Next Article in Journal
Two New Pseudochromadora Species (Nematoda: Desmodorida) from South Korea Based on Morphological and Molecular Evidence
Previous Article in Journal
Climate Change Facilitates the Formation of Natural Barriers in Low-Inflow Estuaries, Altering Environmental Conditions and Faunal Assemblages
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Numerical Investigation of Hybrid Darrieus/Savonius Vertical Axis Wind Turbine Subjected to Turbulent Airflows

by
Rhuandrei Gabriel da Silva Inácio
1,
Igor Almeida da Rosa
1,
Vinicius Heidtmann Avila
1,
Luiz Alberto Oliveira Rocha
2,3,
Liércio André Isoldi
1,
Gustavo da Cunha Dias
1,
Rafael Adriano Alves Camargo Gonçalves
1 and
Elizaldo Domingues dos Santos
1,3,*
1
School of Engineering, Federal University of Rio Grande, Italia Avenue, km 8, Rio Grande 96203-900, RS, Brazil
2
Graduate Program of Mechanical Engineering, Federal University of Rio Grande do Sul (UFRGS), Sarmento Leite St., 425, Porto Alegre 90050-170, RS, Brazil
3
Complex Fluid Systems Lab, Institute of Earth Sciences, Rua Romao Ramalho, 59, 7000-671 Evora, Portugal
*
Author to whom correspondence should be addressed.
J. Mar. Sci. Eng. 2025, 13(10), 1979; https://doi.org/10.3390/jmse13101979
Submission received: 15 September 2025 / Revised: 2 October 2025 / Accepted: 14 October 2025 / Published: 16 October 2025
(This article belongs to the Special Issue Selected Feature Papers in Ocean Engineering)

Abstract

The present work investigated numerically turbulent airflows over a hybrid Darrieus/Savonius vertical axis wind turbine. Firstly, the isolated turbines were validated in comparison to previous studies from the literature. Later, new recommendations were obtained for the simulation of a hybrid turbine subject to turbulent airflow. The numerical simulations consisted of the solution of time-averaged equations of mass and momentum in x and y directions using the finite volume method, available in the commercial code Ansys Fluent (version 2022 R1). For closure of turbulence, the kω SST (Shear Stress Transport) model was employed. For lower magnitudes of tip speed ratio (TSR), the hybrid turbine improved the power coefficient (CP) compared to the Darrieus turbine (e.g., by 70% at TSR = 0.75), thereby demonstrating the self-starting capability of the hybrid configuration. Unexpectedly, at the optimal TSR = 1.5, the hybrid turbine performed about 6.5% better than the Darrieus turbine, indicating that the balance between the additional power generated by the Savonius rotor and losses caused by flow disturbances in the hybrid configuration was positive. As a novelty, results highlighted the role of each rotor (Darrieus and Savonius) for the performance of the hybrid turbine by comparing it with isolated Darrieus and Savonius turbines under the same conditions.

1. Introduction

In recent decades, extreme climate events have intensified as a result of environmental changes, leading to substantial material damage, human and economic losses, and increased health risks [1,2]. Accordingly, growing scientific and societal efforts have been directed toward understanding the drivers of climate imbalance and identifying effective mitigation strategies. A prominent approach to alleviating climate change and maintaining the trend on global energy demand is the diversification of the energy matrix by expanding the share of renewable sources, such as wind, solar, and wave energy, and reducing dependence on fossil fuels. According to the Low Emissions Scenario technical report by [3], it is projected that by 2050, around 80% of global energy consumption will originate from renewable sources. Moreover, wind energy alone is expected to expand up to eightfold relative to its current generation capacity [3].
Among renewable sources, wind energy has been one of the most important alternatives [4]. Consequently, research aimed at improving wind energy exploration remains of considerable importance. Wind turbines perform a main role in this direction, being categorized into two main distinct categories: horizontal axis wind turbines (HAWTs) and vertical axis wind turbines (VAWTs) [5,6]. Although HAWTs dominate large-scale wind farms due to their high efficiency, VAWTs present several advantages, including design simplicity, the absence of a yaw mechanism, ease of maintenance, and the ability to operate under irregular wind conditions, which makes them particularly suitable for urban applications [7]. Moreover, floating VAWTs also presented advantages in comparison with HAWTs, such as higher stability, lower aerodynamic wakes, and easier operation and maintenance, which can be important for the implementation in offshore wind farms [8]. VAWTs can be classified into two subtypes: lift-driven Darrieus turbines and drag-driven Savonius turbines. While Savonius turbines are more suitable for operation at low tip-speed ratios (TSRs), the Darrieus is more efficient at high TSRs [6,9]. Another characteristic of the Darrieus turbine is the difficulty for self-start at low wind velocities [5], which is an advantage for the Savonius turbine [6].
The hybridization of turbines arises from the necessity of benefiting the wide range in magnitudes of velocity of the wind. Therefore, one of the ways to minimize the operation difficulties of some VAWTs consists in the combination of different types such as the Darrieus and Savonius turbines. However, the characterization of the main parameters over turbine performance is necessary for better comprehension and new developments of the device [10].
Considering the importance of the hybridization of the turbines, the Darrieus/Savonius turbine has been recently investigated in the literature. For example, Liang et al. [11] performed a numerical investigation of a hybrid Darrieus with three aerodynamic profiles and a Savonius with two blades. Authors used a kε model in RANS (Reynolds-Averaged Navier–Stokes) to approach for closure of turbulence. Moreover, the influence of the attachment angle (α), number of blades for the Darrieus rotor (N), and radius ratio of the two types of rotor (RR = RSav/RDar) over the power coefficient were investigated. The study showed that self-starting performance, under the investigated conditions, can have a significant improvement by adding the Savonius turbine to the Darrieus turbine, and costing only a small compensation on the power efficiency when geometric parameters are optimized. Later, Asadi and Hassanzadeh [12] performed a similar numerical investigation considering other geometrical parameters, e.g., number of Darrieus blades of ND = 2, a different profile for the Darrieus blade (NACA 0018), and overlap ratio of the Savonius turbine (s/c = 0.2). Moreover, the closure of turbulence was performed with the kω SST (Shear Stress Transport) model. Authors noticed that independent of the free-wind velocity, the hybrid rotor with an attachment angle of α = 45° presented a maximum power coefficient for rotors with TSR = 1.5 and 2.5, and α = 0° led to a better performance for TSR = 3.5. Chegini et al. [6] investigated numerically a hybrid Darrieus/Savonius turbine to enhance the self-starting of the Darrieus turbine and proposed the inclusion of deflectors placed in front of and alongside the hybrid turbine, seeking to increase its efficiency. The modeling of turbulence was also performed with the kω SST model. Authors observed that coupling the Darrieus and Savonius turbines benefited the power coefficient (CP) by around 27% for the lowest investigated TSR = 1.45. However, when TSR is augmented, the hybrid turbine performance decreases due to limitations in aerodynamic performance of the Savonius turbine. The insertion of deflectors proved to be a good strategy for solving this problem, since at the optimum TSR = 2.6, the use of deflectors increased the CP by around 30% by means of increasing wind energy density at the upwind and leeward regions of the turbine. Afterwards, Redchyts et al. [13] performed a numerical analysis of a hybrid Darrieus and Savonius turbine using a specialized package of computational aerodynamics based on the finite volume method (FVM) and treating the turbulence with the one-equation Strain-Adaptive Linear Spalart–Allmaras (SALSA) model. This study focused on the comprehension of the generation of vortex structures and their relationship with the torque generated in the turbine. Results indicated that the Savonius turbine contributed only a small percent of the total torque produced by the installation, being the main contribution achieved by the Darrieus turbine, mainly in the windward section of the trajectory. Moreover, the interaction between the vortices generated by the Savonius rotor and the blades of the Darrieus rotor caused a strong decrease in the torque coefficient in the leeward part of the trajectory. Recently, Arrieta-Gomez et al. [14] conducted thirty-six numerical experiments to investigate the effects of radius ratio (RR), coupling angle (α), and TSR on the turbine performance. Computational fluid dynamics (CFD) simulations to obtain the turbine torque were performed with the commercial code ANSYS Fluent 2024 R2, based on the FVM. Moreover, the kε RNG (Renormalization Group) model was employed for closure of time-averaged momentum equations. To investigate the sensitivity and interaction among the parameters, statistical analysis and five regression models were developed and evaluated using the Akaike Information Criterion (AIC) and Bayesian Information Criterion (BIC), which are statistical criteria to predict new data based on maximum likelihood estimation. Results indicated that the parameter RR had the highest influence on turbine efficiency. It was also noticed in the literature, recent contributions about the distribution of arrangements of hybrid Darrieus/Savonius turbines and optimization of a pair of hybrid turbines investigating various parameters of the arrangement, such as turbine spacing, configuration angle, direction of rotation, relative height, and pitch angle [15,16]. Despite these important investigations, there are no geometrical configurations defined yet for the turbine, as well as computational parameters used in all simulations. To illustrate this situation, Table 1 shows some configurations and geometrical and computational parameters used in different numerical works from the literature.
In addition to the studies related to hybrid Darrieus/Savonius turbines, several important recent contributions have been performed to improve Savonius or Darrieus turbines, e.g., use of deflectors to improve the power coefficient of a pair of Savonius turbines [17], use of aerodynamic profiles for Savonius blades [18], use of double Darrieus turbines [19], and use of the Savonius turbine as a power take-off of an oscillating water column wave energy device [20]. More important insights about improvements in Savonius turbines, Darrieus turbines, and hybrid Darrieus/Savonius turbines are presented in the review works of [10,21,22].
The main objective of the present work is to investigate the influence of a hybrid Darrieus/Savonius turbine on power performance, as well as to identify changes in its behavior compared to an isolated Darrieus configuration under the same airflow conditions. To better understand the contribution of the Savonius rotor in the hybrid setup, simulations were also performed for an isolated Savonius turbine under the same conditions. Furthermore, to gain insight into the local behavior of hybrid and Darrieus turbines, the moment and drag coefficients of each blade were monitored as a function of the rotation angle, which has not been previously investigated in the literature. Another important aspect, which represents a secondary objective, concerns the limited exploration of different computational parameters (e.g., discretization scheme or pressure-coupling treatment) and their influence on turbine performance in the literature. Therefore, distinct numerical parameters for the discretization scheme were tested here, and other parameters, such as pressure–velocity treatment, residuals, meshes, residuals, number of iterations, and time-step are clearly defined to provide recommendations regarding simulation procedures for this type of problem. The best parameters were then applied in the simulation of the hybrid Darrieus/Savonius turbine, which was compared to the isolated Darrieus and Savonius configurations. To the best of the authors’ knowledge, the investigation performed here was not previously performed in the literature.
The remainder of this paper is organized as follows: Section 2 presents the problem description, including the computational domain, boundary conditions, and the performance indicators of the turbines. Section 3 details the mathematical and numerical model, describing the governing equations, the simulation parameters, the mesh parameters employed, and the ANSYS Fluent setup. Section 4 presents the results and discussion, including the influence of the discretization schemes, numerical model validation, mesh independence study, and turbine behavior and performance indicators for the different cases. Finally, Section 5 summarizes the main conclusions of the research.

2. Problem Description

In this section, the description of the computational domain, the number of cases studied, and the performance parameters investigated are detailed.

2.1. Computational Domain and Studied Cases

The computational domain adopted in this study is shown in Figure 1. It is divided into two regions: a rotational zone (gray) and a stationary zone (outside the rotational zone). The diameter of the rotational region is defined to ensure that the simulation results are not influenced by domain boundaries. This value was set to approximately 1.5 times the turbine diameter, and all rotors are modeled to rotate counterclockwise under a constant angular velocity ( θ ˙ ) imposed within the rotating domain to simulate a stable flow condition around the turbine.
The computational domain is defined by a total length of L = 26D, a height of H = 12D, and an upstream length of LI = 12D. In the simulations, the Darrieus and Savonius diameters were DD = 800 mm and DS = 320 mm, respectively, including the hybrid configuration simulations, corresponding to a radius ratio RR = 0.4. These values were defined based on previous studies by [23,24,25]. The rotational zones with turbine dimensions are illustrated in Figure 1b–d. It is worth noting that the no-slip boundary condition is imposed at the blades, i.e., u = v = 0 m/s, related to the rotational domain. The geometric parameters of the Savonius turbine were set as: overlap a = 0 mm, blade thickness e = 1.0 mm, blade chord c = 168.0 mm, and overlap s = 16 mm. The Darrieus rotor consists of three NACA 0018 blades offset by 120°, and a thick airfoil profile known for its higher lift at low rotational speeds. The blade chord length was set to C = 200 mm, with a fixed angle of attack φ = 0°. For the hybrid configuration, as the simulations are conducted in a two-dimensional domain, the Darrieus and Savonius turbines are assumed to have the same height.
Concerning the remaining boundary conditions, at the exit of the domain, it is considered null gauge pressure (Pg = 0 Pa), the symmetry boundary condition is assumed for the upper and lower surfaces of the domain, and a constant velocity is imposed at the inlet of the domain. The boundary conditions adopted at the inlet in all simulations are shown in Table 2, which also presents the main parameters adopted in the different simulations. The parameters used in the hybrid configuration are defined based on the Darrieus characteristics. Despite the same rotational velocity being imposed for comparison among hybrid, Darrieus, and Savonius configurations, the TSR is different for the same angular velocity due to the different dimensions of the radius of the Savonius turbine and the other configurations. The turbulent intensity presented in Table 2 was defined by:
IT % = 100 u 2 ¯ V
where u ¯ represents the fluctuation averaged field (Reynolds stress), and V represents the free-stream horizontal component of the wind flow velocity [26].
The power coefficient obtained in the present study was validated against the experimental results of [5] for the Darrieus configuration, while the Savonius results were compared with the results reported in [23,24,27]. In the present work, a total of 37 simulations were performed for the following investigations:
(i)
4 simulations for investigation of the influence of discretization schemes on the performance of the Savonius turbine (validation/verification case);
(ii)
2 simulations for investigation of the influence of discretization schemes on the performance of the Darrieus turbine (validation/verification case);
(iii)
8 simulations for validation/verification of the Darrieus turbine (used in comparison with the hybrid configuration);
(iv)
7 simulations for validation/verification of the Savonius turbine with ReDS = 8.67 × 105 and TSR = 0.50; 0.75; 1.00; 1.25; 1.50; 1.75, and 2.00;
(v)
8 simulations for new recommendations for hybrid turbines;
(vi)
8 simulations for the Savonius turbine with ReDS and TSRs presented in Table 1 allowing the comparison with the hybrid configuration.

2.2. Turbine Performance Parameters

To assess the aerodynamic performance of the turbine, various parameters can be employed. The power coefficient ( C P ) and moment coefficient ( C m ) are the commonly utilized parameters [6]. The power coefficient is a non-dimensional parameter that evaluates the turbine efficiency, defined as the ratio between the power extracted by the rotor and the available power that could be harnessed. The CP is calculated as follows [28]:
C p   =   2 P ρ A s V 3
where P   ( N   m / s ) represents the turbine’s generated power, ρ   ( kg / m 3 ) is the free stream density, A s   m 2 is the rotor’s swept area, and V   ( m / s ) is the free stream velocity. The turbine power is assessed by the product of generated torque on the rotor and angular velocity [28].
P = T t u r b θ ˙
where T t u r b (Nm) is the generated torque, and θ ˙ ( rad   s 1 ) is the angular velocity. The moment coefficient represents the ratio between the torque generated by the turbine and the torque exerted by the undisturbed flow. The C m is calculated as follows
C m   = 2   T t u r b ρ A s V 2 R S , D
where R S , D (m) is the rotor’s radius. The tip speed ratio T S R is a non-dimensional parameter widely employed to characterize the operating condition of a kinetic turbine. It is defined as the ratio between the tangential velocity and the free stream wind velocity. The TSR is defined as follows:
T S R = R S , D θ ˙ V
The moment coefficient and the power coefficient are related through the T S R . This relationship can be expressed as
C P = T S R ·   C m

3. Mathematical and Numerical Modeling

In this section, the mathematical and numerical procedures employed to simulate the fluid flow are presented. The governing equations, turbulence model, and computational setup are detailed to provide a complete overview of the modeling approach.

3.1. Governing Equations

For a mathematical modeling of turbulent, transient, and incompressible flow within a two-dimensional domain, the conservation of mass and balance of momentum equations are solved. The URANS approach is adopted to solve the momentum equation in terms of mean values. The mass conservation and momentum equations in x and y directions were defined as follows [29]:
u ¯ x + v ¯ y = 0
ρ u ¯ t + u ¯ u ¯ x + v ¯ u ¯ y = p ¯ x + μ + μ t 2 u ¯ x 2 + 2 u ¯ y 2
ρ v ¯ t + u ¯ v ¯ x + v ¯ v ¯ y = p ¯ y + μ + μ t 2 v ¯ x 2 + 2 v ¯ y 2
where u ¯ and v ¯ represents the mean velocity components in the x - and y -directions, respectively; p ¯ is the time-averaged pressure; μ denotes the dynamic viscosity; and μ t is the turbulent viscosity, introduced through the closure process by applying the temporal averaging operator to the advective terms in Equations (8) and (9).
To solve μ t , the kω SST (Shear Stress Transport) turbulence model was employed, which accurately captures the external flow in vortex shedding regions [30,31]. This model provides a gradual transition in the calculation of μ t , using the kω formulation for inner boundary layers (near-wall regions) and the kε model for regions farther from the turbine, where the turbulent flow is more isotropic. According to [31,32], the turbulent viscosity is defined as follows:
μ t = ρ ¯ α 1 k T m a x α 1 ω T , S T F 2
The turbulent viscosity is obtained from the solution of the turbulent kinetic energy transport equations (k) and specific dissipation rate ( ω ) given by the following [31]:
ρ k t + ρ u ¯ i k x i = P ~ k   β * ρ k ω + x i μ + σ k μ t k x i  
ρ ω t + ρ u ¯ i ω x i = α ρ S 2 β ρ ω 2 + x i μ + σ ω μ t ω x i + 2 1 F 1 ρ σ ω 2 1 ω k x i ω x i  
where P ~ k is turbulent kinetic energy production (a function that avoids turbulence generation in stagnant zones), and S is the invariant measure of the strain rate. The constants β * =   0.09 , α 1   =   5 / 9 , β 1   =   0.075 , σ k 1   =   0.85 , σ ω 1   =   0.5 , α 2   =   0.44 , β 2   =   0.0828 , σ k 2   =   1 , and σ ω 2   =   0.856 are standard model constants according to [31,32].
The blending functions F 1 and F 2 are defined by [31], which serve to adjust the solution between kω and kε models, given by the following:
F 1 = t a n h m i n m a x k β * ω y , 500 v y 2 ω , 4 ρ σ ω 2 k C D k w y 2 4
F 2 = t a n h m a x 2 k β * ω y , 500 v y 2 ω 2
where C D k w is given by the following:
C D k w = m a x 2 ρ σ ω 2 1 ω k x i ω x i , 10 10

3.2. Numerical Procedures and Parameters

The Finite Volume Method (FVM) is employed as a discretization procedure. More precisely, it is implemented in the commercial code ANSYS Fluent 2022 R1 [30,33]. Simulations were run on a desktop computer equipped with an Intel® CoreTM i7-5820K CPU 3.30 GHZ (six cores), memory Desktop Gamer DDR4 with 16 GB of RAM, a motherboard LGA 2011-V3 Intel GA-X99-Gaming, and a hard disk 2TB Western Digital Green SataIII 64 MB. Each simulation required approximately 30 h of processing time.
Figure 2a–c shows the spatial discretization employed for hybrid, Savonius, and Darrieus rotors, respectively. The sliding mesh approach was adopted in the interface between the static zone and the rotational region to prescribe the rotational motion of the rotors. This method was commonly used in the literature to simulate the turbine’s motion [4,9,10,12,26,27]. In the rotational motion, the nodes move as a rigid body within the dynamic mesh zone, while the cell zones are coupled through non-conformal interfaces. As the mesh evolves in time, these interfaces are updated to the new relative positions of each zone [30]. The mesh quality parameters are presented in Table 3.
Mesh refinement is applied near the turbine walls, where higher velocity gradients are expected. The mesh near the walls is refined to ensure values of the dimensionless wall distance y +     1 . As can be seen in Table 3, the maximum y+ is superior to 1, but localized to a few points. In general, the y+ is much lower than unity, showing adequate refinement in the near-wall region. This parameter is given as follows [29,31]:
y + = y τ w ρ μ
where τ w represents the wall shear stress (Pa).
For pressure and velocity coupling, the Semi-Implicity Method for Pressure-Linked Equations (SIMPLE) algorithm is employed. The influence of advective terms is investigated for Savonius and Darrieus configurations. These results are confronted with the literature data from [5,23,27]. The first-order implicit scheme is used for the spatial discretization of the transient formulation.
The convergence criterion is set to 10−6. A total of 2000 time-steps are simulated for each case, with a time step size of Δ t   =   0.00175   s , resulting in a total simulation time of 3.5 s. The maximum number of iterations per time step is set to 180. These parameters are the same previously used in the work of dos Santos et al. [24].

4. Results and Discussion

This section presents a study about the influence of discretization schemes used in RANS and kω SST equations on the power coefficient, the verification/validation of isolated Darrieus and Savonius turbines, and the performance assessment of the hybrid turbine configuration in comparison with the Darrieus and Savonius conventional turbines. The results are organized to highlight the power coefficient C P as a performance indicator for the torque behavior of each blade of the hybrid and Darrieus configurations, and flow velocity and total pressure fields to illustrate the flow behavior.

4.1. Influence of Discretization Scheme

This section, is focused on the investigation of the influence of distinct discretization schemes on the power coefficients for one TSR of Savonius and Darrieus turbines investigated in isolated form, and compared with previous results from the literature. This initial simplified investigation is performed since the adopted discretization schemes have not been widely described in the literature, as is the spatial discretization method (where most works use the FVM), the turbulence closure modeling (where kω SST is the most recommended modeling for free shear turbulent basin flows, including the flows over rotational turbines), and the pressure-velocity coupling scheme (where SIMPLE has been the most adopted) [4,9,34].
Firstly, the turbulent airflow over a Savonius turbine with ReDS = 867,000 and TSR = 1.0 was simulated, and the results of CP were compared with the previous experimental and numerical results of [27] and [23], respectively. Table 4 illustrates the power coefficients obtained with the following discretization schemes: first-order upwind, second-order upwind, MUSCL (Monotonic Upstream-centered Scheme for Conservation Laws), and hybrid first- and second-order upwind, available in the software FLUENT 2022 R1 [30,33]. Results of the comparison are also presented in Table 4. As can be seen, the results obtained with first- and second-order upwind led to differences of only 2.2% and 0.4%, respectively, in comparison with the experimental predictions of [27]. Therefore, these two discretization schemes were tested in the simulation of an isolated Darrieus turbine with three blades, profile NACA 0018, ReDD = 438,200 and TSR = 1.00.
For this new investigation, the first- and second-order upwind interpolation functions led to deviations of 0.6% and 33% in comparison to predictions obtained in the literature by [5]. One possible explanation for this behavior is the greater difficulty in achieving convergence with the second-order upwind interpolation function, due to its dependence on a larger number of adjacent finite volumes. The use of a first-order upwind scheme for some turbine simulations may yield better results than a higher-order scheme due to its numerical diffusion. A second-order scheme is theoretically more precise, but it is also highly sensitive to mesh quality and can generate non-physical oscillations and instabilities. The first-order scheme’s diffusion may act as a numerical stabilizer, smoothing out instabilities and, in some cases, providing a more robust and physically representative solution that better matches experimental data. It is worth noting that the study was limited to applying a single discretization scheme throughout each entire simulation. Therefore, based on these comparisons, the first-order upwind is employed in the validation/verification step, along with new recommendations for hybrid turbines.

4.2. Validation/Verification of Numerical Model

In this section, the validation/verification is performed considering the influence of TSR on the time-averaged power coefficient ( C P ¯ ) for Darrieus and Savonius turbines simulated separately. Once validated, this methodology was then applied to simulate a hybrid turbine configuration, generating its CP curve as a function of TSR.
Figure 3 illustrates the curve C P ¯ = f(TSR) for the Darrieus turbine with ReDD = 438,200, being the results compared with those obtained numerically in the work of Elkhoury et al. [5]. It can be observed that there is a very good agreement between the present results and those obtained by Elkhoury et al. [5], with some distinct magnitudes of C P ¯ for TSRs = 0.75 and 1.5, where the present code underestimated and overestimated the magnitudes of Cp, respectively. Another noted difference is in the optimum point of the TSR, which reached TSR = 1.5 with the present code, while the results of Elkhoury et al. [5] estimated TSR = 1.25. Despite these local differences, considering the complexity of the turbulent flows over rotational aerodynamic profiles, it can be assumed that the predictions are similar.
Figure 4 shows the effect of TSR over C P ¯ for the Savonius turbine at turbulent air flows with ReDS = 867,000. The results are compared with the experimental results of Sheldahl et al. [27], and numerical achievements of Akwa et al. [23], and dos Santos et al. [24]. The trend of the C P ¯ = f(TSR) obtained with the numerical methods, including the present method, is like that predicted by the experimental measures. In the region 0.75 ≤ TSR ≤ 1.25, which is the optimal operation region of the Savonius turbine, the present code led to differences inferior to 3.0% in comparison with other numerical/experimental predictions. For TSR = 2.0, the negative magnitude of C P ¯ indicates that the turbine supplies energy to the surrounding airflow instead of obtaining energy from it. At this TSR, it can be noticed that the present method and the method employed by Akwa et al. [23] underestimated the negative magnitude of C P ¯ in comparison with the experimental results of Sheldahl et al. [27]. Due to this behavior, this point is not considered in the hybrid turbine investigation, limiting the TSRS to 1.0.
Overall, considering the difficulties of simulating turbulent flows over complex rotational geometries, as found in the present work, the results obtained with the present numerical method are in agreement with that previously available in the literature for isolated Savonius and Darrieus turbines. Therefore, it is assumed that the present computational methodology for turbine simulation is suitable for the new recommendations obtained with the hybrid Darrieus/Savonius turbine.

4.3. Behavior of Hybrid Darrieus/Savonius Turbine

Before obtaining results about the behavior of the hybrid turbine, a mesh sensitivity analysis was conducted to ensure that the numerical results are independent of the mesh resolution. The hybrid turbine configuration was selected for this analysis, as it represents the most complex mesh and encompasses the other configurations (isolated Savonius and Darrieus configurations). The mesh convergence analysis was performed for TSR = 0.50. The Grid Convergence Index (GCI) method, combined with Richardson extrapolation, was applied following the procedure described by Meana-Fernández [35].
Three meshes with varying refinement in the cross-streamwise direction were considered. Table 5 summarizes the grid characteristics near the airfoil walls and the total number of volumes for each mesh. Once the mesh independence has been confirmed, the finest mesh is used for the subsequent simulations.
Figure 5 shows the variation in the average power coefficient C P ¯ as a function of the grid characteristic size (h). The points are fitted using a polynomial interpolation (dashed blue line). The curve exhibits an asymptotic behavior at C P ¯ = 0.0463, indicated by the Richardson extrapolation point (red circle) in the figure. This confirms that the finest mesh (Mesh 1) provides results sufficiently independent of the mesh resolution.
The hybrid Darrieus/Savonius turbine combines the two designs to complement the strengths of both isolated turbines (Darrieus and Savonius). The Savonius component provides the necessary torque for self-starting and efficient operation at low wind speeds, effectively overcoming the Darrieus rotor’s characteristic low performance in these conditions. To analyze the impact and interaction between the turbines, the Savonius, Darrieus, and hybrid turbine configurations were individually simulated. The transient behavior of CP for each configuration is shown in Figure 6a–d. All turbines illustrated in Figure 6a–d were rotated at constant angular velocities of 10, 20, 25, and 50 rad/s, respectively. These constant angular velocities resulted in different tip speed ratios (TSR) for the Savonius and Darrieus as well as hybrid designs, due to their respective diameters.
The Savonius turbine exhibited lower power coefficient (CP) values compared to the other turbines within the chosen TSR range, except for higher TSRD = 2.50, where TSRD represents the TSR for Darrieus and hybrid configurations. For lower TSR magnitudes, Figure 6a, the presence of the Savonius rotor in the hybrid configuration altered the behavior of CP as a function of time in comparison with the isolated Darrieus turbine, increasing the magnitude of CP by at least a small amount. This effect happened even in a TSRS = 0.20 (Savonius TSR), where its torque still had a lower magnitude than in regions of TSRS~1.00. The overall behavior of the hybrid turbine appears to be primarily guided by the Darrieus design, but with a notable phase shift in TSR caused by the torque influence of the Savonius turbine. This shift, which is quantified by an angular value of nearly 0.13 rad (7.45 degrees), likely depends on the specific TSR and the overall hybrid turbine design. The relative positioning of the Savonius and Darrieus rotor blades, as well as the number of blades, diameters, and blade profiles, may be a significant factor contributing to this behavior.
As the TSR increases, the hybrid turbine’s behavior becomes even more dominated by the Darrieus turbine, see Figure 6b–d. In these scenarios, the aerodynamic forces acting on the Darrieus blades are the primary drivers of the turbine’s response. The specific choice of blade profiles can significantly influence this behavior. This is because the Darrieus turbine is particularly well-suited for high TSRs, which forces the hybrid turbine to adopt similar performance characteristics. These findings emphasize that a Darrieus-like blade design is crucial for the hybrid turbine, especially at higher TSRs. As the TSR increases to TSRS = 1.00 and TSRD = 2.50 (Figure 6d), a different behavior is observed. In this condition, the Savonius still produces a positive power coefficient (CP = 0.233), while the Darrieus and hybrid configurations begin to show negative magnitudes around CP~−0.2. Since the rotational velocity was imposed during the simulation, a negative CP indicates that the turbine is putting energy into the fluid—a scenario that would not occur under real-world operating conditions. This divergent behavior between the Savonius, Darrieus, and hybrid configurations can be related to the operation of the Savonius turbine in a near optimum region of operation (TSRS = 1.00) while the Darrieus turbine operates far from optimum conditions (TSRD = 2.50). The results of Figure 6 also illustrate the importance of the ratio between the Darrieus and Savonius dimensions for the hybrid configuration in terms of the combination of TSR operation of the Darrieus and Savonius rotors. For the present configuration, the peak performance of the Savonius occurred only where the performance of the Darrieus turbine passed from the optimum TSR point.
To investigate more precisely the influence of the association between the Darrieus and Savonius turbines, the instantaneous momentum coefficient (Cm) and drag coefficient (Cd) as a function of the angle of rotation of the turbine’s blade (θb) (see the schematic representation of the blades number and angle of rotation in Figure 7) are illustrated in Figure 8 and Figure 9, respectively, for TSRD = 0.5. More precisely, Figure 8a illustrates the Cm as a function of θb for the three blades of the isolated Darrieus turbine, where θb represents the angle position of each blade that occurs at different times. Figure 8b shows the Cm as a function of θb for the isolated Savonius turbine and the Savonius rotor in the hybrid configuration, and Figure 8c shows the Cm as a function of θb for each blade of the Darrieus hybrid turbine and a comparison with one of the isolated Darrieus turbines. Figure 7 illustrates a schematic representation of the blade number, the angle θb, and the sign of rotation of the turbine.
Figure 8a and Figure 9a show that only slight differences are observed for Cm and Cd as a function of θb, showing a kind of symmetry among the blades. It is also observed that there are peaks of Cm and Cd at each rotation of the turbine when each blade is near the upper position of the rotational region. Figure 8b and Figure 9b illustrate the comparison of Cm and Cd obtained for the isolated Savonius turbine and the same parameters measured in the Savonius rotor (with the same dimensions as the isolated Savonius case) inserted in the hybrid turbine. Results indicated that both Cm and Cd of the Savonius turbine suffered a step decrease from the isolated configuration to the hybrid one. A possible reason for this behavior is related to the perturbation of airflow after passing through the blades of the Darrieus rotor and impinging on the Savonius blades, see the velocity and pressure fields in Figure 10 and Figure 11, with a comparison between Darrieus and hybrid configurations. A Savonius turbine operates primarily due to the difference in drag force between its concave and convex surfaces. The addition of Darrieus blades can significantly alter these drag differences, which in turn affects the hybrid turbine’s performance. In Figure 8c and Figure 9c it can be observed that the inclusion of the Savonius turbine leads to changes in the angle at which the peak for both Cm and Cd happened in comparison to the isolated Darrieus turbine. More precisely, the peaks of the hybrid turbine changed 0.4 rad (22.9 degrees) later than the isolated Darrieus turbine.
To quantify the mean differences between the hybrid and Darrieus turbines, as well as the contribution of the Savonius rotor to the performance of the hybrid turbine, Table 6 and Table 7 show the C m ¯ and C d ¯ for the hybrid configuration, isolated Darrieus and Savonius turbines at TSR = 0.5. As can be seen, the C m ¯ of the Darrieus rotor for the hybrid configuration increases in comparison with the isolated Darrieus configuration (32%). However, the contribution of the Savonius blades conducted the hybrid turbine to a better performance than the Darrieus configuration. It is also observed that there is a step difference between the Savonius rotor in the hybrid configuration and the isolated Savonius rotor. Despite the strong decrease in C m ¯ for the Savonius rotor in the hybrid configuration, it helps the global performance of the hybrid turbine to be higher than the isolated Darrieus configuration.
Figure 12 shows the relationship between the mean CP and TSR for the Savonius, Darrieus, and hybrid turbines. It is reinforced here that, for the same angular velocity ( θ ˙ ) two distinct TSRs are obtained for the Darrieus and Savonius rotors due to difference in radius dimensions. This curve provides crucial insight into the power output characteristics of each turbine type. As expected, the Savonius turbine performs optimally at low TSRs (TSRS = 0.8), while the Darrieus turbine is effective at higher TSRD (TSRD = 1.5). The inclusion of the Savonius rotor in the hybrid design positively influences performance at lower TSR magnitudes, specifically in the range 0.5 ≤ TSRD ≤ 1.0, where the mean Cp is noticeably higher than that reached by the isolated Darrieus configuration. This behavior reinforces the previous observations from the literature that the Savonius turbine increases the torque of the turbine when associated with the Darrieus one. Unexpectedly, the hybrid turbine achieved a maximum C P ¯ that is slightly higher than that of the Darrieus turbine alone (6.5%), a result not typically reported in the literature. While the magnitude of this improvement is small, it highlights a potential benefit of hybrid design. The performance of the hybrid turbine can be benefited by the additional power generated by the Savonius turbine and harmed by the disturbances of fluid flow generated by the Savonius turbine impinging the blades of the Darrieus turbine. For the optimal TSRD = 0.75, this phenomenon was not as pronounced, perhaps due to the specific scale factor between the two turbine components (RR = 0.4). This suggests that the relative size and positioning of the Savonius and Darrieus sections may play a critical role in mitigating the aerodynamic interference. Further investigation into the optimal scale ratio could help to minimize these negative effects and improve the overall efficiency of hybrid vertical-axis wind turbines. It is worth noting that the operation conditions of the TSR magnitudes for each rotor (which are also influenced by the radius ratios of the turbine) can be an important aspect in this issue. Further research is necessary to fully understand this phenomenon and its potential applications, particularly regarding the optimal integration and scaling of the two turbine types to maximize overall efficiency.
Despite some simplifications employed in the present work as the two-dimensional domain and consequent hypothesis of the same height for Darrieus and Savonius rotors in hybrid configuration, the present study obtained promising theoretical recommendations for the application of VAWTs. For instance, results demonstrated that for lower TSRs, the Savonius turbine improved the self-starting of the hybrid turbine. Results of the comparison also indicated that there is a balance between the additional power generated by the Savonius turbine and the additional perturbation of turbulent flow caused by the insertion of the Savonius rotor into the Darrieus one. Considering the VAWTs have been recommended for urban areas and floating offshore wind farms, the use of a hybrid configuration can be a promising strategy to improve the energy generated in these farms.

5. Conclusions

The present work performed a numerical analysis of turbulent airflows over hybrid Darrieus/Savonius vertical axis wind turbines, obtaining new recommendations about numerical procedures for the simulation of this problem and new comparisons among hybrid, Darrieus and Savonius turbines, seeking to understand the influence of the association between Darrieus and Savonius turbines. The numerical simulation was based on the solution of time-averaged equations of mass and momentum in x and y directions using the finite volume method, available in the commercial code Ansys Fluent, and the k−ω SST to tackle the closure of turbulence. Moreover, the sliding mesh methodology was employed for rotational movement of the turbines.
Firstly, an investigation of the reliability of the present computational method for the simulation of rotational Darrieus and Savonius turbines subjected to turbulent flows was conducted. Results indicated that the use of the first-order upwind interpolation function for treatment of advective terms resuled in a accurate prediction of C P ¯ for the simulation of Darrieus and Savonius turbines, with a difference of 0.3% and 2.2% in comparison with studies in the literature [5,23,27]. Despite the very good agreement for the Savonius turbine (0.4% in comparison with experimental results by Sheldahl et al. [27]) obtained with second-order upwind, important discrepancies were noticed for the Darrieus turbine (33% in comparison with the results by Elkhoury et al. [5]). In a second instance, the effect of TSR over C P ¯ for isolated Darrieus and Savonius turbines were simulated, being obtained satisfactorily in agreement with previous findings in the literature considering the difficulties in the simulation of turbulent flows over rotational domains of complex turbine configurations.
Results of the hybrid turbine indicated that, for lower magnitudes of TSR, the hybrid turbine improved the C P ¯ in comparison with the Darrieus turbine. For TSRD = 0.50 and 0.75, for example, differences of 35% and 70% were obtained, respectively. This is a promising indication that the insertion of the Savonius turbine can benefit the self-start of the Darrieus turbine. These results agreed with previous findings in the literature. For the optimal TSRD = 1.5, the hybrid turbine led to a performance nearly 6.5% superior to the Darrieus turbine, which was initially unexpected. The probable reason for this behavior is related to the additional contribution of the Savonius turbine for the power of the system to be higher than the reduction in power in the Darrieus turbine in the hybrid configuration due to new flow disturbances caused by the insertion of the Savonius turbine. This behavior was not previously obtained in the literature and requires further investigation. For TSRD ≥ 1.75, the hybrid turbine obtained an inferior performance in comparison with an isolated Darrieus turbine, as previously described in the literature. Another important observation was the change in the instantaneous operation of the blades of the Darrieus turbines when the Savonius turbine was inserted. Examples are noticed in the change in angle of power peaks for blades of Darrieus turbines. Moreover, the Savonius behavior in a free stream condition and inserted in the Darrieus rotor was also changed.
For future studies, it is recommended to elaborate on the investigation of other geometrical parameters as different radius ratios of Darrieus/Savonius turbines, to try to make the highest performance operation of both rotors compatible in a similar region of TSR.

Author Contributions

Conceptualization, R.G.d.S.I., G.d.C.D., R.A.A.C.G. and E.D.d.S.; methodology, R.G.d.S.I., I.A.d.R., V.H.A., G.d.C.D., R.A.A.C.G. and E.D.d.S.; software, L.A.O.R., L.A.I., R.A.A.C.G. and E.D.d.S.; validation, R.G.d.S.I., I.A.d.R., V.H.A. and R.A.A.C.G.; formal analysis, R.G.d.S.I., I.A.d.R., V.H.A., R.A.A.C.G. and E.D.d.S.; investigation, R.G.d.S.I., I.A.d.R., V.H.A., R.A.A.C.G. and E.D.d.S.; resources, L.A.O.R., L.A.I. and E.D.d.S.; data curation, R.G.d.S.I., I.A.d.R. and R.A.A.C.G.; writing—original draft preparation, I.A.d.R., V.H.A., R.A.A.C.G. and E.D.d.S.; writing—review and editing, I.A.d.R., L.A.O.R., L.A.I. and G.d.C.D.; visualization, I.A.d.R., R.A.A.C.G. and E.D.d.S.; supervision, G.d.C.D., R.A.A.C.G. and E.D.d.S.; project administration, L.A.O.R., L.A.I., R.A.A.C.G. and E.D.d.S.; funding acquisition, I.A.d.R., V.H.A., L.A.O.R., L.A.I. and E.D.d.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was financed in part by the Coordenação de Aperfeicoamento de Pessoal de Nível Superior—Brazil (CAPES)—Finance Code 001; the doctorate scholarship of the author I.A.R. is funded by Conselho Nacional de Desenvolvimento Científico e Tecnológico—Brazil (CNPq) (Process: 177803/2024-0); the research grants of the authors E.D.d.S., L.A.I. and L.A.O.R. are also funded by CNPq (processes 308396/2021-9, 309648/2021-1, and 307791/2019-0); the research was financed by CNPq in the Call CNPq/MCTI No 10/2023-Universal (Process: 403408/2023-7); the research is also financed by Fundação de Amparo à Pesquisa do Estado do Rio Grande do Sul—Brazil (FAPERGS), Public Call FAPERGS 07/2021—Programa Pesquisador Gaúcho—PqG (Process: 21/2551-0002231-0).

Data Availability Statement

The data presented in this study are available upon request from the corresponding author. The data are not publicly available due to privacy reasons.

Acknowledgments

The authors thank the Coordenação de Aperfeicoamento de Pessoal de Nível Superior—Brazil (CAPES)—Finance Code 001 for financial support; I. A. da Rosa thanks Conselho Nacional de Desenvolvimento Científico e Tecnológico—Brazil (CNPq) for the doctorate scholarship (Process: 177803/2024-0); E.D.d.S., L.A.I. and L.A.O.R thank CNPq for research grants (processes 308396/2021-9, 309648/2021-1, and 307791/2019-0); all authors thank CNPq for financial support in the Call CNPq/MCTI No 10/2023-Universal (Process: 403408/2023-7); L.A.I. thanks Fundação de Amparo à Pesquisa do Estado do Rio Grande do Sul—Brazil (FAPERGS), Public Call FAPERGS 07/2021—Programa Pesquisador Gaúcho—PqG (Process: 21/2551-0002231-0).

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations and symbols are used in this manuscript:
AICAkaike information Criterion
BICBayesian information criterion
CFDComputational fluid dynamics
FVMFinite volume method
HAWTHorizontal axis wind turbine
ITIntensity of turbulence
MUSCLMonotonic upstream-centered scheme for conservation laws
NINot informed
RNGRenormalization Group
RANSReynolds-Averaged Navier–Stokes
SIMPLESemi-implicit method for pressure-linked equations
SSTShear stress transport
SALSAStrain-adaptative linear Spalart–Allmaras
TSRTip speed ratio
URANSUnsteady Reynolds-Averaged Navier–Stokes
VAWTVertical-axis wind turbine
A s Rotor’s swept area (mm2)
A Overlap of Savonius blades (mm)
C Savonius blade chord (mm)
C Darrieus blade chord (mm)
C d Drag coefficient
C m Moment coefficient
C p Power coefficient
D D Darrieus diameter (mm)
D S Savonius diameter (mm)
E Thickness of blade (mm)
F 1 , F 2 Blending functions
H Height of domain (mm)
K Turbulent kinetic energy (m2/s2)
L Total length of domain (mm)
N Number of blades on the Darrieus rotor
p - Time-averaged pressure (Pa)
P Turbine’s generated power (W)
P ~ k Turbulent kinetic energy production (N/(m2 s2))
R D a r Darrieus radius (mm)
R e Reynolds number
R R Radius ratio of Savonius and Darrieus rotors
R S a v Savonius radius (mm)
S Invariant measure of the strain rate (1/s)
T Time (s)
T t u r b Generated torque (N m)
u - ,   v - Mean velocity components in the x and y directions (m/s)
u - Fluctuation averaged field (m/s)
V Free current velocity (m/s)
y + Dimensionless wall distance
A Attachment angle (degrees)
Δ t Time step size (s)
ϵ Dissipation rate (m2/s3)
θ b Position angle of the blade (rad)
θ ˙ Angular velocity (rad/s)
Μ Dynamic viscosity (Pa s)
μ t Turbulent viscosity (Pa s)
Ν Cinematic viscosity (m2/s)
Ρ Free stream density (kg/m3)
τ w Wall shear stress (Pa)
Φ Angle of attack (degrees)
Ω Specific dissipation rate (1/s)

References

  1. AR6 Synthesis Report: Summary for Policymakers Headline Statements. Available online: https://www.ipcc.ch/report/ar6/syr/resources/spm-headline-statements (accessed on 23 September 2025).
  2. Otto, F.E.L. Attribution of Extreme Events to Climate Change. Annu. Rev. Environ. Resour. 2023, 48, 813–828. [Google Scholar] [CrossRef]
  3. Statkraft. Low Emissions Scenario; Statkraft: Oslo, Norway, 2023. [Google Scholar]
  4. Sadorsky, P. Wind Energy for Sustainable Development: Driving Factors and Future Outlook. J. Clean. Prod. 2021, 289, 125779. [Google Scholar] [CrossRef]
  5. Elkhoury, M.; Kiwata, T.; Aoun, E. Experimental and Numerical Investigation of a Three-Dimensional Vertical-Axis Wind Turbine with Variable-Pitch. J. Wind. Eng. Ind. Aerodyn. 2015, 139, 111–123. [Google Scholar] [CrossRef]
  6. Chegini, S.; Asadbeigi, M.; Ghafoorian, F.; Mehrpooya, M. An Investigation into the Self-Starting of Darrieus-Savonius Hybrid Wind Turbine and Performance Enhancement through Innovative Deflectors: A CFD Approach. Ocean. Eng. 2023, 287, 115910. [Google Scholar] [CrossRef]
  7. Jacob, J.; Chatterjee, D. Design Methodology of Hybrid Turbine towards Better Extraction of Wind Energy. Renew. Energy 2019, 131, 625–643. [Google Scholar] [CrossRef]
  8. Ghigo, A.; Faraggiana, E.; Giorgi, G.; Mattiazzo, G.; Bracco, G. Floating Vertical Axis Wind Turbines for Offshore Applications among Potentialities and Challenges: A Review. Renew. Sustain. Energy Rev. 2024, 193, 114302. [Google Scholar] [CrossRef]
  9. Dixon, S.L.; Hall, C. Fluid Mechanics and Thermodynamics of Turbomachinery; Butterworth-Heinemann: Oxford, UK, 2013; ISBN 978-0-12-391410-1. [Google Scholar]
  10. Sarma, J.; Jain, S.; Mukherjee, P.; Saha, U.K. Hybrid/Combined Darrieus–Savonius Wind Turbines: Erstwhile Development and Future Prognosis. J. Sol. Energy Eng. 2021, 143, 050801. [Google Scholar] [CrossRef]
  11. Liang, X.; Fu, S.; Ou, B.; Wu, C.; Chao, C.Y.H.; Pi, K. A Computational Study of the Effects of the Radius Ratio and Attachment Angle on the Performance of a Darrieus-Savonius Combined Wind Turbine. Renew. Energy 2017, 113, 329–334. [Google Scholar] [CrossRef]
  12. Asadi, M.; Hassanzadeh, R. Effects of Internal Rotor Parameters on the Performance of a Two Bladed Darrieus-Two Bladed Savonius Hybrid Wind Turbine. Energy Convers. Manag. 2021, 238, 114109. [Google Scholar] [CrossRef]
  13. Redchyts, D.; Fernandez-Gamiz, U.; Tarasov, S.; Portal-Porras, K.; Tarasov, A.; Moiseienko, S. Comparison of Aerodynamics of Vertical-Axis Wind Turbine with Single and Combine Darrieus and Savonius Rotors. Results Eng. 2024, 24, 103202. [Google Scholar] [CrossRef]
  14. Arrieta-Gomez, M.; Vélez-García, S.; Hincapié Zuluaga, D.; Tejada, J.C.; Sanin-Villa, D. Efficiency Optimization of a Hybrid Savonius-Darrieus Hydrokinetic Turbine via Regression Modeling and CFD-Based Design of Experiments. Results Eng. 2025, 27, 105751. [Google Scholar] [CrossRef]
  15. Abdel-razak, M.H.; Emam, M.; Ookawara, S.; Hassan, H. Study the Performance of a Novel Design of Twin Hybrid Darrieus-Savonius Vertical-Axis Wind Turbines Integrated with Building Water Storage Tanks: 3D Optimization Study. Renew. Energy 2026, 256, 124024. [Google Scholar] [CrossRef]
  16. Abdel-razak, M.H.; Emam, M.; Ookawara, S.; Hassan, H. Optimization of the Power Performance for Three Hybrid Darrieus-Savonius Rotors Based on the Taguchi Method. J. Phys. Conf. Ser. 2024, 2857, 012011. [Google Scholar] [CrossRef]
  17. Fatahian, H.; Mishra, R.; Jackson, F.F.; Fatahian, E. Design Optimization of an Innovative Deflector with Bleed Jets to Enhance the Performance of Dual Savonius Turbines Using CFD-Taguchi Method. Energy Convers. Manag. 2023, 296, 117655. [Google Scholar] [CrossRef]
  18. Dinh Le, A.; Nguyen Thi Thu, P.; Ha Doan, V.; The Tran, H.; Duc Banh, M.; Truong, V.-T. Enhancement of Aerodynamic Performance of Savonius Wind Turbine with Airfoil-Shaped Blade for the Urban Application. Energy Convers. Manag. 2024, 310, 118469. [Google Scholar] [CrossRef]
  19. Ahmad, M.; Shahzad, A.; Akram, F.; Ahmad, F.; Shah, S.I.A. Design Optimization of Double-Darrieus Hybrid Vertical Axis Wind Turbine. Ocean. Eng. 2022, 254, 111171. [Google Scholar] [CrossRef]
  20. Santos, A.L.G.; Biserni, C.; da Rosa, I.A.; Gonçalves, R.A.A.C.; Martins, J.C.; Rocha, L.A.O.; Isoldi, L.A.; Souza, J.A.; dos Santos, E.D. Numerical Modeling and Geometrical Investigation of Oscillating Water Column Device into a Full-Scale Wave Flume and Considering Savonius Turbine Insertion. Energy Convers. Manag. 2025, 346, 120468. [Google Scholar] [CrossRef]
  21. Jebelli, A.; Lotfi, N.; Saeid Zare, M.; Yagoub, M.C.E. A Comprehensive Review of Effective Parameters to Improve the Performance of the Savonius Turbine Using a Computational Model and Comparison with Practical Results. Water-Energy Nexus 2024, 7, 266–274. [Google Scholar] [CrossRef]
  22. Shen, Z.; Gong, S.; Zuo, Z.; Chen, Y.; Guo, W. Darrieus Vertical-Axis Wind Turbine Performance Enhancement Approach and Optimized Design: A Review. Ocean. Eng. 2024, 311, 118965. [Google Scholar] [CrossRef]
  23. Akwa, J.V.; Alves da Silva Júnior, G.; Petry, A.P. Discussion on the Verification of the Overlap Ratio Influence on Performance Coefficients of a Savonius Wind Rotor Using Computational Fluid Dynamics. Renew. Energy 2012, 38, 141–149. [Google Scholar] [CrossRef]
  24. dos Santos, A.L.; Fragassa, C.; Santos, A.L.G.; Vieira, R.S.; Rocha, L.A.O.; Conde, J.M.P.; Isoldi, L.A.; dos Santos, E.D. Development of a Computational Model for Investigation of and Oscillating Water Column Device with a Savonius Turbine. J. Mar. Sci. Eng. 2022, 10, 79. [Google Scholar] [CrossRef]
  25. Hosseini, A.; Goudarzi, N. CFD and Control Analysis of a Smart Hybrid Vertical Axis Wind Turbine. In Proceedings of the ASME 2018 Power Conference collocated with the ASME 2018 12th International Conference on Energy Sustainability and the ASME 2018 Nuclear Forum, Lake Buena Vista, FL, USA, 24–28 June 2018. [Google Scholar]
  26. Schlichting (Deceased), H.; Gersten, K. Boundary-Layer Theory; Springer: Berlin/Heidelberg, Germany, 2016; ISBN 978-3-662-52919-5. [Google Scholar]
  27. Sheldahl, R.E.; Blackwell, B.F.; Feltz, L.V. Wind Tunnel Performance Data for Two- and Three-Bucket Savonius Rotors; Sandia Laboratories: Albuquerque, NM, USA, 1977; Volume 2. [Google Scholar]
  28. Leonczuk Minetto, R.A.; Paraschivoiu, M. Simulation Based Analysis of Morphing Blades Applied to a Vertical Axis Wind Turbine. Energy 2020, 202, 117705. [Google Scholar] [CrossRef]
  29. Wilcox, D.C. Turbulence Modeling for CFD, 3rd ed.; DCW Industries: La Cañada, CA, USA, 2006; ISBN 978-1-928729-08-2. [Google Scholar]
  30. ANSYS, Inc. Ansys Fluent: User Guide; ANSYS: Canonsburg, PA, USA, 2022. [Google Scholar]
  31. Menter, F.R.; Kuntz, M.; Langtry, R. Ten Years of Industrial Experience with the SST Turbulence Model. Turbul. Heat Mass Transf. 2003, 4, 625–632. [Google Scholar]
  32. Menter, F.R. Zonal Two Equation Kappa-Omega Turbulence Models for Aerodynamic Flows. In Proceedings of the AIAA Fluid Dynamics Conference, Orlando, FL, USA, 6–9 July 1993. [Google Scholar]
  33. Versteeg, H.K.; Malalasekera, W. An Introduction to Computational Fluid Dynamics: The Finite Volume Method; Pearson Education Limited: London, UK, 2007; ISBN 978-0-13-127498-3. [Google Scholar]
  34. Banh Duc, M.; Tran The, H.; Dinh Duc, N.; Chu Duc, T.; Dinh Le, A. Performance Enhancement of Savonius Wind Turbine by Multicurve Blade Shape. Energy Sources Part A Recovery Util. Environ. Eff. 2023, 45, 1624–1642. [Google Scholar] [CrossRef]
  35. Meana-Fernández, A.; Oro, J.M.F.; Díaz, K.M.A.; Galdo-Vega, M.; Velarde-Suárez, S. Application of Richardson Extrapolation Method to the CFD Simulation of Vertical-Axis Wind Turbines and Analysis of the Flow Field. Eng. Appl. Comput. Fluid Mech. 2019, 13, 359–376. [Google Scholar] [CrossRef]
Figure 1. Computational domain (a) geometry and boundary conditions; (b) view of hybrid turbine; (c) view of Darrieus turbine; and (d) view of Savonius turbine.
Figure 1. Computational domain (a) geometry and boundary conditions; (b) view of hybrid turbine; (c) view of Darrieus turbine; and (d) view of Savonius turbine.
Jmse 13 01979 g001
Figure 2. Discretization of the rotational domain and refinement zones near the turbine walls. (a) hybrid turbine; (b) Darrieus turbine; (c) Savonius turbine.
Figure 2. Discretization of the rotational domain and refinement zones near the turbine walls. (a) hybrid turbine; (b) Darrieus turbine; (c) Savonius turbine.
Jmse 13 01979 g002
Figure 3. Comparison of numerical results from this work with experimental data reported in [5] for the averaged power coefficient of a Darrieus turbine.
Figure 3. Comparison of numerical results from this work with experimental data reported in [5] for the averaged power coefficient of a Darrieus turbine.
Jmse 13 01979 g003
Figure 4. Comparison of numerical results from this work with data reported in [23,24,27] for the averaged power coefficient of a Savonius turbine.
Figure 4. Comparison of numerical results from this work with data reported in [23,24,27] for the averaged power coefficient of a Savonius turbine.
Jmse 13 01979 g004
Figure 5. Richardson extrapolation of the average power coefficient for the three meshes.
Figure 5. Richardson extrapolation of the average power coefficient for the three meshes.
Jmse 13 01979 g005
Figure 6. Temporal series of instantaneous power coefficient of all blades for (a) T S R D   =   0.50 , (b) T S R D   =   1.00 , (c) T S R D   =   1.50 , and (d) T S R D   =   2.50 .
Figure 6. Temporal series of instantaneous power coefficient of all blades for (a) T S R D   =   0.50 , (b) T S R D   =   1.00 , (c) T S R D   =   1.50 , and (d) T S R D   =   2.50 .
Jmse 13 01979 g006
Figure 7. Schematic representation of the hybrid turbine configuration. The three Darrieus blades are identified (blade 01, blade 02, and blade 03), while the Savonius rotor is also indicated. The relative rotation angle between the y-axis and blade 01 θ b is highlighted.
Figure 7. Schematic representation of the hybrid turbine configuration. The three Darrieus blades are identified (blade 01, blade 02, and blade 03), while the Savonius rotor is also indicated. The relative rotation angle between the y-axis and blade 01 θ b is highlighted.
Jmse 13 01979 g007
Figure 8. Instantaneous moment coefficient for TSR = 0.50 and position of blade 01. (a) Darrieus blades, (b) Savonius configuration, and (c) hybrid and Darrieus blade 01.
Figure 8. Instantaneous moment coefficient for TSR = 0.50 and position of blade 01. (a) Darrieus blades, (b) Savonius configuration, and (c) hybrid and Darrieus blade 01.
Jmse 13 01979 g008
Figure 9. Instantaneous drag coefficient for TSR = 0.50 and position of blade 01. (a) Darrieus blades, (b) Savonius configuration, and (c) hybrid and Darrieus blade 01.
Figure 9. Instantaneous drag coefficient for TSR = 0.50 and position of blade 01. (a) Darrieus blades, (b) Savonius configuration, and (c) hybrid and Darrieus blade 01.
Jmse 13 01979 g009
Figure 10. Velocity magnitude field for TSR = 0.50, θ b   =   3.90 π . (a) Darrieus configuration, (b) Darrieus blade 01 zoom in, (c) hybrid configuration, (d) hybrid blade 01 zoom in.
Figure 10. Velocity magnitude field for TSR = 0.50, θ b   =   3.90 π . (a) Darrieus configuration, (b) Darrieus blade 01 zoom in, (c) hybrid configuration, (d) hybrid blade 01 zoom in.
Jmse 13 01979 g010
Figure 11. Pressure field for TSR = 0.50, θ b   =   3.90 π . (a) Darrieus configuration, (b) Darrieus blade 01 zoom in, (c) hybrid configuration, (d) hybrid blade 01 zoom in.
Figure 11. Pressure field for TSR = 0.50, θ b   =   3.90 π . (a) Darrieus configuration, (b) Darrieus blade 01 zoom in, (c) hybrid configuration, (d) hybrid blade 01 zoom in.
Jmse 13 01979 g011
Figure 12. Averaged power coefficient according to TSR value.
Figure 12. Averaged power coefficient according to TSR value.
Jmse 13 01979 g012
Table 1. Examples of configurations and geometric parameters used in the literature for hybrid Darrieus/Savonius turbines.
Table 1. Examples of configurations and geometric parameters used in the literature for hybrid Darrieus/Savonius turbines.
ParameterValues
Liang et al. [11]Asadi and Hassanzadeh [12]Chegini et al. [6]Redchyts et al. [13]Arrieta-Gomez et al. [14]
Darrieus diameter (DD)1500 mm1000 mm1030 mm7200 mm900 mm
Darrieus profileNACA 0012NACA 0018NACA 0021NACA 0020NACA 0018
Darrieus blade chord (C)220 mm220 mm858 mm708 mm200 mm
Number of Darrieus blades (ND)1, 2, 3, 42322
Ratio Darrieus/Savonius radius (DD/DR)0.2, 0.25, 0.330.20.19420.1250.2, 0.5, 0.8
Savonius profileSemicircleSemicircleSemicircleSemicircleSemicircle
Number of Savonius blades (NS)22222
Overlap ratio (s/c)0.10.20.03215 mm0 mm
Separation ratio (a/c)NI *NI *NI *0 mm0 mm
Thickness of Savonius blade (e)NI *0.63 mmNI *10 mmNI *
Attachment angles (α)0°, 45°, 90°30°0°, 45°, 90°, 135°
Wind speed (V)2 m/s5, 10 m/s9 m/s13 m/s1 m/s
Tip-Speed ratio (TSR)0.5–3.751.5, 2.5, 3.51.45–3.23.0 (Darrieus), 0.583 (Savonius)1.5, 2.0, 2.5
Working fluidAirAirAirAirWater
Computational methodFVMFVMFVMFVMFVM
Closure turbulence modelingkεkω SSTkω SSTSALSAkε RNG
Advection interpolation functionNI *2nd-order upwindNI *upwind Rogers-KwakNI *
Pressure-velocity treatmentNI *SIMPLESIMPLEArtificial compressibility modelNI *
Turbine rotation modelNI *Sliding meshSliding meshSliding meshNI *
* NI—Not informed.
Table 2. Boundary conditions and parameters of simulation.
Table 2. Boundary conditions and parameters of simulation.
ParameterSavoniusDarries/Hybrid
V   [ m / s ] 8.008.00
IT [%]1.001.00
R e D = ρ V D S , D / μ 1.73 × 1054.38 × 105
Dynamic   viscosity   μ   P a · s 1.7804 × 10−5
Density ρ [kg/m3]1.225
T S R = θ ˙ R / V 0.20; 0.30; 0.40; 0.50; 0.60; 0.70; 0.80; 1.000.50; 0.75; 1.00; 1.25; 1.50; 1.75; 2.00; 2.50
θ ˙ [rad/s]10; 15; 20; 25; 30; 40; 5010; 15; 20; 25; 30; 35; 40; 50
Time of simulation [s]3.5
Time for statistical analysis [s] 1.75     t     3.55
Table 3. Details of the meshes employed in the present study.
Table 3. Details of the meshes employed in the present study.
ParameterSavoniusDarrieusHybrid
Finite volumes261,832147,496335,989
Finite volume in rotational zone189,54377,988266,481
Inflation number in the near-wall regions484848
Inflation growth factor1.11.11.1
Maximum y+5.4002.4533.118
Minimum y+0.0230.2130.015
Averaged y+0.00360.8980.502
Minimum orthogonal quality0.0690.2340.031
Maximum skewness0.7220.6930.695
Maximum aspect ratio24.60955.46455.464
Table 4. Comparison of power coefficients (CP) obtained with different discretization schemes for one TSR of Savonius and Darrieus turbines.
Table 4. Comparison of power coefficients (CP) obtained with different discretization schemes for one TSR of Savonius and Darrieus turbines.
Parameter/WorkSavoniusDarrieus
R e D = ρ V D S , D / μ 867,000438,200
TSR1.001.00
CP—first-order upwind0.2330.158
CP—second-order upwind0.2270.209
CP—MUSCL0.259----
CP—Hybrid upwind scheme0.257----
Sheldahl et al. (1977) [27]0.228----
Akwa et al. (2012) [23]0.251----
Elkhoury et al. (2015) [5]----0.157
Table 5. Parameters of the mesh independence study for the three computational meshes.
Table 5. Parameters of the mesh independence study for the three computational meshes.
MeshMesh 1Mesh 2Mesh 3
Cells in streamwise direction300300300
Cells in cross-streamwise direction1306532
Grid increasing ratio (cross-streamwise direction)1.11.11.1
Number of volumes334,296286,889253,559
C P ¯   results   for   1.75 t 3.5 s0.046340.046090.04160
TSR0.500.500.50
Table 6. Averaged moment coefficient for TSR = 0.50 over a rotation.
Table 6. Averaged moment coefficient for TSR = 0.50 over a rotation.
Arrangement Moment   Coefficient   C m ¯
Blade 01Blade 02Blade 03Savonius BladeSum
Hybrid0.022230.010280.011920.048790.09323
Darrieus0.023240.023090.02404-0.07038
Savonius---0.310440.31044
Table 7. Averaged drag coefficient for TSR = 0.50.
Table 7. Averaged drag coefficient for TSR = 0.50.
Arrangement Drag   Coefficient   C d ¯
Blade 01Blade 02Blade 03Savonius BladeSum
Hybrid0.269020.249560.265130.188390.97210
Darrieus0.305690.298130.29909-0.90291
Savonius---0.922530.92253
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.

Share and Cite

MDPI and ACS Style

Inácio, R.G.d.S.; da Rosa, I.A.; Avila, V.H.; Rocha, L.A.O.; Isoldi, L.A.; Dias, G.d.C.; Gonçalves, R.A.A.C.; dos Santos, E.D. Numerical Investigation of Hybrid Darrieus/Savonius Vertical Axis Wind Turbine Subjected to Turbulent Airflows. J. Mar. Sci. Eng. 2025, 13, 1979. https://doi.org/10.3390/jmse13101979

AMA Style

Inácio RGdS, da Rosa IA, Avila VH, Rocha LAO, Isoldi LA, Dias GdC, Gonçalves RAAC, dos Santos ED. Numerical Investigation of Hybrid Darrieus/Savonius Vertical Axis Wind Turbine Subjected to Turbulent Airflows. Journal of Marine Science and Engineering. 2025; 13(10):1979. https://doi.org/10.3390/jmse13101979

Chicago/Turabian Style

Inácio, Rhuandrei Gabriel da Silva, Igor Almeida da Rosa, Vinicius Heidtmann Avila, Luiz Alberto Oliveira Rocha, Liércio André Isoldi, Gustavo da Cunha Dias, Rafael Adriano Alves Camargo Gonçalves, and Elizaldo Domingues dos Santos. 2025. "Numerical Investigation of Hybrid Darrieus/Savonius Vertical Axis Wind Turbine Subjected to Turbulent Airflows" Journal of Marine Science and Engineering 13, no. 10: 1979. https://doi.org/10.3390/jmse13101979

APA Style

Inácio, R. G. d. S., da Rosa, I. A., Avila, V. H., Rocha, L. A. O., Isoldi, L. A., Dias, G. d. C., Gonçalves, R. A. A. C., & dos Santos, E. D. (2025). Numerical Investigation of Hybrid Darrieus/Savonius Vertical Axis Wind Turbine Subjected to Turbulent Airflows. Journal of Marine Science and Engineering, 13(10), 1979. https://doi.org/10.3390/jmse13101979

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop