Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 21

Ocean Engineering 287 (2023) 115910

Contents lists available atScienceDirect

Ocean Engineering

journal homepage: www.elsevier.com/locate/oceaneng

An investigation into the self-starting of darrieus-savonius hybrid wind turbine and


performance enhancement through innovative deflectors: A CFD approach
Sahel Chegini a, Mohammadreza Asadbeigi b, Farzad Ghafoorian b, Mehdi Mehrpooya c,*
a
School of Railway Engineering, Iran University of Science and Technology, Tehran, Iran
b
Turbomachinery Research Laboratory, Department of Energy Conversion, School of Mechanical Engineering, Iran University of Science and Technology, Tehran, Iran c
Faculty of New Sciences and Technologies, University of Tehran, Tehran, Iran

Handling Editor: A.I. Incecik Amidst vertical axis wind turbines, Darrieus turbines stand out for their impressive efficiency. However, these turbines
encounter a significant challenge regarding their self-starting ability. Therefore, the current study seeks to explore the
Keywords: effects of combining Savonius and Darrieus turbines on the self-starting capability of the Darrieus turbine and propose a new
Savonius-darrieus hybrid wind turbine wind turbine design with superior starting torque. Additionally, the study investigates the effectiveness of novel deflectors
Self-starting ability placed in front of and alongside the hybrid turbine, focusing on increasing the efficiency of the hybrid turbine. For this
Deflector purpose, numerical simulations are carried out using a two- dimensional computational fluid dynamics (CFD) approach to
CFD simulation simulate the flow field. The SST k- ω turbulence model is employed to provide closure for the Unsteady Reynolds-Averaged
Navier-Stokes (URANS) equations. Also, the modeling validation and verification were performed by sensitivity analysis of
mesh distribution and time step. The research findings indicated that coupling the Darrieus and Savonius turbines improved
the hybrid turbine’s self-starting ability, with a 26.91% increase in power coefficient (Cp) at the lowest tip speed ratio (TSR) of
1.45. However, as TSR increases, the hybrid turbine’s performance decreases due to the Savonius turbine’s limited
aerodynamic performance at high TSRs. This issue is solved using front and side deflectors, improving the hybrid turbine’s
efficiency at the optimum TSR of 2.6 by 30% and 26%, respectively, through increasing wind energy density at the upwind
and leeward regions of the turbine. Furthermore, using both deflectors together, known as the double deflector
configuration, has shown the best performance at both the lowest and optimum TSR.

ARTICLEINFO ABSTRACT

1. Introduction directions (Dixon and Hall, 2010). Based on the aerodynamic


principles that VAWTs employ to capture wind energy, two
Climate change concerns and the critical need to lower categories of VAWTs emerge: lift-driven Darrieus turbines and
greenhouse gas emissions have triggered a worldwide transition drag-driven Savonius turbines. In contrast to Savonius VAWTs,
towards using renewable energy sources. Among these Darrieus VAWTs exhibit superior efficiency at high TSRs.
alternatives, wind energy has become a prominent leader While Darrieus turbines exhibit higher performance
(Sadorsky, 2021). This transition emphasizes the importance of compared to other VAWTs, they encounter challenges in self-
optimizing wind energy technologies to meet rising energy starting ability at low TSRs (Hand and Cashman, 2020).
demands sustainably. Wind turbines play a pivotal role in this Researchers have addressed this issue by studying modifications
endeavor, and they are divided into two primary categories: to blade geometry (Mehrpooya et al., 2023), variable blade pitch
horizontal axis wind turbines (HAWTs) and vertical axis wind angles (Ardaneh et al., 2022), employing J-shaped blades (Celik et
turbines (VAWTs). The latter offers unique advantages that make al., 2022), and hybrid configurations that combine Darrieus and
them particularly suitable for specific applications. VAWTs are a Savonius VAWTs (Liang et al., 2017). The high static torque
more cost-effective, acoustically quieter, and lightweight solution generated by the Savonius turbine at low TSR can be used to
for wind energy. These turbines are suitable for locations with facilitate the Darrieus turbine’s self-starting by incorporating
erratic wind patterns, as they can harness wind from various

* Corresponding author.
E-mail address: mehrpoya@ut.ac.ir (M. Mehrpooya).
https://doi.org/10.1016/j.oceaneng.2023.115910

Received 27 May 2023; Received in revised form 15 September 2023; Accepted 24 September 2023
Available online 4 October 2023
0029-8018/© 2023 Elsevier Ltd. All rights reserved.
S. Chegini et al. Ocean Engineering 287 (2023) 115910
both lift-based and drag-based rotors (Pallotta et al., 2020). SST Shear stress transport Greek
Bhuyan and Biswas (2014)
μ Dynamic viscosity ν Kinematic viscosity

Δθ Azimuthal increment
Nomenclature ω Angular velocity

Symbols ρ Density
V∞ Free stream velocity
Cm Torque coefficient σ Solidity
Cp Power coefficient λ Tip Speed Ratio
CL Lift coefficient
CD Drag coefficient Subscript
P Power t Turbulence
M Torque
Abbreviations
R Rotor radius
URANS Unsteady Reynolds Averaged Naiver
H Rotor height A Swept area
Stokes
FL Lift force
CFD Computational Fluid Dynamic
FD Drag force
HAWT Horizontal Axis Wind
C Blade chord length K Turbulent kinetic energy y+
Turbine VAWT Vertical Axis
Dimensionless wall distance TI Turbulence intensity
Wind Turbine
W Domain width
TSR Tip Speed Ratio
CFL Courant-Friedrichs-Lewy
blade curvature of a Savonius rotor impacts the efficiency of a
conducted a study to assess the hybrid turbine’s self-starting hybrid turbine. The outcomes revealed that through the
ability compared to a Darrieus rotor. The findings of this study optimization of each parameter, it becomes possible to improve
indicate that contrary to the Darrieus turbine, the hybrid VAWT is the hybrid turbine’s performance across both low and high TSR
capable of initiating rotation at any azimuth angle, thus values.
improving its self-starting capabilities. Mohamed (2013) 1. Despite the potential benefits of modifying wind
investigated how the solidity and utilization of a hybrid turbine turbine designs, there are often unintended
affected the Darrieus turbine’s potential to start. The results consequences, including complex shapes that are
revealed that implementing these techniques can lead to challenging to construct, and costly control
significant improvements in the Darrieus rotor’s performance at systems. To address these limitations and increase
low TSRs. Kyozuka (2008) carried out an investigation to explore the efficiency of Savonius and Darrieus wind
the impact of varying attachment angles between Savonius and turbines, researchers have looked into the use of
Darrieus wind turbines on the system’s efficiency. The findings wind guide tools, like guide vanes, nozzles, and
revealed that the rotor’s self-starting ability is profoundly deflectors. Deflector plates, in particular, have
influenced by the attachment angle, which exhibits a notable been shown to be an efficient and affordable
potential for amplifying the initial torque. However, the hybrid method to
turbine’s Cp decreased by up to 70% compared to a Darrieus enhance the performance of VAWTs, as they can
rotor. As was indicated in the prior studies, although the Savonius be installed without requiring modifications to the
turbine plays a pivotal role in increasing the self-starting blades or rotor design (Wong et al., 2018a). By
capability of the Darrieus rotor in hybrid turbines, it also acts as a placing the deflector plate in the vicinity of a multi-
braking mechanism when operating at a higher TSR, which stage Savonius VAWT, its efficiency increased
ultimately led to a reduction in the maximum achievable output between 17 and 42% at different angles (Golecha
power. et al., 2011). Wong et al. (2018b) studied the
The performance of a hybrid turbine at a higher TSR can be impact of a deflector on the Darrieus rotor’s
increased by implementing certain optimizations derived from performance through both experimental and
previous studies aimed at enhancing the efficiency of VAWTs numerical approaches. The findings revealed that
individually (Hosseini and Goudarzi, 2019). These efforts include incorporating the deflector led to enhanced wind
incorporating a cavity on the blade profile (Sobhani et al., 2017), speed around the turbine, leading to a notable
utilizing a blade pitch control mechanism (Shah et al., 2022), and enhancement in the overall performance of the
employing the auxiliary blade (Asadbeigi et al., 2023) to enhance Darrieus rotor. Layeghmand et al. (2020)
the Darrieus VAWT’s efficiency. Additionally, the Savonius investigated the impact of the NACA 0012 airfoil
turbine’s performance has been enhanced through the use of installation angle as a deflector on the efficiency of
twisted blades (Saad et al., 2020), adding fins on the bucket a Savonius VAWT and reported that the highest
(Shouman and Helal, 2023), or by designing the Savonius turbine and lowest power coefficients were obtained at
in the form of a Bach-type (Ibrahim et al., 2020). Asadi and angles of 50 and 90◦, respectively. Qasemi and
Hassanzadeh (2022) examined the impact of internal rotor type Azadani (2020) examined the efficiency of a
on the performance of a hybrid turbine. The findings indicated Darrieus VAWT that was equipped with a deflector
that employing a Bach-type Savonius rotor in the hybrid turbine placed upstream. They found that with optimal
yielded superior efficiency compared to utilizing the conventional operation parameters for the deflector, including
semicircle Savonius rotor. Roshan et al. (2020) conducted its dimensions and position relative to the turbine,
research into how adjusting the arc angle, overlap ratio, and the turbine’s output power could increase by up to

2
S. Chegini et al. Ocean Engineering 287 (2023) 115910
16.42%. Also, by installing the deflector plate at The second step involves performance analysis of the hybrid
the upper and lower regions of the Darrieus rotor turbine with an added deflector. Previous studies investigated
upstream, the efficiency increases by 20% and the influence of
17%, respectively (Chen et al., 2021).
As stated before, the Darrieus VAWT faces limitations in
terms of its self-starting ability. Combining it with a Savonius
turbine offers the potential to enhance torque output at low
TSRs. However, the hybrid turbine’s performance diminishes at
high TSRs compared to the Darrieus turbine. Limited research has
been dedicated to evaluating the effectiveness of such hybrid
rotor systems, a matter of significance for both industry and
academia. After conducting a thorough literature review, several
research gaps have been identified that have spurred current
investigation. These gaps can be summarized as follows.

• Prior research efforts mainly focused on improving the power


output of hybrid turbines at high TSRs by optimizing the
internal turbine geometry. This approach is complex and
costly, involving intricate shapes that are challenging to
construct. Fig. 1. Hybrid wind turbine schematic: (a) without deflector (b) with
• The efficiency of hybrid turbines using a deflector, a double deflector (c) partitioning the hybrid turbine’s trajectory.
potentially more cost-effective enhancement method
Table 1
compared to others has remained largely unexplored in
Geometrical features of Darrieus and Savonius turbines.
previous studies.
• Previous studies, which focused on enhancing the
Geometric specification Quantity
performance of either the Savonius or Darrieus turbines
separately have examined the dimensions and placements of Savonius VAWT Darrieus VAWT
deflectors upstream of VAWTs and neglected to assess their
performance in specific regions of the turbines. Number of blades 2 3
Turbine diameter 0.2 m 1.030 m
Chord length – 0.858 m
Therefore, this study aims to address the challenge of the
Blade profile – NACA0021
Darrieus turbine’s self-starting ability combined with a Savonius Overlap ratio 0.03 m –
turbine and augmenting overall performance through deflector Solidity ratio – 0.25
integration. Additionally, the current study seeks to enhance the
hybrid turbine performance especially, on the upwind, leeward,
and downwind areas of the hybrid turbine by introducing the
deflector parameters on the Darrieus rotor efficiency, including
front, a novel side, and double deflectors.
factors such as angle, distance, and placement. According to
Syawitri et al. (2022), upper deflectors positioned upstream of
2. Study case
the rotor offer higher power enhancement compared to lower
deflectors. Jin et al. (2018) reported that, as the deflector
This research is divided into two sections, with the primary
distance from the turbine increases, the maximum power
goal of conducting a thorough evaluation of the effectiveness of
coefficient tends to decrease. Furthermore, Chen et al. (2021), in
the Savonius- Darrieus hybrid turbine in its bare form, as well as
their exploration of deflector optimization, identified an optimal
examining the impacts of incorporating a deflector on its
configuration characterized by a reduced vertical distance
performance.
between the upper deflector and the rotor’s center, resulting in a
The first part of the investigation involves evaluating the
deflector angle of 30◦. This configuration was shown to enhance
efficiency of the hybrid turbine, the schematic of which is
airflow acceleration within the upwind turbine region. In
depicted in Fig. 1(a). This turbine is constructed by integrating a
consideration of achieving higher aerodynamic efficiency, this
three-bladed Darrieus wind turbine as the outer rotor and a two-
research introduces a front, in this research a front deflector with
bladed Savonius wind turbine as the inner rotor in the middle.
a 30◦ angle, strategically positioned upstream and in close
The Savonius turbine has been modeled based on the empirical
proximity to the turbine. This design, accompanied by dimensions
data presented by Wenehenubun et al. (2015a). This model
selected for their potential to improve aerodynamic
incorporates a two-bladed Savonius rotor with defined
performance, aims to accelerate airflow in the upwind turbine
specifications, including a rotor diameter of 0.2 m, an overlap
area. Additionally, a novel side deflector, matching the
ratio of 0.03 m, and an endplate diameter of 0.37 m. The
dimensions of the front deflector, has been introduced near the
geometric features of the modeled Darrieus VAWT, including the
turbine without impeding the rotor’s rotational space. Notably,
utilization of the NACA 0021 profile and specific dimensions such
this study employs a double deflector that effectively combines
as a diameter of 1.03 m, a blade chord length of 0.0858 m, and a
the benefits of both deflectors for enhanced performance. The
solidity (σ) of 0.25, are derived from the study conducted by deflectors are designed to be 0.8 m long and 0.01 m wide,
Raciti Castelli et al. (2010a). Due to the high lift/drag force ratio respectively. In the mode of employing both deflectors, a
of this profile, the airfoil is efficient at high rotational speeds. An schematic depicting the precise placement of the deflectors is
overview of the geometric specifications of the Darrieus and illustrated in Fig. 1(b).
Savonius turbines is listed in Table 1.

3
S. Chegini et al. Ocean Engineering 287 (2023) 115910
It should be noted that the positioning of each deflector flow behavior in proximity to the wall. Despite the capability of
individually (either just in front or solely next to the turbine) the k-ω model, it is strongly dependent on the turbulence
corresponds to the arrangement shown in Fig. 1(b). condition in the free stream, with small changes in turbulent
3. Numerical setup kinetic energy causing significant variations in turbulent viscosity
and body forces (Kok, 2000). This limitation led to the
3.1. Governing equation introduction of a hybrid model, the shear-stress transport (SST) k-
ω model, which combines the k-ε model in regions away from
To gain insights into the complex fluid flow behavior the wall and the k-ω model near the wall. This turbulence model
surrounding the turbine, the URANS (Unsteady Reynolds- possesses a highly reliable and precise formulation for both the
Averaged Navier-Stokes) equations are employed. These near-wall area and the far free flow regime.
equations extend Reynolds-Averaged Navier-Stokes (RANS) The SST k-ω model has found extensive application in the CFD
equations by introducing controlled unsteadiness, bridging simulation of VAWTs (Menter, 2009; Menter et al., 2005;
steady and fully unsteady simulations. RANS equations build Elkhoury et al., 2015). The present study implements the SST k- ω
upon the Navier-Stokes equations by introducing time-averaging turbulence model for this purpose. In this model, turbulent
to account for turbulence, derived based on Newton’s laws of kinetic energy (k) denotes the energy contained within the
motion and the conservation of mass, and they describe the turbulent eddies and vortices that characterize turbulent motion,
interplay between velocity and pressure in the flow field. The while the specific dissipation rate ( ω) measures the rate of
URANS equations are outlined below (Chowdhury et al., 2016): dissipation of turbulent kinetic energy within the flow due to

∂xi =0 viscous effects. The corresponding transport equations for [ ] k ms22

and ω [s− 1] are defined below (Didane et al., 2019):


( )
[( ) ]
∂ ui ∂ ( ) 1 ∂p ∂ ∂ ui
+ ui uj = − + ν
∂ t ∂xj ∂ ∂ ∂ ρ xi d → → μ
ρ dt (k) +ρ∇.[k( u − ug)] =∇. μ+σ k
t
∇k +Gk − Yk
xj t − ui′uj′ (2) which Eq. (1) shows the

continuity equation and links velocity and


→ug)] σω
pressure variations. Eq. (2) describes the momentum equation, (4)
dt
repre[ ] [ ] senting the dynamic interplay of forces and motion. ui
m m
s and uj s are the mean velocities in the axial system of
Yk and Gk represent the dissipation and generation of
coordinates, providing an [ ] overall representation of fluid flow
[ ]
[] turbulence kinetic energy, respectively. Also, [] Gω is the
characteristics. The terms u′i m
s and u′j m
s represent the
generation of specific dissipation, while Yω signifies the
fluctuating velocities that capture the turbulent and [ ] dissipation of ω. The turbulent Prandtl numbers for k and ω are
unpredictable variations within the fluid. [ ] [ ] p mN2 is the mean denoted as σk[1] and σω
pressure, ν ms2 is the kinematic viscosity and ρ mkg3 is the fluid
density. The term uiuj, commonly known as Reynolds stress, (Sadorsky, 2021The turbulent viscosity ), respectively. [ N.s] is
encapsulates how fluctuations in different velocity components calculated by the equation μt = μt
m2
interact and correlate. It reveals the complex momentum
exchanges induced by turbulence within the fluid. The Reynolds
stress tensor adds complexity to the momentum equations by . The second term in the dominator accounts for
introducing additional unknown terms into the equations, leading the tur-
to a closure problem (Wilcox, 1998). Therefore, the turbulence
bulent shear stress’s transport. In this expression, a∗[1]
model becomes essential to relate Reynolds stress terms and
represents the coefficient of specific turbulent dissipation rate, α1
other flow variables to close the system of equations.
(Sadorsky, 2021) is a constant and SF2 (Sadorsky, 2021) stands for
the shear factor.
3.2. Turbulence model
3.3. Turbine parameters
The turbulence model aims to predict and characterize the
turbulent behavior of the flow through the modeling of transport Various parameters can be utilized to assess the aerodynamic
equations, which encapsulate the complex interplay of fluid efficiency of the turbine. Among these, the moment coefficient
properties as they evolve through space and time. Two-equation and power coefficient are commonly utilized as they illustrate the
eddy viscosity models, such as the various forms of k- ε and k-ω generated torque by the turbine’s blades and its energy
models, are frequently employed for engineering cases. These conversion efficiency. The moment coefficient is calculated as
two models contain two transport equations to describe follows (Leonczuk Minetto and Paraschivoiu, 2020):
turbulent flow properties (Bardina et al., 1997). The k-ε model M[N.m]
employs empirical damping functions within the viscous sublayer, Cm kg 2 =
1
making it less precise in the presence of adverse pressure ρ .D2 [m2 ].H [m].V 2
] [m ] (5)
gradients. In contrast, the k-ω model does not need damping 4 m3 ∞ s2

functions near the wall region, enhancing its ability to predict

4
S. Chegini et al. Ocean Engineering 287 (2023) 115910
where D is the turbine’s diameter, M is the moment, ρ is the less than 2%. Therefore, a choice of 10D for the inlet distance is
density of air, A is the swept area, and V∞ is the free wind appropriate. In Fig. 2(b), deviations in Cp between different outlet
velocity. The power coefficient quantifies the efficiency of a distances become less significant as the inlet distance increases
turbine by comparing the actual power extracted by the rotor to from do = 20D to 40D, with a deviation of only 0.04% observed
the maximum power that could be harnessed from the available within this range. Consequently, selecting d o = 20D is a suitable
wind energy. It is computable using the following equation choice that ensures proper wake development. As illustrated in
(Leonczuk Minetto and Paraschivoiu, 2020): Fig. 2(c), a blockage ratio of 10% leads to a significant
[ ] overestimation of the power coefficient due to flow acceleration
P N.m (Rezaeiha et al., 2017a). To obtain accurate results, a 20D domain
width, corresponding to a blockage ratio of 5%, is selected.
Cp Notably, previous research has shown that the rotating zone
diameter has a negligible effect on the turbine’s performance
here, P is the turbine’s generated power. Moreover, Cp = λCm links (Rezaeiha et al., 2017a). As a result, the rotating zone diameter
was adjusted to 1.5D, in accordance with findings from earlier
the moment coefficient and power coefficient. Here, λ denotes
studies (Beri and Yao, 2011; Balduzzi et al., 2016a; Rossetti and
the tip speed ratio, defined as the ratio of the speed at the tip of
Pavesi, 2013). This adjustment facilitates a gradual increase in
the wind turbine’s rotor blade to the approaching wind speed.
grid size from the airfoils while maintaining an appropriate
This ratio can be measured by (Leonczuk Minetto and
element count.
Paraschivoiu, 2020):

R[m] × ω[s− 1]
3.5. Boundary conditions For boundary conditions, the domain
λ= [m ]
V∞ s inlet and outlet were considered as a uniform velocity inlet of 9
m s− 1 and a static gauge pressure of zero, respectively. The lateral
where ω is the rotor’s angular velocity and R is the radius of the sides of the domain have symmetry boundary conditions as
rotor.
suggested in Balduzzi et al. (2016a). The turbines’ blades were
To comprehend the turbine’s aerodynamic behavior, it’s
essential to define the drag and lift force coefficients acting on defined as a wall with a nonslip condition that rotate with the
the Darrieus blade (Acarer, 2020): rotor’s angular velocity. The stationary domain and rotational
domains are coupled by a nonconformal interface using a sliding
F [N ]
CL =
1
2 ρ]
kg
m3 L
2
.D[[m].V∞2 ]ms(8)
2
mesh technique. Although Castelli et al.‘s experimental
measurements were performed in a low-turbulence wind tunnel,
FD [N ]
[ kg ] [ 2] specifics regarding the wind tunnel’s turbulence characteristics
ρ 3 .D[m].V∞2 ms2
CD =1(9) are not stated. Given the available data and knowledge of the
2 m low blade Reynolds number, a turbulence intensity (TI) of 1% is
considered, with the inlet turbulence length scale set to the
3.4. Computational domain
turbine diameter, i.e., 1 m.
This study was performed within a 2D computational domain
to achieve computational efficiency while capturing essential 3.6. Solver setting
features. The 2D simulations focus on the Darrieus turbine mid-
plane with a large blade aspect ratio (h/c) of 16.97, where 3D tip The transient numerical simulations were performed by
impacts can be avoided (Tescione et al., 2014). The feasibility of a commercial finite volume method-based software ANSYS Fluent
2D investigation is supported by reasonable computational costs 22.2. Due to the incompressibility of the flow, the pressure-based
and the reliability of flow field description through 2D simulations solver was adopted for solving the continuity and momentum
at the turbine’s mid-plane (Bianchini et al., 2017). The equations. The SIMPLE scheme is employed for pressure-velocity
simulations exclude the turbines’ shafts and Darrieus connecting coupling, utilizing a segregated algorithm with a second-order
arms due to their negligible significance on the results (Maître et scheme for spatial and temporal discretization. The residuals’
al., 2013). convergence criteria for the continuity, x-velocity, y- velocity, k,
The 2D computational domain comprises a stationary domain and ω equations were set to 10− 6. During simulation, 30
and a rotational domain in which the Savonius and Darrieus iterations per timestep were configured, ensuring that residuals
turbines are positioned. The schematic representation of the dropped below 10− 6 in each time step. All simulations were
computational domain is depicted in Fig. 1(a). The stationary performed using parallel processing with 8 Intel Core i7-7700 CPU
computational domain dimensions were selected through (4.2 GHz) and 16 GB of RAM.
sensitivity analysis to assess their impact on the averaged power
coefficient. This analysis considered three key parameters: the
3.7. Convergence criterion
distance between the domain inlet and the turbine’s center (d i),
the distance between the turbine’s center and the domain outlet
In the realm of unsteady wind turbine simulations, achieving a
(do), and the blockage ratio (D/W), where ‘W’ represents the
steady flow state is paramount to ensure the extraction of
domain width and ‘D’ denotes the Darrieus turbine diameter. The
reliable data. One widely accepted methodology for achieving
results of the sensitivity analysis are shown in Fig. 2.
this stability is by monitoring the deviation in the averaged power
Observing Fig. 2(a), it’s evident that domains with inlet
coefficient value over a full revolution between two consecutive
distances of di = 5D and 7.5D tend to overestimate the power
cycles. This approach provides an indication of the simulation’s
coefficient. However, the Cp difference with di = 10D and 15D is

5
S. Chegini et al. Ocean Engineering 287 (2023) 115910
convergence and its readiness for accurate data extraction. For a
visual representation, Fig. 3 presents the convergence history of
the power coefficient over a turbine rotation.
From Fig. 3, the power coefficient difference between two
consecutive cycles drops to less than 1% after 7 revolutions and
further reduces to below 0.3% after 9 revolutions. The
convergence criteria adopted in the studies of Rossetti and Pavesi
(2013) and Raciti Castelli et al. (2011) were a Cp difference of 3%
and 1%, respectively, while Rezaeiha et al. (2017a) chose a 0.2%
value. Therefore, based on Fig. 3 and prior studies, data
extraction after 8 to 10 turbine rotations is a prudent choice,
offering a balance between accuracy and computational time.

3.8. Computational grid and mesh independence

Precise spatial discretization of the flow domain plays a


crucial role in achieving accurate CFD code solving. This entails
meticulous mesh refinement to minimize result errors. Fig. 3. Convergence history of Cp for 50 revolutions at λ = 2.5.
Subsequently, the meshing is then
deemed to be verified. Furthermore, the proper mesh structure
for the Darrieus VAWT should effectively capture the complex
flow behavior and high gradients near the airfoil, all while
maintaining reasonable computational cost. In this simulation,
domain discretization is performed by ANSYS-Mesh. While a
structured grid of quadrilateral elements is utilized in the
boundary layer to optimize near-wall precision, both the rotating
and stationary zones are composed of unstructured triangular
elements. Commencing from the airfoil surfaces, grid coarsening
is incrementally implemented towards the rotor-stator interface.
The sliding interface was grided by a non-conformal mesh with a
0.05 c sizing and a growth rate of 1.1 for both inner and outer
circles (Balduzzi et al., 2016a). This was meticulously designed to
progressively augment grid density, resulting in a significant

Fig. 2. Sensitivity analysis of computational domain: (a) Inlet distance effect, (b) Outlet distance effect, (c) Domain width effect.
enhancement of flux calculation precision. Fig. 4 provides a
detailed view of the mesh distribution around the blades and
rotors.
A mesh independence study is carried out for the Darrieus
rotor with the specifications outlined in Table 1 with four
different mesh refinement levels. The grid density around the
airfoils is the primary distinction between the mesh cases. Table
2 lists the specifics of the various grid properties. The coarsest
and finest grids are represented by Grid 1 and Grid 4,
respectively.
The characteristics of the inflation layer for different grids are
selected meticulously to ensure proper coverage of the airfoil’s
physical boundary layer within the inflation layers. The computed
maximum boundary layer’s thickness, with respect to the blade’s
Reynolds

6
S. Chegini et al. Ocean Engineering 287 (2023) 115910
Table 2 + ρwyuτ
Grid independence results. y= μw (10) where y denotes the normal distance
Grid features Grid 1 Grid 2 Grid 3
between the cell center and the wall, √ρw a̅̅̅̅̅̅̅̅̅̅̅ nd ̅ μw are density and
Number of elements ✕103 326 514 808
Number of nodes on airfoil 1580 1735 2170 fluid’s dynamic viscosity at the wall, uτ = τw/ρw is friction velocity
Average airfoil mesh size [mm] 0.1 0.09 0.07 at the wall, and τw = μ∂u/∂y|w corresponds to the wall shear
Airfoil growth rate 1.1 1.07 1.05 stress. The y + ranges for the boundary layer’s laminar sublayer
First layer thickness [mm] 0.0290 0.0203 0.0145
Inflation Growth rate 1.2 1.15 1.1 and buffer layer are as follows: 0< y+ <5 and 5< y+ <30,
Inflation layers 18 24 35
respectively. A y+~1 is crucial for the SST k- ω turbulence model to
Averaged y+ 1.08 0.77 0.56
Maximum y+ 2.9 2 1.05
adequately simulate the viscous sublayer (Marsh et al., 2017). To
Maximum skewness 0.82 0.77 0.79 ensure the sublayer is captured across all grid cases, the heights
Cp 0.460 0.481 0.498 of the first layers have been progressively decreased from Grid 1
number, is 1.9 mm. Note that due to large vorticity gradients to 4, as outlined in Table 3. In all cases, the inflation growth rate
associated with unstable operating conditions at low TSRs and airfoil grid size were correspondingly decreased to provide a
(Balduzzi et al., 2016b), the total height of the prism layers in the finer mesh. An average y+~1 is achieved in Grid 2, Grid 3, and Grid
tested grids was adjusted to 4 mm to ensure the capturing of this 4, while the maximum local y + values of nearly 1 which were
behavior. The comparative details of the mesh distribution at the near the leading edge are gained with Grid 3 and Grid 4.
airfoil’s leading edge for the four analyzed grids are displayed in Importantly, in all grids, the maximum local y + remains within
Fig. 5. the laminar sublayer.
The y + nondimensional parameter serves as the main Maximum skewness and cell aspect ratio metrics were used
determinant among the evaluated meshes. y + represents the to assess the grid quality. For Grid 3, the first element in the
dimensionless distance normal to the wall, scaling with the near-wall mesh,
boundary layer thickness as expressed in Eq. (10) (Lositano and
Danao, 2019˜ ).

Fig. 4. Mesh distribution details for (a) Rotating zone, (b) Around airfoil, (c) Trailing edge, (d) detailed trailing edge.

Fig. 5. Comparative view of mesh refinement near the leading edge.

7
S. Chegini et al. Ocean Engineering 287 (2023) 115910
Table 3
Courant numbers for various spatial and temporal discretization at min and
max TSR.

λ ω Δθ Δt (s) Δx (mm) (rad/ (deg) 0.1 0.09 0.07 0.05


s) (Grid (Grid (Grid (Grid 1) 2) 3) 4)
CFL
1.44 25.2 1 0.0007 90.8 100.9 129.7
181.6
0.5 0.00035 45.4 50.4 64.9
0.1 0.00007 9.1 10.1 13.0
0.05 0.000035 4.5 5.0 6.5
3.28 57.4 1 0.0003 88.7 98.5 126.7
177.4
0.5 0.00015 44.3 49.3 63.3
0.1 0.00003 8.9 9.9 12.7 17.7 0.05 0.000015 4.4 4.9 6.3 8.9

where the highest skewness and aspect ratio are present, the
maximum skewness and cell aspect ratio were 0.79 and 8.2,
respectively. Furthermore, beyond the inflation layers in the first
triangular cells, maximum skewness and cell aspect ratio were Fig. 6. Instantaneous torque coefficient vs. azimuth angle of one blade for
measured as 0.65 and 4.25, respectively, demonstrating excellent grid independence at λ = 2.5.
mesh quality and an appropriate airfoil grid size (Ahmad et al.).
After checking the grid quality, the mean power coefficient
and instantaneous torque coefficient variations for the four grids
at λ = 2.5 were compared. As shown in Table 3, as the total
number of elements increased, the difference in power
coefficient decreased so that the difference in Cp between Grid 3
and Grid 4 was 2%. In Fig. 6, the instantaneous torque coefficient
in the last rotation of blade 1 for different grids is represented in
relation to the azimuth angle.
As shown in Fig. 6, Grid 3 and Grid 4 comply properly, and the
most apparent difference between different grids occurs around
θ = 98◦, where the angle of attack notably exceeds the static stall
angle.
Consequently, according to the mesh study of the Darrieus
rotor, Grid 3 specifications and criteria are adopted with
considerations for subsequent simulations involving the hybrid
turbine and the hybrid turbine with a deflector. These
considerations include computational capacity, processing time,
and result accuracy. In addition, to ensure accurate results from
the hybrid turbine mesh, y + values as a distance along the y-axis
and the y + contour for the hybrid turbine containing both Fig. 7. Wall y + values for the hybrid wind turbine.
Darrieus and Savonius turbines, based on the adopted mesh are
shown in Figs. 7 and 8, respectively.
From Fig. 7, it can be inferred that the maximum y+ of 1.05
occurs on blade 3 for the Darrieus rotor, while the maximum y +
for the Savonius rotor is 0.3. Fig. 8 further demonstrates the
appropriateness of the adopted grid for the Savonius turbine.
Therefore, for the subsequent

Fig. 8. Wall y + contour around the blades.

simulations, that involve the hybrid turbine, the proposed grid


structure with 693,000 cells in the rotating zone and 309,000

8
S. Chegini et al. Ocean Engineering 287 (2023) 115910
cells in the stationary zone is selected. 3.9. Time step
independence

Due to the abrupt changes in the flow behavior around the


airfoil of the Darrieus turbine, particularly at low TSRs where high
vorticity gradients occur, temporal discretization, as well as
spatial discretization, is essential (Balduzzi et al., 2016b).
Therefore, a time step study is conducted for the Darrieus
turbine, and the final choice is applied for the following
simulation. The time step evaluation is performed using Courant-
Friedrichs-Lewy (CFL) criterion, introduced in Eq. (11) (Trivellato
and Raciti Castelli, 2014). uΔt
CFL =
Δx

where, u is the airfoil peripheral velocity, Δt is the time step, and


Δx is the average distance between two cell centroids on the
airfoil.
The courant number (CFL) represents the relationship between
the temporal time step (Δt) and the required time for a particle of
fluid with u velocity to be convected through a cell with Δx Fig. 9. Variation of torque coefficient in relation to time for different time
dimension (Balduzzi et al., 2016b). A courant number on the steps.
order of 10 for viscous turbomachinery flows can provide optimal
error damping (Amano and Sund´en, 2011). The courant number
Table 4
varies with different grid resolutions, time steps, and TSRs. In
Time step sensitivity analysis.
Table 3, variations of the courant number for four different time Δθ [deg] Δt [s] Cm
steps (Δt) and four studied grids at minimum and maximum tip
1 0.0004 0.1991
speed ratios are presented. The different time steps correspond
0.5 0.0002 0.1952
to azimuthal increments of Δθ = 1◦, 0.5◦, 0.1◦ and 0.05◦. 0.1 0.00004 0.1931
As evident from Table 3, the courant numbers are high and 0.05 0.00002 0.1915

undesirable for azimuthal increments of 1 ◦ and 0.5◦. Low courant 3.10. CFD model validation

numbers were achieved in Grids 1 and Grid 2 with Δ θ = 0.1 and
Following a comprehensive solution verification through the
0.05◦, as predicted by Eq. (11). However, simulation accuracy grid and time step sensitivity studies, the validation of the
degrades with increasing grid sizes. A courant number around 10 numerical modeling is essential to evaluate solution accuracy.
is achievable by Grid 3 with Δ θ = 0.1◦ at both the minimum and Consequently, simulation results for both Darrieus and Savonius
maximum tip speed ratios. Fig. 9 and Table 4 compare the torque rotors were independently verified using experimental data. Two
coefficient obtained using Grid 3 as a function of flow time for advantages of separate validation for each rotor are the
four time steps of Δt = 0.0004s, 0.0002s, 0.00004s, and 0.00002s, assurance of accurate results and the provision of data from the
operational tip speed ratio ranges, including their optimal values
corresponding to Δθ = 1◦, 0.5◦, 0.1◦, and 0.05◦, respectively at λ = that are effective in the hybrid rotor. The subsequent sections
2.5. discuss the validations in detail.
The torque coefficient variations for all time steps are similar,
as illustrated in Fig. 9. The torque coefficient peaks become more 3.10.1. Validation of darrieus rotor
closely spaced as the time step is decreased. For the validation of the numerical results of the Darrieus
According to Table 4, the mean torque coefficients for various turbine, the obtained CFD results were compared with both
time steps are nearly identical, with a relative difference of 0.8% empirical and numerical data provided by Raciti Castelli et al.
between Δ θ = 0.1◦ and 0.05. As a result, Grid 3 with a time step (2010b), which studied a Darrieus turbine in the wind tunnel with
that corresponds to azimuthal increments of 0.1 ◦ is an adequate a freestream velocity of 9 m/s. The features of the Darrieus rotor
and reliable selection for all tip speed ratios, offering an utilized for validation are presented in Table 1. Note that the
appropriate simulation time. wind tunnel blockage was not considered in the experimental
data correction, and consequently, this factor was not considered
in the present simulation. The comparison was conducted based
on the Cp across 8 TSRs ranging from 1.45 to 3.3. Notably, the λ =
1.45 is corresponding to a rotational speed of 25.2 rad/s.
The power coefficient comparison in Fig. 10 exhibits good
agreement with the experimental data across all TSRs. The 2D
CFD simulation reproduces the experimental Cp peak at λ = 2.65.
Despite the observed deviations, the analysis results consistently
follow the trend of the experimental results with a nearly
constant discrepancy. Notably, these observed discrepancies and
errors in the 2D model’s results when compared to the
experimental results can be attributed to inherent limitations

9
S. Chegini et al. Ocean Engineering 287 (2023) 115910
associated with 2D modeling, where the blade tip vortices, which
contribute to losses are not considered. Limitation of RANS
modeling especially at low TSRs, where complex dynamic stall
occurs

RANS model and other simplifications discussed in the preceding


section may have contributed to the errors between results.
Finally, two independent validations were conducted to
guarantee the accuracy of the 2D CFD code for hybrid turbine
simulation and to indicate the rotors’ operational conditions.
Fig. 10. Comparison of Darrieus power coefficient from the current 2D Additionally, according to the present work and several studies in
study with experimental and numerical data. the literature review, the 2D model demonstrates the capability
to accurately simulate vertical-axis wind turbines.
justifies overestimation in power coefficient (Rezaeiha et al.,
2017b). Moreover, geometrical simplifications, like omitting 4. Results and discussion
supporting radial arms and connections, lead to an overall
reduction in calculated drag forces and an increase in total power 4.1. Hybrid wind turbine
estimation. Additionally, the lack of specific wind tunnel
characteristics such as turbulence intensity and length scale in The Darrieus wind rotor has been noted to suffer from poor
the experimental literature adds to the complexity of the self- starting ability. Researchers have provided various
comparison. explanations for this issue. There is a zone referred to as the dead
band at low tip speed ratios (0.5–1.5), where the turbine’s
3.10.2. Validation of savonius rotor generated torque is insufficient or even negative, weakening the
The validation of the Savonius rotor results was performed by Darrieus turbine’s self-starting capability (Baker, 1983).
comparing the power coefficient at 4 different TSRs with the Contrarily, the Savonius turbine does not experience the difficulty
experimental results reported by Wenehenubun et al. (2015b). of self-starting since it is drag-driven, and it can always accelerate
The geometric features of the Savonius rotor are listed in Table 1. to the desired TSRs to provide high torque at low TSRs. The start-
The simulations were performed across TSRs ranging from 0.4 to up issue of the Darrieus turbine can be resolved by combining it
0.515, where the experimental peak Cp was observed at λ = 0.47, with the Savonius rotor to produce the necessary torque for self-
starting. This section evaluates the effectiveness of the hybrid
corresponding to U∞ = 10 m/ s and ω = 27.7 rad/s. Fig. 11 presents wind turbine. The power coefficients at various TSRs, as
a comparison of the Cp trends between the experimental illustrated in Fig. 12, are employed to assess the performance of
Savonius rotor results and the present numerical investigation. the hybrid Darrieus-Savonius and Darrieus rotors.
Again, as indicated by Fig. 11, it can be inferred that there is a As illustrated in Fig. 12, the combination of Savonius and
good correlation between the numerical and experimental Darrieus turbines results in a 26.91% increase in Cp at the lowest
results. The 2D tip speed ratio. This enhancement significantly improves
Darrieus’s self-starting ability. The higher efficiency of the hybrid
turbine as compared to the Darrieus rotor at λ = 1.45 is attributed
to the greater torque generation capability of the Savonius
turbine at lower TSRs. This is because the Savonius turbine has a
larger wind-exposed surface area, allowing the hybrid rotor to
generate the necessary torque for self-starting (Roshan et al.,
2020). Tripathi et al. (2022) proposed a novel hybrid turbine to
boost the Darrieus turbine’s self-starting capability. This resulted
in an 8.39% increase in self-starting ability at a TSR of 1.4 for the
novel hybrid wind turbine compared to the conventional hybrid
turbine. Additionally, Pallotta et al. (2020) by employing
optimized blade designs for both Darrieus and Savonius turbines,
propose a hybrid VAWT that enhances the Cp of the hybrid

10
S. Chegini et al. Ocean Engineering 287 (2023) 115910
turbine by 20% at tip speed ratios below 1.5. Based on both prior
research and the current study, it is clear that

Fig. 11. Comparison of Savonius power coefficient from present 2D study with experimental results.
Fig. 12. Power coefficient versus TSR for Darrieus turbine and hybrid turbine.
implementing a hybrid turbine can notably enhance self-starting demonstrates that with an increase in the azimuth angle from 0 ◦
capability. To elaborate further on the superior performance of to 90◦, the Cl of the Darrieus turbine exceeds that of the hybrid
the proposed hybrid wind turbine at low TSR compared to the turbine. Within the subsequent range from 90 ◦ to 270◦, the hybrid
Darrieus wind turbine, the torque coefficients of turbines in one turbine displays analogous behavior to the Darrieus turbine,
indicating a minimal effect of the
rotation at λ = 1.45 are depicted in Fig. 13.
As shown in Fig. 13, the utilization of the hybrid turbine leads
to a remarkable amplification of the torque coefficient when
compared to the Darrieus wind turbine at λ = 1.45. The hybrid
turbine achieves its maximum Cm of 0.34 at an azimuth angle of
77◦, whereas the Darrieus turbine reaches its peak torque
coefficient of 0.26 at θ = 74◦. This signifies a substantial 30%
augmentation in the maximum Cm for the hybrid turbine in
contrast to the Darrieus VAWT. Of particular significance is the
contrast in the zones of negative generated torque around
specific azimuth angles. The hybrid turbine exhibits fewer
negative torque regions, especially around azimuth angles of 0 ◦
and 240◦, in comparison to the Darrieus turbine. This reduction
contributes to a higher Cm for the hybrid turbine, consequently
enhancing its self-starting ability. Although the hybrid turbine
does present a greater area of negative torque around the
azimuth angle of 120◦, its overall averaged Cm remains 20% higher
than that of the Darrieus VAWT. In general, it is observable that
the hybrid turbine demonstrates a broader positive torque range
and a narrower negative torque range across nearly all angles.
This configuration significantly reduces the dead band extension Savonius VAWT on the efficiency of the Darrieus rotor. During the
for the hybrid turbine in comparison to the Darrieus rotor, azimuth angle range of 270◦–310◦, the hybrid turbine
thereby mitigating the adverse effect of the negative torque demonstrates an enhanced lift coefficient compared to the
generated by the Darrieus turbine (Asadi and Hassanzadeh, Darrieus turbine. However, after passing 310 ◦, the Cl of the hybrid
2021). turbine experiences a decline. Overall, the hybrid turbine
While the hybrid turbine excels in generating higher torque experiences a lift coefficient reduction at most angles due to the
than the Darrieus turbine at low TSRs, as illustrated in Fig. 12, its existence of the Savonius rotor and its limited aerodynamic
power coefficient decreases as the TSR rises (Pallotta et al., performance at high TSRs (Bhuyan and Biswas, 2014).
2020). This outcome leads to a 23% decrease in the hybrid Consequently, this contributes to a decrease in its power
turbine’s output power as compared to the Darrieus rotor at the coefficient relative to the Darrieus turbine.
optimal TSR of 2.6. To explore the hybrid turbine’s aerodynamic Fig. 14(b) reveals that the drag coefficient of blade 1 in the
performance at this optimal TSR, Fig. 14 presents a comparative hybrid rotor consistently remains lower compared to that of the
assessment of the lift (Cl) and drag coefficients (Cd) for blade 1 of Darrieus turbine across most angles. This reduction is most
both Darrieus and hybrid turbines throughout one complete
rotation. pronounced at θ = 270◦, when the blade is positioned behind the
As is commonly recognized, the Darrieus wind turbine utilizes Savonius turbine. However, the drag coefficient of blade 1 in the
the lift principle to induce rotational motion, and its performance hybrid rotor surpasses that of the Darrieus turbine when not
is optimized when the lift-to-drag ratio is maximized. Fig. 14(a) influenced by the presence of the Savonius turbine. This is

11
S. Chegini et al. Ocean Engineering 287 (2023) 115910
◦ ◦ azimuth angle of 0◦, the pressure distribution is nearly identical
particularly evident within the angle ranges of 180 –240 and
for both turbine types. With increasing azimuth angle from 45 ◦ to
310◦–360◦. 135◦, the pressure gradient between the lower and upper
To comprehend the cause for the variations in drag and lift surfaces of the Darrieus turbine’s blade is higher than the blade
coefficients between the hybrid and Darrieus turbines, a of the hybrid turbine. This phenomenon occurs because, within
meticulous analysis of the flow patterns near the airfoils has been these angles, the lower pressure side of the hybrid turbine’s
conducted. This analysis involves comparing pressure contours blade 1 experiences higher pressure due to the high pressure
for both turbine types at angles where the performance field produced by the presence of the Savonius rotor. As a result,
differences are most significant. This visualization analysis is this reduces the pressure variation across the blades of the
presented in Fig. 15, offering a comparative view of the flow hybrid turbine. Consequently, the favorable pressure difference
patterns for Darrieus turbine (Fig. 15(a)) and hybrid turbine (Fig. on the Darrieus rotor’s blades leads to enhanced lift force
generation and correspondingly, greater torque generation in
15(b)) at azimuthal angles of 0◦, 45◦, 90◦, 135◦, 225◦, and 285◦ at λ =
2.6. this turbine. From the angle of 225 ◦–285◦, the high- pressure field
Fig. 15 illustrates that coupling a Savonius with a Darrieus around the Savonius turbine becomes advantageous, aiding
turbine generally leads to a pressure gradient reduction along the blade 1 in producing a pressure difference that generates more
blade as compared to the Darrieus turbine operating alone. At an lift force compared to the Darrieus turbine alone. This effect
reaches its peak

12
S. Chegini et al. Ocean Engineering 287 (2023) 115910

Fig. 13. Variation of torque coefficient over a complete rotation for Darrieus turbine and hybrid turbine at λ = 1.45.
at 285◦, as evident when referring to Fig. 14(a). The vorticity distribution surrounding the hybrid turbine at different azimuth angles at a
TSR of 2.6 is presented in Fig. 16.
As is evident, the Savonius turbine’s presence within the hybrid turbine creates vortices, which have the potential to adversely
impact the Darrieus turbine’s efficiency. Specifically, vortices form around the blade of the Savonius turbine between azimuth angles of
0◦ and 135◦, significantly diminishing its contribution to the resultant torque coefficient. Also, at an azimuth angle of 225 ◦ and 285◦, the
vorticity generated on the convex side of the Savonius turbine’s blade decreases, resulting in a higher drag coefficient at these azimuth
angles.
As mentioned earlier, although the hybrid turbine has superior self- starting ability, it is poor in performance at higher TSRs
compared to the Darrieus. To address this issue, deflectors have been incorporated in front of and next to the hybrid wind turbine to
enhance its overall performance.

4.2. Hybrid wind turbine with deflector

The rotation trajectory of a wind turbine is segmented into four distinct paths based on the azimuth angle, namely windward,
upwind, leeward, and downwind, as depicted in Fig. 1(c). Within the upwind region of the hybrid turbine, as observed in Fig. 17, the
generated torque
Fig. 14. Variation of (a) lift coefficient, and (b) drag coefficients throughout a complete rotation for blade 1 of the Darrieus and hybrid turbine at λ = 2.6.

Fig. 15. Pressure contour for a blade of (a) Darrieus turbine and (b) hybrid turbine at azimuth angles of 0 ◦, 45◦, 90◦, 135◦, 225◦, and 285◦.

13
S. Chegini et al. Ocean Engineering 287 (2023) 115910

Fig. 16. Vorticity field around the hybrid turbine at the angles of 0◦, 45◦, 90◦, 135◦, 225◦, and 240◦ for λ = 2.6.

turbine’s upwind region, while the side deflector has been


employed and investigated to redirect the wind towards the
leeward and partial of the downwind side, thereby enhancing the
turbine’s performance in these zones. Additionally, the
effectiveness of the hybrid turbine’s performance is also
examined by using double deflectors, which employ both
deflectors simultaneously. The Cp of a hybrid turbine equipped
with side, front, and double deflectors, as well as a bare Darrieus
turbine, in relation to the TSR is presented in Fig. 18.
The findings show that the implementation of a front deflector

Fig. 17. Variation of torque coefficient of blade 1 over a rotation of


Darrieus and hybrid turbines at λ = 2.6.

of Darrieus’s blade is significantly lower in magnitude when


contrasted with that of the Darrieus rotor at a TSR of 2.6. Also, it
has been determined that in the leeward and downwind regions
of both the Darrieus and hybrid turbine, the produced torque is
significantly lower compared to the upwind side due to the
dissipation of wind energy (Ferreira, 2009; Zidane et al., 2023).
In this part, to enhance the generated power of the hybrid
turbine in the upwind region, leeward, and downwind regions, Fig. 18. Power coefficient as a function of TSR for bare Darrieus turbine,
front and side deflectors are introduced, respectively. The front bare hybrid turbine, and hybrid turbine with front, side, and double
deflector aims to increase the power generated in the hybrid deflectors.

14
S. Chegini et al. Ocean Engineering 287 (2023) 115910
increased the maximum Cp of the hybrid wind turbine by 30% at As previously stated, the deflector has proven to significantly
the optimal TSR. However, this deflector can weaken the self- enhance the torque coefficient, particularly in the case of the
starting capability of a hybrid turbine by 40%. On the other hand, double deflector. Fig. 20 illustrates that within the azimuth angle
the hybrid turbine with a side deflector exhibited good
range of 30◦–60◦, the hybrid turbine generates more torque than
performance at a low TSR, with also significant 26%
the hybrid turbine with a front deflector. This observation aligns
augmentation in the Cp of the hybrid turbine at optimal TSR. The
with Fig. 19(b), which demonstrates that the presence of a low-
most optimal result is achieved with a double deflector, raising
speed zone caused by the deflector curtails wind velocity in that
the hybrid turbine’s Cp by 55% and 17.36% at a TSR of 2.6,
compared to the hybrid and bare Darrieus turbines, respectively. region. A similar phenomenon occurs within the 300 ◦–360◦ range.
It is noteworthy that the double deflectors, in contrast to the Additionally, the front deflector contributes to an increased
other cases, yielded their maximum Cp at a TSR of 3.07, leading to torque coefficient between 60◦ and 140◦ compared to the hybrid
an extension of the turbine’s operational range (Pallotta et al., turbine. For the case of the side deflector, up to
2020; Zidane et al., 2023). The double deflectors also enhanced
the hybrid turbine’s self-starting ability with a 10% increase in Cp.
Indeed, employing a double deflector enhances the hybrid
turbine’s self-starting capability and significantly improves the
performance of the Darrieus rotor at high TSR. Eshagh Nimvari et
al. (2020) introduced an innovative deflector arrangement
positioned in front of the Savonius rotor, resulting in a 10%
enhancement in the rotor’s maximum Cp. Qasemi and Azadani
(2020) investigated how a flat plate deflector impacts the
performance of a Darrieus rotor and the result indicated that it
can improve the Cp of the rotor by 16.42% as compared to the
bare Darrieus rotor. In contrast to the aforementioned studies,
the current research yields a more favorable result by employing
a novel double deflector, which achieves a superior outcome
with an improvement in Cp of 17.36%, and 55% compared to the
bare Darrieus and hybrid turbines, respectively.
To undertake a thorough analysis of the hybrid turbine with
deflectors performance, it is imperative to conduct a detailed
investigation of the flow aerodynamics behavior surrounding the
turbine. In this regard, Fig. 19 presents the velocity distribution
contours at an azimuth angle of 0◦ for the hybrid turbine with
front, side, and double deflectors, as well as the hybrid turbine
without a deflector.
It is observable that utilizing a deflector can increase wind
speed, consequently resulting in a power output enhancement.
This can be attributed to the reduction of wind velocity upon
colliding with the deflector, causing the establishment of a zone
with a low velocity. In the vicinity of this area, the wind flow that
has been diverted from the upper and lower edges of the
deflector moves at a velocity that is 20% greater than that of the
incoming wind (Wong et al., 2018a). Also, the interaction
between the deflected and oncoming flow creates a non-
horizontal flow. This angled flow can decrease the adverse
impact of the wake produced by the upwind region of the rotor,
resulting in higher power output (Orlandi et al., 2015). The
utilization of the front deflector results in the formation of a
region with high wind speed in the hybrid turbine’s upwind
section. This is achieved by redirecting the wind that passes along
the lower edge of the front deflector. Similarly, through the
redirection of the wind departing from the upper edge of the side
deflector, a high-speed region is formed in the leeward section
and a portion of the downwind section relative to the hybrid
turbine. The dual benefits of both deflectors are concurrently
harnessed by the double deflector. As a result, by utilizing the
double deflector, the upwind, leeward, and downwind regions
exhibit higher wind speeds compared to the hybrid turbine,
thereby enhancing the turbine’s overall performance in these
zones.
Fig. 20 illustrates the torque coefficient variation in relation to
the azimuthal angle for a hybrid turbine operating at λ = 2.6, both
with no deflectors and with front, side, and double deflectors.

15
S. Chegini et al. Ocean Engineering 287 (2023) 115910

Fig. 19. Velocity contours for the hybrid turbine (a) without deflector, (b) with front deflector, (c) with side deflector, and (d) with double deflector.
Fig. 20. Variation of torque coefficient of blade 1 over a rotation of the Fig. 21(b), as the azimuth angle increases to 45 ◦, the pressure
Darrieus turbine, hybrid turbine, and hybrid with front, side, and double gradient between the lower and upper surface of the hybrid
deflector at λ = 2.6. turbine’s blade with a front deflector decreases. This leads to
reduced lift force and a decrease in torque coefficient, which is
an azimuth angle of 80 ◦, the trend of Cm for the hybrid turbine consistent with the findings in Fig. 20. Moving to an azimuth
with the side deflector remains the same with the bare hybrid angle of 90◦, a significant variation in the local pressure field is
turbine. This indicates that the side deflector does not affect the observed, particularly on the airfoil’s suction surface in the hybrid
hybrid turbine in this area. Subsequently, the side deflector by turbine with the front deflector. This variation contributes to
increasing the wind velocity improves the hybrid turbine’s increased torque generation.
The pressure contour around blade 1 in the bare hybrid
performance from the azimuth angle of 80 ◦–240◦, which
turbine and the hybrid turbine equipped with a side deflector at
encompasses the leeward and partial downwind regions. The
double deflector enhances torque generation and overall λ = 2.6, for azimuth angles of 180 ◦, 240◦, and 270◦ in the leeward
performance in the upwind areas (akin to the front deflector) and and windward regions, is
the leeward and partial downwind regions (similar to the side
deflector).
Further examinations were focused only on the front and side
deflectors, as it was observed that the double deflector
performed like a front deflector between 0 ◦ and 120◦ angles and
like a side deflector between 120◦ and 360◦ angles. The pressure
distribution around blade 1 in the hybrid turbine without a
deflector and the hybrid turbine equipped with a front deflector,
both operating at λ = 2.6, at azimuth angles of 0 ◦, 45◦ and 90◦ in
the upwind region is illustrated in Fig. 21.
At an azimuth angle of 0◦, the pressure distribution around
the hybrid turbine’s airfoil illustrates a high-pressure area at the
leading edge, resulting in an increased drag force and a reduced
lift-to-drag ratio, thereby decreasing the torque generation. In

16
S. Chegini et al. Ocean Engineering 287 (2023) 115910

Fig. 21. Pressure contours for blade 1 of (a) hybrid turbine, and (b) hybrid turbine with front deflector at θ = 0◦, 45◦, and 90◦.
illustrated in Fig. 22.
At azimuth angles of 180◦ and 240◦, the hybrid turbine with a
side deflector exhibits a higher pressure difference exists
between the airfoil’s lower and upper surfaces compared to the
bare hybrid turbine. This results in a higher efficiency of the
hybrid turbine equipped with a side deflector, compared to the
bare hybrid turbine, as shown in Fig. 20. At the azimuth angle of
270◦, the differential in pressure between the airfoil’s lower and
upper surfaces becomes more significant for the hybrid turbine
without a deflector.
To assess the airflow behavior around the hybrid turbine with
the front deflector, the vorticity distributions at θ = 180◦, and 240◦
are shown in Fig. 23.
Contrary to the bare hybrid wind turbine, blade 1 of the
hybrid turbine with a front deflector experiences boundary layer
separation at an azimuth angle of 45 ◦. This indicates that the
airflow near the blade surface loses its momentum and comes to
a stop, causing flow separation (Simpson, 1989). An adverse
pressure gradient within the boundary layer slows down the flow
and reverses its direction. At the same time, at an azimuth angle
of 165◦, blade 2 experiences a vortical disturbance on its surface,
and the airflow separates from the blade surface, leading to the
vortex forming in the wake of the blade. Furthermore, at θ = 90◦,
blade 1 in Fig. 23(b) experiences intense flow separation on its
surface compared to the bare hybrid turbine’s blade. Flow
aerodynamics significantly influence the hybrid turbine’s
efficiency, evident from the declining Cm at azimuth angles of 45 ◦
and 90◦. Although the torque generated by the front deflector at
90◦ is higher than that of the hybrid turbine, it experiences a
significant decline due to increased flow separation at this
azimuth angle. Furthermore, at an azimuth angle of 165 ◦, the
turbine’s torque output diminishes as a result of dynamic stalls
occurring on the airfoil.
To investigate the airflow behavior around the hybrid turbine
with a side deflector which presents a higher Cm in the leeward
and downwind region in comparison to the hybrid turbine
without a deflector, the vorticity contours at the angles of 180 ◦,
and 240◦ are shown in Fig. 24.
As can be seen, both hybrid VAWT with the side deflector and
hybrid turbine without deflector, exhibit similar behavior for the
blade at an angle of 180◦. However, for blade 1 in Fig. 24 (b) at θ =
240◦, an augmented boundary layer separation is observed in
contrast to the hybrid turbine without a deflector. This
phenomenon is supported by the decline of Cm at θ = 240◦, as

17
S. Chegini et al. Ocean Engineering 287 (2023) 115910
demonstrated in Fig. 20, which is attributable to the start of flow aerodynamic performance of this hybrid turbine. The findings
separation. indicate that the proposed model effectively resolves the main
challenge of self-starting encountered by the Darrieus VAWT.
5. Conclusion Through the integration of the Savonius rotor, the hybrid
turbine’s efficiency at λ = 1.44 can be improved by 26.91%,
To enhance the self-starting ability of the Darrieus Vertical
thereby enhancing its self-starting ability. However, at the
Axis Wind Turbine, a hybrid turbine configuration combining a 2-
optimum TSR, the hybrid turbine’s Cp decreases by 23% compared
blade Savonius turbine and a 3-blade Darrieus turbine has been
to the Darrieus turbine. To address this issue, flat plate deflectors
proposed. A 2D computational fluid dynamics approach utilizing
placed in the front,
the SST k-ω turbulence model was adopted to investigate the

Fig. 22. Pressure contour for blade 1 of (a) hybrid turbine, and (b) hybrid turbine with side deflector at θ = 180◦, 240◦, and 270◦.

Fig. 23. Vorticity field around the (a) hybrid turbine without deflector, and (b) hybrid turbine with front deflector at θ = 180◦, and 240◦.

18
S. Chegini et al. Ocean Engineering 287 (2023) 115910

Fig. 24. Vorticity field around the (a) hybrid turbine without deflector, and (b) hybrid turbine with side deflector at θ = 180◦, and 240◦.
side, or on both sides, referred to as double deflectors, are economic efficiency and the structure’s lifetime, crucial in
simulated to enhance the hybrid turbine’s efficiency in the practical terms. Consequently, the focus of future experimental
upwind, leeward, and downwind regions. The results study is to consider the hybrid turbine structure, solve the fixed
demonstrate: deflector limitation through the exploration of passive deflectors
coupled with the shaft to the turbine, and conduct an economic
- The inclusion of a deflector in front of a hybrid turbine can analysis for the mentioned hybrid turbine system.
lead to a 30% increase in its Cp at the optimal TSR.
Furthermore, the use of a front deflector can result in higher CRediT authorship contribution statement
torque production in the turbine’s upwind region compared
to a hybrid turbine without the deflector. However, the Sahel Chegini: Conceptualization, Methodology, Software,
hybrid turbine’s self-starting ability with the front deflector Formal analysis, Investigation, Data curation, Writing – original
can be significantly reduced by 24% when compared to a draft, Writing – review & editing, Visualization. Mohammadreza
Darrieus rotor. Asadbeigi: Conceptualization, Methodology, Software,
- By incorporating a side deflector in a hybrid turbine, a notable Validation, Formal analysis, Investigation, Writing – original draft,
improvement in its performance in the leeward and partial Writing – review & editing. Farzad Ghafoorian: Writing – original
downwind regions can be attained. The inclusion of this draft, Writing – review & editing, Resources. Mehdi Mehrpooya:
deflector next to the hybrid turbine can lead to a substantial Methodology, Supervision, Writing – review & editing,
increase of 26% in its coefficient of power at λ = 2.6. Resources, Project administration, Funding acquisition.
- The double deflector configuration appears to be the most
efficient in terms of improving the Cp of a hybrid turbine. This Declaration of competing interest
configuration can lead to a 40% and 17.63% increase at the
lowest and optimal tip speed ratio, respectively, compared to The authors declare that they have no known competing
a Darrieus turbine. Furthermore, this setup can substantially financial interests or personal relationships that could have
elevate the hybrid turbine’s Ct in the upwind, leeward, and appeared to influence the work reported in this paper.
partial downwind regions.
Data availability
Future work
All data have been reported in the paper.
The findings of this comprehensive numerical study clearly
demonstrate that the incorporation of double deflectors into a
References
hybrid wind turbine yields a substantial enhancement in output
power across all ranges of TSRs. This innovation stands as a Acarer, S., 2020. Peak lift-to-drag ratio enhancement of the DU12W262 airfoil by
practical solution for augmenting the performance of previously passive flow control and its impact on horizontal and vertical axis wind turbines.
Energy 201, 117659. https://doi.org/10.1016/j.energy.2020.117659. T. Ahmad, S.
low-performing Darrieus-Savonius hybrid wind turbines. In the
L. Plee, and J. P. Myers, “Fluent User’s Guide”.
current two-dimensional study, many parameters such as the Amano, R., Sunden, B., 2011. Computational Fluid Dynamics and Heat Transfer:´
effect of the existence of shaft and other components of the Emerging Topics, vol. 23. WIT Press.
turbines on the hybrid turbine’s aerodynamic performance have Ardaneh, F., Abdolahifar, A., Karimian, S.M.H., 2022. Numerical analysis of the
not been considered. Additionally, employing fixed deflectors in pitch angle effect on the performance improvement and flow characteristics
of the 3-PB Darrieus vertical axis wind turbine. Energy 239, 122339.
VAWTs confines turbine performance to specific wind directions, https://doi.org/10.1016/j. energy.2021.122339.
a concern addressed in prior research (Tian et al., 2022). It should Asadbeigi, M., Ghafoorian, F., Mehrpooya, M., Chegini, S., Jarrahian, A., 2023. A
be noted that the complexity of the hybrid turbine’s structure, 3D study of the darrieus wind turbine with auxiliary blades and economic
which includes deflectors, makes many parameters, such as

19
S. Chegini et al. Ocean Engineering 287 (2023) 115910
analysis based on an optimal design from a parametric investigation. Kok, J.C., 2000. Resolving the dependence on freestream values for the k-
Sustainability 15 (5), 4684. turbulence model. AIAA J. 38 (7), 1292–1295.
https://doi.org/10.3390/su15054684. https://doi.org/10.2514/2.1101.
Asadi, M., Hassanzadeh, R., 2021. Effects of internal rotor parameters on the Kyozuka, Y., 2008. An experimental study on the darrieus-savonius turbine for
performance of a two bladed Darrieus-two bladed Savonius hybrid wind the tidal current power generation. JFST 3 (3), 439–449.
turbine. https://doi.org/10.1299/jfst.3.439.
Energy Convers. Manag. 238, 114109 https://doi.org/10.1016/j. Layeghmand, K., Ghiasi Tabari, N., Zarkesh, M., 2020. Improving efficiency of
enconman.2021.114109. Savonius wind turbine by means of an airfoil-shaped deflector. J. Braz. Soc.
Asadi, M., Hassanzadeh, R., 2022. On the application of semicircular and Bach- Mech. Sci. Eng. 42 (10), 528. https://doi.org/10.1007/s40430-020-02598-7.
type blades in the internal Savonius rotor of a hybrid wind turbine system. J. Leonczuk Minetto, R.A., Paraschivoiu, M., 2020. Simulation based analysis of
Wind Eng. Ind. Aerod. 221, 104903 morphing blades applied to a vertical axis wind turbine. Energy 202, 117705.
https://doi.org/10.1016/j.jweia.2022.104903. https://doi.org/ 10.1016/j.energy.2020.117705.
Baker, J.R., 1983. Features to aid or enable self starting of fixed pitch low solidity Liang, X., Fu, S., Ou, B., Wu, C., Chao, C.Y.H., Pi, K., 2017. A computational study
vertical axis wind turbines. J. Wind Eng. Ind. Aerod. 15 (1–3), 369–380. of the effects of the radius ratio and attachment angle on the performance
https://doi.org/ 10.1016/0167-6105(83)90206-4. of a Darrieus- Savonius combined wind turbine. Renew. Energy 113, 329–
Balduzzi, F., Bianchini, A., Maleci, R., Ferrara, G., Ferrari, L., 2016a. Critical issues 334. https://doi.org/ 10.1016/j.renene.2017.04.071.
in the CFD simulation of Darrieus wind turbines. Renew. Energy 85, 419– Lositano, I.C.M., Danao, L.A.M., 2019. Steady wind performance of a 5 kW three-
435. https://doi. org/10.1016/j.renene.2015.06.048. bladed ˜ H-rotor Darrieus Vertical Axis Wind Turbine (VAWT) with cambered
Balduzzi, F., Bianchini, A., Ferrara, G., Ferrari, L., 2016b. Dimensionless numbers tubercle leading edge (TLE) blades. Energy 175, 278–291.
for the assessment of mesh and timestep requirements in CFD simulations https://doi.org/10.1016/j.
of Darrieus wind turbines. Energy 97, 246–261. energy.2019.03.033.
https://doi.org/10.1016/j.energy.2015.12.111. Maître, T., Amet, E., Pellone, C., 2013. Modeling of the flow in a Darrieus water
Bardina, J.E., Huang, P.G., Coakley, T.J., 1997. Turbulence Modeling Validation. turbine: wall grid refinement analysis and comparison with experiments.
and development, testing. Renew. Energy 51, 497–512. https://doi.org/10.1016/j.renene.2012.09.030.
Beri, H., Yao, Y., 2011. Effect of camber airfoil on self starting of vertical Axis wind Marsh, P., Ranmuthugala, D., Penesis, I., Thomas, G., 2017. The influence of
turbine. J. of Environmental Science and Technology 4 (3), 302–312. turbulence model and two and three-dimensional domain selection on the
https://doi. org/10.3923/jest.2011.302.312. simulated performance characteristics of vertical axis tidal turbines. Renew.
Bhuyan, S., Biswas, A., 2014. Investigations on self-starting and performance Energy 105, 106–116. https://doi.org/10.1016/j.renene.2016.11.063.
characteristics of simple H and hybrid H-Savonius vertical axis wind rotors. Mehrpooya, M., Asadbeigi, M., Ghafoorian, F., Farajyar, S., 2023. Investigation
Energy Convers. Manag. 87, 859–867. and optimization on effective parameters of a H-rotor darrieus wind turbine,
https://doi.org/10.1016/j.enconman.2014.07.056. using CFD method. Iran. J. Chem. Chem. Eng. (Int. Engl. Ed.).
Bianchini, A., Balduzzi, F., Bachant, P., Ferrara, G., Ferrari, L., 2017. Effectiveness https://doi.org/10.30492/ ijcce.2023.562396.5610.
of two-dimensional CFD simulations for Darrieus VAWTs: a combined Menter, F.R., 2009. Review of the shear-stress transport turbulence model
numerical and experimental assessment. Energy Convers. Manag. 136, 318– experience from an industrial perspective. Int. J. Comput. Fluid Dynam. 23
328. https://doi.org/ 10.1016/j.enconman.2017.01.026. (4), 305–316. https://doi.org/10.1080/10618560902773387.
Celik, Y., Ingham, D., Ma, L., Pourkashanian, M., 2022. Design and aerodynamic Menter, F.R., Langtry, R., Volker, S., Huang, P.G., 2005. Transition modelling for
performance analyses of the self-starting H-type VAWT having J-shaped general ¨ purpose CFD codes. In: Engineering Turbulence Modelling and
aerofoils considering various design parameters using CFD. Energy 251, Experiments, vol. 6.
123881. https://doi. org/10.1016/j.energy.2022.123881. Elsevier, pp. 31–48. https://doi.org/10.1016/B978-008044544-1/50003-0.
Chen, W.-H., Wang, J.-S., Chang, M.-H., Mutuku, J.K., Hoang, A.T., 2021. Efficiency Mohamed, M.H., 2013. Impacts of solidity and hybrid system in small wind
improvement of a vertical-axis wind turbine using a deflector optimized by turbines performance. Energy 57, 495–504.
Taguchi approach with modified additive method. Energy Convers. Manag. https://doi.org/10.1016/j.energy.2013.06.004.
245, 114609 https://doi.org/10.1016/j.enconman.2021.114609. Orlandi, A., Collu, M., Zanforlin, S., Shires, A., 2015. 3D URANS analysis of a
Chowdhury, A.M., Akimoto, H., Hara, Y., 2016. Comparative CFD analysis of vertical axis wind turbine in skewed flows. J. Wind Eng. Ind. Aerod. 147, 77–
Vertical Axis Wind Turbine in upright and tilted configuration. Renew. 84. https://doi.org/ 10.1016/j.jweia.2015.09.010.
Energy 85, 327–337. https://doi.org/10.1016/j.renene.2015.06.037. Pallotta, A., Pietrogiacomi, D., Romano, G.P., 2020. Hybri – a combined Savonius-
Didane, D.H., Rosly, N., Zulkafli, M.F., Shamsudin, S.S., 2019. Numerical Darrieus wind turbine: performances and flow fields. Energy 191, 116433.
investigation of a novel contra-rotating vertical axis wind turbine. Sustain. https:// doi.org/10.1016/j.energy.2019.116433.
Energy Technol. Qasemi, K., Azadani, L.N., 2020. Optimization of the power output of a vertical
Assessments 31, 43–53. https://doi.org/10.1016/j.seta.2018.11.006. axis wind turbine augmented with a flat plate deflector. Energy 202, 117745.
Dixon, S.L., Hall, C.A., 2010. Fluid Mechanics and Thermodynamics of https://doi.org/ 10.1016/j.energy.2020.117745.
Turbomachinery, sixth ed. Butterworth-Heinemann/Elsevier, Burlington, MA. Raciti Castelli, M., Ardizzon, G., Battisti, L., Benini, E., Pavesi, G., 2010a. Modeling
Elkhoury, M., Kiwata, T., Aoun, E., 2015. Experimental and numerical strategy and numerical validation for a darrieus vertical Axis micro-wind
investigation of a three-dimensional vertical-axis wind turbine with variable- turbine. In:
pitch. J. Wind Eng. Ind. Fluid Flow, Heat Transfer and Thermal Systems, Parts A and B, Vancouver, vol.
Aerod. 139, 111–123. https://doi.org/10.1016/j.jweia.2015.01.004. 7.
Eshagh Nimvari, M., Fatahian, H., Fatahian, E., 2020. Performance improvement ASMEDC, Canada, pp. 409–418. https://doi.org/10.1115/IMECE2010-39548.
of a Savonius vertical axis wind turbine using a porous deflector. Energy British Columbia.
Convers. Manag. 220, 113062 Raciti Castelli, M., Ardizzon, G., Battisti, L., Benini, E., Pavesi, G., 2010b. Modeling
https://doi.org/10.1016/j.enconman.2020.113062. strategy and numerical validation for a darrieus vertical Axis micro-wind
Ferreira, C.S., 2009. The Near Wake of the VAWT-2D and 3D Views of the VAWT turbine. In: Fluid Flow, Heat Transfer and Thermal Systems, Parts A and B,
Aerodynamics. Vancouver, vol. 7.
Golecha, K., Eldho, T.I., Prabhu, S.V., 2011. Influence of the deflector plate on the ASMEDC, Jan., Canada, pp. 409–418. https://doi.org/10.1115/IMECE2010-
performance of modified Savonius water turbine. Appl. Energy 88 (9), 3207– 39548. British Columbia.
3217. https://doi.org/10.1016/j.apenergy.2011.03.025. Raciti Castelli, M., Englaro, A., Benini, E., 2011. The Darrieus wind turbine:
Hand, B., Cashman, A., 2020. A review on the historical development of the lift-
proposal for a new performance prediction model based on CFD. Energy 36
type vertical axis wind turbine: from onshore to offshore floating
(8), 4919–4934. https://doi.org/10.1016/j.energy.2011.05.036.
application. Sustain.
Rezaeiha, A., Kalkman, I., Blocken, B., 2017a. CFD simulation of a vertical axis
Energy Technol. Assessments 38, 100646. https://doi.org/10.1016/j.
wind turbine operating at a moderate tip speed ratio: guidelines for
seta.2020.100646.
minimum domain size and azimuthal increment. Renew. Energy 107, 373–
Hosseini, A., Goudarzi, N., 2019. Design and CFD study of a hybrid vertical-axis 385. https://doi.org/10.1016/j. renene.2017.02.006.
wind turbine by employing a combined Bach-type and H-Darrieus rotor
Rezaeiha, A., Kalkman, I., Blocken, B., 2017b. Effect of pitch angle on power
systems. Energy Convers. Manag. 189, 49–59.
performance and aerodynamics of a vertical axis wind turbine. Appl. Energy
https://doi.org/10.1016/j.enconman.2019.03.068.
197, 132–150. https://doi.org/10.1016/j.apenergy.2017.03.128.
Ibrahim, K., Djanali, V.S., Ikhwan, N., 2020. Numerical study of bach-bladed
Roshan, A., Sagharichi, A., Maghrebi, M.J., 2020. Nondimensional parameters’
Savonius wind turbine with varying blade shape factor. JMES Int. Journal of
effects on hybrid darrieus–savonius wind turbine performance. J. Energy
Mech. Eng. Sc. 4 (2), 12. https://doi.org/10.12962/j25807471.v4i2.7839.
Resour. Technol. 142 (1), 011202 https://doi.org/10.1115/1.4044517.
Jin, X., Wang, Y., Ju, W., He, J., Xie, S., 2018. Investigation into parameter
Rossetti, A., Pavesi, G., 2013. Comparison of different numerical approaches to
influence of upstream deflector on vertical axis wind turbines output power
the study of the H-Darrieus turbines start-up. Renew. Energy 50, 7–19.
via three- dimensional CFD simulation. Renew. Energy 115, 41–53.
https://doi.org/ 10.1016/j.renene.2012.06.025.
https://doi.org/10.1016/j. renene.2017.08.012.

20
S. Chegini et al. Ocean Engineering 287 (2023) 115910
Saad, A.S., El-Sharkawy, I.I., Ookawara, S., Ahmed, M., 2020. Performance
enhancement of twisted-bladed Savonius vertical axis wind turbines. Energy
Convers. Manag. 209, 112673
https://doi.org/10.1016/j.enconman.2020.112673.
Sadorsky, P., 2021. Wind energy for sustainable development: driving factors and
future outlook. J. Clean. Prod. 289, 125779 https://doi.org/10.1016/j.
jclepro.2020.125779.
Shah, O.R., Jamal, M.A., Khan, T.I., Qazi, U.W., 2022. Experimental and numerical
evaluation of performance of a variable pitch vertical-Axis wind turbine. J.
Energy Resour. Technol. 144 (6), 061303 https://doi.org/10.1115/1.4051896.
Shouman, M.R., Helal, M.M., 2023. Numerical investigation of improvement of
counter rotating Savonius turbines performance with curtaining and fin
addition on blade. Alex. Eng. J. 75, 233–242.
https://doi.org/10.1016/j.aej.2023.05.002.
Simpson, R.L., 1989. Turbulent boundary-layer separation. Annu. Rev. Fluid
Mech. 21
(1), 205–232. https://doi.org/10.1146/annurev.fl.21.010189.001225.
Sobhani, E., Ghaffari, M., Maghrebi, M.J., 2017. Numerical investigation of
dimple effects on darrieus vertical axis wind turbine. Energy 133, 231–241.
https://doi.org/ 10.1016/j.energy.2017.05.105.
Syawitri, T.P., Yao, Y., Yao, J., Chandra, B., 2022. Optimisation of straight plate
upstream deflector for the performance enhancement of vertical axis wind
turbine at low, medium and high regimes of tip speed ratios. Wind Eng. 46
(5), 1487–1510. https://doi.org/10.1177/0309524X221084980.
Tescione, G., Ragni, D., He, C., Simao Ferreira, C.J., van Bussel, G.J.W., 2014. Near
wake ˜ flow analysis of a vertical axis wind turbine by stereoscopic particle
image velocimetry. Renew. Energy 70, 47–61. https://doi.org/10.1016/j.
renene.2014.02.042.
Tian, W., Bian, J., Yang, G., Ni, X., Mao, Z., 2022. Influence of a passive upstream
deflector on the performance of the Savonius wind turbine. Energy Rep. 8,
7488–7499. https://doi.org/10.1016/j.egyr.2022.05.244.
Tripathi, A., Das, P., Aggarwal, T., Sahil, 2022. Efficiency enhancement of a hybrid
Vertical axis wind turbine by utilizing Optimum parameters. Mater. Today:
Proc. 62, 3582–3588. https://doi.org/10.1016/j.matpr.2022.04.406.
Trivellato, F., Raciti Castelli, M., 2014. On the Courant–Friedrichs–Lewy criterion
of rotating grids in 2D vertical-axis wind turbine analysis. Renew. Energy 62,
53–62. https://doi.org/10.1016/j.renene.2013.06.022.
Wenehenubun, F., Saputra, A., Sutanto, H., 2015a. An experimental study on the
performance of Savonius wind turbines related with the number of blades.
Energy
Proc. 68, 297–304. https://doi.org/10.1016/j.egypro.2015.03.259.
Wenehenubun, F., Saputra, A., Sutanto, H., 2015b. An experimental study on the
performance of Savonius wind turbines related with the number of blades.
Energy Proc. 68, 297–304. https://doi.org/10.1016/j.egypro.2015.03.259.
Wilcox, D.C., 1998. Turbulence Modeling for CFD, vol. 2. DCW industries La
Canada, CA.
Wong, K.H., Chong, W.T., Poh, S.C., Shiah, Y.-C., Sukiman, N.L., Wang, C.-T.,
2018a. 3D CFD simulation and parametric study of a flat plate deflector for
vertical axis wind turbine. Renew. Energy 129, 32–55.
https://doi.org/10.1016/j.renene.2018.05.085.
Wong, K.H., et al., 2018b. Experimental and simulation investigation into the
effects of a flat plate deflector on vertical axis wind turbine. Energy Convers.
Manag. 160, 109–125. https://doi.org/10.1016/j.enconman.2018.01.029.
Zidane, I.F., Ali, H.M., Swadener, G., Eldrainy, Y.A., Shehata, A.I., 2023. Effect of
upstream deflector utilization on H-Darrieus wind turbine performance: an
optimization study. Alex. Eng. J. 63, 175–189. https://doi.org/10.1016/j.
aej.2022.07.052.

21

You might also like