You are on page 1of 22

applied

sciences
Article
Analysis of the Effect of Vortex Generator Spacing on
Boundary Layer Flow Separation Control
Xin-kai Li 1, *, Wei Liu 1 , Ting-jun Zhang 1 , Pei-ming Wang 1 and Xiao-dong Wang 2
1 China Huadian Engineering Co., LTD (CHEC), Beijing 100160, China; liuw@chec.com.cn (W.L.);
zhangtj@chec.com.cn (T.-j.Z.); wangpm@chec.com.cn (P.-m.W.)
2 College of Energy, Power and Mechanical Engineering, North China Electric Power University,
Beijing 102206, China; wangxd@ncepu.edu.cn
* Correspondence: lixinkai@chec.com.cn; Tel.: +86-010-63918174

Received: 31 October 2019; Accepted: 9 December 2019; Published: 13 December 2019 

Abstract: During the operation of wind turbines, flow separation appears at the blade roots, which
reduces the aerodynamic efficiency of the wind turbine. In order to effectively apply vortex generators
(VGs) to blade flow control, the effect of the VG spacing (λ) on flow control is studied via numerical
calculations and wind tunnel experiments. First, the large eddy simulation (LES) method was used to
calculate the flow separation in the boundary layer of a flat plate under an adverse pressure gradient.
The large-scale coherent structure of the boundary layer separation and its evolution process in
the turbulent flow field were analyzed, and the effect of different VG spacings on suppressing the
boundary layer separation were compared based on the distance between vortex cores, the fluid
kinetic energy in the boundary layer, and the pressure loss coefficient. Then, the DU93-W-210 airfoil
was taken as the research object, and wind tunnel experiments were performed to study the effect of
the VG spacing on the lift–drag characteristics of the airfoil. It was found that when the VG spacing
was λ/H = 5 (H represents the VG’s height), the distance between vortex cores and the vortex core
radius were approximately equal, which was more beneficial for flow control. The fluid kinetic energy
in the boundary layer was basically inversely proportional to the VG spacing. However, if the spacing
was too small, the vortex was further away from the wall, which was not conducive to flow control.
The wind tunnel experimental results demonstrated that the stall angle-of-attack (AoA) of the airfoil
with the VGs increased by 10◦ compared to that of the airfoil without VGs. When the VG spacing
was λ/H = 5, the maximum lift coefficient of the airfoil with VGs increased by 48.77% compared to
that of the airfoil without VGs, the drag coefficient decreased by 83.28%, and the lift-to-drag ratio
increased by 821.86%.

Keywords: vortex generators; wind turbine; airfoil; aerodynamic performance; large eddy simulation

1. Introduction
During the operation of wind turbines, flow separation occurs at the blade root, which reduces
the wind turbine efficiency [1]. Flow control technologies, such as VGs [2], plasma flow control [3],
microflaps [4], microtabs [5], blowing and suction [6], synthetic jets [7], and flexible walls [8], have
been increasingly applied to the design or optimization of wind turbine blades, aiming to improve
their aerodynamic performance. Barlas [9] and Johnson [10] compared these flow control methods,
while Lin [11] and Wang [12] found that VGs are one of the most effective devices for improving the
aerodynamic performance of blades.
In 1947, Taylor [13] was the first to apply VGs to wing flow control. The basic principle of
flow separation control using VGs is that when the fluid passes through a VG, a concentrated vortex
will be generated downstream. Under the action of the concentrated vortex, the high-energy fluid

Appl. Sci. 2019, 9, 5495; doi:10.3390/app9245495 www.mdpi.com/journal/applsci


Appl. Sci. 2019, 9, 5495 2 of 22

outside the boundary layer will enter the bottom of the boundary layer, increasing the mixing of
the fluids inside and outside the boundary layer. This will increase the kinetic energy of the fluid
at the bottom of the boundary layer and delay its separation [14,15]. Due to their simple structure
and convenient installation, VGs have been widely applied in fields such as flow control [16–18]
and heat transfer [19,20]. In 1995, Reuss et al. [21] conducted two-dimensional wind tunnel tests
on the National Advisory Committee for Aeronautics (NACA) 4415 airfoil in the subsonic wind
tunnel of the Aerospace Research Laboratory. The aim of these tests was to record the cross-sectional
lift and bending moment characteristics under different airflow conditions. The effect of the airfoil
leading-edge roughness and VGs on the aerodynamic performance of the airfoil was investigated.
In 2006, Godard and Stanislas [22–24] experimentally studied the control effect of triangular VGs on
the boundary layer flow separation on a curved plate. Their experimental set-up could change the
installation angle of VGs and measure the wake structure of small wings. In 2010, Yang et al. [25]
performed numerical simulations to study a blunt trailing edge airfoil with and without VGs, and
the performance of the airfoil was analyzed. The interaction between vortices generated by the VGs
and wake vortices or separation vortices was investigated, while a mechanism for suppressing the
boundary layer separation using VGs was revealed. In 2011, Zhen et al. [26] studied the effect of VGs on
the aerodynamic performance of airfoils using numerical simulations, where the Spalart–Allmaras (SA)
turbulence model and the standard wall function were employed. Their study results demonstrated
that when the VGs were close to the separation point, the airfoil had a high maximum lift coefficient,
and the performance of the rectangular and curved edge VGs was better than that of the triangular VGs.
Lishu et al. [27] designed two types of VG combinations and conducted wind tunnel experiments in
order to investigate airfoil flow control in depth. Their results revealed that the reasonable combination
of a flap and VGs can significantly improve the aerodynamic performance of the airfoil, providing
an ideal form of combined control. In 2012, Delnero et al. [28] experimentally investigated the effects
of triangular VGs on airfoil aerodynamic performance. The VGs were located 10% and 20% away
from the airfoil leading edge. The results showed that VGs can improve the airfoil aerodynamic
characteristics, and the effect is better when the VGs are arranged 20% away from the leading edge.
In 2013, Velte and Hansen [29] performed wind tunnel experiments using three-dimensional particle
image velocimetry to observe the effect of VGs on the near-stall flow behavior of the DU91-W2-250
airfoil. Their aim was to study the flow structures induced by the VGs on the wing. It was found that
the VGs transfer the high momentum fluid to the near-wall region, and the variation of the velocity
component direction in the boundary layer along the flow increases. In 2014, Sørensen et al. [30]
conducted numerical simulations on the FFA-W3-301 and FA-W3-360 airfoils. The calculated results
were compared with wind tunnel measurements, and the variation of lift and drag with the angle
of attack was obtained. Although their method was not able to capture the flow details accurately,
it can be used to qualitatively compare the effects of different VG settings on airfoil aerodynamic
performance. In 2015, Manolesos and Voutsinas [31] employed passive VGs as a simple and effective
method to delay or suppress airfoil flow separation, and carried out an experimental study on the
optimized blades. Flow visualization of triangular VGs through three-dimensional particle image
velocimetry was carried out. Overall, the results demonstrated that the maximum lift increased by
44%, while the drag increased by 0.2%. Lin et al. [17] experimentally evaluated a boundary layer
separation control method using miniature surface VGs. Wind tunnel tests were carried out in a
low-turbulence pressure tunnel at the NASA (The National Aeronautics and Space Administration)
Langley research center. The measured parameters included lift, drag, surface pressure, wake profile,
and surface heat flux fluctuations. The results indicated that small VGs can effectively reduce the
flow separation in the boundary layer of the flaps. Gao et al. [32] performed computational fluid
dynamics (CFD) simulations to study the effect of the physical characteristics and magnitudes of flow
from VGs on the aerodynamic performance of the DU97-W-300 blunt-trailing-edge wing. The effect
of the VG size was analyzed in terms of the trailing edge thickness and the VG length. The results
demonstrated that the drag loss was more sensitive to the increase in the VG’s height than lift, while
Appl. Sci. 2019, 9, 5495 3 of 22

an increase in the VG’s length had a negative impact on both lift and drag. Analysis of the streamlines
and vortices further revealed the flow field characteristics in the wake region. In 2016, Omar et
al. [33] experimentally studied the NACA 4415 airfoil equipped with VGs. The results showed that
triangular VGs were more suitable for controlling the boundary layer separation, while micro VGs can
effectively control the flow. Zhang et al. [34] studied the effect of VGs on the aerodynamic performance
of thick blades. In particular, the effects of height and chord position of VGs and double-row VGs
on airfoil aerodynamic performance were investigated. The results demonstrated that after the VGs
were installed on the airfoil leading edge, the maximum lift coefficient increased. In addition, the
lift coefficient of the double-row VGs could be further improved at high angles of attack due to the
different installation positions and sizes. In 2017, Wang et al. [12] studied the aerodynamic performance
of the S809 airfoil with and without VGs using CFD simulations. The results showed that the VGs
can effectively improve the aerodynamic performance of the S809 airfoil, reduce the thickness of the
boundary layer, and delay stall. The double-row VG arrangement demonstrated a good performance
in controlling flow separation, which further improved the aerodynamic performance of the S809
airfoil. Martínez-Filgueira et al. [35] studied the motion of micro VGs in the boundary layer of a flat
plate using numerical simulations. The installation angle of the VGs was 18.5◦ . The trajectory and size
of the vortices on the plate were analyzed. Brüderlin et al. [36] improved airfoil aerodynamic efficiency
by means of passive VGs placed on the control surface. First, the aerodynamic characteristics of the
airfoil were simulated using a Reynolds-averaged Navier–Stokes solver (RANS). Then, through the
parameterization study of different VG sizes, the effectiveness of the VGs under certain flow conditions
were studied. In 2018, Gageik et al. [37] used a numerical visualization method to investigate whether
appropriate micro VGs can suppress the pressure wave. The flow around the VGs was characterized
using numerical schlieren visualization. Baldacchino et al. [38] performed wind tunnel experiments
on the DU97-W-300 airfoil, where the oil flow visualization method was used to verify the effect of
VGs on airfoil stall separation suppression. The results demonstrated that VGs can delay stall, while
under a stall condition, the presence of VGs increases the load fluctuation. In 2019, Kundu et al. [39]
used the numerical simulation method to study the performance of the S1210 airfoil with VGs under
improved trailing edge conditions. The results showed that by adding counter-rotating VGs near
the airfoil trailing edge, the lift coefficient of the wing can be increased by 17% and the stall angle
can be delayed from 10◦ to 12◦ . Moreover, a rounder and thicker trailing edge can increase the lift
coefficient by 13.5%, thus improving the performance. Nowadays, VGs are becoming an important
part of wind turbine blade design. However, the challenges involved in calculating the flow field
around VGs have not yet been satisfactorily addressed. Mereu et al. [40] used the scale decomposition
method to simulate the stall behavior of the DU97-W-300 airfoil at high Reynolds numbers. In addition,
a detailed sensitivity analysis was performed to evaluate the effects of time integration, grid width
resolution, and region width on the simulation accuracy. Using this method, the flow on the airfoil
surface with VGs was numerically simulated. Manolesos et al. [41] proposed a numerical model of
VGs for RANS simulations. The performance and application of the VG numerical model were also
investigated. A wind turbine impeller was taken as an example to study the reliability of the model.
Without considering the mesh-induced errors, the results showed that the vortices generated in the
numerical simulation were weaker than those in the fully analytical calculation.
The technical development of VGs as a means of flow control has been analyzed since the
1990s. Although many scholars have investigated the flow control mechanism of VGs, in the practical
application of VGs, there are still some problems to be solved. When VGs are used to control the flow
of wind turbine blades, an optimal VG spacing needs to be determined since there is flow interference
between VGs and the spacing between VGs may affect the development and dissipation of concentrated
vortices. The effect of VG spacing on vortex characteristics and wind turbine airfoil aerodynamic
characteristics has not yet been reported. Therefore, in this paper, the effect of VG spacing on vortex
characteristics and the aerodynamic characteristics of special airfoils for wind turbines is investigated
using numerical simulations and wind tunnel experiments.
Appl. Sci. 2019, 9, 5495 4 of 22

2. Methods

2.1. Numerical Conditions

2.1.1. Large Eddy Simulation


A large eddy simulation (LES) is a spatial average of turbulent fluctuations, that is, large-scale
vortices and small-scale vortices are separated using a filtering function [42]. More specifically,
large-scale vortices are simulated directly, while small-scale vortices are closed using models. The
theoretical basis of an LES is the existence of inertial sub-scale vorticity in turbulent flows with high
Reynolds numbers. The vorticity of such scale has statistical isotropy. It transfers the energy of vortices
with an energy scale to vortices with a dissipation scale. After filtering the Navier–Stokes equation with
a filter function, the governing equation of the LES method for an incompressible flow can be derived:

∂ρ ∂ρui
+ =0 (1)
∂t ∂xi

∂(ρui ) ∂(ρui u j ) ∂ ∂σij ∂p ∂τij


!
+ = µ − − (2)
∂t ∂x j ∂x j ∂x j ∂xi ∂x j
where σij is the stress tensor and τij is the subgrid stress.
In this paper, the Smagorinsky–Lilly model was used for the subgrid model. This model was first
proposed by Smagorinsky [43]. The eddy viscosity can be defined as follows:

µt = ρL2s s (3)
q  1
where Ls is the mixed length of the grid and s 2sij sij , Ls = min kd, Cs V 3 , where k is a constant, d is
the nearest distance from the wall, Cs is the Smagorinsky constant (in this study, Cs = 0.1), and V is the
volume of the calculation unit.

2.1.2. Mesh Development


The ICEM software (ANSYS 16.0, Canonsburg, PA, USA) was employed to mesh the computational
domain. The number of mesh nodes corresponding to the length, width, and height of the computational
domain was 220 × 120 × 100, respectively, and the total number of elements was 2,640,000. The
height of the first layer mesh was 0.001 mm and Y+ < 1, which met the calculation requirements. The
mesh around the VGs consisted of 50 nodes along the length direction and 40 nodes along the height
direction. The mesh distribution can be seen in Figure 1.
The applied boundary conditions were the velocity inlet and pressure outlet. The upper boundary
of the computational domain was set as an Euler wall, the bottom boundary and the VGs were set
as adiabatic non-slip walls, and the two sides of the computational domain were set as symmetrical
boundaries. The dimensions of the computational domain are shown in Figure 2. A turbulent velocity
boundary condition was set at the inlet, where the inlet turbulence velocity type is given as follows:
  
y 1/7
 U δ

 y<δ
u= , (4)
 U
 y≥δ

where u is the local velocity, δ is the thickness of the boundary layer, U is the main velocity, and y is the
normal height of the wall.
Appl. Sci. 2020, 10, x FOR PEER REVIEW 5 of 24
Appl. Sci. 2019, 9, 5495 5 of 22

Figure 1. Computational domain mesh and detail of the mesh around the vortex generators (VGs).

The applied boundary conditions were the velocity inlet and pressure outlet. The upper
boundary of the computational domain was set as an Euler wall, the bottom boundary and the VGs
were set as adiabatic non-slip walls, and the two sides of the computational domain were set as
symmetrical boundaries. The dimensions of the computational domain are shown in Figure 2. A
turbulent velocity boundary condition was set at the inlet, where the inlet turbulence velocity type is
given as follows:

1/7
  y
U y <δ ,
u =   δ  (4)
 U y ≥δ

where u is the local velocity, δ is the thickness of the boundary layer, U is the main velocity, and y is
Computational
Figureheight
the normal
Figure 1.1.Computational domainmesh
of the wall.
domain meshand
anddetail
detailof
ofthe
themesh
mesharound
aroundthe
thevortex
vortexgenerators
generators(VGs).
(VGs).

The applied boundary conditions were the velocity inlet and pressure outlet. The upper
boundary of the computational domain was set as an Euler wall, the bottom boundary and the VGs
were set as adiabatic non-slip walls, and the two sides of the computational domain were set as
symmetrical boundaries. The dimensions of the computational domain are shown in Figure 2. A
turbulent velocity boundary condition was set at the inlet, where the inlet turbulence velocity type is
given as follows:

1/7
  y
U   y and
Figure 2. Computational domaindimensions < δ ,applied boundary conditions.
Figure 2. Computational domain δ 
u = dimensions and applied boundary conditions. (4)
2.1.3. Computational Setup  U y ≥δ

2.1.3.TheComputational
FLUENT software Setup (ANSYS 16.0, Canonsburg, PA, USA) was used for the simulation. The
where u is the local velocity, δ is the thickness of the boundary layer, U is the main velocity, and y is
finiteThevolume
FLUENT method was used to discretize the equations PA,and the was
finiteused
centered difference scheme
the normal height ofsoftware
the wall.(ANSYS 16.0, Canonsburg, USA) for the simulation. The
was used to discretize the space. Because for LES computing, the SIMPLE
finite volume method was used to discretize the equations and the finite centered difference scheme (Semi-Implicit Method
for
wasPressure Linked Equations)
used to discretize the space.algorithm
Because for saves
LESmore computing
computing, resources
the SIMPLE than the PISO
(Semi-Implicit (Pressure
Method for
Implicit
Pressure with LinkedSplitting of Operators)
Equations) algorithm algorithm,
saves more the computing
pressure–velocity
resources coupling
than the was based
PISO on the
(Pressure
SIMPLE
Implicit algorithm
with Splittingin thisof paper. The time
Operators) step was
algorithm, the as ∆t = 3 × 10−5 s.coupling
setpressure–velocity After statistical
was based stability
on thein
the
Appl. flow
SIMPLE field10,was
algorithm
Sci. 2020, achieved,
x FOR in time-averaged
thisREVIEW
PEER paper. The time statistics
step waswere performed.
set as Δt = 3 × 10In s.
−5 particular, the six cycles
After statistical 6 after
stability
of in
24
the flow
flow field
fieldwas
wasstabilized
achieved,were time-averaged
time-averaged with the
statistics werefluid sweepingIn
performed. across the lower
particular, thewall of the
six cycles
of the
flat
afterplateflatflow
the asplate as was
a field
cycle.a Figure
cycle. Figure 3 illustrates
3 illustrates
stabilized were the structural
the structural
time-averaged parameters
parameters
with the fluid VGs,ofwhere
ofsweeping VGs, lwhere
across the lVG
is the is the
lower VG
length,
wall
length, H is the VG height, S is the distance between VGs, and λ is the spacing
H is the VG height, S is the distance between VGs, and λ is the spacing of the VGs (see Figure 3 for of the VGs (see Figure
3the
fordifference
the difference
between between
S andSλ). and λ).present
The The present
studystudy
focusedfocused
on the oneffect
the effect
of theofVGthespacing
VG spacing on
on flow
flow separation control. Table 1 lists the geometric parameters of the VGs, while
separation control. Table 1 lists the geometric parameters of the VGs, while the inflow velocity was set the inflow velocity
was
as U set= 82asm/s,
U =Figure
82
andm/s,
the and the Reynolds
Reynolds
2. Computationalnumber number
domainbased based
on the VG
dimensions on height
and the VGwas
applied height was
about
boundary 3 about
× 104 . 3 × 104.
conditions.

2.1.3. Computational Setup


The FLUENT software (ANSYS 16.0, Canonsburg, H
PA, USA) was used for the simulation. The
λ
l
finite volume method was used to discretize the equationsS and the finite centered difference scheme
β
was used to discretize the space. Because for LES computing, the SIMPLE (Semi-Implicit Method for
Pressure Linked Equations) algorithm saves more computing resources than the PISO (Pressure
Implicit with Splitting of Operators) algorithm, the pressure–velocity coupling was based on the
SIMPLE algorithm in this paper. The time step was set as Δt = 3 × 10−5 s. After statistical stability in
Figure 3. VG geometric model.
the flow field was achieved, time-averaged
Figure 3. statistics were
VG geometric performed. In particular, the six cycles
model.
after the flow field was stabilized were time-averaged with the fluid sweeping across the lower wall
Table 1. VG geometric parameters.

Setting Angle β (°) Length l (mm) Height H (mm) S/H λ/H


20 17 6 7 3, 5, 7, 9
Appl. Sci. 2019, 9, 5495 6 of 22

Table 1. VG geometric parameters.

Setting Angle β (◦ ) Length l (mm) Height H (mm) S/H λ/H


20 17 6 7 3, 5, 7, 9

The pressure loss coefficient at the inlet and outlet of the calculation domain was defined as C∆p ,
which can be calculated according to Equation (5), where Pinlet is the total pressure at the inlet and
Poutlet the total pressure at the outlet. The pressure difference coefficient reflects the magnitude of
pressure difference loss in the computational domain. For flow control, the smaller the pressure loss
difference, the smaller the corresponding energy loss.

(Pinlet − Pinlet )
C∆p = (5)
Pinlet

In order to verify the mesh independence, three meshes with different densities were developed
with 1,320,000 (coarse), 1,980,000 (medium), and 2,640,000 (fine) elements. The Richardson extrapolation
method was used to verify the convergence of the mesh, and the results are presented in Table 2, where
RE is the value obtained using a Richardson extrapolation, p is the order of accuracy, and R is the
ratio of errors. The convergence conditions are 0 < R < 1 (monotonic convergence), R < 0 (oscillation
convergence), and 1 < R (divergence). As it can be seen, the R = 0.22 result indicates monotonic
convergence. Therefore, in this paper, the finest mesh was used for the numerical simulations.

Table 2. Mesh independence study results.

Mesh Richardson Extrapolation


Variables
Coarse Medium Fine RE p R
C∆p 0.3635 0.3628 0.3622 0.3620 2.75 0.22

2.2. Experimental Setup


An experiment of an airfoil with VGs was carried out in the wind tunnel of North China Electric
Power University (NCEPU), which contains a 2-D airfoil measurement section. The size of the test
section is 1.5 × 3 × 4.5 m3 , and the maximum wind speed of the wind tunnel is 62 m/s. The measurement
section is equipped with 256- and 64-channel electronic scanning valves and a PXI (Platform PCI
eXtensionsfor Instrumentation) data acquisition system. A pitot tube was installed on the wall of the
upstream wind tunnel of the airfoil and provided a measurement of the total pressure of the inflow.
The sidewall of the pitot tube was punctured to provide air holes, which were used to measure the
static pressure of the incoming flow. There were 96 pressure taps in the middle section of the airfoil to
measure the static pressure distribution on the surface of the airfoil. The static pressure coefficient (Cp )
was the time-averaged result found with a sampling time interval of 0.1 seconds. The experimental
sampling time was 10 seconds for each working condition. There were 111 total pressure piezometric
holes and 5 static pressure piezometric holes (called the wake rake) downstream of the airfoil, which
were used to calculate the drag coefficient of the airfoil. The geometric dimensions of the wind tunnel
and test airfoils are shown in Figure 4.
The test airfoil used in the experiment is a special wind turbine airfoil: the DU93-W-210 airfoil. The
airfoil was developed and designed by Delft University. The DU93-W-210 airfoil is very representative
of the aerodynamic characteristics of wind turbine airfoils in general. The thickness of the airfoil was
21% of C, the chord length of the tested airfoil was 0.8 m, and the span of the airfoil was 1.5 m. The
VGs were arranged at 0.2 times the chord length from the leading edge of the airfoil. Figure 5 shows
the test airfoil with the VGs installed. The geometric dimensions of the VGs are shown in Table 3. The
tunnel wind speed was set to 18.2 m/s, and the Reynolds number based on the chord of the airfoil was
1 × 106 . The AoA of the airfoil ranged from −10◦ to 25◦ .
The static pressure coefficient (Cp) was1.5the 0.8 m
m time-averaged result found Trailwith
rake a sampling time
interval of 0.1 seconds. The experimental sampling time was 10 seconds for each working condition.
There were 111 total pressure Windpiezometric holesPitot tubes
and 5 static pressure
Pressure taps piezometric
0.75 m holes (called the
wake rake) downstream of the airfoil, which were used to calculate the drag coefficient of the airfoil.
Umax = 62 m/s Turntable
The Sci.
Appl. geometric dimensions
2019, 9, 5495 of theintensity
Turbulence wind =tunnel
0.3 % and test airfoils are shown in Figure 4. 7 of 22
Measuring section
1.5 ×3 ×4.5 m3

4.5 m
Movable along axis
Test airfoil
3m
0.8 m
1.5 m Trail rake
Figure 4. Wind tunnel measurement section size and test airfoil.
Wind Pitot tubes
Pressure taps 0.75 m
The test airfoil used in the experiment is a special wind turbine airfoil: the DU93-W-210 airfoil.
Umax = 62 m/s Turntable
The airfoil was developed and designed by Delft University. The DU93-W-210 airfoil is very
Turbulence intensity = 0.3 %
representative of the aerodynamic characteristics of wind turbine airfoils in general. The thickness
Measuring section
of the airfoil was 21% of C,1.5 ×3
the×4.5
chordm3 length of the tested airfoil was 0.8 m, and the span of the

airfoil was 1.5 m. The VGs were arranged at 0.2 times the chord length from the leading edge of the
airfoil. Figure 5 shows the test airfoil with the VGs installed. The geometric dimensions of the VGs
are shown in Table 3. The tunnel wind speed was set to 18.2 m/s, and the Reynolds number based
Figure 4. Wind tunnel measurement section size and test airfoil.
on the chord of the airfoil was 1 × 106. The AoA of the airfoil ranged from −10° to 25°.
Figure 4. Wind tunnel measurement section size and test airfoil.

The test airfoil used in the experiment is a special wind turbine airfoil: the DU93-W-210 airfoil.
The airfoil was developed and designed by Delft University. The DU93-W-210 airfoil is very
representative of the aerodynamic characteristics of wind turbine airfoils in general. The thickness
of the airfoil was 21% of C, the chord length of the tested airfoil was 0.8 m, and the span of the
airfoil was 1.5 m. The VGs were arranged at 0.2 times the chord length from the leading edge of the
airfoil. Figure 5 shows the test airfoil with the VGs installed. The geometric dimensions of the VGs
are shown in Table 3. The tunnel wind speed was set to 18.2 m/s, and the Reynolds number based
on the chord of the airfoil was 1 × 106. The AoA of the airfoil ranged from −10° to 25°.

Figure 5. Test airfoil with VGs.


Figure 5. Test airfoil with VGs.
Table 3. VG geometric parameters and arrangement.

Installation Angle β (◦ ) Table 3. VGl (mm)


Length geometric Height
parameters and arrangement.
H (mm) Spacing S/H Pitch Distance λ/H
20Angle β (°)
Installation 17 l (mm)
Length Height5H (mm) Spacing5 S/H 3, 5, 7, 9 λ/H
Pitch Distance
20 17 5 5 3, 5, 7, 9
Figure 6 shows the wind tunnel test results of the DU93-W-210 airfoil without the VGs. NCEPU
provided the test results in this paper. NPU is the wind tunnel test results from the Northwest
Polytechnic University, and Delft is the wind tunnel test results from Delft University. By comparing
the results of the three wind tunnels, it can be seen that the wind tunnel test results in this paper were
close to those obtained by NPU and Delft Figure 5. Test
before airfoilAoA
a stall ◦ ) for the airfoil lift coefficient. After the
with(8VGs.
stall AoA, the lift coefficient measured in the wind tunnel was slightly higher than that measured in
the other two wind tunnels.Table 3. VG
For the draggeometric parameters
coefficient, and stall
before the arrangement.
AoA, the drag coefficient matched
the results of the other
Installation Anglewind
β (°)tunnels.
LengthHowever, after the
l (mm) Height stall AoA,
H (mm) the wind
Spacing S/H tunnel test results
Pitch Distance λ/Hin this
paper were closer20 to the Delft University
17 wind tunnel test5 results. Since5 the aerodynamic characteristics
3, 5, 7, 9
of airfoils fluctuate after a stall, the lift and drag coefficients must be revised, and the mechanism
of each correction method is different. However, the present study is focused on the aerodynamic
characteristics of the airfoil before the stall.
measured in the other two wind tunnels. For the drag coefficient, before the stall AoA, the drag
coefficient matched the results of the other wind tunnels. However, after the stall AoA, the wind
tunnel test results in this paper were closer to the Delft University wind tunnel test results. Since the
aerodynamic characteristics of airfoils fluctuate after a stall, the lift and drag coefficients must be
revised, and the mechanism of each correction method is different. However, the present study
Appl. Sci. 2019, 9, 5495
is
8 of 22
focused on the aerodynamic characteristics of the airfoil before the stall.

a) b) 0.35
NCEPU
1.5 NPU
0.30
Delft
1.0 0.25

Cl
NCEPU 0.20

Cd
0.5 NPU
Delft 0.15
0.0 0.10
-15-10 -5 0 5 10 15 20 25 30
0.05
-0.5 α (°)
0.00
-15-10 -5 0 5 10 15 20 25 30
-1.0 α (°)
Figure 6. Comparison of experiments: (a) lift coefficient and (b) drag coefficient.
Figure 6. Comparison of experiments: (a) lift coefficient and (b) drag coefficient.
3. Results and Discussion
3. Results and Discussion
3.1. Analysis of the CFD Calculation Results
3.1. Analysis
Figure of 7 the CFD time-averaged
shows Calculation Results axial velocity contours on the symmetrical plane of the
computational
Figure 7 domain. As it can be observed,
shows time-averaged the channel
axial velocity expanded
contours between
on the 0–35H, which
symmetrical planeledoftothean
adverse pressure gradient (APG), and a continuous decrease in fluid velocity.
computational domain. As it can be observed, the channel expanded between 0–35H, which led to an Under the action of the
APG
adverseandpressure
wall viscosity, flow(APG),
gradient separation
and aoccurred
continuousat the lower wall,
decrease and velocity.
in fluid a low-velocity
Under zone
the appeared.
action of
The flow separation
the APG and wall position
viscosity,was
flowapproximately at x1 ≈ 20H.
separation occurred at Since the free-slip
the lower wall, and wallaboundary
low-velocity condition
zone
was adopted on the upper wall of the computational domain, no flow
appeared. The flow separation position was approximately at x1 ≈ 20H. Since the free-slip wall separation occurred. In the
35–135H
boundarysection, there was
condition was aadopted
zero pressure
on thegradient
upper(ZPG),wall and flowcomputational
of the reattachment occurred domain, at xno
2 ≈ flow
60H.
The axial velocity contours with VGs revealed that the VGs inhibited flow
separation occurred. In the 35–135H section, there was a zero pressure gradient (ZPG), and flow separation in all four cases.
When the VG occurred
reattachment spacing was at xλ/H = 7, aThe
2 ≈ 60H. small low-speed
axial velocity region
contours appeared
with VGs at x revealed
≈ 30H. When the VGs
that the VGs
spacing
inhibitedwas λ/Hseparation
flow = 9, the area
in of
allthe low-speed
four cases. Whenregion at VG
the the same
spacing location
was λ/H increased.
= 7, a This
smallshowed
low-speedthat
when λ/H =
region appeared at x ≈ 30H. When the VGs spacing was λ/H = 9, the area of the low-speed regionthe
9, the effectiveness of the concentrated vortices at this location became weak, while at
hydrodynamic
the same location energy was already
increased. low under
This showed thatthe
wheninverse
λ/Hpressure gradient due of
= 9, the effectiveness to the
the large spacing
concentrated
between
vortices atVGs.
thisNo low-velocity
location becamezoneweak,appeared
while the at the other two installations
hydrodynamic energy was with a smaller
already spacing.
low under the
Figure 8 shows the distribution of the pressure coefficient
inverse pressure gradient due to the large spacing between VGs. No low-velocity (C p ) on the bottom wall
zone appeared atof the
computational domain. In the
the other two installations noaVG
with the inflection point of Cp appeared at x ≈ 20H, and the slope
case, spacing.
smaller
of the Cp change decreased. In the cases where the boundary layer flow was controlled by the VGs, the
Cp on the wall increased in the APG section. Since no flow separation occurred, no alteration in the
slope of Cp change in the APG section was observed. The Cp in most positions of the ZPG segment
was basically 0. The difference between different VG spacings was primarily reflected in the turning
position of the Cp curve. In particular, the larger the VG spacing, the smaller the slope of the wall
pressure drop and the worse the flow control effect. After the fluid passed through the expansion
section, the Cp increased. According to Figure 7, the flow separation occurred at x ≈ 20H. After the
flow separation, the slope of the Cp drop decreased, while after the flow reattachment, no significant
changes in Cp were observed in the ZPG section.
Appl. Sci. 2019, 9, 5495 9 of 22
Appl. Sci. 2020, 10, x FOR PEER REVIEW 9 of 24

Figure 7. Spanwise mean velocity contours on the plane of symmetry of the computational domain.

Figure 8 shows the distribution of the pressure coefficient (Cp) on the bottom wall of the
computational domain. In the no VG case, the inflection point of Cp appeared at x ≈ 20H, and the
slope of the Cp change decreased. In the cases where the boundary layer flow was controlled by the
VGs, the Cp on the wall increased in the APG section. Since no flow separation occurred, no
alteration in the slope of Cp change in the APG section was observed. The Cp in most positions of the
ZPG segment was basically 0. The difference between different VG spacings was primarily reflected
in the turning position of the Cp curve. In particular, the larger the VG spacing, the smaller the slope
of the wall pressure drop and the worse the flow control effect. After the fluid passed through the
expansion section, the Cp increased. According to Figure 7, the flow separation occurred at x ≈ 20H.
After Figure
the flow separation, the slope of the Cp drop decreased, while after the flow reattachment, no
7. Spanwise mean velocity contours on the plane of symmetry of the computational domain.
significant changes in Cmean
Figure 7. Spanwise p were observed
velocity in the
contours onZPG section.
the plane of symmetry of the computational domain.

Figure 8 shows the distribution of the pressure coefficient (Cp) on the bottom wall of the
λ/H = 5
computational domain. In the -0.8no VG case, the inflection point ofλ/HC=p7 appeared at x ≈ 20H, and the
λ/H
slope of the Cp change decreased. In the cases where the boundary layer =9 flow was controlled by the
λ/H = 11
-0.6
VGs, the Cp on the wall increased in the APG section. SinceNonocontrolflow separation occurred, no
alteration in the slope of Cp change in the APG section was observed. The Cp in most positions of the
Cp

ZPG segment was basically -0.40. The difference between different VG spacings was primarily reflected
in the turning position of the Cp curve. In particular, the larger the VG spacing, the smaller the slope
of the wall pressure drop and -0.2the worse the flow control effect. After the fluid passed through the
expansion section, the Cp increased. According to Figure 7, the flow separation occurred at x ≈ 20H.
After the flow separation, the slope of the Cp drop decreased, while after the flow reattachment, no
0.0
significant changes in Cp were observed in the ZPG section.
0 40 80 120
x/H
λ/H = 5
-0.8
Figure 8. Distribution of Cp on the bottom wall of the computational
λ/H = 7 domain.
Figure 8. Distribution of Cp on the bottom wall of the computational
λ/H = 9 domain.
Although in some simple λ/H =be
11determined by intuition and
-0.6 flows, the existence of vortices canNo control
visualization, in three-dimensional viscous flows, especially in complex flows, a large number of
Cp

experimental or numerical simulation


-0.4 data are required to show the structure, evolution, and interaction
of vortices. Therefore, it is necessary to give an objective discrimination of the vortices’ criteria. At
present, the Q criterion is more commonly used, where the region of Q > 0 is defined as a vortex, that
-0.2
is kΩk2 > kEk2 . Its physical meaning is that the fluid rotation in the region of the vortex plays a leading
role in comparing the strain rate. The specific formula is as follows:
0.0
1 2 2
0 Q = 2 (kΩk
40 − kEk80) 120 (6)
x/H

Figure 8. Distribution of Cp on the bottom wall of the computational domain.


plays a leading role in comparing the strain rate. The specific formula is as follows:

1 2 2
Q= (Ω − E ) (6)
2

and Ω 2 = Ωij Ω ji = 1 ω 2 , eij , Ω ij are the strain rate tensor and the vorticity
2
2019,=9,e 5495
Appl. Sci. E 10 of 22
ij e ji
where
2
tensor, respectively.
where kEk2 = eij e ji and kΩk2 = Ωij Ω ji = 12 |ω|2 , eij , Ωij are the strain rate tensor and the vorticity
In this
tensor, study, the Q criterion was used to identify the three-dimensional vortices in the flow
respectively.
field. In this study, the Q criterion was used to identify the three-dimensional vortices in the flow field.
Figure 9 shows the Q isosurfaces colored according to the velocity in the flow field (Q = 1 × 105 ).
Figure 9 shows the Q isosurfaces colored according to the velocity in the flow field (Q = 1 × 105).
Many disordered vortices were observed in the turbulent flow field without VGs. However, typical
Many disordered vortices were observed in the turbulent flow field without VGs. However, typical
turbulent coherent structures could be distinguished. In the initial flow stage, under the fluid viscosity
turbulent coherent structures could be distinguished. In the initial flow stage, under the fluid
and shear stress, the vortices were rolled up near the wall (as indicated by A). Due to self-rotation of the
viscosity and shear stress, the vortices were rolled up near the wall (as indicated by A). Due to
vortices, the radius of the vortex cores increased gradually downstream. At the same time, disturbed
self-rotation of the vortices, the radius of the vortex cores increased gradually downstream. At the
by the three-dimensional flow field, the vortices fluctuated extensively, twisted, and protruded upward
same time, disturbed by the three-dimensional flow field, the vortices fluctuated extensively,
along the normal direction. The protrusion was induced by the self-rotation of the spanning vortices,
twisted, and protruded upward along the normal direction. The protrusion was induced by the
which intensified the protrusion height (as indicated by B). Under the action of a shear flow, the vortices
self-rotation of the spanning vortices, which intensified the protrusion height (as indicated by B).
continued to stretch and lift, and gradually developed into closed upper end hairpin vortices with an
Under the action of a shear flow, the vortices continued to stretch and lift, and gradually developed
opening at the lower end (as indicated by C). Moving backwards, the legs of the vortices extended and
into closed upper end hairpin vortices with an opening at the lower end (as indicated by C). Moving
the head of the hairpin vortices became larger, forming a local high-shear layer. Under the action of the
backwards, the legs of the vortices extended and the head of the hairpin vortices became larger,
mainstream fluid, the vortex was stretched along the flow direction into a tube (as indicated by D).
forming a local high-shear layer. Under the action of the mainstream fluid, the vortex was stretched
This indicated that in turbulent flows, the average length of the vortex tube was generally increased.
along the flow direction into a tube (as indicated by D). This indicated that in turbulent flows, the
The energy transport process from larger to smaller vortices and the viscous dissipation into heat is
average length of the vortex tube was generally increased. The energy transport process from larger
called the cascade principle.
to smaller vortices and the viscous dissipation into heat is called the cascade principle.

Figure 9. Q isosurfaces (Q = 1 × 105 ) colored according velocity without VGs: (a) computational
Figure
domain9.and
Q isosurfaces (Q = 1 × 105) colored according velocity without VGs: (a) computational
(b) inlet section.
domain and (b) inlet section.
Figure 10 shows Q isosurfaces colored according to the velocity in the flow field (Q = 1 × 105 )
as aFigure 10the
result of showsVG Q isosurfaces
control. First,colored according
by comparing to the9 velocity
Figures and 10, it incan
the be
flow fieldthat
found (Q =without
1 × 105) the
as
aVG
result of the VG control. First, by comparing Figures 9 and 10, it can be found that
control, many irregular vortices were generated downstream of the computational domain, while without the VG
control, manywas
this number irregular
reduced vortices
when were generated
the flow downstream
was controlled of the
by VGs. Bycomputational
comparing thedomain,
vortices while
under
this number
different VGwas reduced
spacings, it when
can bethe flowthat
found waswith
controlled by VGs.
the increase of By
the comparing
VG spacing, thethe
vortices
heightunder
of the
different VG spacings, it can be found that with the increase of the VG spacing, the
vortices increased. As it can be seen in Figure 10b, the presence of VGs created longitudinal vortices, height of the
vortices
and thisincreased. As itthe
fact organized canlocal
be seen in Figure
vortex 10b,
structure thefurther
and presence of VGsthe
hindered created longitudinal
development vortices,
and movement
of the spanwise vortices. The vortices generated between two VGs rotated in opposite directions, thus
the outward vortices of the VGs were “suppressed.” On the other hand, the upward forces at the inner
side of the VGs were toward the center. Under the action of the two concentrated vortices, the inner
extended vortices were “raised” to form hairpin-like vortices, and as the flow developed downstream,
the concentrated vortices gradually merged with the leg of the hairpin-like vortices. The legs of the
concentrated and hairpin-like vortices played an important role in the exchange of kinetic energy and
energy dissipation near the wall. When λ/H = 3, the centralized vortices generated by the two VGs
were so close to each other that the spanning vortices inside the VGs were not “raised,” but instead
were fused together at a downstream location under a reverse pressure gradient, forming hairpin-like
vortices. When λ/H = 5, the outer spreading vortices of the VGs were “suppressed,” while the inner
spreading vortices were “raised,” forming regular hairpin-like vortices that moved downstream. When
vortices. The legs of the concentrated and hairpin-like vortices played an important role in the
exchange of kinetic energy and energy dissipation near the wall. When λ/H = 3, the centralized
vortices generated by the two VGs were so close to each other that the spanning vortices inside the
VGs were not “raised,” but instead were fused together at a downstream location under a reverse
pressure
Appl. gradient,
Sci. 2019, 9, 5495 forming hairpin-like vortices. When λ/H = 5, the outer spreading vortices11 ofofthe
22
VGs were “suppressed,” while the inner spreading vortices were “raised,” forming regular
hairpin-like vortices that moved downstream. When λ/H = 7 and λ/H = 9, no regular hairpin-like
λ/H = 7 and
vortices wereλ/H = 9, no regular
developed on thehairpin-like
inner side vortices were
due to the developed
large on theSuch
VG spacing. innerdisordered
side due to turbulent
the large
VG
structures were not conducive to the energy exchange in the near-wall region, which became in
spacing. Such disordered turbulent structures were not conducive to the energy exchange the
prone
near-wall region, which became prone to flow separation.
to flow separation.

a)

?/H=3
?/H = 3 ?/H=5
?/H =5

?/H=7
?/H = 7 ?/H=9
?/H = 9

b) Inside Outside
? ? ? ?

?/H=3
?/H = 3 ?/H=5
?/H =5

?/H =7
?/H=7 ?/H=9
?/H =9
Figure 10. Q isosurfaces (Q = 1 × 105 ) colored according to velocity with VGs: (a) computational
domain and (b) inlet section.

Figure 11 shows vorticity isolines and spatial streamlines at five cross-sectional locations
downstream of the VGs. First, as it can be seen in Figure 11a, the concentrated vortices generated by
the VGs scrolled around the vortex core center and rotated in spirals. When λ/H = 7, at x > 20H, the
streamlines began to bend due to the counter-pressure gradient effect, and the flow became unsmooth.
When λ/H = 9, at x > 20H, the streamlines were twisted, which indicated that the VGs were not able to
adequately suppress the flow separation. In Figure 11b, it can be observed that with the increase of
the VG spacing, the distance between the two vortices gradually increased. Downstream of x/H =
15, when λ/H = 7 and λ/H = 9, the shape of the vortices was irregular with no obvious vortex center,
while the area surrounding the vortices increased. When λ/H = 3 and λ/H = 5, the two vortices were
closer to each other, and the shape of the concentrated vortices was still intact. Since the concentrated
unsmooth. When λ/H = 9, at x > 20H, the streamlines were twisted, which indicated that the VGs
were not able to adequately suppress the flow separation. In Figure 11b, it can be observed that with
the increase of the VG spacing, the distance between the two vortices gradually increased.
Downstream of x/H = 15, when λ/H = 7 and λ/H = 9, the shape of the vortices was irregular with no
obvious vortex center, while the area surrounding the vortices increased. When λ/H = 3 and λ/H = 5,
Appl. Sci. 2019, 9, 5495 12 of 22
the two vortices were closer to each other, and the shape of the concentrated vortices was still intact.
Since the concentrated vortices rotated upward at the inner side of the VGs and were subjected to
upwardrotated
vortices forces,upward
the highest
at thevortex core of
inner side height from
the VGs andthe wall
were was observed
subjected to upwardat λ/H = 3.the
forces, When
highest the
distance
vortex corebetween
height fromthe vortices
the wallwaswassmall
observed at λ/H
enough, the=upward
3. Whenrotation forcebetween
the distance made thethe vortices
vortices liftwas
off
the wall.
small This the
enough, implies
upwardthat rotation
in order force
to restrain
made the
the flow separation
vortices using
lift off the VGs,
wall. Thisthe vortices
implies need
that to be
in order
closer
to to the
restrain the wall to give a stronger
flow separation using VGs,energy exchange
the vortices effect
need to beincloser
the near-wall
to the wallarea, anda the
to give more
stronger
beneficial
energy the flow
exchange effectcontrol. When λ/H
in the near-wall = and
area, 3, the
thespacing was somewhat
more beneficial small, which
the flow control. When wasλ/H =not 3,
conducive
the spacing to was flow control. small, which was not conducive to flow control.
somewhat

a)

λ/H=3 λ/H=5

λ/H=7 λ/H=9

Figure 11.
Figure (a) Vorticity
11. (a) Vorticity streamlines
streamlines and
and contours
contours in
in the
the flow
flow domain
domain and
and (b)
(b) vorticity contours at
vorticity contours at
different cross-sections.
different cross-sections.

Figure 12 shows the relationship between the vortex core radius and vortex core spacing with
Figure 12 shows the relationship between the vortex core radius and vortex core spacing with
flow distance, where r represents the vortex radius and ∆z represents the distance between the two
flow distance, where r represents the vortex radius and Δz represents the distance between the two
vortex cores. With the increase of the VG spacing, the ∆z increased gradually. The vortex core radius
represents the radial range of action. Theoretically, when VGs are used to suppress flow separation,
the optimal distance is when the vortex radius r is equal to the distance between the two vortex cores
∆z/2 (half the distance). However, the radius of the vortices increased along the flow direction, and the
vortices slightly deviated toward the flow direction. Therefore, the radius of the vortices and spacing
between vortices cannot be equal everywhere along the flow direction. When the distance between
vortices was ∆z/2 < r, VGs hindered the development of vortices. When ∆z/2 > r, a region where the
vortices could not act was developed, which led to the decrease of the added value of fluid kinetic
the optimal distance is when the vortex radius r is equal to the distance between the two vortex cores
Δz/2 (half the distance). However, the radius of the vortices increased along the flow direction, and
the vortices slightly deviated toward the flow direction. Therefore, the radius of the vortices and
spacing between vortices cannot be equal everywhere along the flow direction. When the distance
between vortices
Appl. Sci. 2019, 9, 5495was Δz/2 < r, VGs hindered the development of vortices. When Δz/2 > r, a region 13 of 22
where the vortices could not act was developed, which led to the decrease of the added value of
fluid kinetic energy in the boundary layer. By comparing the four different VG spacing results, it can
energy
be seen in theexcept
that, boundary layer.
at x/H = 1,Bythe
comparing
distance the four different
between VG spacing
the vortices Δz wasresults,
smalleritthan
can be theseen that,
vortex
except at x/H = 1, the distance between the vortices ∆z was smaller than
radius r at the other locations when λ/H = 3. This shows that this VG spacing hindered the the vortex radius r at the other
locations when
development of λ/H = 3. at
vortices This shows
most that this
positions VG the
along spacing
flowhindered
direction,the
anddevelopment
that the VGofspacing
vorticeswasat most
too
positions along the flow direction, and that the VG spacing was too small
small to facilitate flow control. When λ/H = 7 and 9, Δz > r for each direction, which reduced theto facilitate flow control.
When λ/H
effective = 7 of
range andthe ∆z > r for
9, vortices each
and direction,
was which reduced
not conducive the effective
to flow control. When range
λ/H of theit vortices
= 5, was found and
wasΔz
that not> conducive
r before x/H to flow
= 10, control. When
while after x/Hλ/H = 5,
= 10, Δzit <was found that ∆z >
r. Consequently, r before
a VG x/H =
spacing of10,
λ/H while after
= 5 was
x/H = 10, ∆z < r. Consequently,
more suitable for flow control. a VG spacing of λ/H = 5 was more suitable for flow control.

r Δz/2
2
λ/H = 3
z/H

1
0 5 10 15 20
2 λ/H = 5
z/H

1
0 5 10 15 20
4
λ/H = 7
z/H

2
0 5 10 15 20
6
z/H

4 λ/H = 9
2
0 5 10 15 20
x/H
Figure 12. Relationship between the vortex core radius and vortex core spacing with flow direction.
Figure 12. Relationship between the vortex core radius and vortex core spacing with flow direction.
Second-order spatial correlation is a statistical method commonly used to study turbulence.
Second-order
Two-point spatial spatial correlation
correlation is a the
can reflect statistical
spatialmethod commonly
relationship of oneused to study
or more turbulence.
pulsations in the
Two-point
Appl. Sci. 2020, 10, xspatial
FOR correlation
PEER REVIEW can reflect the spatial relationship of one or more pulsations
turbulent flow field. This method can be used to analyze the correlation of the structural characteristics in the 15 of 24
turbulent flow field.
of the downstream VGThis method
vortices. can be used
The correlation to analyze
function theascorrelation
is defined follows: of the structural
characteristics of the downstream VG vortices.
that they are completely unrelated. For Duniform turbulence, The correlation function is defined as follows: coefficient is
the correlation
E
0 0
independent of the coordinate position, Rij (r, xand
) = related
ui (x, t)u j (only
x + r,to
t) ,the correlation distance. In (7) order to
Rij (r , x ) = momentum
investigate the correlation of the turbulent ui' ( x, t )u 'j ( x + downstream
r, t) , of the VGs, the correlation
(7)
where ui and uj are components of the pulsating velocity, and x and r are vectors. When i = j, this
characteristics
expresses
of the turbulent momentum
theuautocorrelation of of
a certain
v′ alongand
pulsation,
the flowi direction
j, it rmeans
atthat
certain points at different
where ui and j are components the pulsating velocity,when and x,and the
are vectors. two pulsations
When i = j, this
heights oncorrelated.
are a pair
expresses
of The
centrosymmetric
autocorrelation
the autocorrelation
planes
diagram
of a certain
of the VGs
is shown
pulsation, and
werei compared.
in when
Figure 13.
≠ j, it means that the two pulsations
are correlated. The autocorrelation diagram is shown in Figure 13.
The above formula can be transformed into dimensionless correlation coefficients as follows:

ui' ( x, t )u 'j ( x + r , t )
Rij = . (8)
ui'2 ( x, t )u 'j2 ( x + r , t )

The correlation coefficient magnitude represents the statistical similarity between two random
variables or the probability that the fluctuation takes the same sign at different positions in the flow
Figure
field. Rij' = ±1 denotes that two 13. Autocorrelation
random of pulse velocity
variables are completely u.
correlated, while Rij' = 0 denotes
Figure 13. Autocorrelation of pulse velocity u.
The above formula can be transformed into dimensionless correlation coefficients as follows:
Figure 14 illustrates the autocorrelation Rv' 'v ' ( r1 ) for the calculation of the
 coefficient curve 
u0i (x, t)u0j (x + r, t)
fluctuating velocity v′ at different positions
Rij = r on the central symmetry . plane of the domain.(8) Five
u02
reference points at different heights and different
i
)u02
(x, tpoints
j
(x +along
r, t) the flow direction were selected for
coherence analysis (Figure 14a). The minimum correlation distance was 0.5 mm, the maximum
correlation distance was 675 mm, and the VGs were placed at the 0 position. According to Figure
14b, the coherence coefficient at the reference point is 1, indicating that the reference point was
correlated with itself. In addition, the larger the correlation coefficient near the reference point, the
larger the correlation coefficient amplitude, and the higher the correlation between the fluctuation
and the reference point. A comparison between the following four cases demonstrated that when
Appl. Sci. 2019, 9, 5495 14 of 22

The correlation coefficient magnitude represents the statistical similarity between two random
variables or the probability that the fluctuation takes the same sign at different positions in the flow
field. R0ij = ±1 denotes that two random variables are completely correlated, while R0ij = 0 denotes
that they are completely unrelated. For uniform turbulence, the correlation coefficient is independent
of the coordinate position, and related only to the correlation distance. In order to investigate the
correlation of the turbulent momentum downstream of the VGs, the correlation characteristics of the
turbulent momentum v0 along the flow direction at certain points at different heights on a pair of
centrosymmetric planes of the VGs were compared.
Figure 14 illustrates the autocorrelation coefficient curve R0v0 v0 (r1 ) for the calculation of the
fluctuating velocity v0 at different positions on the central symmetry plane of the domain. Five
reference points at different heights and different points along the flow direction were selected for
coherence analysis (Figure 14a). The minimum correlation distance was 0.5 mm, the maximum
correlation distance was 675 mm, and the VGs were placed at the 0 position. According to Figure 14b,
the coherence coefficient at the reference point is 1, indicating that the reference point was correlated
with itself. In addition, the larger the correlation coefficient near the reference point, the larger the
correlation coefficient amplitude, and the higher the correlation between the fluctuation and the
reference point. A comparison between the following four cases demonstrated that when λ/H = 3, the
magnitude of the correlation coefficient was the largest at most locations, which indicated that at this
VG spacing, the structure of turbulence in the downstream domain preserved a high organization.
When λ/H = 7 and 9, the magnitude of the correlation coefficient was larger at most locations, while that
of λ/H = 5 was closer to R0v0 v0 (r1 ) = 0 at most locations. The correlation between the turbulent vortices
downstream of the VGs and the reference point was poor, indicating that this VG spacing (λ/H = 5) was
more suitable for suppressing flow separation. The width between the correlation coefficient curve and
the two intersections reflected the scale of the coherent structure in the downstream turbulent flow field.
It can be observed that the scale of the vortices increased gradually as the flow developed downstream.
Figure 15 shows the average velocity distribution at different positions along the downstream
direction. <u>/U∞ represents the ratio of the time-averaged velocity to the mainstream velocity and
y/H is the ratio of the normal wall height to the VG height. It can be observed that the increase in
the kinetic energy of the fluid at the bottom of the boundary layer corresponded to a decrease in the
kinetic energy of the fluid above the boundary layer. The S-type velocity profile took the center of
the vortex core as the symmetrical point, which is the main principle of using VGs to control flow
separation. In particular, due to the rotation of the concentrated vortices, the high-energy fluid outside
of the boundary layer was transferred into the bottom of the boundary layer, inhibiting the boundary
layer separation. By comparing the velocity profiles developed under the four VG spacings, it can be
seen that the kinetic energy of the fluid at the bottom of the boundary layer increased with the decrease
of the VG spacing. However, when λ/H = 3, downstream of x/H = 15, the center of the vortex core was
higher than the wall, and the high-energy fluid in the boundary layer was also higher than the wall
height. When λ/H = 5, the kinetic energy at the bottom of the boundary layer was the largest.
v 'v ' 1

between the turbulent vortices downstream of the VGs and the reference point was poor, indicating
that this VG spacing (λ/H = 5) was more suitable for suppressing flow separation. The width
between the correlation coefficient curve and the two intersections reflected the scale of the coherent
structure in the downstream turbulent flow field. It can be observed that the scale of the vortices
Appl. Sci. 2019, 9, 5495 15 of 22
increased gradually as the flow developed downstream.

x/H
Appl. Sci. 2020, 10, x FOR PEER REVIEW 16 of 24
(a) Reference points of self-correlation
?/H = 3 ?/H = 5 ?/H = 7 ?/H = 9
y/H = 0.5 x/H = 1

y/H =1 x/H = 5

y/H = 2 x/H = 10

y/H = 4 x/H = 15

y/H = 6 x/H = 20

(b) Correlation coefficients

Figure 14. Correlation coefficient distribution of fluctuating velocity v0 .


Figure 14. Correlation coefficient distribution of fluctuating velocity v′.

Figure 15 shows the average velocity distribution at different positions along the downstream
direction. <u>/U∞ represents the ratio of the time-averaged velocity to the mainstream velocity and
y/H is the ratio of the normal wall height to the VG height. It can be observed that the increase in the
kinetic energy of the fluid at the bottom of the boundary layer corresponded to a decrease in the
kinetic energy of the fluid above the boundary layer. The S-type velocity profile took the center of
the vortex core as the symmetrical point, which is the main principle of using VGs to control flow
separation. In particular, due to the rotation of the concentrated vortices, the high-energy fluid
outside of the boundary layer was transferred into the bottom of the boundary layer, inhibiting the
boundary layer separation. By comparing the velocity profiles developed under the four VG
spacings, it can be seen that the kinetic energy of the fluid at the bottom of the boundary layer
increased with the decrease of the VG spacing. However, when λ/H = 3, downstream of x/H = 15, the
center of the vortex core was higher than the wall, and the high-energy fluid in the boundary layer
was also higher than the wall height. When λ/H = 5, the kinetic energy at the bottom of the boundary
Appl. Sci.
Appl. Sci. 2019,
2020, 9,
10,5495
x FOR PEER REVIEW 17 of 22
16 of 24
Appl. Sci. 2020, 10, x FOR PEER REVIEW 17 of 24

Figure 15.
Figure 15. Spanwise
Spanwise mean
mean velocity
velocity profiles.
profiles.
Figure 15. Spanwise mean velocity profiles.
Figure 16
Figure 16 shows
shows the
the relationship
relationship between
between thethe pressure
pressure loss
loss coefficient
coefficientCC∆P
ΔP and the VG spacing.

The Figure
The 16 shows
percentage
percentage the relationship
decrease
decrease ΔP with between
of CC∆P
of the pressure
VGs compared to thatloss coefficient
without VGs C(clean)
ΔP and the VG spacing.
can be observed.
The
Whenpercentage
VGs weredecrease
installed,of C
the
ΔP Cwith inVGs
the compared
calculation to that
domain without VGs
decreased
When VGs were installed, the C∆P in the calculation domain decreased compared to that
ΔP (clean)
compared cantobe observed.
that without
without VGs.
When
Among VGs
VGs. Among were
them, the installed,
them,
C∆Pthe
was CΔP the
thewasC in the calculation
when λ/H
the smallest
ΔP
smallest when domain
= 5,λ/H
and= the decreased
5, and compared
the maximum
maximum to
C∆P wasCreduced that without
ΔP was reduced
by 30.95%.by
VGs. Among
30.95%. When them,
λ/H the
= C
9, the
ΔP was
C ΔPthe smallest
was the when
largest, λ/H
which = 5, and
was the maximum
When λ/H = 9, the C∆P was the largest, which was reduced by 2.37% compared to that without VGs.
reduced by C
2.37% was reduced
ΔPcompared to by
that
30.95%.
withoutWhen
Therefore, λ/H = 9,more
λ/HTherefore,
VGs. = 5 was the
λ/HCsuitable
=ΔP5 was for
theflow
more largest,
suitablewhich
control. wascontrol.
for flow reduced by 2.37% compared to that
without VGs. Therefore, λ/H = 5 was more suitable for flow control.
0.45
0.45
0.40
0.40 -2.37%
0.35 -2.37%
?P
C?PC

0.35
-11.44%
0.30 -22.58%
-11.44%
0.30 -22.58%
0.25
0.25 -30.95%
-30.95%
0.20
0.20 Clean ?/H=3
Clean ?/H ?/H=5
= 3 ?/H ?/H=7
= 5 ?/H = 7 ?/H=9
?/H = 9
Clean ?/H=3
Clean ?/H = 3 ?/H = 5 ?/H = 7 ?/H=9
?/H=5 ?/H=7 ?/H = 9
Figure 16. Relationship between the pressure loss coefficient C∆P and VGs spacings.
Figure 16. Relationship between the pressure loss coefficient CΔP and VGs spacings.
3.2. AnalysisFigure
of the Experimental Results
16. Relationship between the pressure loss coefficient CΔP and VGs spacings.
3.2. Analysis
Figure 17ofshows the Experimental
the lift–drag Results
coefficient and lift–drag ratio curves of the airfoil with and without
3.2. Analysis of the Experimental Results
VGs. Figure
Figure 17 17ashows demonstrates the liftcoefficient
the lift–drag coefficientand curve. It can be
lift–drag seencurves
ratio that theof lift
thecoefficients
airfoil with of and
the
airfoilFigure
without withVGs. 17 Figure
and shows
without the
17a VGs lift–drag coefficient
were almost
demonstrates thethe andfor
liftsame lift–drag
coefficient ancurve. ratio
angle-of-attack
It can curves ofthat
(AoA)
be seen the
below
the 8◦ .coefficients
airfoil
lift with
Above 8◦ ,
and
without
the
of the VGs.
couldFigure
VGsairfoil with 17a without
effectively
and demonstrates
increase the
thewere
VGs lift
lift of coefficient
the
almost airfoil.
the same curve.
In forIt 4,
Table can
an be seen that that
itangle-of-attack
can be seen the(AoA)
lift
when λ/H =8°.
coefficients
below 9,
of the airfoil ◦ ◦
the
Abovestall AoA
8°, thewith of
VGstheand
could without
airfoil was 16 VGs
effectively were under
, while
increase almost the
the
the lift same
other
of VGforspacings,
the airfoil. anInangle-of-attack
Tablethe stall
4, it canAoA(AoA)
be wasbelow
seen 8°.
18 .when
that The
Above
maximum 8°, the
lift VGs could
coefficient effectively
was increase
obtained when theλ/H
lift= of5,the
whichairfoil.
was
λ/H = 9, the stall AoA of the airfoil was 16°, while under the other VG spacings, the stall AoA was In Table
increased 4, it
bycan be
48.77% seen that
comparedwhen to
λ/H
that =
18°. of 9, the stall
themaximum
The AoA
clean airfoil. of the
liftWhen λ/H = 3,
airfoil
coefficient was
wasλ/Hobtained
= 7, and when
16°, while λ/H = 9,
under the
λ/H other
the=maximum VG spacings,
5, which lift the
wascoefficient stall AoA
increasedincreased was
by 48.77%by
18°.
45.67%,The48.32%,
compared maximum
to that and lift
of 30.20%,coefficient
the clean airfoil.was
respectively.
When obtained
λ/H = 3,when
Therefore, =λ/H
the maximum
λ/H 7, and= 5,λ/H which
lift was
=coefficient
9, the increased liftby
of the airfoil
maximum 48.77%
was the
coefficient
compared
largest
increasedwhen to λ/H
bythat45.67%, = the
of clean
5. 48.32%,
On the airfoil.
other Whenthe
hand,
and 30.20%, λ/H = 3,coefficient
drag
respectively.λ/H =Therefore,
7, andwasλ/H the
the= worst
9, the maximum
maximum whenliftλ/Hcoefficient
=lift9.coefficient
In Table 3,
of the
increased
itairfoil
can bewas bythe
seen 45.67%,
that
largest 48.32%,
when whenλ/H λ/H
= 3,=30.20%,
and the drag
5. On respectively.
coefficient
the other hand, ofTherefore,
the
the airfoil the maximum
decreased
drag coefficient by lift
was thecoefficient
82.06%
worst compared
whenof theto
λ/H
airfoil
that ofwas
= 9. In the the
Table 3,largest
clean itairfoil
can bewhen an λ/H
atseen that=when
AoA of On◦ .λ/H
5. 18 the other
In addition,
= 3, thehand, the drag coefficient
it decreased
drag coefficient byof 83.28% was
when
the airfoil theλ/H = 5,when
worst
decreased λ/H
by 78.43%
82.06%
=when
comparedλ/H =
9. In Table to3,7,that
itand
can be
ofby seen
the13.06%
clean that when
when
airfoil λ/H
at an =
λ/H =9.3,The
AoA the drag Incoefficient
lift–drag
of 18°. ratio of
addition, itofdecreased
theairfoil
the airfoilbydecreased
increased
83.28% by by741.03%
when82.06%
λ/H =
compared
compared
5, by 78.43% to
tothatthatofofλ/H
when the
theclean
=clean airfoil
7, and airfoilatat
anan
by 13.06% AoA
AoA
when of of
18°.
λ/H

18=In9. addition,
when λ/Hit=decreased
The lift–drag 3, ratio of by
by 821.86%the83.28%
airfoil λ/H =λ/H
when
whenincreased 5, by=
by
5, by
612.06% 78.43%when when λ/H =
λ/H 7,= 7,
and and
by by 13.06%
55.06% whenwhen λ/H =
λ/H =
9. 9. The lift–drag
741.03% compared to that of the clean airfoil at an AoA of 18° when λ/H = 3, by 821.86% when λ/H = ratio of the airfoil increased by
741.03% compared
5, by 612.06% when toλ/H
that=of 7,the
andclean airfoil when
by 55.06% at an AoA
λ/H =of9.18° when λ/H = 3, by 821.86% when λ/H =
5, by 612.06% when λ/H = 7, and by 55.06% when λ/H = 9.
Appl. Sci. 2019, 9, 5495 17 of 22
Appl. Sci. 2020, 10, x FOR PEER REVIEW 18 of 24

a) 2.5 b) Clean
λ/H = 3
c) Clean
λ/H = 3
0.3 λ/H = 5 λ/H = 5
2.0 2
λ/H = 7 λ/H = 7
λ/H = 9 λ/H = 9
1.5
Cl

Cd
0.2
Clean
1.0 λ/H = 3
1

Cl
λ/H = 5
0.5 λ/H = 7
λ/H = 9 0.1
0.0 0
-10 0 10 20 30 0.0 0.1 0.2 0.3
α (°) Cd
-0.5
0.0
-10 0 10 20 30
-1.0 α (°) -1

17.(a)(a)
Figure 17.
Figure Curves
Curves of the
of the lift lift coefficient,
coefficient, (b) drag
(b) drag coefficient,
coefficient, and (c)and (c) lift–drag
lift–drag ratio
ratio with andwith and
without
without
VGs. VGs.

Table 4. Relative variation of Cl , Cd , and lift–drag ratio under different VG spacings.


Table 4. Relative variation of Cl, Cd, and lift–drag ratio under different VG spacings.
Stall Increase of Decrease of Cd at Increase of Lift–Drag
Case Cd◦ at 18° IncreaseRatio
Case Stall Angle-of-Attack Increase of Maximum
Angle-of-Attack Maximum ClCl Decrease of 18 at 18◦
of Lift–Drag Ratio at 18°
Clean Clean 8° 8◦ 0% 0% 0% 0% 0% 0%
λ/H = 3 λ/H = 318° 18◦ 45.67% 45.67% 82.06%
82.06% 741.03%
741.03%
λ/H = 5 λ/H = 518° 18◦ 48.77% 48.77% 83.28%
83.28% 821.86%
821.86%
λ/H = 7 λ/H = 718° 18◦ 48.32% 48.32% 78.43%
78.43% 612.06%
612.06%
λ/H = 9 λ/H = 916° 16◦ 30.20% 30.20% 13.06%
13.06% 55.06%
55.06%

In
In Figure
Figure 18,
18,the
thesurface
surfacepressure
pressurecoefficient
coefficient(C(Cp) distribution of the airfoil is presented. It can be
p ) distribution of the airfoil is presented. It can
seen thatthat
be seen thethe
curves
curves of of
CpCdistribution on the airfoil surface with and without VGs coincided at 0°◦
p distribution on the airfoil surface with and without VGs coincided at 0
and
and 5 AoA. At AoAs of 10 , 15 , and 18◦ , the
5°◦ AoA. At AoAs of 10°,◦ 15°,◦ and 18°, the VGs
VGs postponed
postponedthe thepressure
pressure plateau
plateauto tosome
someextent.
extent.
However,
However, whenwhen λ/H
λ/H ==9, 9, the
the effect
effect of
of the
the VGs
VGs flow
flow control
control was was the
the worst,
worst, and
and the
the pressure
pressure plateau
plateau
appeared
appeared earlier
earlier than
than inin the
the other
other three
three cases.
cases. It
It can
can bebe observed
observed thatthat the
the CCpp curves
curves ofof the
the other
other three
three
VG spacings basically coincided. At an AoA of 20°,
◦ the C p curves and the position of the pressure
VG spacings basically coincided. At an AoA of 20 , the Cp curves and the position of the pressure
platform
platform ofof λ/H
λ/H== 99 were
weresimilar
similartotothose
thoseofof
thethe clean
clean airfoil,
airfoil, while
while the the Cp curves
Cp curves of other
of the the other
threethree
VGs
VGs spacings were similar. At 22°, the C p curves of λ/H = 3 and λ/H = 5 basically coincided with the
spacings were similar. At 22 , the Cp curves of λ/H = 3 and λ/H = 5 basically coincided with the clean

clean airfoil,
airfoil, whilewhile
λ/H =λ/H = 7 exhibited
7 exhibited the effect.
the best best effect.
When When the AoA
the AoA was was equalequal
to 25to
◦ , 25°,
the Cthe Cp curves of
p curves of the
the airfoil
airfoil withwith
VGsVGs coincided
coincided withwith
thosethose of the
of the clean
clean airfoil.
airfoil.
Figure 19 shows the curve of the total pressure coefficient distribution at the wake rake. It can be
-1.5
seen that at AoAs of α = 0◦ , 5◦ , and 10◦ , forClean
an airfoil with -2.0VGs, the total pressure loss at the wake
Clean rake
λ/H = 3 λ/H = 3
was larger
-1.0 because the generators increased λ/H =the
5 thickness
-1.5 of the boundary layer and thus λ/H increased
=5
the viscous losses. When the AoA was 15◦λ/H and= 7 18◦ , the total pressure loss of the airfoil withλ/H =7
VGs was
-0.5 λ/H = 9 -1.0 λ/H = 9
smaller than that of the clean airfoil. Among the
α = 0° four VG spacings, the total pressure loss of the
α = 5° airfoil
-0.5
Cp
Cp

with VGs 0.0 was the largest when λ/H = 9, and under the other three spacings, the total pressure loss was
0.0
similar. At α = 20◦ , the total pressure loss was the largest when λ/H = 9 and the smallest when λ/H = 7.
0.5 0.5
At α = 22◦ , the total pressure loss was the largest when λ/H = 7, which was larger than that of the clean
1.0
1.0
airfoil, while the total pressure loss of λ/H = 7 was the smallest. When α = 25◦ , the total pressure loss
was the same, 1.5
0.0 with
0.2or without
0.4 VGs. 0.8
0.6 1.0 0.0 0.2 0.4 0.6 0.8 1.0
x/C x/C
appeared earlier than in the other three cases. It can be observed that the Cp curves of the other three
VG spacings basically coincided. At an AoA of 20°, the Cp curves and the position of the pressure
platform of λ/H = 9 were similar to those of the clean airfoil, while the Cp curves of the other three
VGs spacings were similar. At 22°, the Cp curves of λ/H = 3 and λ/H = 5 basically coincided with the
clean airfoil,
Appl. Sci. 2019, 9,while
5495 λ/H = 7 exhibited the best effect. When the AoA was equal to 25°, the Cp curves
18 of of
22
the airfoil with VGs coincided with those of the clean airfoil.

-1.5
Clean -2.0 Clean
λ/H = 3 λ/H = 3
-1.0 λ/H = 5 -1.5 λ/H = 5
λ/H = 7 λ/H = 7
-0.5 λ/H = 9 -1.0 λ/H = 9
α = 0° α = 5°
-0.5

Cp
Cp

0.0
0.0
0.5 0.5
1.0
1.0
Appl. Sci. 2020, 0.0
10, x FOR 1.5
0.2PEER0.4
REVIEW
0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
19 of 24
x/C x/C

-4 Clean -6 Clean
VGs installation location λ/H = 3 λ/H = 3
λ/H = 5 λ/H = 5
-3 Δx≈88H λ/H = 7 Δx≈46.6H λ/H = 7
-4
λ/H = 9 λ/H = 9
-2 α = 10° α = 15°

Cp
Flow separation position of Clean
Δp -2
Cp

-1 (kpa  m -1)
≈ −0.77
Δx Δp
≈ −1.62
Δx
0 0
1
2
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
x/C x/C
-8 -8
Clean Clean
λ/H = 3 λ/H = 3
-6 λ/H = 5 -6 λ/H = 5
λ/H = 7 λ/H = 7
Δx≈30H λ/H = 9 Δx≈20H λ/H = 9
-4 α = 18° -4 α = 20°
Cp
Cp

-2 Δp
-2 Δp
≈ −2.4 ≈ −2.96
Δx Δx
0 0

2 2
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
x/C x/C
-8
Clean -6 Clean
λ/H = 3 λ/H = 3
-6 λ/H = 5 λ/H = 5
λ/H = 7 λ/H = 7
-4
Δx≈10H λ/H = 9 λ/H = 9
-4 α = 22° α = 25°
Flow separation position ≈ 0.18C
Cp

Cp

-2
-2
Δp 0
0 ≈ −3.95
Δx

2 2
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
x/C x/C

Figure 18. Cp distribution on the airfoil surface.


Figure 18. Cp distribution on the airfoil surface.

Figure 19 shows the curve of the total pressure coefficient distribution at the wake rake. It can
be seen that at AoAs of α = 0°, 5°, and 10°, for an airfoil with VGs, the total pressure loss at the wake
rake was larger because the generators increased the thickness of the boundary layer and thus
increased the viscous losses. When the AoA was 15° and 18°, the total pressure loss of the airfoil with
VGs was smaller than that of the clean airfoil. Among the four VG spacings, the total pressure loss of
the airfoil with VGs was the largest when λ/H = 9, and under the other three spacings, the total
pressure loss was similar. At α = 20°, the total pressure loss was the largest when λ/H = 9 and the
smallest when λ/H = 7. At α = 22°, the total pressure loss was the largest when λ/H = 7, which was
larger than that of the clean airfoil, while the total pressure loss of λ/H = 7 was the smallest. When α =
25°, the total pressure loss was the same, with or without VGs.
Appl. Sci. 2019, 9, 5495 19 of 22
Appl. Sci. 2020, 10, x FOR PEER REVIEW 20 of 24

Clean λ/H = 3 λ/H = 5 λ/H = 7 λ/H = 9 Clean λ/H = 3 λ/H = 5 λ/H = 7 λ/H = 9

α = 0° α = 5° α = 10° α = 15°
1.8 1.8 1.8 1.8
1.5 1.5 1.5 1.5
1.2 1.2 1.2 1.2

z (m)
z (m)

z (m)

z (m)
0.9 0.9 0.9 0.9
0.6 0.6 0.6 0.6
0.3 0.3 0.3 0.3
0.0 0.0 0.0 0.0
0.8 0.9 1.0 0.8 0.9 1.0 0.7 0.8 0.9 1.0 1.1 0.6 0.8 1.0
Cp Cp Cp Cp
Clean λ/H = 3 λ/H = 5 λ/H = 7 λ/H = 9 Clean λ/H = 3 λ/H = 5 λ/H = 7 λ/H = 9

α = 18° α = 20° α = 22° α = 25°


1.8 1.8 1.8 1.8
1.5 1.5 1.5 1.5
1.2 1.2 1.2 1.2

z (m)
z (m)
z (m)
z (m)

0.9 0.9 0.9 0.9


0.6 0.6 0.6 0.6
0.3 0.3 0.3 0.3
0.0 0.0 0.0 0.0
0.4 0.6 0.8 1.0 0.4 0.6 0.8 1.0 0.0 0.3 0.6 0.9 1.2 0.0 0.4 0.8 1.2
Cp Cp Cp Cp

Figure19.
Figure Distributionofoftotal
19.Distribution totalpressure
pressurecoefficient
coefficientatatthe
thewake
wakerake.
rake.

4. Conclusions
4. Conclusions
When VGs are used for the flow control of wind turbine blades, they are always installed under a
When VGs are used for the flow control of wind turbine blades, they are always installed under
certain arrangement. In this paper, the LES method was used to simulate the boundary layer separation
a certain arrangement. In this paper, the LES method was used to simulate the boundary layer
flow with and without VGs in an adverse pressure gradient environment. The large-scale coherent
separation flow with and without VGs in an adverse pressure gradient environment. The large-scale
structure of the boundary layer separation caused by the adverse pressure gradient and its evolution
coherent structure of the boundary layer separation caused by the adverse pressure gradient and its
process in the turbulent flow field were analyzed. The effect of different VG spacings on the suppression
evolution process in the turbulent flow field were analyzed. The effect of different VG spacings on
of the boundary layer separation were compared through the distance between vortex cores, the kinetic
the suppression of the boundary layer separation were compared through the distance between
energy of the fluid in the boundary layer, and the pressure loss coefficient. Subsequently, the effect of
vortex cores, the kinetic energy of the fluid in the boundary layer, and the pressure loss coefficient.
VG spacings on the lift–drag characteristics, airfoil surface pressure distribution, and airfoil pressure
Subsequently, the effect of VG spacings on the lift–drag characteristics, airfoil surface pressure
loss were investigated in a wind tunnel using the DU93-W-210 airfoil as the research object.
distribution, and airfoil pressure loss were investigated in a wind tunnel using the DU93-W-210
According to the relationship between the vortex core radius and the distance between the vortex
airfoil as the research object.
cores, when the VG spacing was λ/H = 3, the distance between the vortex cores ∆z/2 was smaller than
the vortex
According to the rrelationship
core radius in most positions
betweenalong thethevortex
flow direction.
core radiusTherefore,
and thethis spacingbetween
distance inhibitedthe the
development of vortices. When λ/H = 7 and 9, ∆z/2 was always larger
vortex cores, when the VG spacing was λ/H = 3, the distance between the vortex cores Δz/2 wasthan r. When λ/H = 5, at x/H
= 10, the
smaller ∆z/2the
than the r were
andvortex core approximately
radius r in most equal.
positions the λ/H
Thus, along the= 5flow
spacing was more
direction. suitablethis
Therefore, for
flow control.
spacing Thethe
inhibited increase in fluid of
development kinetic energy
vortices. in the
When λ/Hboundary
= 7 and 9,layer
Δz/2was
wasinversely
always largerproportional
than r.
to theλ/H
When VG =spacing.
5, at x/HHowever,
= 10, the when
Δz/2 andthe VGthe spacing was λ/H = 3, the
r were approximately distance
equal. Thus,between
the λ/H vortices
= 5 spacingwas
small
was andsuitable
more the vortexforcore
flowwas far from
control. Thethe wall, resulting
increase in the high-energy
in fluid kinetic energy in the fluid in the downstream
boundary layer was
boundaryproportional
inversely layer being higher fromspacing.
to the VG the wall.However,
In the near-wall
when theregion, the kinetic
VG spacing wasenergy
λ/H = of 3, the fluid was
distance
the largest when λ/H = 5, especially at the x/H = 35 cross-section. When
between vortices was small and the vortex core was far from the wall, resulting in the λ/H = 5, the C between
∆Phigh-energy the
inlet in
fluid andthe outlet of the computational
downstream boundary layerdomain washigher
being the smallest, which
from the was
wall. Inby 30.95%
the lowerregion,
near-wall than that thein
the computational domain without VGs.
kinetic energy of the fluid was the largest when λ/H = 5, especially at the x/H = 35 cross-section.
WhenInλ/H the=wind
5, thetunnel experimental
CΔP between the inletresults, it wasof
and outlet found that the stall AoA
the computational domainof the airfoil
was the with VGs
smallest,
was increased
which by 10◦ compared
was by 30.95% lower thantothatthatinofthe
thecomputational
clean airfoil. When
domainλ/Hwithout
= 5, theVGs.
maximum lift coefficient
of the airfoil with VGs increased by 48.77%, which was the largest increase. Furthermore, at an AoA
of 18In
◦ , the wind
when λ/Htunnel experimental
= 5, the results,
drag coefficient it was found
decreased that the
by 83.28%, stall was
which AoAtheof the airfoil
largest with VGs
reduction. At
was increased by 10° compared to that of the clean airfoil. When λ/H
an AoA of 18 , the lift–drag ratio of the airfoil increased by 741.03% when λ/H = 3, 821.86% when
◦ = 5, the maximum lift
coefficient of the airfoil with VGs increased by 48.77%, which was the largest increase. Furthermore,
at an AoA of 18°, when λ/H = 5, the drag coefficient decreased by 83.28%, which was the largest
reduction. At an AoA of 18°, the lift–drag ratio of the airfoil increased by 741.03% when λ/H = 3,
Appl. Sci. 2019, 9, 5495 20 of 22

λ/H = 5, 612.06% when λ/H = 7, and 55.06% when λ/H = 9. Consequently, it can be deduced that,
when the VG spacing was λ/H = 5, the comprehensive effect of airfoil flow control was the best. The
optimization of the VG spacing was essentially the optimization of vortex scale and vortex core spacing.
The vortex scale and vortex core spacing were related to the VGs installation angle, geometric height,
and arrangement spacing.

Author Contributions: X.-k.L., W.L., T.-j.Z., and P.-m.W. conceived the research idea. X.-k.L. and X.-d.W.
performed the numerical simulation. P.-m.W. and T.-j.Z. analyzed data and numerical results. All authors
contributed to the writing, editing, and revising of this manuscript.
Funding: This work was jointly supported by the National Natural Science Foundation of China (no. 51806221)
and the Development Plan of Key Scientific and Technological Projects of China Huadian Engineering Co., LTD
(CHEC) (no. CHECKJ 19-02-03).
Conflicts of Interest: The authors declare no conflict of interest.

Nomenclature
∆t Unsteady computational physical time step (s)
β Angle between vortex generators and incoming flow (◦ )
l Vortex generators chord length (mm)
H Vortex generators height (mm)
S Distance between the trailing edge of the vortex generators in the opposite direction
λ Distance between the trailing edge of the vortex generators in the same direction
C∆p Pressure loss coefficient at the inlet and outlet of the calculation domain
Pinlet Total pressure at the inlet
Poutlet Total pressure at the outlet
RE The value obtained using the Richardson extrapolation
R The ratio of errors
p The order of accuracy
Cp Airfoil surface static pressure coefficient
Cl Lift coefficient of airfoil
Cd Drag coefficient of airfoil
Ulocal Local velocity
U Inlet velocity
Q Vortex core criterion
Ω Vortex core vorticity
r Vortex core radius
∆z The distance between the two vortex cores
<u> Time-averaged velocity
Clean Airfoil without VGs

Abbreviations
VGs Vortex generators
LES Large eddy simulation
AoA Angle-of-attack
CFD Computational fluid dynamics
NCEPU North China Electric Power University
NPU Northwest Polytechnic University
Delft Delft University
APG Adverse pressure gradient
ZPG Zero pressure gradient
Appl. Sci. 2019, 9, 5495 21 of 22

References
1. Wang, X.; Ye, Z.; Kang, S.; Hu, H. Investigations on the Unsteady Aerodynamic Characteristics of a
Horizontal-Axis Wind Turbine during Dynamic Yaw Processes. Energies 2019, 12, 3124. [CrossRef]
2. Li, X.; Yang, K.; Wang, X. Experimental and Numerical Analysis of the Effect of Vortex Generator Height on
Vortex Characteristics and Airfoil Aerodynamic Performance. Energies 2019, 12, 959. [CrossRef]
3. Ma, L.; Wang, X.; Zhu, J.; Kang, S. Dynamic Stall of a Vertical-Axis Wind Turbine and Its Control Using
Plasma Actuation. Energies 2019, 12, 3738. [CrossRef]
4. Macquart, T.; Maheri, A.; Busawon, K. Microtab dynamic modelling for wind turbine blade load rejection.
Renew. Energy 2014, 64, 144–152.
5. Ebrahimi, A.; Movahhedi, M. Wind turbine power improvement utilizing passive flow control with microtab.
Energy 2018, 150, 575–582. [CrossRef]
6. Kametani, Y.; Fukagata, K.; Orlu, R.; Schlatter, P. Effect of uniform blowing/suction in a turbulent boundary
layer at moderate Reynolds number. Int. J. Heat Fluid Flow 2015, 55, 132–142. [CrossRef]
7. Ziadé, P.; Feero, M.A.; Sullivan, P.E. A numerical study on the influence of cavity shape on synthetic jet
performance. Int. J. Heat Fluid Flow 2018, 74, 187–197. [CrossRef]
8. Yang, H.; Jiang, L.-Y.; Hu, K.-X.; Peng, J. Numerical study of the surfactant-covered falling film flowing down
a flexible wall. Eur. J. Mech. Fluids 2018, 72, 422–431. [CrossRef]
9. Barlas, T.K.; van Kuik, G.A.M. Review of state of the art in smart rotor control research for wind turbines.
Prog. Aerosp. Sci. 2010, 46, 1–27. [CrossRef]
10. Johnson, S.J.; Baker, J.P.; Dam, C.P.V. An overview of active load control techniques for wind turbines with
an emphasis on microtabs. Wind Energy 2010, 13, 239–253. [CrossRef]
11. Lin, J.C. Review of research on low-profile vortex generators to control boundary-layer separation. Prog.
Aerosp. Sci. 2002, 38, 389–420. [CrossRef]
12. Wang, H.; Zhang, B.; Qiu, Q.; Xu, X. Flow control on the NREL S809 wind turbine airfoil using vortex
generators. Energy 2017, 118, 1210–1221. [CrossRef]
13. Taylor, H.D. The Elimination of Diffuser Separation by Vortex Generators; Report No. R-4012-3; United Aircraft
Corporation: Moscow, Russia, June 1947.
14. Khoshvaght-Aliabadi, M.; Sartipzadeh, O.; Alizadeh, A. An experimental study on vortex-generator insert
with different arrangements of delta-winglets. Energy 2015, 82, 629–639. [CrossRef]
15. Skullong, S.; Promvonge, P.; Thianpong, C.; Pimsarn, M. Thermal performance in solar air heater channel
with combined wavy-groove and perforated-delta wing vortex generators. Appl. Therm. Eng. 2016, 100,
611–620. [CrossRef]
16. Rao, D.; Kariya, T. Boundary-layer submerged vortex generators for separation control—An exploratory
study. In Proceedings of the 1st National Fluid Dynamics Conference, Cincinnati, OH, USA, 25–28 July 1998;
American Institute of Aeronautics and Astronautics: Washington, DC, USA, 1988; pp. 839–846.
17. Lin, J.C.; Robinson, S.K.; McGhee, R.J.; Valarezo, W.O. Separation control on high-lift airfoils via micro-vortex
generators. J. Aircr. 1994, 31, 1317–1323. [CrossRef]
18. Bragg, M.B.; Gregorek, G.M. Experimental study of airfoil performance with vortex generators. J. Aircr. 1987,
24, 305–309. [CrossRef]
19. Yanagihara, J.I.; Torii, K. Enhancement of laminar boundary layer heat transfer by a vortex generator. JSME
Int. J. 1992, 35, 400–405. [CrossRef]
20. Fiebig, M.; Valencia, A.; Mitra, N.K. Local heat transfer and flow losses in fin-and-tube heat exchangers with
vortex generators: A comparison of round and flat tubes. Exp. Therm. Fluid Sci. 1994, 8, 35–45. [CrossRef]
21. Reuss, R.L.; Hoffman, M.J.; Gregorek, G.M. Effects of surface roughness and vortex generators on the NACA
4415 airfoil. Vortex Augment. Turbines 1995, 12, 442–472.
22. Godard, G.; Stanislas, M. Control of a decelerating boundary layer. Part 1: Optimization of passive vortex
generators. Aerosp. Technol. 2006, 10, 181–191. [CrossRef]
23. Godard, G.; Stanislas, M. Control of a decelerating boundary layer. Part 2: Optimization of slotted jets vortex
generators. Aerosp. Sci. Technol. 2006, 10, 394–400. [CrossRef]
24. Godard, G.; Stanislas, M. Control of a decelerating boundary layer. Part 3: Optimization of round jets vortex
generators. Aerosp. Sci. Technol. 2006, 10, 455–464. [CrossRef]
Appl. Sci. 2019, 9, 5495 22 of 22

25. Yang, K.; Zhang, L.; Xu, J.Z. Simulation of aerodynamic performance affected by vortex generators on blunt
trailing-edge airfoils. Sci. China Ser. E Technol. Sci. 2010, 53, 1–7. [CrossRef]
26. Zhen, T.K.; Zubair, M.; Ahmad, K.A. Experimental and Numerical Investigation of the Effects of Passive
Vortex Generators on Aludra UAV Performance. Chin. J. Aeronaut. 2011, 24, 577–583. [CrossRef]
27. Lishu, H.; Zhide, Q. Experimental Research on Control Flow over Airfoil Based on Vortex Generator and
Gurney Flap. Acta Aeronaut. Astronaut. Sin. 2011, 32, 1429–1434.
28. Delnero, J.S.; Leo, J.M.D.; Camocardi, M.E. Experimental study of vortex generators effects on low Reynolds
number airfoils in turbulent flow. Int. J. Aerodyn. 2012, 2, 50–72. [CrossRef]
29. Velte, C.M.; Hansen, M.O.L. Investigation of flow behind vortex generators by stereo particle image
velocimetry on a thick airfoil near stall. Wind Energy 2013, 16, 775–785. [CrossRef]
30. Sørensen, N.N.; Zahle, F.; Bak, C. Prediction of the Effect of Vortex Generators on Airfoil Performance. J. Phys.
Conf. Ser. 2014, 524, 012019. [CrossRef]
31. Manolesos, M.; Voutsinas, S.G. Experimental investigation of the flow past passive vortex generators on an
airfoil experiencing three-dimensional separation. J. Wind Eng. Ind. Aerodyn. 2015, 142, 130–148. [CrossRef]
32. Gao, L.; Zhang, H.; Liu, Y.; Han, S. Effects of vortex generators on a blunt trailing-edge airfoil for wind
turbines. Renew. Energy 2015, 76, 303–311. [CrossRef]
33. Fouatih, O.M.; Medale, M.; Imine, O.; Imine, B. Design optimization of the aerodynamic passive flow control
on NACA 4415 airfoil using vortex generators. Eur. J. Mech. B/Fluids 2016, 56, 82–96. [CrossRef]
34. Zhang, L.; Li, X.; Yang, K.; Xue, D. Effects of vortex generators on aerodynamic performance of thick wind
turbine airfoils. J. Wind Eng. Ind. Aerodyn. 2016, 156, 84–92. [CrossRef]
35. Martínez-Filgueira, P.; Fernandez-Gamiz, U.; Zulueta, E.; Errasti, I.; Fernandez-Gauna, B. Parametric study
of low-profile vortex generators. Int. J. Hydrog. Energy 2017, 42, 17700–17712. [CrossRef]
36. Brüderlin, M.; Zimmer, M.; Hosters, N.; Behr, M. Numerical simulation of vortex generators on a winglet
control surface. Aerosp. Sci. Technol. 2017, 71, 651–660. [CrossRef]
37. Gageik, M.; Nies, J.; Klioutchnikov, I.; Olivier, H. Pressure wave damping in transonic airfoil flow by means
of micro vortex generators. Aerosp. Sci. Technol. 2018, 81, 65–77. [CrossRef]
38. Baldacchino, D.; Simao Ferreira, C.; De Tavernier, D.A.M. Experimental parameter study for passive vortex
generators on a 30% thick airfoil. Wind Energy 2018, 21, 118–132. [CrossRef]
39. Kundu, P.; Sarkar, A.; Nagarajan, V. Improvement of performance of S1210 hydrofoil with vortex generators
and modified trailing edge. Renew. Energy 2019, 142, 643–657. [CrossRef]
40. Mereu, R.; Passoni, S.; Inzoli, F. Scale-resolving CFD modeling of a thick wind turbine airfoil with application
of vortex generators: Validation and sensitivity analyses. Energy 2019, 187, 115969. [CrossRef]
41. Manolesos, M.; Papadakis, G.; Voutsinas, S.G. Revisiting the assumptions and implementation details of the
BAY model for vortex generator flows. Renew. Energy 2020, 146, 1249–1261. [CrossRef]
42. Li, X.; Yang, K.; Hu, H.; Wang, X.; Kang, S. Effect of Tailing-Edge Thickness on Aerodynamic Noise for Wind
Turbine Airfoil. Energies 2019, 12, 270. [CrossRef]
43. Smagorinsky, J. General Circulation Experiments with the Primitive Equations, Part 1, Basic Experiments.
Mon. Weather Rev. 1963, 91, 99–164. [CrossRef]

© 2019 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

You might also like