1. Introduction
The number of articles tackling Active Flow Control (AFC) applied to airfoils is rising sharply [
1,
2,
3,
4,
5]. The AFC technology basically consists of adding or subtracting momentum to/from the main flow in order to interact with the boundary layer and delay its separation. Its main advantage over passive flow control approaches is that it is not creating a drag penalty in off-design conditions.
The different AFC techniques were divided by Cattafesta and Sheplak [
6] into three categories: (1) moving body actuators [
7], (2) plasma actuators [
8,
9,
10,
11], and (3) fluidic actuators (FA). From these three categories, FA are the only ones capable of generating large momentum coefficients, and this is why they were chosen for the present research. In some particular FA designs, the origin of the self-sustained oscillations was recently unveiled in [
12,
13,
14]. From the different FA, Zero Net Mass Flow Actuators (ZNMFA), also called synthetic jet actuators (SJA), have the advantage of not requiring external fluid supply, being particularly effective in controlling the separation of the boundary layer [
15,
16,
17,
18]. The use of pulsating flow has the advantage of coupling with the boundary’s layer natural instabilities and so being more energetically efficient [
19,
20,
21]. In fact, in order to minimize the energy required in AFC implementations, it is needed to somehow optimize the different parameters associated with each particular AFC application; nowadays, the most common methodology employed to tune the AFC-associated parameters is a parametric optimization, although the use of optimizers is increasingly accepted [
22,
23,
24].
Active Flow Control implementation in Vertical (VAWT) and Horizontal Axis Wind Turbines (HAWT) is recently gaining momentum [
1,
5,
25,
26,
27,
28,
29,
30,
31,
32], although so far, the bases to optimize the five AFC-associated parameters along the blade are not being established. Among the different devices employed in AFC applications on (WT), synthetic jets (SJ) appear to be particularly effective, as they combine the use of periodic forcing (blowing/suction) with a sufficiently large momentum coefficient
, which is defined as
, where
h is the jet width,
and
are the jet and far field densities, respectively,
is the maximum jet velocity,
C is the airfoil chord,
is the free-stream velocity, and
is the jet inclination angle with respect to the adjacent surface, a combination that appears to be particularly effective in delaying the boundary layer separation. The non-dimensional frequency associated with periodic forcing is defined as
, with
f being the boundary layer natural dimensional frequency.
A thorough experimental work employing SJ on wind turbine blade sections was undertaken by [
25], where it was established that AFC-SJ was capable of delaying airfoil stall and proved that a significant reduction of wind turbine start-up velocity was possible. In [
26,
33], SJ were employed to control the fluid flow vibration induced on turbine blades; they observed that the boundary layer reattachment obtained when applying AFC was having associated lower amplitude lift and drag dynamic coefficients and with it smaller structural vibrations. The use of SJ to enhance the performance of VAWT was experimentally addressed in [
27]; they obtained a considerable reduction of the noise generated while increasing the power generated by the turbine as well as its safety. Similar experimental studies using SJ but on a VAWT S809 and a HAWT S809 airfoil were performed by [
28] and [
31], respectively. The former study demonstrated that lift, drag, and pitching moment hysteresis could be significantly reduced. In the later research they observed that a 10.6% power input reduction to drive the rotor was achieved when a total of 20 SJA along each blade were employed.
Due to the high cost associated with experimental studies and the increasing trustworthiness of the computational simulations, the current tendency is to simulate the performance of any device. In this direction are going the 2D-URANS numerical simulations on VAWT performed by [
1,
29]. In both cases a straight-bladed WT was investigated, although the second one was of Darrieus type. Three different SJ locations were investigated in the later research, while in the former, not just the location but also the number of holes of the SJ were considered. A net power increase was obtained in both investigations; a power coefficient increase of 15.2% was obtained in [
1]. A numerical investigation on the S809-HAWT airfoil aerodynamic enhancement using SJ was undertaken by [
30]. Lift increase was observed for a wide range of injected flow angles. As previously performed by [
1], Wang et al. [
32] placed the SJA at the airfoil trailing edge and numerically studied the aerodynamic characteristics improvement on a straight-bladed VAWT. They evaluated the performance of several configurations of dual synthetic jets actuators (DSJA) to improve the power capacity of the WT and observed that the power coefficient could be increased by 58.87% when the juxtaposition-type DSJA was used. One of the latest pieces of research on AFC applications to HAWT with S809 airfoils is the one performed experimentally by Maldonado et al. [
5]. Two rotors with blade aspect ratios of 7.79 and 9.74 and having 20 SJA distributed along the span of each blade were analyzed at four different turning speeds and three pitch angles. The rotor with a higher aspect ratio was found to be more aerodynamically efficient, and thanks to the use of AFC, unsteady bending and unsteady torsion could be reduced by 36.5% and 22.9%, respectively.
In all the research just presented on WT, the five AFC parameters—SJ position, width, injection angle, frequency, and momentum coefficient—were chosen without performing a previous optimization, which appears to be the necessary task if minimum energy for the actuation is to be used. In fact, when aiming to optimize the different AFC parameters, the following considerations gathered from previous investigations are needed.
When experimentally studying the flow control on the NACA-0025 airfoil at
=
and AoA =
, Goodfellow et al. [
34] observed that the momentum coefficient was the primary control parameter. When analyzing the effect of jet position in controlling the boundary layer separation [
35], it was observed that the maximum effectiveness was obtained when placing the jet groove close to the boundary layer separation point, either downstream or upstream. The same observations were made by Amitay et al. [
36] and Amitay and Glezer [
37] which observed that locating the actuator near the boundary layer separation point, a lower momentum coefficient was needed to reattach the separated flow. Furthermore, Feero et al. [
38] observed that the required
to reattach the boundary layer was the lowest for excitation frequencies in the range of the vortex shedding frequency. The same observation was made by Tuck and Soria [
39] then they reported optimal actuation frequencies of
= 0.7 and
, the highest one being most effective in combination with an optimum momentum coefficient of
= 0.0123. Kitsios et al. [
40] conducted a LES study on a NACA-0015 airfoil at
=
and noticed that the optimal pulsating frequency coincided with the baseline shedding frequency (
). This was also experimentally confirmed by Buchmann et al. [
41] and corroborated using (3D-DNS) by Zhang and Samtaney [
42]. Numerical studies on the NACA-23012 airfoil at
=
were performed by [
43,
44] and concluded that maximum lift was obtained when
= 1, the jet was placed nearby the boundary layer separation point, and tangential injection/suction was observed to be particularly effective.
The numerical work performed in the present manuscript describes a novel methodology based on parametric analysis optimization to maximize the aerodynamic efficiency of a given airfoil section. The method is applied to an airfoil section belonging to the DTU10MW Horizontal Axis Wind Turbine (HAWT). By considering the methodology here outlined and extending it to all Wind Turbine (WT) airfoil sections where the boundary layer is separated, the energy required to attach the boundary layer in all WT sections shall be minimized, therefore maximizing the energy increase obtained by any WT. It is particularly interesting to highlight that the attachment of the boundary layer along the blade span not only increases the aerodynamic efficiency of the blade and the wind turbine power but also decreases the downstream wake generated by the wind turbine. Many recent studies [
45,
46,
47] are analyzing the power loss suffered by wind turbines placed in wind farms due to the wake generated by turbines located upstream; therefore, the wind turbine wake reduction proposed in the present research may lead to a decrease of the distance consecutive wind turbines can be located in wind farms as well as a power increase of wind turbines placed downstream.
The remainder of the paper is structured as follows: The problem formulation, numerical methods, and mesh independence study are presented in
Section 2.
Section 3 serves to define the WT’s main characteristics and the airfoils selected to be simulated via CFD. The baseline case study of the different sections, the complete parametric analysis of the chosen airfoil, and the energy assessment are to be found in
Section 4. The research performed is summarized in
Section 5.
2. Numerical Method
2.1. Governing Equations and Turbulence Model
The Navier–Stokes (NS) equations under incompressible flow conditions take the following form:
As Reynolds number increases, the mesh and the associated time step NS equations require drastic decreases in order to reach the Kolmogorov length and time scales, but reaching such scales involves extremely large computational times; such simulations are called Direct Numerical Simulations (DNS). Due to its drawbacks, nowadays DNS is just used for research purposes and for relatively small Reynolds numbers. For highly turbulent flow and to obtain reasonably accurate results, turbulence models are still needed. The use of Large Eddy simulation (LES) as a turbulence model is a very good option in 3D flows, yet it still requires very fine meshes and large computational times; therefore, LES simulations require supercomputers to shorten the computational time. As a result, simulations at large Reynolds numbers still need to be performed using Reynolds Averaged Navier Stokes (RANS) or URANS unsteady-RANS turbulence models. Their precision is not as accurate as LES or DNS, but the computational time needed shortens drastically.
Under incompressible flow conditions, the only variables associated with the NS equations are the pressure and the three velocity components, which are generically called
. In order to be able to apply URANS models, each variable from the NS equations needs to be substituted by its average
and a fluctuation
term.
After the substitution, the resulting NS equations (considering just two dimensions) take the following form:
It is now relevant to consider the concept of the apparent Reynolds stress tensor (), which comprises a symmetric matrix encompassing the time-averaged fluctuations in the x, y, and z directions. It is important to observe that in the case of two-dimensional flows, this matrix explicitly consists of four terms, whereas in three dimensions, it expands to a symmetric matrix.
Some RANS turbulence models are based on solving the Navier–Stokes equations by incorporating the concept of turbulence viscosity (
), which can be mathematically implemented into the momentum equations described above using the subsequent definition (called the Boussinesq hypothesis):
The X and Y momentum terms of the NS equations can therefore be expressed as follows:
We have chosen to use the
k-
SST turbulence model, which relies on using the
k-
model near the wall, the
k-
model far away from the object, and a blending function between these two. Mathematically,
is calculated as follows:
As stated, this model is a two-transport equation model in order to solve
k and
. The transport equations used to define each parameter are as follows:
where the turbulent stress tensor
is defined as
The SST model combines the constants of the
k-
and
k-
turbulence models, defined as
and
, respectively, using a blending function
, which depends on the distance of the first cell to the wall. The generic values taken by the different constants
are blended as follows:
where
is the blending function defined as
being the argument arg1 given by
The constants
and
employed by the corresponding turbulent model are having the following values:
,
,
,
,
,
.
,
,
,
. For an in-depth understanding of the proposed
k-
SST turbulence model, the reader should refer to [
48,
49].
2.2. Non-Dimensional Parameters
Reynolds Number
The Reynolds number is defined as the ratio of inertial to viscous forces. Mathematically, it is expressed as follows:
where
is the relative velocity,
C the chord length, and
the fluid absolute viscosity.
Aerodynamic Coefficients
The most common aerodynamic coefficients are the lift
and drag
ones, which are mathematically expressed as follows:
L and D represent the lift and drag forces, respectively.
The aerodynamic efficiency is the ratio between the lift and drag coefficients:
Pressure Coefficient
It characterizes the relative pressure field () along the body.
It is mathematically expressed as
where (
) is defined as the dynamic pressure.
Friction Coefficient
The friction coefficient, also known as the skin friction coefficient or tangential friction coefficient, evaluates the frictional force per unit area
acting on a surface in contact with the fluid.
Momentum Coefficient
It represents the momentum associated with the jet divided by the incoming fluid momentum. According to reference [
34],
is the primary AFC parameter. Mathematically it is expressed as follows:
where (
h) is the jet width, (
) is the maximum jet velocity, and (
) is the jet inclination angle with respect to the airfoil surface.
Forcing Frequency
The forcing frequency, given as a non-dimensional number, is defined as follows:
where
f is an arbitrary frequency and
is the vortex shedding frequency (the frequency at which vortices are shed from the airfoil).
Courant–Friedrichs–Levy Number
The Courant number, often denoted as
, quantifies how the numerical scheme/setting correctly captures the propagation of the information in the flow due to its velocity. Mathematically, it is defined as the fluid velocity measured at any given mesh cell multiplied by the simulation time step and divided by the mesh cell length scale (
).
For time-dependent simulations, the Courant number is an important parameter in order to ensure the stability of the system, and it is recommended to keep it below 1.
Dimensionless Wall Distance
The dimensionless wall distance
represents the distance from the wall to the first mesh cell,
y, normalized by a characteristic velocity
and the relative viscosity of the fluid
. Mathematically, it is expressed as follows:
where
is the friction velocity, defined as
2.3. Numerical Domain and Boundary Conditions
For computational domains relatively small in respect to the bluff-body characteristic length, the flow around the bluff-body is affected by the boundary conditions defined at the boundaries. Conversely, if the domain is chosen to be too large, there may be excessive mesh cells, resulting in longer computational times and poor simulation efficiency. Therefore, it is crucial to choose an appropriate domain size that balances the need for accurate results with computational efficiency.
Based on the previous research undertaken by [
22,
24,
50], the computational domain employed in the present research has the following dimensions:
from the domain inlet to the airfoil leading edge,
from the trailing edge to the domain outlet, and
from the body surface to both the upper and lower limits of the domain,
C being the airfoil chord; see
Figure 1.
Two points,
P and
Q, delimiting the inlet and outlet boundaries, have been implemented in
Figure 1. The outlet boundary extends from
P to
Q in a clockwise direction, and from
P to
Q in a counterclockwise direction is to be found the inlet boundary. The curvilinear shape of the domain at the inlet has been selected to facilitate the simulation of various angles of attack without altering the mesh geometry. It should be noted that the airfoil has been horizontally positioned, and by adjusting the inlet velocity components, any required angle of attack can be obtained. This approach not only allows flexibility in the simulation but also ensures that the mesh remains consistent throughout the analysis.
The present research focuses on simulating external aerodynamics with transient turbulent flow (aerodynamic forces change over time) and at large Reynolds numbers of the order of . Additionally, the angle of attack is relatively high for the two baseline case sections studied, which, particularly for the section placed nearest the blade root (Section 54 of the DTU-10MW-HAWT), is expected to produce the separation of the boundary layer at some point along the chord, generating the corresponding vortical structures and vortex shedding. Moreover, since the relative velocity in both sections studied does not exceed the incompressible flow limit, which is commonly accepted as , the flow will be considered as incompressible across the entire study. The flow solver contained in the OpenFOAM package, which fulfills the needs for the present research, is PisoFoam (Pressure implicit with splitting operator). This solver is particularly well suited for simulating unsteady turbulent flow and is widely used in aerodynamic simulations due to its accuracy and efficiency.
Section 54 with a wind speed of 10 m/s was chosen to perform the mesh independence study. Notice that the Reynolds number is almost the same in both sections, and the angle of attack is much larger in this particular one. The main data relative to the two sections studied, Sections 54 and 81, are presented in
Table 1, which includes the airfoil type, the radius versus the turbine main axis, the section number, the chord length, the Reynolds number, the angle of attack, and the relative velocity, which is the vectorial magnitude obtained from the combination of the wind and turning speeds and when considering the axial (a) and angular (a’) induction factors.
Figure 2 introduces the approximate location of the two sections under study as well as their respective profile.
The different boundary conditions employed to simulate section 54, regardless of the mesh density considered, are outlined in
Table 2. It’s worth noting that these specific values are based on the
k-
SST turbulence model, aligned with the corresponding chord length and a turbulence intensity at the inlet of
and according to [
48,
49]. In this paper, all simulations were performed using air at sea level based on the International Standard Atmosphere (ISA). The two relevant air parameters, density and kinematic viscosity, are summarized in
Table 3.
2.4. Mesh Assessment
The blade considered in the present research is the DTU-10MW offshore HAWT and consists of six different airfoil types, which, seen from the central axis to the blade tip, see
Figure 3, are FFA-W3-600; FFA-W3-480; FFA-W3-360; FFA-W3-301; FFA-W3-241; and NACA0015. Note that DTU implemented a Gurney flap at the trailing edge of the FFA-W3-600 airfoil type, so in terms of meshing, this modification must be considered. The mesh independence study was performed using Section 54, which belongs to the FFA-W3-301 airfoil type. For a wind speed of 10 m/s, the Reynolds number associated with this section is
, and considering the angle of attack associated, boundary layer separation is expected to occur.
For the FFA-W3-301 airfoil family, a C-type body-fitted structured mesh has been chosen for the region close to the airfoil. The C-type mesh comprises a curvilinear line surrounding the solid geometry, which is closed on the leading edge of the airfoil but open on the trailing edge. The region between the airfoil and the C-line is where the structured mesh is allocated; this region is usually called the “halo.” The halo thickness employed in the present research and measured perpendicular to the airfoil surface is about
of the chord; this value was kept constant for the two airfoils considered. A general view of the computational domain with the hybrid mesh employed is presented in
Figure 4a. A zoomed view of the structured mesh zones at the airfoil’s leading and trailing edges is presented in
Figure 4b, where the details of the mesh at the trailing edge are observed. A small, rectangularly structured domain has been employed at the trailing edge region to properly capture the vortical structures due to possible flow detachment. The remaining portion of the computational domain has been designated with a non-structured mesh; see
Figure 4a. This choice is rooted in the understanding that a certain level of reduced accuracy in regions distant from the airfoil surface is acceptable. Furthermore, opting for a hybrid mesh enables a more favorable balance between computational efficiency and accuracy of the results. Notwithstanding this, a heightened cell density has been strategically implemented in the vicinity of the airfoil to attain the desired level of precision.
The implementation of the AFC technology using synthetic jets (SJ) requires the modification of the mesh in the location where the SJ is to be implemented.
Figure 5b presents a zoom view of the mesh around the SJ, and
Figure 5a introduces the main SJ-associated parameters that will be tuned via the corresponding parametric optimization.
To ensure maximum accuracy in the simulated results, a thorough investigation of the mesh density is required, particularly in the halo region near the airfoil surface where the boundary layer develops and it is likely to separate. In this region, a drastic velocity and shear stress evolution in the normal direction to the airfoil surface is to be expected. Due to the lack of prior experimental investigations considering the Reynolds number and the airfoil type chosen and noting that the previous numerical investigations performed by reference [
51] presented the time-averaged value for
and
, a mesh independence study was performed to ensure a dissociation between mesh resolution and the outcomes produced, thus establishing a robust foundation for further analysis. The assessment for mesh independence involves generating four distinct meshes, all sharing identical overall geometries (depicted in
Figure 4a), while varying the targeted
values. For the purpose of result validation, the approach chosen involves examining the time-averaged
and
, as well as the distributions of
and
along the chord for each mesh resolution. This comparative study seeks to identify the
value at which substantial result deviations are absent.
The simulations carried out encompassed a
duration, wherein the average lift and drag coefficients were derived from a sampling window spanning the concluding
s of the simulation. The numerical outcomes of the averaged aerodynamic coefficients, coupled with the corresponding relative error denoted as Error
in percentage and versus the value obtained by [
51], are presented comprehensively in
Table 4. The number of cells used for each mesh, the minimum and maximum
at the airfoil surface, the airfoil efficiency
, as well as the boundary layer separation point denoted as
, are presented in the same table. Upon careful examination, it becomes evident that as the maximum
is reduced, the error in the lift coefficient drastically drops, becoming negligible between the two finest meshes. For instance, the discrepancy stands at
between the
obtained for maximum
and the value obtained by [
51]; such discrepancy is reduced to
when comparing the reference value and the one obtained in meshes C and D. It is relevant to note that meshes C and D generate identical results for
and for the boundary layer separation
, and minor discrepancies are observed in the
values generated by these two meshes. A discrepancy of around 32% is observed when comparing the reference
with the one obtained in mesh C; this large discrepancy is understood when noting that the airfoil chord used in reference [
51] was the unity, the Reynolds number being
. In the present study, the chord associated with Section 54 is 6.2 m, and the Reynolds number is
, therefore clarifying why the drag coefficient, is a bit larger in the present study. Considering the notable accuracy achieved by Mesh C and the manageable computational time entailed in its simulation, it has been judiciously chosen as the preferred candidate for subsequent simulations.
For a visual and perhaps more precise comparison of the outcomes, an illustrative representation has been generated to showcase the variation in pressure and friction coefficients across the chord length and for the four distinct tested meshes, as exhibited in
Figure 6a,b. Examining the pressure distribution over both the upper and lower surfaces of the airfoil reveals a remarkably uniform behavior across all the meshes. Notably, there exists a consistent and gradual pressure recovery on the upper surface. The sole discernible divergence in pressure coefficients among all the meshes manifests near the
m
mark along the chord. Nevertheless, the graph depicting
presents a strikingly analogous pattern for all the meshes. In the realm of the friction coefficient graph, as depicted in
Figure 6b, distinctions emerge among the meshes. Notably, the lower
meshes, namely Mesh A and B, exhibit some variations in the boundary layer separation point,
. Conversely, Mesh C closely mirrors the
trends observed in Mesh D, aligning with the minor errors observed in the averaged aerodynamic coefficients. This alignment renders Mesh C the optimal selection for future simulations.
3. Main Parameters Definition of the DTU 10MW RWT
In order to increase the accuracy of the 2D-CFD simulations, it is important to consider the flow’s three-dimensional effects over the airfoil section under study. Two induction factors, axial
and tangential (angular)
, which consider some flow three-dimensional effects, are used to more precisely calculate the axial and radial velocities for 2D simulations [
52]. The decrease in streamwise velocity from the free-stream conditions to the rotor plane is called the axial velocity deficit, and the fraction by which it is reduced is called the axial induction factor (
a). The tangential induction factor
considers the increase of the tangential speed in the rotor plane.
Figure 7 introduces the typical velocity and force components acting on a generic HAWT airfoil, and considering the corresponding induction factors, the main angles associated are also depicted. The aerodynamic forces (lift and drag) are decomposed into radial and tangential directions. For the present study, the values of the induction factors corresponding to each of the two airfoil sections studied, Sections 54 and 81, were obtained from [
53].
Additionally, for the computation of essential metrics such as net force, torque, angle of attack, and relative wind speed for each blade section, as described by Formulas (27)–(30) from [
54,
55], the framework necessitates the establishment of two distinct angles. These angles serve as pivotal factors in the analytical processes. The initial angle, referred to as
, corresponds to the angle formed between the section chord and the plane of rotation, designated as the section twist angle. This angle, which is different for each section, is defined in the design process of each HAWT and cannot be modified. In fact, and in order to obtain the maximum power for each given wind speed, HAWT traditionally uses the blade pitch angle, not depicted in
Figure 7; note that the pitch angle is constant for all blade sections and just depends on the wind speed. For the present DTU10MW RWT, when the wind speed is 10 m/s, the pitch angle is zero. Lastly, the other angle that plays a crucial role in the calculations is labeled as
and characterizes the inflow angle. This angle quantifies the divergence between the relative wind speed of each blade section and the plane of rotation, as delineated in
Figure 7. Notice that the angle of attack (AoA) is given as
.
Based on the information just presented, the relation between the relative velocity
, the free stream velocity
, and the tangential one
,
r being the section (airfoil) radius and
standing for the wind turbine turning speed, is given as Equation (
27).
The equation defining the angle of attack (given in degrees) for any blade section takes the following form:
For a given section, the resultant force acting on the rotor plane direction
and the force acting in a plane perpendicular to it
, are expressed as.
From Equation (
29), several insights can be derived regarding the significance of lift and efficiency across different parts of the blade: when dealing with high inflow angles at the root region (close to 90 deg.), the contribution of the lift coefficient becomes more significant than that of drag in torque generation. Conversely, at the tip, the drag coefficient (and hence the aerodynamic efficiency) plays a more prominent role, making it crucial to enhance efficiency in this region. Based on the Angles of Attack (AoA) associated with each section and the aerodynamic data provided by reference [
56], it is estimated that approximately
of airfoils will experience separated flow at a wind speed of 10 m/s, making it a good regime for enhancing the aerodynamic capabilities via AFC implementation. It should be noted that other wind velocities also have a high number of airfoils in post-stall conditions, which opens a door into improving the WT efficiency and power generated.