Next Article in Journal
A New Exoskeleton Prototype for Lower Limb Rehabilitation
Previous Article in Journal
Feature-Based Bearing Fault Classification Using Taylor–Fourier Transform
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Power Generation Enhancement of Horizontal Axis Wind Turbines Using Bioinspired Airfoils: A CFD Study

by
Hamid R. Kaviani
1 and
Mohammad Moshfeghi
2,*
1
School of Mechanical Engineering, Malayer University, Malayer P.O. Box 14395-515, Iran
2
Department of Mechanical Engineering, Sogang University, Seoul 04107, Republic of Korea
*
Author to whom correspondence should be addressed.
Submission received: 25 September 2023 / Revised: 24 October 2023 / Accepted: 25 October 2023 / Published: 30 October 2023
(This article belongs to the Special Issue Recent Developments in Offshore Wind Turbines)

Abstract

:
This research investigates the performance implications of employing a bioinspired airfoil (seagull’s wing cross-section) in horizontal-axis wind turbines. Specifically, we replaced the S809 airfoil from NREL Phase VI with an airfoil modeled after seagull wings. Initially, we calibrated four coefficients of the GEKO turbulence model for both the S809 and the bioinspired airfoil, utilizing experimental data. Subsequently, using the calibrated generalized k-ω (GEKO) model, we conducted a comparative analysis between the S809 and the seagull airfoils, revealing the considerable superiority of the seagull airfoil in terms of lift and drag coefficients. Furthermore, we numerically simulated the original NREL Phase VI turbine and a modified version where the S809 airfoil was replaced with the seagull airfoil using 3D computational fluid dynamics (CFD) with the airfoil-based-calibrated GEKO turbulence model. This investigation spanned a wide range of air speeds, including 7 m/s, 13 m/s, and 25 m/s. At these wind speeds, we observed a substantial increase in turbine power generation, with enhancements of 47.2%, 204.4%, and 103.9%, respectively. This study underscores the significant influence of nature’s designs in advancing energy extraction within industries, particularly within the wind energy sector.

1. Introduction

Until a few decades ago, horizontal-axis wind turbines (HAWTs) primarily employed airfoils originally developed for aviation applications, including the well-known NACA and NASA series [1]. However, the escalating interest in wind energy has prompted research institutions worldwide, such as the National Renewable Energy Laboratory (NREL) in the United States [2], the Delft Wind Energy Institute (DU-WIND) [3], the Risø DTU National Laboratory in Denmark [4], and the Energy Centre of the Netherlands (ECN) in the Netherlands [5], to engineer an array of specialized airfoils tailored to diverse wind turbine blade operating conditions. These airfoils, along with their refined iterations, have become the cornerstone for the design of modern industrial wind turbines.
Optimizing airfoil shapes serves multiple objectives, with the most critical one for wind turbines being the improvement of energy extraction from the wind [6,7,8]. However, it is essential to recognize that optimization has its inherent constraints, including:
(a)
Biasing toward the initial values, or the base airfoil shape [9].
(b)
Constraints imposed by the optimization algorithm, which encompass limitations associated with the dimensions of the search space [10], the constraints and unpredictability of mutations [11], and the probability for convergence at a locally optimal point [12].
(c)
Inherent limitations in geometry parameterization methods, which are employed to reduce the number of variables and expedite the optimization process [13].
Geometry parameterization methods play a crucial role in the optimization process by reducing the number of variables involved. To achieve this, mathematical functions, often of degree 6 or 7, are employed to represent the intricate airfoil shape. This results in a significant reduction in optimization variables, typically down to seven coefficients for a 6th-degree polynomial, from the multitude of points required to define an airfoil, which can exceed 60. Nonetheless, it is important to acknowledge that using polynomials to precisely represent airfoil shapes comes with five inherent limitations: completeness, flawlessness, orthogonality, parsimony, and intuitiveness [14]. These limitations manifest in practical applications. For instance, attempts to parameterize the cross-section of a seagull wing airfoil using the CST method, a well-established technique in this field, have proven to be challenging, yielding errors that exceed expectations. Similarly, alternative methods for parameterizing geometry also come with their own set of constraints. Thus, the limitations inherent in these approaches underscore the value of nature as a rich source of inspiration for the ongoing development of highly efficient airfoils.
In computational fluid dynamics (CFD) simulations, particularly concerning airfoils operating under stall conditions, the primary challenges revolve around accurately predicting the separation point and determining the size of the separated bubble. Given the significant cost associated with direct numerical simulation (DNS) and large eddy simulation (LES) methods [15], most engineering applications resort to utilizing Reynolds-averaged Navier–Stokes (RANS) equations, often coupled with Boussinesq’s hypothesis. It is worth noting that the predictions of RANS turbulence models exhibit variations in terms of the onset of flow separation [16]. These differing approaches have a substantial impact on the calculated aerodynamic coefficients, including drag and lift forces [17]. Additionally, there are notable disparities in modeling free shear flows [18]. While these distinctions may not significantly affect the modeling of simple flows, their consequences become apparent in complex flow scenarios [17]. Furthermore, certain turbulence models, such as SST-k-ω, demonstrate sensitivity to specific parameters, such as near-wall spacing [19].
RANS turbulence models underwent evaluation through research conducted by the AIAA Fluid Dynamics Technical Committee (FDTC) in a coordinated program. This study investigated the efficacy of nine models in predicting the aerodynamic coefficients of the NACA-0012 airfoil under stall conditions, which included models like SA, SST, and LRR/SSG. Notably, all models exhibited close predictions of aerodynamic coefficients at low angles of attack. The combined models, like the SST-k-ω and LRR/SSG, showed a slight advantage in lift coefficient prediction. However, when the angles of attack exceeded 17 degrees, all models faced limitations and failed to provide satisfactory results [20].
To address these challenges, Menter et al. [21] developed a novel model based on the k-ω formulation, referred to as the generalized k-ω (GEKO) turbulence model. They demonstrated that, through the calibration of specific coefficients referred to as “free coefficients,” GEKO offers a flexible and robust approach for turbulence modeling in a broad range of problems, particularly excelling at high angles of attack [21].
This paper aims to investigate enhancements in power generation using a bioinspired airfoil for a HAWT. To achieve this objective, the seagull’s wing airfoil has been substituted for the S809 airfoil in the NREL Phase VI HAWT, and the power generation of these turbines is compared. The information presented in the remainder of this article is structured as follows: Section 2 provides details about the GEKO turbulence model and its coefficients. It also includes information related to the calibration of GEKO for both the S809 and the bioinspired (seagull) airfoil, utilizing CFD simulations and experimental validation of the results. Section 3 describes the 3D CFD domain and the mesh arrangement used for CFD simulations of both the original 3D NREL Phase VI turbine and a new turbine constructed using the bioinspired airfoil. Section 4 is dedicated to the results of 3D CFD simulations and discussions about how the seagull airfoil enhances the power generation of this HAWT. Finally, conclusions and references are presented in Section 5.
It is important to note that the S1223 airfoil is inherently thin and may not be the default choice for multi-megawatt turbines. However, in the context of the present research, which primarily focuses on enhancing the aerodynamic performance of the wind turbine, the strategy employed here is rational and well-justified.

2. Generalized K-Omega (GEKO)

In a recent study, Menter et al. [21] introduced a revolutionary turbulence model known as the generalized k-ω model, abbreviated as GEKO. This model incorporates six parameters, commonly referred to as “free parameters.” These free parameters play a pivotal role in shaping the behavior of the model to match the flow characteristics in various regions, such as turbulent boundary layers, shear flows, and flow in corners, among others. The inclusion of these calibration coefficients, or free parameters, renders the GEKO turbulence model an adaptable and robust approach for RANS-based turbulence modeling, suitable for a broad spectrum of problems [21].
In the shear stress transfer model, turbulence kinetic energy (k) and specific dissipation rate (ω) are obtained from the following transport equations:
t ρ k + x j ρ k u j = x j ( μ + 1 σ k μ t ) k x j β * ρ k ω + P k
t ρ ω + x j ρ ω u j = x j ( μ + 1 σ ω μ t ) ω x j + C ω 1 F 1 ω k P k C ω 2 F 2 ρ ω 2 + ρ F 3 2 σ ω 1 ω k x j ω x j
where β* σ k , and  σ ω  represent empirical coefficients. Additionally,  u j   denotes velocity components, ρ stands for density, t represents time, μ represents molecular viscosity, and S represents the added source term. The turbulent eddy viscosity ( μ t ) is defined as:
μ t = ρ v t = ρ k m a x ω , S / C R e a l i z e d
where S is the invariant measure of the strain rate, and  C R e a l i z e d   is an empirical coefficient.
The limiting factor for turbulence production in stagnated areas concerning turbulence kinetic energy (Equation (1)) is defined as:
P k = τ i j U i x j
and
τ i j = τ i j E V C C o r n e r 12 μ t max 0.3 ω 0.5 S 2 Ω 2 × ( S i k Ω k j Ω i k S k j )
where  C C o r n e r  is another empirical coefficient. The term  τ i j E V   can be calculated via:
τ i j E V = ρ u i u j ¯ = μ t 2 S i j 2 3 ρ k δ i j
In the above equations, other variables are defined as:
S i j = 0.5 U i x j + U j x j ,       Ω i j = 0.5 U i x j U j x j i ,     S = 2 S i j S i j   ,             Ω = 2 Ω i j Ω i j  
The six free coefficients of the GEKO model are implemented through functions ( F 1 ,   F 2 , F 3 ). The default values used for these free coefficients are presented in Table 1.
It is worth noting that the separation parameter ( C S e p ) stands out as the most critical among the free parameters that significantly impact the behavior of the GEKO model, particularly when dealing with an airfoil in stall conditions. To ensure the utmost precision in our current research, we not only calibrated this pivotal parameter but also compared the accuracy of the GEKO method in simulating deep stall conditions with some other established turbulence models. Subsequently, we harnessed the GEKO method, incorporating the calibrated ( C S e p ) coefficient, to conduct flow calculations around the NREL Phase VI HAWT.

3. Case Studies Using S809 and S1223 Airfoils

In this research, the simulations of the NREL Phase VI turbine are conducted using both S809 and S1223 airfoils. The S809 airfoil serves as the primary airfoil for the NREL Phase VI HAWT. In contrast, the second airfoil, S1223, bears a close resemblance to a seagull’s wing cross-section and hence is employed for calibration of the seagull’s wing cross-section. The latter is then used throughout the 3D investigation in this study to illustrate its potential for enhancing the power generation of the NREL Phase VI turbine.
The S809 airfoil is notably characterized by its relatively thick profile and has been proven to exhibit remarkable performance at high angles of attack (AOAs), as well as in the near-stall and post-stall regimes. Furthermore, it maintains a commendably low drag coefficient, falling within the range of 0.44 to 0.77. This wide operational range makes it well-suited for power generation across various attack angles [22]. In contrast, the S1223 airfoil stands out as a thin airfoil with significant camber and high-lift characteristics. It has been subjected to testing in wind tunnels at both the University of Illinois [22] and the University of Santa Catarina [23]. The pronounced curvature of its camber reduces pressure on the suction surface while concurrently increasing pressure on the lower surface, leading to the generation of a high lift coefficient. Figure 1 illustrates the geometries of the three airfoils: S809, S1223, and the seagull airfoil.

3.1. GEKO Calibration for S809 and S1223 Airfoils

To calibrate the GEKO coefficients, two-dimensional simulations were conducted to evaluate the turbulence models’ accuracy and estimate the aerodynamic coefficients of airfoils in stall conditions. For the S1223 airfoil, the highest available Reynolds (Re) number for the lift coefficient experimental data was obtained from Selig and Guglielmo [22] at Re = 200,000. However, for the S809 airfoil, higher Re numbers were accessible. In the current study, Re = 2,000,000 was employed, which is in the range of the physical Reynolds number related to the NREL Phase VI blade in operation.
To facilitate this, a C-type mesh was employed in the CFD domain surrounding the airfoils. The domain’s height was maintained uniformly at 20C, with the inlet positioned at a distance of 10C from the airfoil and a downstream domain extending 30C. To enhance cell resolution in regions proximate to the airfoils, the CFD domains were subdivided into three distinct subregions: inner, middle, and outer.
A mesh sensitivity study was conducted using 2000 points along the chordwise direction of the airfoils. Meshes comprising 155,000 and 153,000 cells were found to exhibit mesh independence for the S809 and S1223 airfoils, respectively. The distribution of points along the chordwise direction was chosen such that mesh density at the leading and trailing edges was 20 times greater than at the airfoil’s midpoint. A 30-layer mesh was implemented for simulating the flow within the boundary layer, with the first layer’s height set at 5 × 10−6 (m) and an expansion ratio of 1.2. Cell sizes were adjusted to 0.008 C in the near area, 0.04 C in the middle area, and 0.16 C in the far area. Figure 2 provides a comprehensive view of the mesh arrangement near the airfoil and within the domain.
The inlet featured a constant velocity, while the outlet maintained a constant pressure equivalent to 1 atm. The top and bottom sides, along with the two normal-to-surface planes, were assumed to be symmetric. Spatial discretization methods employed the least squares cell-based method for gradients and the second-order method for pressure equations. For spatial discretization of momentum, intermittency, turbulence kinetic energy (k), and its specific dissipation rate (ω), a third-order MUSCLE approach was chosen.
The simulations were performed using steady-state assumptions. To expedite convergence, high-order term relaxation methods were employed. Convergence was deemed achieved when the results showed minimal fluctuations, with the absolute convergence criterion set to 10−5 for the residuals.
The GEKO turbulence model was calibrated for the separation parameter ( C S e p ) for application to the S809 and S1223 airfoils. Given the significance of model accuracy at high angles of attack (AOAs), calibration was performed at AOA = 20°, where both airfoils experience substantial separation, and the experimental data are available [22,24,25]. As detailed in Table 2 and Table 3, the GEKO turbulence model exhibited exceptional precision, with a separation factor of  C S e p  = 2.5 for the S809 airfoil and  C S e p    = 0.9 for the S1223 airfoil. It is worth noting that the performance of GEKO surpasses that of the SST-k-ω and Spalart–Almaras (SA) turbulence models significantly.
In Figure 3, a comparison is presented between the experimental data [22,24] and the computed values for the lift, drag, and lift-to-drag coefficients of the S809 and S1223 airfoils. The computed results align well with the experimental data, indicating a high level of agreement between the two.
Figure 4 compares the vortices generated by the GEKO model, employing the previously calibrated separation coefficients, and other turbulence models around the S809 and S1223 airfoils at AOA = 20°. These visual representations vividly showcase the distinctions in flow separation predictions across diverse turbulence models, offering valuable insights into the disparities in the lift and drag values predicted by each turbulence model.
The Q-criterion is employed in this research to visualize vortices. It identifies eddies as regions where the flow’s rotation component exceeds the strain [26] The Q-criterion is calculated as follows:
Q = 1 2 ( Ω i j Ω i j S i j S i j )
where  Ω i j  is the vorticity tensor and  S i j  is the strain rate tensor, which is defined as:
Ω i j = 1 2 ( u i x j u j x i )
S i j = 1 2 ( u i x j + u j x i )
Figure 5 elucidates the vortices generated by diverse turbulence models surrounding the S809 and S1223 airfoils. These vortices are visualized using the Q-criterion and color-coded based on velocity values. Notably, within both airfoils, we observe the presence of a discernible boundary layer characterized by blue to green regions, while the external flow exhibits hues ranging from green to red. The majority of methods have forecasted a heightened turbulence level within the boundary layer of the S1223 airfoil in comparison to the S809 airfoil. This contrast can be attributed to the more pronounced curvature along the centerline (camber line) of the S1223 airfoil, setting it apart from the S809 airfoil.
A conspicuous observation in Figure 5 is the severe reduction in vortex intensity and boundary layer thickness surrounding the S1223 airfoil, particularly at high angles of attack (AOA = 15° and 20°). This phenomenon implies an enhanced pressure recovery on the upper surface of the S1223 airfoil and a decrease in drag attributed to flow separation at elevated attack angles. The configuration of the boundary layer has a direct impact on velocity and pressure profiles, consequently influencing the airfoil’s aerodynamic coefficients. A thinner boundary layer facilitates a more favorable alignment of flow streamlines with the airfoil’s curvature, ultimately augmenting lift coefficients. Furthermore, the curved configuration of airfoils such as the S1223 provides an ideal surface for aerodynamic loading, further contributing to increased forces.
In Table 4 one can find a tabulation of lift ratios and lift-to-drag coefficients for the S1223 airfoil to the S809 airfoil across different angles of attack (AOAs). Table 5 presents a similar comparison of the seagull airfoil with respect to the S1223 airfoil.
As demonstrated in Figure 1, the S1223 and seagull airfoils are geometrically close to each other. This can also be seen in the aerodynamic results shown in Figure 6.

3.2. Validation of Simulation and Power Improvement 3D Wind Turbine

Domain Size, Mesh Density, and CFD Setting

Following the establishment of the seagull airfoil’s superiority over the S809, the subsequent phase entails the replacement of the S809 airfoil in the NREL Phase VI with the seagull airfoil, followed by a comprehensive comparison of the outcomes for the two wind turbines. The NREL Phase VI is characterized by a two-bladed horizontal-axis turbine, with the rotor positioned upstream of the tower. The turbine’s blade radius measures 5.05 m, and specific details regarding the local twist angles and chord lengths of the blades can be found in Table 6 [27].
The blades of both HAWTs were simulated within CFD domains of similar dimensions. In order to mitigate computational expenses, only one half of each turbine, effectively representing a single blade, underwent simulation (Figure 7). As it can be seen from this figure, the CFD model encompasses two distinct sub-domains: an inner subdomain encircling the blade and an outer subdomain. The dimensions of the domain were set based on a sensitivity analysis, not presented here, and the final length of the CFD domain was set to 70 m (=14R), with inner and outer subdomain radii of 7.5 m (=1.5R) and 15 m (=3R), respectively.
Regarding the boundary condition, a periodic boundary condition was imposed on the model. The inlet was defined by a constant velocity ranging from 7 to 25 m/s, while the outlet was treated as a constant pressure outlet (P = 1 atm).
In general, a higher mesh resolution was applied to the inner subdomain, while the outer subdomain featured a relatively lower mesh density. During the investigation for mesh independence, various mesh parameters were explored until it was determined that further increasing mesh density would not substantially influence the power output of the turbines. Specifically, cell counts along the airfoil’s perimeter were tested, ranging from 60 to 90 cells. It was observed that augmenting the number of cells beyond 80 did not result in any significant impact on the power output. To further refine the mesh resolution near the leading and trailing edges, a biasing technique with a factor of 20 was implemented.
In the final mesh arrangement, the first layer of the boundary layer was positioned at a distance of  5 × 10 6  m from the blade surface. Subsequently, 15 layers were generated with an expansion ratio of 1.6. With these settings, the total number of elements in the mesh was approximately  11 × 10 6 .
To ensure that the mesh satisfied the requisite near-wall spacing, the Y+ value was assessed on the blade surface at a nominal speed of 12 m/s. As illustrated in Figure 8, the maximum Y+ value on the blade surface was 0.5, which is below the stipulated threshold of one. Furthermore, another criterion involved an examination of the vortices generated around the turbine blade. This was conducted using the Q-criterion contour with a consistent value of 100,000 [s−2)] based on velocity values, as visualized in Figure 9. The presence of these vortices signifies the high quality of grid cells and adherence to the selected convergence criteria.

4. Results and Discussions

Initially, to validate the accuracy of the current CFD results, a comparison is made between the output power of the NREL turbine at various wind speeds obtained from this simulation and the results from other researchers [28,29], as well as the NREL experiment [27], as illustrated in Figure 10. The CFD results stemming from the current simulation, utilizing the S809 airfoil and the calibrated GEKO model, exhibit excellent agreement with the findings of other researchers and the experimental data (indicated by dashed lines). This alignment underscores the precision of the ongoing CFD simulation.
Subsequently, to explore the impact of the seagull airfoil and juxtapose it with the S809, the power generated by the two blades is compared in Figure 10 and Table 7. It is crucial to note that the separation coefficient ( C S e p ) in the GEKO model is set at 0.9 when utilizing the seagull wing airfoil, owing to its resemblance to the S1223 airfoil. In contrast, when employing the S809 airfoil,  C S e p  is set to 2.5. The results reveal that, in comparison to the original NREL Phase VI, the blade featuring the seagull airfoil demonstrates a substantial increase in the turbine’s output power. The power generated by the seagull airfoil-equipped blade exhibits a maximum enhancement of approximately 204% compared to the original NREL Phase VI, aligning with the superior performance of the seagull airfoil observed in 2D simulations, as detailed in the previous section.
The power coefficients ( C p ) of these two turbines (implanting the S809 and seagull airfoils) were computed using:
C p = p o w e r   0.5 ρ A v 3  
where  ρ  denotes the air density,  A  represents the area of the turbine’s rotor disk, and  v  indicates the wind speed. The maximum power that can be extracted from the wind by a turbine is known as the Betz limit, which has been determined to be 59.3% [30]. While the NREL Phase VI reached 39%, the same turbine achieved a maximum  C p  of 54% when employing the seagull airfoil (see Figure 11).
To delve deeper into the aforementioned discussion, a comparison of the pressure contours on the suction surface and the pressure surface of the S809 and seagull airfoils is provided. Figure 12, Figure 13, Figure 14 and Figure 15 visually portray the transformation in pressure distribution brought about by the alteration of the airfoil. This change results in an increase in pressure on the pressure surface and a decrease in pressure on the suction surface. These shifts in pressure distribution significantly contribute to the enhancement of the turbine’s aerodynamic power.

5. Concluding Remarks

In this study, the potential of bioinspired airfoils in enhancing the performance of wind turbines, particularly in the context of the NREL Phase VI wind turbine, was explored. To achieve this, the conventional S809 airfoil was substituted with a bioinspired airfoil modeled after a seagull’s wing. Remarkably, the shape of this seagull airfoil bore a striking resemblance to the S1223 airfoil.
The investigation commenced with a rigorous examination of the precision of using the GEKO turbulence model for the numerical modeling of flow over the S809 and S1223 airfoils, particularly in deep stagnation flow scenarios. For that, a comprehensive comparative analysis was conducted, pitting four turbulence models: Spalart–Almaras, SST-k-ω, IDDES, and the GEKO turbulence model—against each other. The results unveiled the superior performance of the GEKO method, specifically using a separation factor of  C S e p  = 2.5 for the S809 airfoil and  C S e p  = 0.9 for the S1223 airfoil.
Building upon these promising findings, a bold step was taken by replacing the S809 airfoil with the seagull airfoil in the NREL Phase VI wind turbine blade. The 3D fluid dynamics simulations, executed with the GEKO turbulence model, yielded results that underscored the transformative potential of this modification. A substantial increase in power generation was observed, with a 2.8 kW (47.2%) boost at a wind speed of 7 m/s, an astonishing 15.4 kW (204.4%) surge at a wind speed of 13 m/s, and an impressive 11.12 kW (103.9%) rise at the maximum operational speed of 25 m/s.
This research serves as a testament to the significant impact bioinspired airfoils can have on improving energy extraction in the wind energy industry. Furthermore, it underscores the reliability and value of the GEKO turbulence model for aerodynamic simulations, particularly for scenarios up to an AOA of 20°.

Author Contributions

Both H.R.K. and M.M. have contributed equally to all parts of this process of this research “Conceptualization, methodology, software, validation, formal analysis, investigation, resources, data curation, writing—original draft preparation, writing—review and editing, visualization, supervision, and project administration.” All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Data is unavailable due to privacy restrictions.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Timmer, W. An overview of NACA 6-digit airfoil series characteristics with reference to airfoils for large wind turbine blades. In Proceedings of the 47th AIAA Aerospace Sciences Meeting Including the New Horizons Forum and Aerospace Exposition, Orlando, FL, USA, 5–8 January 2009; p. 268. [Google Scholar]
  2. Tangler, J.; Somers, D. NREL Airfoil Families for Hawts; NREL/TP-442-7109; National Renewable Energy Laboratory: Golden, CO, USA, 1995.
  3. Bertagnolio, F.; Sørensen, N.; Johansen, J.; Fuglsang, P. Wind Turbine Airfoil Catalogue; Office of Scientific and Technical Information (OSTI), United States Department of Energy: Washington DC, USA, 2001.
  4. Fuglsang, P.; Bak, C. Development of the Risø wind turbine airfoils. Wind. Energy Int. J. Prog. Appl. Wind. Power Convers. Technol. 2004, 7, 145–162. [Google Scholar] [CrossRef]
  5. Grasso, F. Design of a family of new advanced airfoils for low wind class turbines. J. Phys. Conf. Ser. 2014, 555, 012044. [Google Scholar] [CrossRef]
  6. Akbari, V.; Naghashzadegan, M.; Kouhikamali, R.; Afsharpanah, F.; Yaïci, W. Multi-Objective Optimization and Optimal Airfoil Blade Selection for a Small Horizontal-Axis Wind Turbine (HAWT) for Application in Regions with Various Wind Potential. Machines 2022, 10, 687. [Google Scholar] [CrossRef]
  7. Huang, X.; Yang, J.; Gao, Z.; Sha, C.; Yang, H. Output Power and Wake Flow Characteristics of a Wind Turbine with Swept Blades. Machines 2022, 10, 876. [Google Scholar] [CrossRef]
  8. Li, G.; Dai, J.; Zhang, F.; Zuo, C. New Two-BWT Blade Aerodynamic Design and CFD Simulation. Machines 2023, 11, 399. [Google Scholar] [CrossRef]
  9. Liu, J.; Chen, R.; Lou, J.; Hu, Y.; You, Y. Deep-learning-based aerodynamic shape optimization of rotor airfoils to suppress dynamic stall. Aerosp. Sci. Technol. 2023, 133, 108089. [Google Scholar] [CrossRef]
  10. Nejat, A.; Kaviani, H.R. Aerodynamic optimization of a megawatt class horizontal axis wind turbine blade with particle swarm optimization algorithm. Modares Mech. Eng. 2017, 16, 1–11. [Google Scholar]
  11. Maxwell, M.J.; Sober, E. Gradualism, natural selection, and the randomness of mutation–fisher, Kimura, and Orr, connecting the dots. Biol. Philos. 2023, 38, 16. [Google Scholar] [CrossRef]
  12. Kaviani, H.R.; Moshfeghi, M. Multi-Megawatt Horizontal Axis Wind Turbine Blade Optimization Based on PSO Method. Aerospace 2023, 10, 158. [Google Scholar] [CrossRef]
  13. Kontogiannis, K.; Sóbester, A.; Taylor, N. Efficient parameterization of waverider geometries. J. Aircr. 2017, 54, 890–901. [Google Scholar] [CrossRef]
  14. Sripawadkul, V.; Padulo, M.; Guenov, M. A comparison of airfoil shape parameterization techniques for early design optimization. In Proceedings of the 13th AIAA/ISSMO Multidisciplinary Analysis and Optimization Conference, Fort Worth, TX, USA, 13–15 September 2010. [Google Scholar]
  15. Engelmann, L.; Hasslberger, J.; Baik, S.-J.; Klein, M.; Kempf, A. Direct numerical simulation of an unsteady wall-bounded turbulent flow configuration for the assessment of large-eddy simulation models. Sci. Rep. 2023, 13, 11202. [Google Scholar] [CrossRef] [PubMed]
  16. Madaliev, E.U.; Madaliev, M.E.U.; Mullaev, I.I.; Shoev, M.A.U.; Ibrokhimov, A.R.U. Comparison of Turbulence Models for the Problem of an Asymmetric Two-Dimensional Plane Diffuser. Middle Eur. Sci. Bull. 2021, 18, 119–127. [Google Scholar]
  17. Sogukpinar, H.; Bozkurt, I. Implementation of different turbulence model to find proper model to estimate aerodynamic properties of airfoils. In Proceedings of the Turkish Physical Society 33rd International Physics Congress (TPS33), Bodrum, Turkey, 6–10 September 2017; p. 020003. [Google Scholar]
  18. Singh, A.; Aravind, S.; Srinadhi, K.; Kannan, B. Assessment of Turbulence Models on a Backward Facing Step Flow Using OpenFOAM®. IOP Conf. Ser. Mater. Sci. Eng. 2020, 912, 042060. [Google Scholar] [CrossRef]
  19. Moshfeghi, M.; Song, Y.J.; Xie, Y.H. Effects of near-wall grid spacing on SST-K-ω model using NREL Phase VI horizontal axis wind turbine. J. Wind Eng. Ind. Aerodyn. 2012, 107, 94–105. [Google Scholar] [CrossRef]
  20. Langley Research Center Official Website. Available online: https://turbmodels.larc.nasa.gov/naca0012_val.html (accessed on 15 July 2023).
  21. Menter, F.; Lechner, R.; Matyushenko, A. Best Practice: Generalized k-ω Two-Equation Turbulence Model in ANSYS CFD (GEKO); Technical Report; ANSYS: Darmstadt, Germany, 2019; p. 27. [Google Scholar]
  22. Selig, M.S.; Guglielmo, J.J. High-lift low Reynolds number airfoil design. J. Aircr. 1997, 34, 72–79. [Google Scholar] [CrossRef]
  23. Abdelrahman, A.; Emam, A.; Adam, I.; Hassan, H.; Ookawara, S.; El-wardany, A. A Numerical Study into the Influence of Leading Edge Tubercles on the Aerodynamic Performance of a Highly Cambered High Lift Airfoil Wing at Different Reynolds Numbers. In Proceedings of the ASME International Mechanical Engineering Congress and Exposition, Online, 16–19 November 2020; p. V004T004A001. [Google Scholar]
  24. Somers, D. Design and Experimental Results for the S809 Airfoil; sr-440-6918; Technical Report; National Renewable Energy Lab (NREL): Golden, CO, USA, 1997.
  25. Reuss Ramsay, R.; Gregorek, G.M.; Hoffmann, M.J. Effects of Grit Roughness and Pitch Oscillations on the S810 Airfoil; Airfoil Performance Report, Revised (12/99); The Ohio State University: Columbus, OH, USA, 1999. [Google Scholar]
  26. Hunt, J.C.; Wray, A.A.; Moin, P. Eddies, streams, and convergence zones in turbulent flows. In Studying Turbulence Using Numerical Simulation Databases—II. Proceedings of the 1988 Summer Program; NASA Ames Research Center: Mountain View, CA, USA, 1988. [Google Scholar]
  27. Jonkman, J.M. Modeling of the UAE Wind Turbine for Refinement of FAST {_} AD; National Renewable Energy Lab.: Golden, CO, USA, 2003.
  28. Elfarra, M.A.; Sezer-Uzol, N.; Akmandor, I.S. NREL VI rotor blade: Numerical investigation and winglet design and optimization using CFD. Wind Energy 2014, 17, 605–626. [Google Scholar] [CrossRef]
  29. Duque, E.P.; Burklund, M.D.; Johnson, W. Navier-Stokes and comprehensive analysis performance predictions of the NREL phase VI experiment. J. Sol. Energy Eng. 2003, 125, 457–467. [Google Scholar] [CrossRef]
  30. Anderson, C. Wind Turbines: Theory and Practice; Cambridge University Press: Cambridge, UK, 2020. [Google Scholar]
Figure 1. Comparison of S1223, S809, and the seagull airfoil profiles.
Figure 1. Comparison of S1223, S809, and the seagull airfoil profiles.
Machines 11 00998 g001
Figure 2. Mesh arrangement around the airfoils and inside the boundary layer: (a) mesh layers perpendicular to the airfoil wall; (b) S1223 airfoil; and (c) S809 airfoil.
Figure 2. Mesh arrangement around the airfoils and inside the boundary layer: (a) mesh layers perpendicular to the airfoil wall; (b) S1223 airfoil; and (c) S809 airfoil.
Machines 11 00998 g002
Figure 3. Comparison between (a) lift coefficients versus AOA and (b) Cl versus Cd of S809 and S1223 at −15° < AOA < 20° and comparison between experimental results of Selig and Gugelielmo [22] Somers [24].
Figure 3. Comparison between (a) lift coefficients versus AOA and (b) Cl versus Cd of S809 and S1223 at −15° < AOA < 20° and comparison between experimental results of Selig and Gugelielmo [22] Somers [24].
Machines 11 00998 g003
Figure 4. Comparison between separation over the S809 and S1223 predicted by different turbulence models at AOA = 20°.
Figure 4. Comparison between separation over the S809 and S1223 predicted by different turbulence models at AOA = 20°.
Machines 11 00998 g004
Figure 5. Comparison between the level of separation over the S809 and S1223 at different AOAs predicted by the GEKO turbulence model.
Figure 5. Comparison between the level of separation over the S809 and S1223 at different AOAs predicted by the GEKO turbulence model.
Machines 11 00998 g005
Figure 6. Plot of (a) lift coefficient versus AOA and (b) lift versus drag the seagull airfoil and comparison with experimental data of Selig and Guglielmo [22].
Figure 6. Plot of (a) lift coefficient versus AOA and (b) lift versus drag the seagull airfoil and comparison with experimental data of Selig and Guglielmo [22].
Machines 11 00998 g006
Figure 7. The domain shape used for the simulation of the two wind turbines.
Figure 7. The domain shape used for the simulation of the two wind turbines.
Machines 11 00998 g007
Figure 8. Contour of Y+ values on the pressure (a) and suction (b) surfaces of the NREL Phase VI.
Figure 8. Contour of Y+ values on the pressure (a) and suction (b) surfaces of the NREL Phase VI.
Machines 11 00998 g008
Figure 9. Vortices behind the NREL Phase VI at constant values of Q = 100,000 [ s 2 ] colored with velocity values.
Figure 9. Vortices behind the NREL Phase VI at constant values of Q = 100,000 [ s 2 ] colored with velocity values.
Machines 11 00998 g009
Figure 10. A comparison between the CFD results of NREL Phase VI with S809 airfoil and the same turbine using seagull airfoil. The experimental and CFD results are extracted from Jonkman [27] Elfarra et al. [28] and Duque et al. [29].
Figure 10. A comparison between the CFD results of NREL Phase VI with S809 airfoil and the same turbine using seagull airfoil. The experimental and CFD results are extracted from Jonkman [27] Elfarra et al. [28] and Duque et al. [29].
Machines 11 00998 g010
Figure 11. The power coefficient of NREL Phase VI with S809 airfoil and the same turbine using seagull airfoil and comparison with the experimental results of Jonkman [27].
Figure 11. The power coefficient of NREL Phase VI with S809 airfoil and the same turbine using seagull airfoil and comparison with the experimental results of Jonkman [27].
Machines 11 00998 g011
Figure 12. Pressure contours on the suction surface of NREL-based blade with S809 airfoil.
Figure 12. Pressure contours on the suction surface of NREL-based blade with S809 airfoil.
Machines 11 00998 g012
Figure 13. Pressure contours on the suction surface of NREL-based blade with S1223 airfoil.
Figure 13. Pressure contours on the suction surface of NREL-based blade with S1223 airfoil.
Machines 11 00998 g013
Figure 14. Pressure contours on the pressure surface of NREL-based blade with S809 airfoil.
Figure 14. Pressure contours on the pressure surface of NREL-based blade with S809 airfoil.
Machines 11 00998 g014
Figure 15. Pressure contours on the pressure surface of NREL-based blade with S1223 airfoil.
Figure 15. Pressure contours on the pressure surface of NREL-based blade with S1223 airfoil.
Machines 11 00998 g015
Table 1. Default values of free coefficients of the GEKO model.
Table 1. Default values of free coefficients of the GEKO model.
  C C u r v   C C o r n e r   C J e t   C M i x   C N W   C S e p
1.01.00.91.00.51.75
Table 2. Calibration of the separation parameter ( C S e p ) of the GEKO for S809 and comparison with experimental data.
Table 2. Calibration of the separation parameter ( C S e p ) of the GEKO for S809 and comparison with experimental data.
Turbulence ModelLift Coef.Error (%)
1Experiment [24]0.8860
2GEKO,  C S e p  = 11.73695.95
3GEKO,  C S e p  = 1.751.28344.78
4GEKO,  C S e p  = 2.50.9537.60
5GEKO,  C S e p  = 3.250.771−12.96
6k-ω SST 1.41860.06
7Spalart–Almaras (2D)1.25942.07
Table 3. Calibration of the separation parameter ( C S e p ) of the GEKO for S1223 and comparison with experimental data.
Table 3. Calibration of the separation parameter ( C S e p ) of the GEKO for S1223 and comparison with experimental data.
Turbulence ModelLift Coef.Error (%)
1Experiment [22]2.040
2GEKO,  C S e p  = 0.42.88141.23
3GEKO,  C S e p  = 0.92.0812.01
4GEKO,  C S e p  = 11.733−15.05
5GEKO,  C S e p  = 1.751.351−33.77
6GEKO,  C S e p  = 2.51.339−34.36
7k-ω SST1.413−30.74
8Spalart–Almaras (2D)1.819−10.83
Table 4. Increment ratio of lift and lift-to-drag coefficients for S1223 compared to S809 at different AOAs.
Table 4. Increment ratio of lift and lift-to-drag coefficients for S1223 compared to S809 at different AOAs.
AOA (α=)Lift Increment Ratio
(%)
Lift-to-Drag Increment Ratio
(%)
−14°108%103%
−5°129%103%
675%106%
117%−47%
10°97%1%
15°109%75%
20°118%100
Table 5. Increment ratio of lift and lift-to-drag coefficients for seagull airfoil compared to S1223 at different AOAs.
Table 5. Increment ratio of lift and lift-to-drag coefficients for seagull airfoil compared to S1223 at different AOAs.
AOA (α=)Lift Increment Ratio
(%)
Lift-to-Drag Increment Ratio
(%)
−14°50.577.6
−5°127.4369.7
−0.6142.6
0.9150.0
10°−4.5119.2
15°−4.068.9
20°21.7163.1
Table 6. Local chord and twist angle of the NREL Phase VI at different radial locations.
Table 6. Local chord and twist angle of the NREL Phase VI at different radial locations.
Radius (m)Nondimensional Radius (r/R)Local Chord Length
(m)
Local Twist Angle
(deg.)
00.000.2180
0.5080.100.2180
0.660.130.2180
0.8830.180.1830
1.0080.200.3496.7
1.0670.210.4419.9
1.1330.230.54413.4
1.2570.250.73720.04
1.3430.270.72818.074
1.510.300.71114.292
1.9520.390.6667.979
2.2570.450.6365.308
2.3430.470.6274.715
2.5620.510.6053.425
2.8670.570.5742.083
3.1720.630.5431.15
3.1850.630.5421.115
3.4760.690.5120.494
4.0230.800.457−0.381
4.0860.810.451−0.475
4.6960.930.389−1.352
4.780.950.381−1.469
5.0291.000.355−1.815
Table 7. The power output of NREL phase VI HAWT.
Table 7. The power output of NREL phase VI HAWT.
Wind Speed (m/s)Experiment [28]Present CFD (S809 Airfoil)Present CFD (Seagull Airfoil)Power
Improvement (%)
76.015.928.7247.23
1010.109.8817.7779.93
139.807.5422.94204.38
158.788.5619.87132.19
208.359.3021.14127.30
2511.0711.6623.78103.93
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

Kaviani, H.R.; Moshfeghi, M. Power Generation Enhancement of Horizontal Axis Wind Turbines Using Bioinspired Airfoils: A CFD Study. Machines 2023, 11, 998. https://0-doi-org.brum.beds.ac.uk/10.3390/machines11110998

AMA Style

Kaviani HR, Moshfeghi M. Power Generation Enhancement of Horizontal Axis Wind Turbines Using Bioinspired Airfoils: A CFD Study. Machines. 2023; 11(11):998. https://0-doi-org.brum.beds.ac.uk/10.3390/machines11110998

Chicago/Turabian Style

Kaviani, Hamid R., and Mohammad Moshfeghi. 2023. "Power Generation Enhancement of Horizontal Axis Wind Turbines Using Bioinspired Airfoils: A CFD Study" Machines 11, no. 11: 998. https://0-doi-org.brum.beds.ac.uk/10.3390/machines11110998

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