additionally, Rt is the distancec between the drum sha R represents the distance
between the center of the blade chord and the blade drum when the blades are opened or
open turbine
closed. blade.Rt The
Additionally, angle between
is the distance of rotation
the drumis shaft
denoted
and the by
fullyθ in turbine
open this system
angle between
blade. The anglethe flow isand
of rotation theby
denoted drum tosystem,
θ in this the center
and θ1 isof
thethe
angleblade.
between the
flow and the drum to the center of the blade.

Figure 1. The definition of parameters.

Figure 1.ToThe
demonstrate the turbine’s function, two dimensionless parameters are defined. The
definition of parameters.
flow coefficient is calculated using Equation (1):

ωRc
To demonstrate the turbine’s
C f =function, two dimensionless paramete
(1)
U
The flow coefficient is calculated using Equation (1):
In Equation (2), the power coefficient (Cp ) is defined as the ratio of the actual power
𝜔𝑅water flow ( 12 ρAU3 ). The
output (P) of the turbine to the power available in the incoming
parameters A, ρ, U, and T represent the cross-sectional𝐶 area
= of the turbine, water density,
the velocity of the incoming flow, and the produced torque on 𝑈 the turbine, respectively.

In Equation (2), the power


Cp =
coefficient
P
=
ω p) is defined as the ratio of th
T ×(C
(2)
1/2ρAU 3 1/2ρAU 3
output (P) of the turbine to the power available in the incoming water flow
parameters A, ρ, U, and T represent the cross-sectional area of the turbine
the velocity of the incoming flow, and the produced torque on the turbine
𝑃 𝑇×𝜔
Water 2023, 15, x FOR PEER REVIEW 5 of 19

Water 2023, 15, 2312 5 of 18

In the numerical simulation, by knowing U and 𝜔 and by solving the Reynolds-


averaged Navier–Stokes
In the equations by
numerical simulation, with Ansys CFX
knowing U and software
ω and by to find T, Cthe
solving p can be calcu-
Reynolds-
lated. However,
averaged in the experimental
Navier–Stokes test,Ansys
equations with the water in the testing
CFX software to findtunnel is motionless,
T, Cp can the
be calculated.
barge
However, in the experimental test, the water in the testing tunnel is motionless, the bargethe
that holds the turbine moves with a specific speed, and the produced torque on
turbine is driven
that holds from the
the turbine movesdigital
withsetup installed
a specific speed,onand thethe
turbine.
produced torque on the turbine
Figurefrom
is driven 2 illustrate thesetup
the digital design parameters
installed on theof the Hunter turbine and the blade posi-
turbine.
tions atFigure
various angles. the
2 illustrate It should
design be noted that
parameters angle
of the θ represents
Hunter turbine andthethe
angle
bladebetween
positionsthe
flow direction
at various and the
angles. end of be
It should thenoted
nearest
thatblade
angletoθ it.
represents the angle between the flow
direction and the end of the nearest blade to it.

Figure 2. The position of blades at different angles.


Figure 2. The position of blades at different angles.
3. Setup
3. Setup
The turbine assembly is composed of various components, such as a lengthy main
Thea turbine
shaft, assembly
cylindrical is six
body, and composed of various
blades. Figure components,
3 depicts such has
that the turbine as aalengthy
diametermain
of
shaft, a cylindrical
100 mm, body,
a height of and and
120 mm, six blades. Figure
an aspect 3 depicts
ratio (H/D) that
of 1.2. thetesting
The turbine has aused
channel diameter
for
of experiments had a length
100 mm, a height of 500and
of 120 mm, m and a depthratio
an aspect of 4.5 m, asof
(H/D) illustrated in Figure
1.2. The testing 3.
channel used
for experiments had a length of 500 m and a depth of 4.5 m, as illustrated in Figure 3.
Water 2023, 15, 2312 6 of 18
Water 2023, 15, x FOR PEER REVIEW 6 of 19

(a) (b)
Figure 3.
Figure 3. Turbine
Turbine laboratory
laboratory (a),
(a), Test
Testsetup
setup(b).
(b)

4. Numerical
4. Numerical Setup
Setup
TheANSYS
The ANSYSCFX CFXsoftware
software used
used in this
in this study
study is a completely
is a completely implicit
implicit flow based
flow solver solver
on the RANS
based on thesolver
RANSmethod. This solver
solver method. Thisallows
solverforallows
a simultaneous solution ofsolution
for a simultaneous the equations
of the
of motion and
equations continuity,
of motion and resulting
continuity, in aresulting
more accurate
in a more and accurate
efficient solution.
and efficient Additionally,
solution.
itAdditionally,
eliminates the it need for a pressure
eliminates the needcorrection term correction
for a pressure to maintainterm masstoconservation,
maintain mass which
con-
is beneficialwhich
servation, in simulating boundary
is beneficial layer flows
in simulating typically
boundary layerfound
flowsintypically
vertical axis
found turbines.
in verti-
Theturbines.
cal axis k-ω turbulence model was chosen for this study to analyze the Reynolds stresses
in theTheRANS equations because
k-ω turbulence model was thechosen
discretization method
for this study for solving
to analyze the the problem
Reynolds was
stresses
first order
in the RANSupwind. Yang because
equations proved that the k-ω modelmethod
the discretization is suitable forfor a flowthe
solving field simulation
problem was
with a high angle of attack. Therefore, the k- ω turbulence model was
first order upwind. Yang proved that the k-ω model is suitable for a flow field simulation selected to estimate
the
withReynolds
a high anglestresses of theTherefore,
of attack. RANS equations. In addition,model
the k- ω turbulence the k-ωwasturbulence
selected tomodel
estimateis
known
the Reynolds stresses of the RANS equations. In addition, the k-ω turbulence model It
for its ability to provide accurate predictions of turbulent flow phenomena. is
is particularly
known effective
for its ability in capturing
to provide turbulent
accurate boundary
predictions layers; furthermore,
of turbulent flow phenomena. the k-ω
It is
model excelseffective
particularly at modeling near-wallturbulent
in capturing flows, which is crucial
boundary in applications,
layers; furthermore,such as turbine
the k-ω model
design. Furthermore, compared to more advanced turbulence models,
excels at modeling near-wall flows, which is crucial in applications, such as turbine such as Large Eddy
de-
Simulation
sign. Furthermore, compared to more advanced turbulence models, such as Large good
(LES) and Direct Numerical Simulation (DNS), the k-ω model offers a Eddy
balance
Simulationbetween
(LES)accuracy
and Direct andNumerical
computational efficiency
Simulation [38]. the k-ω model offers a good
(DNS),
balance between accuracy and computational efficiency [38].and tested at six different ro-
The turbine was designed using SolidWorks software
tationTheangles:
turbine 0, 10,
was20, 30, 40, using
designed and 50SolidWorks
degrees. The bladeand
software positions
tested atat six
each angle were
different rota-
determined based on experimental tests conducted by Yang and Yawn.
tion angles: 0, 10, 20, 30, 40, and 50 degrees. The blade positions at each angle were deter- The computational
domain of the turbine is illustrated in Figure 4.
mined based on experimental tests conducted by Yang and Yawn. The computational do-
The boundary condition for the computational domain (Figure 4) for the left, right,
main of the turbine is illustrated in Figure 4.
and lower surface is Wall. For the inlet and outlet, the boundary condition is Normal Speed
and Average Relative Static Pressure: 0 Pa, relatively. The boundary condition of the upper
surface is Opening.
In order to simulate the flow around the Hunter turbine, it was essential to generate
appropriate grids for the computational domain. Both structured and unstructured grids
were tested, but due to the complex geometry of the turbine and its curvature, especially
during the closing, creating a structured grid proved to be complicated and difficult. There-
fore, unstructured grids with tetrahedral cells were used to discretize the computational
domain. The number of cells on the blades and the body of the turbine is illustrated in
Figures 5 and 6, respectively.
Water2023,
Water 15, x2312
2023,15, FOR PEER REVIEW 7 of 18 7 of

Figure 4. The computational domain of the turbine.

The boundary condition for the computational domain (Figure 4) for the left, righ
and lower surface is Wall. For the inlet and outlet, the boundary condition is Norm
Speed and Average Relative Static Pressure: 0 Pa, relatively. The boundary condition o
the upper surface is Opening.
In order to simulate the flow around the Hunter turbine, it was essential to genera
appropriate grids for the computational domain. Both structured and unstructured grid
were tested, but due to the complex geometry of the turbine and its curvature, especial
during the closing, creating a structured grid proved to be complicated and difficul
Therefore, unstructured grids with tetrahedral cells were used to discretize the comput
tional domain. The number of cells on the blades and the body of the turbine is illustrate
Figure 4. The computational domain of the turbine.
in Figure
Figure 4. The5computational
and Figure domain
6, respectively.
of the turbine.

The boundary condition for the computational domain (Figure 4) for the left, righ
and lower surface is Wall. For the inlet and outlet, the boundary condition is Norm
Speed and Average Relative Static Pressure: 0 Pa, relatively. The boundary condition o
the upper surface is Opening.
In order to simulate the flow around the Hunter turbine, it was essential to genera
appropriate grids for the computational domain. Both structured and unstructured grid
were tested, but due to the complex geometry of the turbine and its curvature, especiall
during the closing, creating a structured grid proved to be complicated and difficul
Therefore, unstructured grids with tetrahedral cells were used to discretize the computa
tional domain. The number of cells on the blades and the body of the turbine is illustrate
in Figure 5 and Figure 6, respectively.

Figure 5. Grids in the computational domain.


Figure 5. Grids in the computational domain.

To ensure that simulation results are independent of mesh numbers, a mesh conver-
gence test was performed at θ = 0, and the torque results as a function of grid number are
presented in Figure 7. It can be observed that for more than 1,400,000 grids, the resulting
answer becomes almost uniform, indicating that the simulation results are consistent and
not significantly influenced by the mesh size.

Figure 5. Grids in the computational domain.


Water
Water 2023, 15,15,x 2312
2023, FOR PEER REVIEW 8 of 18 8 of 19

Figure 6. The number of cells on the blades and the body of the turbine.

To ensure that simulation results are independent of mesh numbers, a mesh conver-
gence test was performed at θ = 0, and the torque results as a function of grid number are
presented in Figure 7. It can be observed that for more than 1,400,000 grids, the resulting
answer becomes almost uniform, indicating that the simulation results are consistent and
notFigure
significantly influenced
6. The number of cellsby
onthe
themesh
bladessize.
and the body of the turbine.
Figure 6. The number of cells on the blades and the body of the turbine.

0.85
To ensure that simulation results are independent of mesh numbers, a mesh conver
gence test was performed at θ = 0, and the torque results as a function of grid number are
presented in Figure 7. It can be observed that for more than 1,400,000 grids, the resulting
0.8becomes almost uniform, indicating that the simulation results are consistent and
answer
Torque (N.m)

not significantly influenced by the mesh size.


0.75
0.85

0.7
0.8
Torque (N.m)

0.65
0.750 0.5 1 1.5 2
Grid number (million)
Figure 7. 0.7
Grid independency.
Figure 7. Grid independency.
5. Results and Discussion
5.1. The0.65
Performance of a Stand-Alone Turbine
5. Results and Discussion
To ensure 0 the validity of 0.5 the simulation results, 1 a comparison 1.5was made between 2 the
5.1. The Performance
numerical simulation of adata
Stand-Alone Turbine test results. The turbine was tested at six
and experimental
Grid number (million)
To ensure
different the validity
flow velocities of 0.4of m/s,
the simulation
0.7 m/s, 0.8results,
m/s, 1 m/s,a comparison
1.2 m/s, and was1.4made
m/s, andbetween
the the
turbine was
numerical submerged
simulation to aand
data depth of 2D. The flow
experimental testcoefficients
results. Thecorresponding
turbine was to tested
each ofatthesix dif-
six inlet
Figure velocities
7. Grid were
independency.0.25, 0.32, 0.40, 0.46, 0.49, and 0.52.
ferent flow velocities of 0.4 m/s, 0.7 m/s, 0.8 m/s, 1 m/s, 1.2 m/s, and 1.4 m/s, and the turbine
The comparison
was submerged of the overall
to a depth of 2D.performance between thecorresponding
The flow coefficients experiments andtothe numeri-
each of the six
cal
5. simulation
Results and for a submerged
Discussion depth of 2D is shown in Figure 8. Both experimental and
inlet velocities were 0.25, 0.32, 0.40, 0.46, 0.49, and 0.52.
numerical results show that the turbine power coefficient first increases with increasing
5.1.The
flow Thecomparison
Performance
coefficient, ofofthe
reaching overall performance
a aStand-Alone
maximum Turbine
value of Cp , between
and then the experiments
decreases. The power and coeffi-
the numer-
ical simulation
cient To ensure
obtained forthe
from a the
submerged
validity
numerical depth
of the offor
2Da is
simulation
solution shown
results,
flow in Figure
of 0.478.was
a comparison
coefficient Both
was experimental
0.185,made the andthe
whilebetween
numerical
experimentalresults show
data yielded
numerical simulation that
data the
a power turbine power coefficient
coefficient of 0.177.
and experimental first
The turbine
test results. increases
reaches its
The turbine waswith increasing
maximum
tested at six dif
flow coefficient,
performance at reaching
a flow a maximum
coefficient of 0.47 value
for a of C
ferent flow velocities of 0.4 m/s, 0.7 m/s, 0.8 m/s, 1 m/s, 1.2 m/s, and 1.4 m/s,power
p, anddepth
submerged then decreases.
of 2D. The The
comparison
and theofcoeffi-
turbine
generated
cient power by the
changing the flow velocityfor
for athe experimental test and the numerical
was submerged to a depth of 2D. The flow coefficients corresponding to each ofwhile
obtained from numerical solution flow coefficient of 0.47 was 0.185, the six
thesimulation
experimental is presented
data in Table
yielded a 1, whichcoefficient
power shows a small
inlet velocities were 0.25, 0.32, 0.40, 0.46, 0.49, and 0.52. of difference
0.177. The betweenreaches
turbine the experi-
its maxi-
mummental and simulation
performance results.
at a flow The difference
of 0.47between the experimental
depth and numerical
The comparison
simulations is because of thecoefficient
of two overall performance
main factors:
for abetween
submerged
1. experimental the experiments
errors;
of 2D.
2. assumptions
The the
and compar-
and numer
ison
icalofsimulation
generated for power by changing
a submerged depththeofflow
2D is velocity
shownfor the experimental
in Figure test and the
8. Both experimental and
boundary conditions in the computational domain.
numerical results show that the turbine power coefficient first increases with increasing
flow coefficient, reaching a maximum value of Cp, and then decreases. The power coeffi
cient obtained from the numerical solution for a flow coefficient of 0.47 was 0.185, while
the experimental data yielded a power coefficient of 0.177. The turbine reaches its maxi
mum performance at a flow coefficient of 0.47 for a submerged depth of 2D. The compar
numerical simulation is presented in Table 1, which shows a small difference between the
experimental and simulation results. The difference between the experimental and nu-
Water 2023, 15, 2312 merical simulations is because of two main factors: 1. experimental errors; 2. assumptions
9 of 18
and boundary conditions in the computational domain.

0.35
Numerical
Experimental

0.25
Cp

0.15

0.05
0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55
Cf
Figure 8. Comparison between calculated and experimental amounts of power coefficient.
Figure 8. Comparison between calculated and experimental amounts of power coefficient.
Table 1. Comparison of experimental and numerical generated power.
Table 1. Comparison of experimental and
Velocity (m/s) numerical
Power generated power.Power (w)—Nu.
(w)—Exp.

Velocity0.4
(m/s) Power0.017(w)—Exp. 0.09 (w)—Nu.
Power
0.7 0.11 0.17
0.40.8 0.017
0.19 0.29 0.09
0.7 1 0.11
0.44 0.53 0.17
0.81.2 0.58
0.19 0.78 0.29
1.4 0.89 1.06
1 0.44 0.53
1.2 0.58 0.78
The velocity contours and pressure contours at different angles and coefficients are
1.4
displayed in Figures 9 and 10, respectively.
0.89 1.06
As depicted in Figure 10, the pressure on Blade 2 is considerably higher at θ = 0 and
θ =The velocity
50, and contours
it increases withand pressure
the flow contours
coefficient. at different
However, angles
at θ = 20, and
there is no coefficients
significant are
displayed
pressurein Figure 9on
difference and Figure
either side10,
of respectively.
Blade 2. The pressure difference across Blade 2 is
utilized to generate torque, which causes the blade to rotate around the hinge pin when
it is enlarged. The generated torque is observed to be greater at rotation angles of 0 and
50 compared to 20. At θ = 0, blades 3, 4, and 5 are almost entirely closed and contribute little
to the output torque, while only Blade 2 is fully opened and contributes significantly to the
torque. Similarly, at θ = 50, most of the output torque is produced by Blade 1, which is fully
opened at this position. Additionally, raising the flow coefficient increases the pressure
difference between the two blade surfaces, as shown in Figures 9 and 10 [13].

5.2. Effect of Turbine Submerged Depth on Its Performance


In this step, the performance of the Hunter turbine was evaluated at different sub-
merged depths to determine the optimal vertical position for achieving the highest perfor-
mance. Six flow velocities were tested, with corresponding flow coefficients of 0.25, 0.32,
0.40, 0.46, 0.49, and 0.52. The results, shown in Figure 11, indicate that as the flow coefficient
increases, the turbine power coefficient initially increases and then decreases after reaching
the maximum value of Cp. Additionally, as the vertical distance from the water surface
increases, the turbine performance also increases, but only up to a certain height. Beyond
this height, the power that can be obtained from the stream remains constant, and this
height represents the optimal position for placing the turbine. The results indicate that the
best position is at a depth of 3D, where the maximum power coefficient achieved by the
turbine is 0.218 at a flow coefficient of 0.49.
Water 2023, 15, 2312 10 of 18
Water 2023, 15, x FOR PEER REVIEW 10 of 19

Figure9.9.Velocity
Figure Velocitycontours
contours forfor
3D3Dfromfrom θ =θ0,= 20, 50 and
0, 20, Cf 0.4
50 and Cf and 0.46.0.46.
0.4 and ((a):((a):
θ = 0,
θ Cf
= 0,=Cf0.4;= (b): = 20,
0.4; θ(b): θ=
Cf
20, Cf = 0.4; (c): θ = 50, Cf = 0.4; (d): θ = 0, Cf = 0.46; (e): θ = 20, Cf = 0.46; (f): θ = 50, Cf = 0.46). θ 20,
= 0.4; (c): θ = 50, Cf = 0.4; (d): θ = 0, Cf = 0.46; (e): θ = 20, Cf = 0.46; (f): θ = 50, Cf = 0.46). θ = 0, = 0,
50
20,and Cf 0.49
50 and and and
Cf 0.49 0.52.0.52.
((g):((g):
θ = θ0,=Cf 0, =Cf0.49; (h):(h):
= 0.49; θ =θ20, CfCf
= 20, = 0.49; (i):(i):
= 0.49; θθ = =50,
50,CfCf= =0.49;
0.49;(j):
(j):θθ==0,0,
Cf
Cf==0.52;
0.52;(k):
(k):θθ==20,
20,Cf
Cf== 0.52;
0.52; (l):
(l): θθ == 50,
50, Cf
Cf == 0.52).
Water 2023, 15, 2312 11 of 18
Water 2023, 15, x FOR PEER REVIEW 11 of 19

Figure 10. Pressure


Figure Pressurecontours
contoursfor for3D3Dfrom
from θ =θ0,= 20, 50 and
0, 20, Cf 0.4
50 and Cfand
0.4 0.46. ((a): θ((a):
and 0.46. = 0, Cfθ == 0,
0.4;Cf(b): θ
= 0.4;
= 20,θ Cf
(b): = 0.4;
= 20, Cf (c): θ =(c):
= 0.4; 50,θCf = 0.4;
= 50, Cf =(d): θ (d):
0.4; = 0, θCf==0,0.46;
Cf =(e): θ (e):
0.46; = 20,θ Cf = 0.46;
= 20, Cf =(f): θ =(f):
0.46; 50,θ Cf = 0.46).
= 50, θ=
Cf = 0.46).
θ0,=20,
0, 50
20,and Cf 0.49
50 and and 0.52.
Cf 0.49 ((g): θ((g):
and 0.52. = 0,θCf= =0,0.49;
Cf =(h): θ =(h):
0.49; 20, θCf==20,
0.49;
Cf (i): θ = 50,
= 0.49; (i):Cf
θ= = 0.49;
50, Cf (j):= θ0.49;
=
0, Cf = 0.52; (k): θ = 20, Cf = 0.52; (l): θ = 50, Cf = 0.52).
(j): θ = 0, Cf = 0.52; (k): θ = 20, Cf = 0.52; (l): θ = 50, Cf = 0.52).
coefficient increases, the turbine power coefficient initially increases and then decreases
after reaching the maximum value of Cp. Additionally, as the vertical distance from the
water surface increases, the turbine performance also increases, but only up to a certain
height. Beyond this height, the power that can be obtained from the stream remains con-
Water 2023, 15, 2312 stant, and this height represents the optimal position for placing the turbine. The 12
results
of 18
indicate that the best position is at a depth of 3D, where the maximum power coefficient
achieved by the turbine is 0.218 at a flow coefficient of 0.49.

Figure 11. The turbine performances in different submerged depths.


Figure 11. The turbine performances in different submerged depths.

Figure11.
Figure 11.illustrates
illustrates the
the pressure
pressure contours
contours of the
of the turbine
turbine at different
at different vertical
vertical posi-
positions,
tions, a rotation angle of 50, and a flow coefficient of 0.46. It is evident from the
a rotation angle of 50, and a flow coefficient of 0.46. It is evident from the figure that as the figure that
as the height increases, the static pressure, total pressure, low-pressure
height increases, the static pressure, total pressure, low-pressure area, and back pressure area, and back
Water 2023, 15, x FOR PEER REVIEW 13 of 19
pressure
behind thebehind
turbinethe turbine
also alsodue
increase increase
to the due toof
effect thestatic
effect of static The
pressure. pressure. Thecontours
pressure pressure
ofcontours of the
the turbine turbine in
in different different
vertical verticalwith
positions, positions, with
a rotation a rotation
angle is shownangle is shown
in Figure 12. in
Figure 12.

Figure 12. The pressure contours of the turbine in different vertical positions, with a rotation angle of
50Figure
and a 12.
flowThe pressureof
coefficient contours
0.46. of the turbine in different vertical positions, with a rotation angle
of 50 and a flow coefficient of 0.46.
In submerged depths of 2D and 3D, by increasing the flow coefficient from 0.25 to 0.52,
In submerged
the maximum torquedepths of 2D
produced by and 3D, by increases
the turbine increasingfrom
the flow coefficient
1.1 Nm from
to 2.1 Nm and0.25
fromto
0.52,
1.3 Nm thetomaximum torque produced
2.2 Nm, respectively. by the
However, in turbine
depths increases
of 4D andfrom 1.1 maximum
5D, the Nm to 2.1 Nm and
torque
from 1.3 Nm to 2.2 Nm, respectively. However, in depths of 4D and 5D, the maximum
torque produced by the turbine is relatively constant, ranging from 1.7 Nm to 1.9 Nm and
from 1.9 Nm to 2 Nm, respectively, for flow coefficients ranging from 0.25 to 0.52.
It can be observed from Figures 13–16 that the maximum torque produced by the
turbine occurs at a rotational angle of 10 degrees in all flow coefficients and submerged
Water 2023, 15, 2312 13 of 18

produced by the turbine is relatively constant, ranging from 1.7 Nm to 1.9 Nm and from
1.9 Nm to 2 Nm, respectively, for flow coefficients ranging from 0.25 to 0.52.
It can be observed from Figures 13–16 that the maximum torque produced by the
turbine occurs at a rotational angle of 10 degrees in all flow coefficients and submerged
depths. At a rotational angle of 30 degrees, the torque produced by the turbine is minimum.
Furthermore, the results show that the optimum submerged depth for the turbine is 3D,
as it produces the highest torque in all flow coefficients. This result is consistent with the
Water 2023,
Water 2023, 15,
15, xx FOR
FOR PEER
PEER REVIEW
REVIEW 11
findings in the previous section, which showed that the best position for placing the turbine
is at a depth of 3D.

Figure
Figure
Figure 13.
13.13. The
TheThe trend
trend
trend of change
of change
of change of
of totalof totalwith
total
torque torque
torque with
anglewith anglewith
angle
of rotation of rotation
of rotation with
with
various flow variousin
various
coefficients flow coeffi
flow coeffi
in 1D depth.
in depth.
1D 1D depth.

Figure
Figure
Figure 14.
14.14. The
TheThe trend
trend
trend of change
of change
of change of
of totalof totalwith
total
torque torque
torque with
anglewith anglewith
angle
of rotation of rotation
of rotation with
with
various flow variousin
various
coefficients flow coeffi
flow coeffi
in
in 2D
2D depth.
depth.
2D depth.
Water 2023, 15, 2312 14 of 18
Figure 14. The trend of change of total torque with angle of rotation with various flow coeffi
in 2D depth.

Water 2023, 15, x FOR PEER REVIEW 1


Figure
Figure 15.15.
TheThe trend
trend of change
of change of total of totalwith
torque torque
anglewith anglewith
of rotation of rotation with
various flow various inflow coeffi
coefficients
in depth.
3D 3D depth.

Figure
Figure 16.16.
TheThe trend
trend of change
of change of total of totalwith
torque torque
anglewith anglewith
of rotation of rotation with
various flow various inflow coeffi
coefficients
in 4D depth.
4D depth.

Figures
Figures 17–20 depict
17–20 the turbine
depict power coefficient
the turbine in different in
power coefficient submerged
differentdepths as a
submerged dep
function of the rotational angle and flow coefficient. It is evident that the turbine power
a function of the rotational angle and flow coefficient. It is evident that the turbine
coefficient increases for all flow coefficients from θ = 0◦ to θ = 10◦ and reaches its maximum
coefficient
value, increases
as the turbine for allthe
generates flow coefficients
highest torque atfrom
θ = 10θ◦ .=Then,
0° tothe
θ =trend
10° and reaches
changes, and its max
value,
the power ascoefficient
the turbine generates
decreases the highest
by increasing torque atangle,
the rotational θ = 10°. Then,
reaching its the trend change
minimum
◦ . Subsequently, the turbine power coefficient increases again from θ = 30◦
the power coefficient decreases by increasing the rotational angle, reaching its min
value at θ = 30
tovalue ◦ . The flow coefficient of 0.32 results in the best turbine power coefficient, whereas
θ = 60at θ = 30°. Subsequently, the turbine power coefficient increases again from θ
the worst performance is observed in the flow coefficient of 0.52. Furthermore, the power
to θ = 60°. The flow coefficient of 0.32 results in the best turbine power coefficient, w
coefficient initially increases with an increase in the submerged depth and becomes equal
thedepths
for worstofperformance
3D and 4D. is observed in the flow coefficient of 0.52. Furthermore, the
coefficient initially increases with an increase in the submerged depth and becomes
for depths of 3D and 4D.
the power coefficient decreases by increasing the rotational angle, reaching its min
value at θ = 30°. Subsequently, the turbine power coefficient increases again from
to θ = 60°. The flow coefficient of 0.32 results in the best turbine power coefficient, w
the worst performance is observed in the flow coefficient of 0.52. Furthermore, the
Water 2023, 15, 2312 coefficient initially increases with an increase in the submerged depth 15 of 18become
and
for depths of 3D and 4D.

Water 2023, 15, x FOR PEER REVIEW 1


Figure
Figure 17.17.
TheThe trend
trend of change
of change of power
of power coefficient
coefficient with anglewith anglewith
of rotation of rotation withcoeffi-
various flow various flow
cientsinin
cients 1D1D depth.
depth.

Figure
Figure 18.18.
TheThe trend
trend of change
of change of power
of power coefficient
coefficient with anglewith angle with
of rotation of rotation withcoeffi-
various flow various flow
cients in 2D depth.
cients in 2D depth.
Water 2023, 15, 2312 16 of 18
Figure 18. The trend of change of power coefficient with angle of rotation with various flow
cients in 2D depth.

Water 2023, 15, x FOR PEER REVIEW 17


Figure
Figure 19.19.
TheThe trend
trend of change
of change of power
of power coefficient
coefficient with anglewith angle with
of rotation of rotation withcoeffi-
various flow various flow
cients
cients in in
3D3D depth.
depth.

Figure 20. The trend of change of power coefficient with angle of rotation with various flow coeffi-
Figure
cients in20.
4D The trend of change of power coefficient with angle of rotation with various flow
depth.
cients in 4D depth.
6. Conclusions
The present study aimed to analyze a vertical axis tidal turbine (VATT) both exper-
6. Conclusions
imentally and numerically. Initially, an experimental model of the Hunter turbine was
The with
designed present study
an aspect aimed
ratio of 1.2to analyze
and a avertical
tested in axis tidal
water channel. turbinedata
Laboratory (VATT)
were both ex
mentally and numerically. Initially, an experimental model of the Hunter turbine wa
used to verify the RANS equations, which were implemented around the turbine using
CFD software.
signed with an The effectratio
aspect of theof
turbine’s
1.2 andsubmerged
tested in adepth
waterfrom the water
channel. surface on data
Laboratory its were
performance was then investigated. The turbine was analyzed at four different depths: 1D,
to verify the RANS equations, which were implemented around the turbine using
2D, 3D, and 4D. The numerical solution was reasonably close to the experimental data. The
software. The effect of the turbine’s submerged depth from the water surface on its
formance was then investigated. The turbine was analyzed at four different depths
2D, 3D, and 4D. The numerical solution was reasonably close to the experimental
The results demonstrated that as the vertical distance from the water surface incre
Water 2023, 15, 2312 17 of 18

results demonstrated that as the vertical distance from the water surface increased, the
energy that could be harnessed from the water flow also increased, resulting in improved
turbine performance. However, there was a specific height beyond which increasing the
vertical distance did not enhance the turbine’s performance. In other words, there was a
minimum depth required for achieving maximum power output, which was found to be
3D. Moreover, this study revealed that the maximum power coefficient of the turbine was
observed in smaller flow coefficients with increased depth. The highest power coefficient
achieved was 0.218 at a flow coefficient of 0.49 for submerged depths of 3D and 4D.

Author Contributions: Conceptualization, E.G. and H.K.; methodology, E.G., H.K. and S.H.; software,
E.G. and H.K.; validation, E.G. and H.K.; formal analysis, E.G., H.K. and S.H.; investigation, E.G.,
H.K. and S.H.; resources, E.G., H.K. and S.H.; data curation, E.G. and H.K.; writing—original draft
preparation, E.G. and H.K.; writing—review and editing, E.G., H.K. and S.H.; visualization, E.G.,
H.K. and S.H.; supervision, S.H.; project administration, E.G., H.K. and S.H. All authors have read
and agreed to the published version of the manuscript.
Funding: This research received no external funding.
Data Availability Statement: All data used to support the findings of this study are included in
the article.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Harries, T.; Kwan, A.; Brammer, J.; Falconer, R. Physical testing of performance characteristics of a novel drag-driven vertical axis
tidal stream turbine; with comparisons to a conventional Savonius. Int. J. Mar. Energy 2016, 14, 215–228. [CrossRef]
2. Kariman, H.; Hoseinzadeh, S.; Heyns, S.; Sohani, A. Modeling and exergy analysis of domestic MED desalination with brine tank.
Desalin. Water Treat. 2020, 197, 1–13. [CrossRef]
3. Mansour, A.E.; Pedersen, P.T.; Paik, J.K. Wave energy extraction using decommissioned ships. Ships Offshore Struct. 2013,
8, 504–516. [CrossRef]
4. Kariman, H.; Hoseinzadeh, S.; Heyns, P.S. Energetic and exergetic analysis of evaporation desalination system integrated with
mechanical vapor recompression circulation. Case Stud. Therm. Eng. 2019, 16, 100548. [CrossRef]
5. Derakhshan, S.; Kasaeian, N. Optimization, Numerical, and Experimental Study of a Propeller Pump as Turbine. J. Energy Resour.
Technol.-Trans. Asme 2014, 136, 012005. [CrossRef]
6. Kariman, H.; Hoseinzadeh, S.; Shirkhani, A.; Heyns, P.S.; Wannenburg, J. Energy and economic analysis of evaporative vacuum
easy desalination system with brine tank. J. Therm. Anal. Calorim. 2020, 140, 1935–1944. [CrossRef]
7. Khan, M.J.; Bhuyan, G.; Iqbal, M.T.; Quaicoe, J.E. Hydrokinetic energy conversion systems and assessment of horizontal and
vertical axis turbines for river and tidal applications: A technology status review. Appl. Energy 2009, 86, 1823–1835. [CrossRef]
8. Martin-Short, R.; Hill, J.; Kramer, S.C.; Avdis, A.; Allison, P.A.; Piggott, M.D. Tidal resource extraction in the Pentland Firth,
UK: Potential impacts on flow regime and sediment transport in the Inner Sound of Stroma. Renew. Energy 2015, 76, 596–607.
[CrossRef]
9. Sanchez, M.; Carballo, R.; Ramos, V.; Iglesias, G. Tidal stream energy impact on the transient and residual flow in an estuary: A
3D analysis. Appl. Energy 2014, 116, 167–177. [CrossRef]
10. Chen, L.; Lam, W.H. A review of survivability and remedial actions of tidal current turbines. Renew. Sustain. Energy Rev. 2015,
43, 891–900. [CrossRef]
11. Ma, Y.; Zhang, L.; Ma, L.; Chen, Z. Developing status and development trend of vertical axis turbine-type tidal current energy
power generation device. Keji Daobao Sci. Technol. Rev. 2012, 30, 71–75.
12. Batten, W.M.J.; Bahaj, A.S.; Molland, A.F.; Chaplin, J.R. The prediction of the hydrodynamic performance of marine current
turbines. Renew. Energy 2008, 33, 1085–1096. [CrossRef]
13. Li, Y.; Calisal, S.M. Numerical analysis of the characteristics of vertical axis tidal current turbines. Renew. Energy 2010, 35, 435–442.
[CrossRef]
14. Copping, A.; Hanna, L.; Whiting, J.; Geerlofs, S.; Grear, M.; Blake, K.; Coffey, A.; Massaua, M.; Brown-Saracino, J.; Battey, H.
Environmental Effects of Marine Energy Development around the World; Annex IV Final Report; IEA Ocean Energy Systems Initiative,
Annex IV: Richland, WA, USA, 2013.
15. Jacobson, P.T.; Amaral, S.V.; Castro-Santos, T.; Giza, D.; Haro, A.J.; Hecker, G.; McMahon, B.; Perkins, N.; Pioppi, N. Environmental
Effects of Hydrokinetic Turbines on Fish: Desktop and Laboratory Flume Studies; Electric Power Research Institute: Palo Alto, CA,
USA, 2012.
16. Yang, B.; Lawn, C. Fluid dynamic performance of a vertical axis turbine for tidal currents. Renew. Energy 2011, 36, 3355–3366.
[CrossRef]
Water 2023, 15, 2312 18 of 18

17. Chen, B.; Cheng, S.B.; Su, T.C.; Zhang, H. Numerical investigation of channel effects on a vertical-axis tidal turbine rotating at
variable speed. Ocean Eng. 2018, 163, 358–368. [CrossRef]
18. Jing, F.M.; Sheng, Q.H.; Zhang, L. Experimental research on tidal current vertical axis turbine with variable-pitch blades. Ocean
Eng. 2014, 88, 228–241. [CrossRef]
19. Sun, J.J.; Huang, D.G. Numerical investigation on aerodynamic performance improvement of vertical-axis tidal turbine with
super-hydrophobic surface. Ocean Eng. 2020, 217, 107995. [CrossRef]
20. Derakhshan, S.; Ashoori, M.; Salemi, A. Experimental and numerical study of a vertical axis tidal turbine performance. Ocean
Eng. 2017, 137, 59–67. [CrossRef]
21. Maduka, M.; Li, C.W. Experimental evaluation of power performance and wake characteristics of twin flanged duct turbines in
tandem under bi-directional tidal flows. Renew. Energy 2022, 199, 1543–1567. [CrossRef]
22. Arcos, F.Z.D.; Vogel, C.R.; Willden, R.H.J. A parametric study on the hydrodynamics of tidal turbine blade deformation. J. Fluids
Struct. 2022, 113, 103626. [CrossRef]
23. Moreau, M.; Germain, G.; Maurice, G.; Richard, A. Sea states influence on the behaviour of a bottom mounted full-scale twin
vertical axis tidal turbine. Ocean Eng. 2022, 265, 112582. [CrossRef]
24. Sun, K.; Yi, Y.; Zhang, J.S.; Zhang, J.H.; Zaidi, S.S.H.; Sun, S.H. Influence of blade numbers on start-up performance of vertical
axis tidal current turbines. Ocean Eng. 2022, 243, 110314. [CrossRef]
25. Ma, Y.; Hu, C.; Li, L. Hydrodynamics and wake flow analysis of a Π-type vertical axis twin-rotor tidal current turbine in surge
motion. Ocean Eng. 2021, 224, 108625. [CrossRef]
26. Satrio, D.; Utama, I.K.A.P. Experimental investigation into the improvement of self-starting capability of vertical-axis tidal current
turbine. Energy Rep. 2021, 7, 4587–4594. [CrossRef]
27. Chen, Y.L.; Sun, J.; Lin, B.L.; Lin, J.; Guo, J.X. Spatial evolution and kinetic energy restoration in the wake zone behind a tidal
turbine: An experimental study. Ocean Eng. 2021, 228, 108920. [CrossRef]
28. Xie, J.M.; Chen, J.Y. Vertical-axis ocean current turbine design research based on separate design concept. Ocean Eng. 2019,
188, 106258. [CrossRef]
29. Han, J.; Jung, J.; Hwang, J.H. Optimal configuration of a tidal current turbine farm in a shallow channel. Ocean Eng. 2021,
220, 108395. [CrossRef]
30. Manolesos, M.; Chng, L.; Kaufmann, N.; Ouro, P.; Ntouras, D.; Papadakis, G. Using vortex generators for flow separation control
on tidal turbine profiles and blades. Renew. Energy 2023, 205, 1025–1039. [CrossRef]
31. Wang, P.Z.; Zhao, B.W.; Cheng, H.T.; Huang, B.; He, W.S.; Zhang, Q.; Zhu, F.W. Study on the performance of a 300W counter-
rotating type horizontal axis tidal turbine. Ocean Eng. 2022, 255, 111446. [CrossRef]
32. Samadi, M.; Hassanabad, M.G.; Mozafari, S.B. Performance enhancement of low speed current savonius tidal turbines through
adding semi-cylindrical deflectors. Ocean Eng. 2022, 259, 111873. [CrossRef]
33. Khanjanpour, M.H.; Javadi, A.A. Optimization of a Horizontal Axis Tidal (HAT) turbine for powering a Reverse Osmosis (RO)
desalination system using Computational Fluid Dynamics (CFD) and Taguchi method. Energy Convers. Manag. 2021, 231, 113833.
[CrossRef]
34. Yang, B.; Lawn, C. Three-dimensional effects on the performance of a vertical axis tidal turbine. Ocean Eng. 2013, 58, 1–10.
[CrossRef]
35. Wang, X.L.; Qiao, D.S.; Jin, L.X.; Yan, J.; Wang, B.; Li, B.B.; Ou, J.P. Numerical investigation of wave run-up and load on heaving
cylinder subjected to regular waves. Ocean Eng. 2023, 268, 113415. [CrossRef]
36. Liang, H.Z.; Qiao, D.S.; Wang, X.Z.; Zhi, G.N.; Yan, J.; Ning, D.Z.; Ou, J.P. Energy capture optimization of heave oscillating buoy
wave energy converter based on model predictive control. Ocean Eng. 2023, 268, 113402. [CrossRef]
37. Wang, Z.M.; Qiao, D.S.; Yan, J.; Tang, G.Q.; Li, B.B.; Ning, D.Z. A new approach to predict dynamic mooring tension using LSTM
neural network based on responses of floating structure. Ocean Eng. 2022, 249, 110905. [CrossRef]
38. Nazarieh, M.; Kariman, H.; Hoseinzadeh, S. Numerical simulation of fluid dynamic performance of turbulent flow over Hunter
turbine with variable angle of blades. Int. J. Numer. Methods Heat Fluid Flow 2023, 33, 153–173. [CrossRef]

Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.

You might also like