You are on page 1of 11

International Journal of Aeronautical and Space Sciences

https://doi.org/10.1007/s42405-020-00346-8

ORIGINAL PAPER

Aerodynamic Performance Enhancement of Tiltrotor Aircraft Wings


Using Double‑Row Vortex Generators
Hao Chen1 · Bo Chen2

Received: 3 August 2020 / Revised: 24 November 2020 / Accepted: 18 December 2020


© The Korean Society for Aeronautical & Space Sciences 2021

Abstract
To improve the aerodynamic performance of tiltrotor aircraft, the flow characteristics of the V-22 airfoil without and with
one, two parallel rows of vortex generators (VGs) were investigated using numerical simulation methods. The results were
obtained with three-dimensional compressible Reynolds-averaged Navier–Stokes equations, and the turbulence was simulated
with the SA-based DES model. The influence of the chordwise installation of VGs was emphatically discussed. First, the
single-row VGs are, respectively, mounted at the 10, 20, and 40% chord positions, denoted as VGs1, VGs2, and VGs3. The
results indicate that with the addition of the optimal configuration (VGs1), the maximum lift coefficient is extended from
1.43 to 2.44, and the stall angle of attack increases from 10° to 22° with respect to the clean airfoil. The drag coefficient of the
airfoil with VGs1 is 60.8% lower than that of the airfoil without VGs at an angle of attack of 24°. On this basis, the double-
row VGs’ arrangements are located at the 10 and 30% chord positions (VGs4), the 10 and 50% chord positions (VGs5) and
the 10 and 70% chord positions (VGs6). Compared to single-row VGs, double-row VGs have the greater potential to suppress
the flow separation. Due to the effect of VGs4, the maximum value of lift coefficient has increased to 2.61. However, the
case of VGs6 degrades the overall aerodynamic performance as compared to VGs1. Further research suggest that with the
increase in the height of the second row of vortex generators (VGs7), the aerodynamic characteristics of airfoil A821201
can be effectively enhanced at large angles of attack.

Keywords Tiltrotor aircraft · Vortex generators · Aerodynamic characteristics · Numerical simulation

1 Introduction devices are necessary to improve aerodynamic performance


of tiltrotor aircraft.
Tiltrotor aircraft such as the XV-15 and the V-22 are unique In the last few decades, various flow control techniques
in that they combine the vertical take-off and landing capa- have been proposed to reenergize the boundary layer, among
bility of the helicopter with the efficient high-speed cruise which the vortex generators are much cheaper and easier to
performance of fixed-wing turboprop. Due primarily to install. A vortex generator is a small vane which is typically
structural issues, the span of these wings is relatively small, attached to the suction side of airfoil, where it causes local
and the airfoil has to be thick enough to allow sufficient mixing in the boundary layer and thereby can prevent flow
internal volume for the drive mechanism to the tip-mounted separation. Extensive studies have been conducted to assess
and hinged rotors. These high thickness airfoils are particu- the effects of VGs. In the aspect of experimental investiga-
larly susceptible to stall and the wing stall affects the width tions [2–5], abundantly measured data had been obtained
of the conversion corridor and limits the flight path man- which provided great assistance to validations and evalua-
agement during climb and transition [1]. Thus, flow control tions about the aerodynamic characteristics of VGs. How-
ever, these tests are costly and time-consuming. Alterna-
* Hao Chen tively, one can use advanced computational fluid dynamics
chenhao@hyit.edu.cn (CFD) methods to understand and quantify the influence of
VGs. Although there exist some low-fidelity VG calculation
1
Huaiyin Institute of Technology, Huaian 223003, Jiangsu, model, such as the vortex source model [6, 7] and the lifting
China
force approach [8], there are not able to accurately model
2
Water Conservancy Bureau of Hongze District, the combined effect of multiple vortex generators [9]. Such
Huaian 223100, Jiangsu, China

13
Vol.:(0123456789)
International Journal of Aeronautical and Space Sciences

analysis demands high-fidelity methods, i.e., methods based 𝜕W 𝜕f 𝜕g 𝜕q 𝜕R 𝜕S 𝜕T


on a solution of the Navier–Stokes equations that compute + + + = + + , (1)
𝜕t 𝜕x 𝜕y 𝜕z 𝜕x 𝜕y 𝜕z
both vortices and the pressure field accurately. Some notable
works include those of Namura et al. [10], Lee and Kwon where W is the vector of conservative variables, and defined
[11], and Suarez et al. [12]. as:
Different VG parameters play important roles in the
streamwise vortex evolution downstream of VGs, and even- ⎡𝜌 ⎤
⎢ ⎥
tually impact the aerodynamic performance of airfoils. God- ⎢ 𝜌u ⎥
ard and Stanislas [13] found that the counter-rotating device W=⎢ 𝜌v ⎥, (2)
⎢ ⎥
appears more effective than the co-rotating one to suppress ⎢ 𝜌w⎥
the flow separation. Namura et al. [14] observed that the VG ⎢ ⎥
⎣ 𝜌E ⎦
spacing significantly affects the shock wave location. If the
spacing is narrow, the shock wave moves downstream, which f, g, and q are the convective flux vectors defined as:
increases the lift coefficient at a high angle of attack. Tai
[15] used the Reynolds-averaged Navier–Stokes method to ⎡ 𝜌u ⎤ ⎡ 𝜌v ⎤ ⎡ 𝜌w ⎤
calculate the flow over the wing–fuselage–nacelle configu- ⎢ 2 ⎥ ⎢ 𝜌vu ⎥ ⎢ 𝜌wu ⎥
⎢ 𝜌u + p ⎥ ⎢ ⎥ ⎢ ⎥
ration in forward flight with and without VGs. The results
f = ⎢ 𝜌uv ⎥, g = ⎢ 𝜌v2 + p ⎥, q = ⎢ 𝜌wv ⎥.
demonstrated that the chordwise location and incidence ⎢ ⎥ ⎢ ⎥ ⎢ ⎥
angle are two important parameters to achieve the desired ⎢ 𝜌uw ⎥ ⎢ 𝜌vw ⎥ ⎢ 𝜌w + p ⎥
2
⎢ ⎥ ⎢ ⎥ ⎢ ⎥
separation alleviation. The influence of VG height has been ⎣ 𝜌uE + up⎦ ⎣ 𝜌vE + vp⎦ ⎣ 𝜌wE + wp⎦
studied by Li et al. [16]. They found that the lift and drag (3)
coefficients basically have a logarithmic relationship with In the above formulas, p, ρ, and E are the pressure, the
the height of the VGs. Bevan et al. [17] used an adaptive sur- density, and the total energy per unit mass, respectively; u,
rogate model technique to optimize VG geometry to alleviate v, and w are the Cartesian velocity in the directions x, y, z; R,
the buffet phenomena of a representative tiltrotor aircraft. S, and T are the viscous flux vectors. Pressure p is obtained
However, most research of the VGs were arranged in a from:
single-row layout and few studies were conducted on the [
multiple-row vortex generators. The effect of multiple-row 1( )]
p = (𝛾 − 1)𝜌 E − u2 + v2 + w2 . (4)
VGs on the S809 airfoil were first investigated by Wang 2
et al. [18]. Zhu et al. [19] used double-row VGs to control The second-order accuracy central difference scheme
the dynamic stall of the wind turbine airfoil. However, the with adaptive artificial viscosity was used for the spatial
mechanism of flow control with multiple-row vortex genera- discretization [21] and the dual time-stepping approach was
tors remains unclear due to the lack of systematic studies. employed to perform time integration. In this approach, the
In this paper, we present numerical simulations of A821201 solution is marched forward in pseudo-time to steady state
airfoil without and with one, two parallel rows of VGs. The through an implicit LU-SGS scheme at each physical time
objective of current work is to provide a deep understanding level [22]. The Detached Eddy Simulation (DES) method in
of the aerodynamic interference of double-row VGs using combination with the Spalart–Allmaras one-equation turbu-
CFD approach. The remainder of the paper is organized as lence model [23] was used to simulate the massively sepa-
follows. Section 2 illustrates the numerical simulation meth- rated flows. For boundary conditions, the no-slip boundary
ods and describes the vortex generator configurations. The condition was applied on the airfoil surface, in which the VG
downstream VGs should experience the turbulent flow from was replaced by a surface with zero thickness. At the far-
upstream VGs; therefore, Sect. 3 mainly discusses the effects field boundary, the characteristic boundary condition with
of the chordwise position between two parallel rows of VGs. the Riemann invariants was applied. Suppose the spanwise
Conclusions are given in Sect. 4. of the wing is infinite, the left and right are defined as peri-
odic boundary conditions in the spanwise direction. Using
periodic boundary conditions can reduce the total number
2 Physical and Numerical Modeling of grid, thereby achieving high computational efficiency.
Details on the periodic boundary conditions can be found
2.1 Flow Solver in Ref. [24].

In the Cartesian coordinate system, the three-dimensional


compressible RANS equations in a conservative differential
form can be written as [20]:

13
International Journal of Aeronautical and Space Sciences

2.2 Geometric Description and Mesh Generation

The delta-shaped vane-type VGs are implemented on a rep-


resentative V-22 tiltrotor airfoil in a counter-rotating con-
figuration. The airfoil was developed at Bell Helicopter spe-
cifically for use on the V-22 wing, and is designated as Bell
A821201 airfoil which has a thickness ratio of 23% [25].
The VG arrangement, as visualized in Fig. 1, is determined
based on six parameters: height h, length L, intra-pair spac-
ing d, inter-pair spacing D, angle of incidence to a local flow
β, and their locations xVG with respect to the airfoil leading
edge. Figure 2 shows the geometry of A821201 airfoil with Fig. 2  Geometry of airfoil A821201 with double-row VGs
double-row installation of VGs. The computational domain
was meshed with hexahedral structured grids and the far-
field boundary was located at a distance approximately equal is shown in Fig. 5. The computed aerodynamic coefficients
to 30 chord lengths. The height of first cell layer around the of airfoil without and with VGs are in good accordance with
airfoil is about 1 × ­10−6c (c is the chord length of airfoil), experimental data. This indicates that present numerical
and the typical computational mesh near VG surface is dis- approach can be used to obtain the accurate aerodynamic
played in Fig. 3. For clean airfoil, VGs’ surface is set the characteristics of the airfoil with vortex generators.
inner flow surface; for VGs’ case, VGs’ surface is set zero
thickness wall.
3 Results and Discussion
2.3 Validation of the Numerical Simulations
The chordwise position of the VGs determines the con-
2.3.1 V‑22 Wing‑Alone Case trol range of the streamwise vortices, and, therefore, has
a significant influence on the aerodynamic characteristics.
The first V-22 wing-alone case is considered simply to vali- This work puts emphasis on the effect of relative chord-
date the present CFD results with available experimental wise position of the two-row layout of VGs. The other main
data. In this case, the freestream Mach number is 0.13 with parameters throughout this paper are primarily from previ-
a Reynolds number of 2.0 million, and the angle of attack ous studies [28, 29] to ensure the efficiency. Table 1 pre-
is 7°. The computed chordwise pressure distribution is pre- sents the geometric parameters of seven delta VGs’ models,
sented in Fig. 4 along with the experimental measurements including the one-row VGs at 10, 20, and 40% chordwise
[26]. The comparison between the calculated results and the locations, denoted as VGs1, VGs2, and VGs3, respectively.
experimental data is quite good. The double-row VGs’ arrangements are located at the 10 and
30% chord positions (VGs4), the 10 and 50% chord posi-
2.3.2 DU97‑W‑300 Airfoil with Vortex Generators tions (VGs5), and the 10 and 70% chord positions (VGs6).
The last case VGs7 is employed to explore the impact of the
Due to the lack of experimental results of the V-22 airfoil height of the second-row VGs on the control effect of bound-
equipped with VGs, present numerical modeling is validated ary-layer flow. In this paper, the chord length of A821201
against the experimental data of the DU97-W-300 airfoil. airfoil is 650 mm and the spanwise length is 35 mm. The
The VG parameters and flow conditions can be found in Ref.
[27]. Comparison of numerical results and experimental data

Fig. 1  Geometric parameters


for vane-type VGs of triangular
shape

13
International Journal of Aeronautical and Space Sciences

Fig. 4  Surface pressure coefficient distribution

Fig. 3  Mesh in the local region around VGs

freestream Mach number is 0.3, and the chord-based Reyn- computations. Therefore, the medium grid was chosen for
olds number is equal to 4.5 × ­106. the present computations.
Unsteady simulations require a proper setting of both the
time-step size and the convergence criteria within each time-
3.1 Mesh and Time‑Step Dependency Study step. Within each time-step, iterations were performed until
the solution no longer changed. The results of the numerical
To prove the adequacy of the mesh resolution, mesh simulation converge when the residual is less than 1 × ­10–6.
dependency computation was carried out. Three sets of The time step was found to be sufficiently small to give rela-
grids with different densities were calculated: 2,271,255 tively good convergence for the equation residuals, and both
(coarse), 3,760,515 (medium), and 5,016,255 (fine). From the computing hardware and computing time must be con-
the comparison of the computed results depicted in Table 2, sidered. As a result, the physical time step was selected as
the differences between medium grid and fine grid are 0.01 s in this study (refer to Table 3).
less than 1%, which is thought to be adequate for reliable

Fig. 5  Aerodynamic performance of DU97-W-300 airfoil with and without VGs

13
International Journal of Aeronautical and Space Sciences

Table 1  Main parameters of Case number xVG ∕c (%) h (mm) L (mm) d (mm) D (mm) 𝛽 (◦ )
VGs in this study
VGs1 10 5 15 17.5 35 15
VGs2 20 5 15 17.5 35 15
VGs3 40 5 15 17.5 35 15
VGs4 10 and 30 5 15 17.5 35 15
VGs5 10 and 50 5 15 17.5 35 15
VGs6 10 and 70 5 15 17.5 35 15
VGs7 10 and 30 10 7.5 17.5 35 15

Table 2  Mesh dependency study of the A821201 airfoil with VGs4 a 12° increase in the stall angle of the airfoil. For VGs2, the
at α = 10° stall angle is 20°, whereas for VGs3, it is 16°. It can be seen
Mesh Total number of grid Cl Cd that positioning VGs too far downstream (VGs3) can lead to
an early abrupt stall, which draws the same conclusion as in
Coarse 2,271,255 1.6458 0.0309
Ref. [3]. When VGs are installed at 10%c (VGs1), the maxi-
Medium 3,760,515 1.6535 0.0291
mum lift coefficient is highest in the three cases. The Cl,max
Fine 5,016,255 1.6571 0.0289
increments of VGs1, VGs2, and VGs3, are 70.6%, 62.2%,
and 40.6%, respectively. Compared to the clean airfoil, drag
coefficients of the airfoil with VGs increase at small angles
of attack. For large angles of attack, VGs not only increase
Table 3  Time-step study of the Physical Cl Cd the lift coefficient but also decrease the drag coefficient
A821201 airfoil with VGs4 at time-step (s) because of suppression of boundary-layer separation. The
α = 10°
drag coefficient of the airfoil with the optimal configuration
0.001 1.6534 0.0291
(VGs1) is 60.8% lower than that of the airfoil without VGs
0.01 1.6535 0.0291
at an angle of attack of 24°.
0.1 1.6522 0.0295
Figure 7 shows the comparison of center section stream-
lines without and with VGs at typical angle of attack.
When the angle of attack is small (Fig. 7a), the flow is fully
3.2 Influence of First‑Row VG Location attached. However, after stall, as shown in Fig. 7b, the flow
over the clean airfoil is completely separated and the sepa-
Figure 6 shows the effect of the first-row VG location on ration point appears at the chord position x/c = 0.36. The
the airfoil lift and drag coefficients. The results indicate that addition of the VGs1 or VGs2 can transfer momentum from
the chordwise installation of VGs plays a critical role in the the outer flow to the near wall region, and therefore, the flow
stall behavior of the airfoil. The stall angle of the clean case can be attached well to the surface (Fig. 7c and d). The case
is 10°, while that of the airfoil with VGs1 is 22°, so there is of VGs3 (Fig. 7e) has no effect on delaying flow separation

Fig. 6  Effect of the chordwise position of the first row of VGs on lift and drag coefficients

13
International Journal of Aeronautical and Space Sciences

Fig. 7  Center section instantaneous streamlines of uncontrolled and controlled airfoil flow

at the angle of attack of 18°, because the separation point is aircraft wings. In this work, the first row of VGs is fixedly
ahead of the VGs. The pressure distributions of the airfoil installed at the 10% chord positions, and the second row of
illustrate this phenomenon in Fig. 8. VGs is placed at 30, 50, and 70% chord positions, denoted as
VGs4, VGs5, and VGs6, respectively. As can be seen from
3.3 Influence of Second‑Row VG Location Fig. 9, when the angle of attack is small, the influence of the
longitudinal position of the second-row VGs is not obvious.
The two-row layout is calculated to validate whether it can Compared to the single-row layout (VGs1), the lift coef-
further improve the aerodynamic performance of tiltrotor ficients of VGs4 are improved significantly when the angle

13
International Journal of Aeronautical and Space Sciences

of attack is between 18° and 24°. However, for VGs6, both


the stall angle and the maximum lift coefficient decrease
slightly. In this regard, inappropriate installation of the sec-
ond-row VGs may degrade the overall aerodynamic perfor-
mance of A821201 airfoil. Generally, the drag coefficient
curves show nearly the same rising trends for all cases. The
quantitative results show that the maximum lift coefficient
of the airfoil with VGs4 increases by 82.5% with respect to
the clean airfoil, and the drag coefficient of the airfoil with
VGs4 is 59.4% lower than that of the airfoil without VGs at
an angle of attack of 24°.
The vorticity contours in the wake region of VGs are pre-
sented for two cases (VGs1 and VGs4). As shown in Fig. 10,
four different chordwise positions x/c = 0.15, x/c = 0.20,
x/c = 0.25, and x/c = 0.35 are selected. The strong vortices
Fig. 8  Comparison of pressure coefficients with and without VGs at were created by counter-rotating vortex generators. The
α = 18°

Fig. 9  Effect of the chordwise position of the second row of VGs on lift and drag coefficients

Fig. 10  Distribution of vorticity contour downstream of VGs

13
International Journal of Aeronautical and Space Sciences

Fig. 11  Geometries of the second row of VGs

high momentum of the outer region is mixed with the low length is halved so as to keep the same VG reference area,
momentum region near the airfoil surface, making the flow as shown in Fig. 11.
more resistant to separation at the adverse pressure gradi- As can be seen in Fig. 12, it is easy to demonstrate that
ent. For the case of VGs4, due to the effect of second row the larger height of the second-row VGs plays strong posi-
of VGs, the vorticity magnitude at the position of x/c = 0.35 tive role for enhancing the aerodynamic characteristics after
increases significantly. the angle of attack of 22°. The stall angle of VGs7 is up to
The height of VGs is often related to the local boundary- 24°, and the drag coefficient is greatly reduced by 60.2%
layer thickness. According to the boundary-layer theory, the with respect to the clean airfoil at the angle of attack of
thickness of the boundary layer increases along the chord 24°. Figure 13 indicates that the flow separation is highly
line. Therefore, the boundary-layer thickness is higher at suppressed by VGs7. For VGs7, the velocity of fluid is fur-
the position of the second row of VGs. To further improve ther increased at the suction surface and the high-momen-
the aerodynamic performance, the effect of the height of the tum fluid is transported from the outer flow down into the
second row of VGs has been investigated. For the case of boundary layer near the surface in the downstream region, as
VGs7, the height of the second-row VGs is doubled and the shown in Fig. 14. Figure 15 shows the influence of VGs on

Fig. 12  Effect of the height of the second row of VGs on lift and drag coefficients

13
International Journal of Aeronautical and Space Sciences

Fig. 13  Center section instantaneous streamlines of uncontrolled and controlled airfoil flow

Fig. 14  Comparison of velocity contours along the wing center section

13
International Journal of Aeronautical and Space Sciences

4 Conclusions

In this paper, the aerodynamic characteristics of the airfoil


A821201 without and with one, two parallel rows of vortex
generators were investigated using CFD simulations. For all
the cases studied, the effect of VGs is visible on the delay
boundary-layer separation. The lift coefficients, drag coef-
ficients, and stall angles of attack were analyzed qualitatively
and quantitatively. Based on the results obtained, the main
findings can be summarized as follows:

1. For single-row VGs, chordwise installation significantly


affects the stall behavior. The results indicate that the
more rearward the VG location, the more abrupt the stall
characteristic. When the VGs are installed at the 10%
chord positions, the maximum lift coefficient is extended
Fig. 15  Comparison of pressure coefficients with and without VGs at from 1.43 to 2.44, and the stall angle increases from 10°
α = 24°
to 22° with respect to the airfoil without VGs.
2. Compared to single-row VGs, the application of double-
the pressure coefficients of the airfoil at the angle of attack row VGs can achieve better aerodynamic characteristics.
of 24°. Due to the effects of VGs, the integration areas of When the VGs are located at the 10 and 30% chord posi-
the pressure coefficients increase in comparison to the air- tions, the maximum value of lift coefficient increases
foil without VGs. Table 4 gives the relative variation in the by 82.5% relative to the airfoil without VGs. However,
stall angle, maximum lift coefficient, drag coefficient, and inappropriate installation of the second-row VGs may
lift–drag ratio. The double vortex generator arrangements degrade the overall aerodynamic performance.
show better performance in the control of flow separation 3. Increasing the height of the second row of VG can fur-
and can further improve the aerodynamic performance of ther result in a remarkable lift enhancement and drag
the airfoil A821201. reduction at large angles of attack.

Table 4  Relative variation of Clean VG1 VG2 VG3 VG4 VG5 VG6 VG7
aerodynamic coefficients
Stall angle of attack 10° 22° 20° 16° 22° 22° 20° 24°
Increment of ­Cl,max 0% 70.6% 62.2% 40.6% 82.5% 74.8% 66.4% 85.3%
Decrease of Cd at 24° 0% 60.8% 51.9% 7.89% 59.4% 59.1% 57.6% 60.2%
Increment of lift–drag ratio at 24° 0% 481.7% 318.8% 12.6% 496.4% 477.6% 414.8% 532.1%

13
International Journal of Aeronautical and Space Sciences

Acknowledgements This work was supported by the Scientific 14. Namura N, Jeong S (2013) Parametric study of vortex generators
Research Foundation of Huaiyin Institute of Technology (no. on a super critical infinite-wing to alleviate shock-induced separa-
Z301B19511). tion. Trans Jpn Soc Aeronaut Space Sci 56(5):293–302
15. Tai TC (2003) Effect of midwing vortex generators on V-22 air-
Data Availability The data used to support the findings of this study are craft forward-flight aerodynamics. J Aircraft 40(4):623–630
available from the corresponding author upon request. 16. Li XK, Yang K, Wang XD (2019) Experimental and numerical
analysis of the effect of vortex generator height on vortex charac-
teristics and airfoil aerodynamic performance. Energies 12(5):19
Compliance with Ethical Standards 17. Bevan RLT, Poole DJ, Allen CB, Rendall TCS (2017) Adaptive
surrogate-based optimization of vortex generators for a tiltrotor
Conflicts of Interest The authors declare that there is no conflict of geometry. J Aircraft 54(3):1011–1024
interest regarding the publication of this paper. 18. Wang HP, Zhang B, Qiu QG, Xu X (2017) Flow control on the
NREL S809 wind turbine airfoil using vortex generators. Energy
118:1210–1221
19. Zhu CY, Chen J, Wu JH, Wang TG (2019) Dynamic stall control
References of the wind turbine airfoil via single-row and double-row passive
vortex generators. Energy 189:11
1. Alli P, Nannoni F, Cicale M (2003) Erica: the European tiltrotor 20. Huang W, Lu ZL, Guo TQ, Xue F, Zhang M (2012) Numerical
design and critical technology projects. In:AIAA international air method of static aeroelastic correction and jig-shape design for
and space symposium and exposition: the next 100 years, p 2515, large airliners. Sci China Technol Sci 55(9):2447–2452
Dayton, Ohio, USA 21. Jameson A, Schmidt W, Turkel E (1981) Numerical solution of
2. Kelley CL, Corke TC, Thomas FO, Patel M, Cain AB (2016) the Euler equations by finite volume methods using Runge Kutta
Design and scaling of plasma streamwise vortex generators for time stepping schemes. In: 14th AIAA fluid and plasma dynamic
flow separation control. AIAA J 54(11):3397–3408 conference, p 1259, Palo Alto, California, USA
3. Mueller-Vahl H, Pechlivanoglou G, Nayeri CN, Paschereit CO 22. Yoon S, Jameson A (1988) Lower-upper Symmetric-Gauss-Sei-
(2012) Vortex generators for wind turbine blades: a combined del method for the Euler and Navier-Stokes equations. AIAA J
wind tunnel and wind turbine parametric study. In: ASME Turbo 26(9):1025–1026
Expo 2012: turbine technical conference and exposition, p 69197, 23. Spalart PR, Deck S, Shur ML, Squires KD, Strelets MKh, Travin
Copenhagen, Denmark A (2006) A new version of detached-eddy simulation, resist-
4. Fouatih OM, Medale M, Imine O, Imine B (2016) Design opti- ant to ambiguous grid densities. Theoret Comput Fluid Dyn
mization of the aerodynamic passive flow control on NACA 4415 20(3):181–195
airfoil using vortex generators. Eur J Mech B/Fluids 56:82–96 24. Namura N, Shimoyama K, Obayashi S (2015) Multipoint design
5. Zhang L, Li XX, Yang K, Xue DY (2016) Effects of vortex gen- of vortex generators on a swept infinite-wing under cruise and
erators on aerodynamic performance of thick wind turbine airfoils. critical condition. In: 33rd AIAA applied aerodynamics confer-
J Wind Eng Ind Aerodyn 156:84–92 ence, p 2726, Dallas, Texas, USA
6. May NE (2001) A new vortex generator model for use in complex 25. Felker FF, Shinoda PR, Hefferman RM, Sheehy HF (1990) Wing
configuration CFD solvers. In: 19th AIAA applied aerodynamics force and surface pressure data from a hover test of a 0.658-scale
conference, p 2434, Anaheim, California, USA V-22 rotor and wing, NASA Tech. Rep. 102244
7. Yi JS, Kim C, Lee BJ (2012) Adjoint-based design optimiza- 26. Tai TC (1996) Simulation and analysis of V-22 tiltrotor aircraft
tion of vortex generator in an S-shaped subsonic inlet. AIAA J forward-flight flowfield. J Aircraft 33(2):369–376
50(11):2492–2507 27. Baldacchino D, Manolesos M, Ferreira C et al (2016) Experi-
8. Dudek JC (2011) Modeling vortex generators in a Navier-Stokes mental benchmark and code validation for airfoils equipped with
code. AIAA J 49(4):748–759 passive vortex generators. J Phys Conf Ser 753(2):022002
9. Jirasek A (2005) Vortex-generator model and its application to 28. Baldacchino D, Ferreira C, Tavernier DD, Timmer WA, van Bus-
flow control. J Aircraft 42(6):1486–1491 sel GJW (2018) Experimental parameter study for passive vortex
10. Namura N, Shimoyama K, Obayashi S, Ito Y, Koike S, Nakakita generators on a 30% thick airfoil. Wind Energy 21(9):745–765
K (2019) Multipoint design optimization of vortex generators on 29. Zhu CY, Wang TG, Wu JH (2019) Numerical investigation of
transonic swept wings. J Aircraft 56(4):1291–1302 passive vortex generators on a wind turbine airfoil undergoing
11. Lee HM, Kwon OJ (2019) Numerical simulation of horizontal pitch oscillations. Energies 12(4):15
axis wind turbines with vortex generators. Int J Aeronaut Space
Sci 20(2):325–334 Publisher’s Note Springer Nature remains neutral with regard to
12. Suarez JM, Flaszynski P, Doerffer P (2018) Application of rod jurisdictional claims in published maps and institutional affiliations.
vortex generators for flow separation reduction on wind turbine
rotor. Wind Energy 21(11):1202–1215
13. Godard G, Stanislas M (2006) Control of a decelerating boundary
layer. Part 1: optimization of passive vortex generators. Aerosp
Sci Technol 10:181–191

13

You might also like