Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

29th AIAA Applied Aerodynamics Conference AIAA 2011-3354

27 - 30 June 2011, Honolulu, Hawaii

Steady Aerodynamics of Miniature Trailing-Edge


Devices in Transonic Flows
K. Richter∗ and H. Rosemann†
German Aerospace Center (DLR), Institute of Aerodynamics and Flow Technology,
Göttingen, Germany

The aerodynamic effect of “Miniature Trailing-Edge Devices” (MiniTEDs) on a super-


critical airfoil was investigated in transonic flow both experimentally and numerically. The
investigations were performed for the Gurney flap, the split flap and the divergent trailing
edge, and showed that the MiniTEDs operate by increasing the effective airfoil camber and
by spreading the pressure distribution on the rear part of the airfoil. MiniTEDs show a
strong influence on the supersonic flow regime, influencing the lift and drag characteristics.
At constant lift coefficient a redistribution of the lift generation from the front to the rear
part of the airfoil takes place. Total drag reductions can be achieved by the reduction of
wave drag. The maximum lift-to-drag ratio of the airfoil can be increased and the oper-
ational range can be extended to higher lift coefficients. The influences of the different
types of MiniTEDs are shown. The investigation of the influence of geometry parameters
showed that the variation of the height of the Gurney flap and of the deflection angle of the
split flap have similar effects. Perforation of the MiniTEDs showed only a small influence.
The investigation of the influence of the free stream parameters indicated that increases in
Mach number and Reynolds number increase the MiniTED effect in a similar manner and
cause a shift of the low-drag range of operation to higher lift coefficients. Moreover, it was
shown that a drag-optimal application is possible with adaptive MiniTEDs, and the same
performance is achieved by an adaptive Gurney flap and an adaptive split flap. A compar-
ison with a flexible trailing edge could identify performance advantages of the MiniTEDs,
indicating that an adaptive MiniTED could also be used as a control surface.

Nomenclature

c = Chord, m
cd = Drag coefficient
cl = Lift coefficient
cm = Pitching moment coefficient
cp = Pressure coefficient
cl /cd = Lift-to-drag ratio
h = Height, m
hGF = Gurney flap height, m
hM iniT ED = MiniTED height, m
htr = Height of transition tripping, m
l = Length, m
lF T E = Length of flexible trailing edge, m
lSF = Length of split flap, m
M = Mach number
p = Pressure, Pa
Re = Reynolds number
t = Thickness, m
∗ Research Scientist, Dept. Helicopters, AIAA Member.
† Research Scientist, Head Dept. High Speed Configurations.

1 of 19
Copyright © 2011 by Kai Richter, Henning Rosemann, German Aerospace Center. Published by the American Institute of Aeronautics and Astronautics, Inc., with permission.
American Institute of Aeronautics and Astronautics
x, y = Coordinates, m
xSF = Split flap hinge location, m
xtr = Transition tripping location, m
y+ = Normalized wall distance
Symbols
α = Angle of attack, deg
αT E = Flow angle at the trailing edge, deg
∆(cl /cd ) = Lift-to-drag ratio increment, ∆(cl /cd ) = (cl /cd )M iniT ED − (cl /cd )Baseline
∆cd = Drag increment, ∆cd = cd,M iniT ED − cd,Baseline
∆cl = Lift increment, ∆cl = cl,M iniT ED − cl,Baseline
∆cm = Pitching moment increment, ∆cm = cm,M iniT ED − cm,Baseline
δ = Deflection angle, deg
δF T E = Deflection angle of flexible trailing edge, deg
δSF = Split flap deflection angle, deg

I. Introduction
he operating costs of transport aircraft are largely affected by the aerodynamic performance of the wing.
T Due to rising fuel costs, it has become a requirement to exploit the full aerodynamic potential of the wing
for the duration of the entire cruise flight. Future wings could meet those requirements with a permanent
drag-optimal adaptation of the wing flow, with adaptive elements already taken into account in the wing
design. Possible adaptive elements are, for example, camber control measures which could be implemented
on the wing in the form of conventional trailing-edge flaps or “MiniTEDs”.
“Miniature Trailing-Edge Devices” (MiniTEDs) are a group of mechanical devices designed for flow
control at the trailing edge of airfoils and wings, and are very small compared to conventional control
surfaces. While a conventional trailing edge flap has a size of 10% to 30% of the local wing chord, MiniTEDs
are usually limited to sizes of 2% or less of chord. The MiniTEDs include several elements of different
geometries, of which the Gurney flap, the split flap and the divergent trailing edge are the best known, and
these are shown in Fig. 1.
The use of MiniTEDs for flow control is well known. The Gurney flap, named after the American race car
driver Dan Gurney, was first scientifically described by Liebeck1 in 1978 and a patent for a very similar mini
flap from Zaparka2 dates back to 1935. The principle of the split flap goes back to Gruschwitz and Schrenk3
in 1932 and the divergent trailing edge is based on a patent by Boyd4 from the year 1985. Experimental
and numerical investigations of the aerodynamic effects of these devices have been shown in great numbers,
but almost exclusively for subsonic flow speeds up to M ≈ 0.25. The investigations focused mainly on the
Gurney flap and give results for flow around two dimensional single5–7 and multi-element airfoils8–10 , rotor
blade airfoils11, 12 , turbine blades13 , three-dimensional wings14, 15 and delta wings.16
For cruise flight relevant transonic Mach numbers only a small number of results have been published.
Henne and Gregg17 , Köster18 and Thompson and Lotz19 reported on the steady design of transonic airfoils
with a divergent trailing edge and their experimental and numerical simulation. Bechert et al.20 reported
briefly on an experimental investigation on the CAST10-2 airfoil with a Gurney flap at M = 0.73. The
authors of this paper published early experimental results of the effect of Gurney flaps and divergent trailing
edges on the VC-Opt airfoil at M = 0.75521 and demonstrated the numerical simulation of a split flap on
a transport aircraft configuration for cruise flight.22 A comprehensive study of the aerodynamic effects of
MiniTEDs at transonic speeds does not appear in the literature.
Since the effect of MiniTEDs at cruise flight conditions is important for the realization of future adaptive
wings, this paper deals with extensive experimental and numerical investigations on the steady aerodynamic
effect of MiniTEDs on a supercritical airfoil at transonic speeds. The effects of MiniTEDs on the two-
dimensional airfoil flow and the aerodynamic coefficients will be shown for the Gurney flap, the split flap
and the divergent trailing edge. The influences of different geometric parameters such as the height of the
Gurney flap, the deflection angle of the split flap and perforations are discussed. An adaptive use of the
MiniTEDs will be presented for drag-optimal application and compared to the performance of a flexible
trailing edge. Further, the influence of flow parameters such as Mach number and Reynolds number will be
discussed.

2 of 19

American Institute of Aeronautics and Astronautics


Gurney Flap Split Flap Divergent Trailing Edge

δSF
hGF hDT E
lSF
xSF lDT E
Figure 1. Miniature Trailing-Edge Devices (MiniTEDs)

II. Experimental and Numerical Tools


Wind tunnel experiments and numerical simulations were performed to investigate the influence of Mini-
TEDs on the steady transonic flow around the two-dimensional airfoil VC-Opt.

II.A. Transonic Wind Tunnel Göttingen and VC-Opt airfoil model


Measurements were conducted with the two-dimensional airfoil model VC-Opt in the Transonic Wind Tunnel
Göttingen (DNW-TWG) at Mach numbers of M ∈ [0.755, 0.775, 0.790] and a chord-based Reynolds number
of Re = 5 × 106 . The DNW-TWG is a continuously operating closed-loop wind tunnel for exerperiments on
two- and three-dimensional wind tunnel models.23, 24 The operating range is 0.3 ≤ M ≤ 2.21 and Reynolds
numbers of up to Re1m = 18×106 can be achieved by total pressure variation. The DNW-TWG provides three
interchangeable test sections for different flow speed regimes, each with a cross-section of 1 m × 1 m. For the
present investigation, the adaptive wall test section with two-dimensional contour-adaptive top and bottom
walls was used, which is particularly suitable for the investigation of two-dimensional airfoil models. The
wall adaptation was performed based on the measured top and bottom wall pressure distributions following
the two-dimensional Cauchy-method by Amecke.25 Remaining interferences of the upper and lower walls
still present at the position of the wind tunnel model after the adaptation were taken into account in the
form of corrections to the freestream dynamic pressure and model angle of attack. For the measurements
presented, those were mostly in the order of ∆M = 0.005 and ∆α = −0.06◦ .
The airfoil VC-Opt (Variable Camber Optimized) is a turbulent supercritical airfoil that could be used
in the outboard wing section of modern transport aircraft. The airfoil was designed to reach optimum
performance at cruise flight conditions in the Mach number range 0.757 ≤ M ≤ 0.784 by the use of variable
camber with a trailing edge flap. The VC-Opt airfoil model used in this investigation had a chord of
c = 400 mm and a span of 1180 mm and was made from steel. The model was equipped with 154 pressure
taps with a diameter of 0.3 mm for steady pressure measurements in three sections on the upper side and
one section on the lower side. Figure 2 shows a schematic of the airfoil model and a photograph of the model
installed in the DNW-TWG adaptive wall test section. Transition was tripped at xtr /c = 10% on both
model sides using a zig-zag tape with a height of 100 µm. The effectiveness of the transition tripping for a
Reynolds number of Re = 5.0 × 106 was experimentally demonstrated by a Reynolds number variation.21
The model was positioned with an accuracy in the angle of attack of ∆α = 0.02◦ , the measurement of
the model incidence was carried out with an accuracy of ∆α = 0.01◦ . The aerodynamic coefficients for lift,
drag and pitching moment were computed based on steady pressure measurements on the model surface and
in the wake. A wake rake was used at a distance of 1.5 chord to the model trailing edge, containing 136 total
pressure taps and eight static pressure taps. Pressures were measured with an accuracy of ∆p = 52 Pa and
the repeatability of a test point with respect to lift and drag coefficients was ∆cl = 0.005 and ∆cd = 0.0002
for test points with attached flow.

II.B. DLR-TAU code and numerical grids


Two-dimensional steady simulations of the VC-Opt airfoil were conducted using the DLR-TAU code. The
DLR-TAU code is a software package for solving the Reynolds-averaged Navier-Stokes equations on hybrid
grids26 used at universities, research institutions and in industry. The software package includes several
modules to preprocess, run and postprocess a flow simulation. The solver is based on a three-dimensional

3 of 19

American Institute of Aeronautics and Astronautics


220. 120. 60.

Main Body Upper Side Cover

Trailing-Edge Part
Lower Side Cover

25. 200. all dimensions in [mm]


400.

Figure 2. Schematic of the VC-Opt airfoil model and the model installed in the test section

finite volume scheme integrating the RANS equations. In this work, inviscid fluxes were calculated by em-
ploying a central method with scalar dissipation and viscous fluxes were discretized using central differences.
The gradients of the flow variables were determined by a Green-Gauss formula and the discretization of the
temporal gradients used an explicit multi-step Runge-Kutta scheme. For convergence acceleration to steady
state, residual smoothing and a multigrid technique were employed. A grid adaptation module is also part
of the TAU software package enabling the adaptation of the grid depending on the flow solution. Both the
adaptation of height and distribution of the prismatic layers within the boundary layer (‘y + ’-adaptation)
and the refinement of the flow field at locations of high gradients were used.
The simulations were performed on hybrid grids generated with the commercial grid generation software
package CENTAURTM .27 The grids had a circular far field with a radius of 1000 chord and the airfoil contour
was resolved with approx. 400 points in the initial grid. For the wall normal resolution of the boundary layer
30 structured layers were used, the height of the first layer adjusted to a normalized distance from the wall of
y + ≈ 1 as a function of Mach number and Reynolds number. The simulations were carried out as a sequence
of several successive individual computations and grid adaptations applied in between. Each computation
was conducted to reach the convergence of the aerodynamic coefficients, with a maximum change of the drag
coefficient of ∆cd < 0.0001, which is equal to half the repeatability of the drag measurement in the DNW-
TWG wind tunnel experiment. Grid adaptations were carried out as ‘y + ’-adaptation and refinement. The
grid refinement was based on the local gradients of the flow field variables velocity, density, total pressure and
total enthalpy. The sequence of computations and grid adaptations was performed until the entire simulation
sequence fulfilled the convergence criteria and thus grid convergence was achieved. Depending on the model
configuration and the flow condition a maximum of seven adaptations were needed. The computational grids
typically reached about 150000 grid points.
For the selection of a suitable turbulence model, the impact of different models on the steady solution of
the airfoil was investigated with and without MiniTEDs. For attached airfoil flow, only a small dependence
of the flow solution was found for several one-equation or two-equation models implemented in the TAU
code. The main features of the airfoil flow and the MiniTED effect were predicted approximately in the
same way and a good convergence behavior could be observed even when configurations with high MiniTEDs
were computed. Taking into account past experience with the TAU code in the field of transonic airfoils,
the one-equation model of Spalart-Allmaras with modifications according to Edwards28 was used for all two-
dimensional steady TAU simulations in this work. To match the experiment, the simulations were performed
with laminar-turbulent transition at xtr /c = 10% on both airfoil sides.

II.C. Model configurations


Twenty-eight model configurations were investigated both experimentally and numerically. The configura-
tions are listed in Table 1 and described below.

Baseline
The baseline airfoil represents the clean VC-Opt wind tunnel model with fixed transition at xtr /c = 10% on
upper and lower sides.

4 of 19

American Institute of Aeronautics and Astronautics


Table 1. Configurations of the VC-Opt airfoil model

Model Configurations
No. Type Length Height Deflection Hinge Perforation Remark
l/c [%] h/c [%] δ [◦ ] x/c [%]
1 Baseline - - - - -
2-5 GF - 0.25, 0.50, 0.75, 1.00 - 100.0 -
6 - 11 SF 2.0 - 0, 7.5, 15, 22.5, 30, 37.5 98.5 -
12 SF 2.0 - 15 98.1 - flush with TE
13 SF 2.0 - 15 99.0 - a

14 SF 2.0 - 30 98.3 - flush with TE


15 SF 1.0 - 90 99.0 - a

16 SF 2.0 - 15 98.5 75 ver. slots


17 SF 2.0 - 15 98.5 11 hor. slots
18 SF 2.0 - 15 98.5 75 drill-holes
19 DTE 0.5 0.5 - - - a

20 DTE 1.0 1.0 - - - a

21 DTE 1.9 0.5 - - - No. 12 filled


22 DTE 1.7 1.0 - - - No. 14 filled
23 - 28 FTE 15.0 - 0, 1, 2, 3, 4, 5 85.0 -
M ∈ [0.755, 0.775, 0.790], Re = 5 × 106 , −3◦ ≤ α ≤ 5◦ , xtr /c = 10%, c = 400 mm

Gurney flap (GF)


Four Gurney flaps (GF) (config. no. 2 - 5) with heights 0.25% ≤ hGF /c ≤ 1.00% were investigated. The
Gurney flaps were made from folded rectangular brass plates with a thickness of t/c = 0.075%, shown in
Fig. 3 (left), and were flush mounted to the airfoil trailing edge on the lower side of the baseline airfoil. The
leg used for bonding had a constant length of l/c = 5% for all variants, whereas the bent leg was varied
according to the desired height of the Gurney flap. For the attachment an adhesive film with a thickness of
t/c = 0.0125% was used.

Split flap (SF)


Thirteen variants of the split flap (SF) (config. no. 6 - 18) were investigated. Split flaps were manufactured in
the same way as Gurney flaps from folded brass plates. The bent leg forming the split flap was manufactured
in different lengths and deflection angles. Three split flaps (config. no. 16 - 18) were perforated with vertical
slots, horizontal slots and drill-holes. These configurations were investigated with perforation in the model
center line symmetrical to the model center. Figure 3 (right) shows a perforated split flap with vertical slots.
a Results are not shown in this paper.

Figure 3. A Gurney flap (left) and a perforated split flap (right) used in the wind tunnel experiments

5 of 19

American Institute of Aeronautics and Astronautics


-1.4
-1.2 -0.4
DNW-TWG
-1.0 2D TAU
-0.8 -0.2
-0.6 Airfoil VC-Opt
0.0 GF, hGF/c = 0.50%
-0.4 6
M ≈ 0.755, Re = 5 x 10 DNW-TWG

cp
cp

-0.2 α ≈ 0°, xtr/c = 10% 2D TAU


0.0 0.2
0.2 Airfoil VC-Opt
0.4 GF, hGF/c = 0.50% 0.4
6
0.6 M ≈ 0.755, Re = 5 x 10
α ≈ 0°, xtr/c = 10% Airfoil shape
0.8 0.6
1.0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 0.75 0.80 0.85 0.90 0.95 1.00 1.05
x/c x/c
Figure 4. Measured and numerical pressure distributions of the VC-Opt airfoil with Gurney flap
(hGF /c = 0.50%) at Re = 5 × 106 , M ≈ 0.755 and α ≈ 0◦

Divergent trailing edge (DTE)


Four variants of divergent trailing edges (DTE) (config. no. 19 - 22) were examined. Divergent trailing edges
were made by filling out the split flaps that were flush mounted to the model trailing edge or by modelling
a ramp upstream of a Gurney flap.

Flexible trailing edge (FTE)


A flexible trailing edge with a length lF T E /c = 15% was designed, built and used on the airfoil model as a
comparison to the MiniTEDs. Unlike a conventional trailing edge flap, the flexible trailing edge provided a
continuous skin of glass fiber reinforced plastic on the model upper side, so that discontinuities of the slope
of the contour were avoided in the intersection between the main model part and the trailing edge flap. The
intersection was formed by a thin sheet of fiber glass that operates as a leaf spring element. The gap on the
model lower side was closed with a spring steel sheet. Because of the narrow space available, the flap driving
system could not be integrated into the model. Instead, four external drive units with streamlined fairings
were distributed equidistantly over the model span. Six configurations (config. no. 23 - 28) were investigated
with deflections 0◦ ≤ δF T E ≤ 5◦ .

III. Results and Discussion


The steady aerodynamics of MiniTEDs are well known for subsonic applications up to free stream Mach
numbers of M ≈ 0.25. The transonic airfoil flow is significantly different from these subsonic cases due to
the much higher flow velocities, especially due to the existence of local supersonic flow and the occurrence of
shocks. Therefore results obtained at low subsonic Mach numbers are not directly transferable to transonic
cruise speeds. The results presented in this paper therefore aim to provide an overview of the steady
aerodynamics of MiniTEDs at transonic speeds. Results are discussed with respect to the mechanisms
whereby the MiniTEDs influence the flow, the influence of the MiniTED geometry, the influence of the free
stream parameters and a comparison with a flexible trailing edge. The discussion is based on experimental
and numerical results as it could be demonstrated that the TAU code is able to predict the two-dimensional
airfoil flow with MiniTEDs accurately. Figure 4 shows the good agreement of experimental and numerical
pressure distributions for the airfoil with a Gurney flap.

III.A. MiniTED influence in transonic flow


The influence of MiniTEDs on the transonic airfoil flow has fundamental similarities to the results known
for low subsonic speeds but also important differences due to the presence of local supersonic flow. The
Gurney flap, split flap and divergent trailing edge cause the spreading of the pressure distribution on the
rear part of the airfoil and lead to an increase in effective airfoil camber due to the turning of the flow at
the trailing edge both in subsonic and transonic applications. This leads to an increase in lift and drag for
constant angle of attack, and to a decrease in the pitching moment. However, for transonic flow the presence

6 of 19

American Institute of Aeronautics and Astronautics


-1.4 -1.4
Airfoil VC-Opt Airfoil VC-Opt
-1.2 M ≈ 0.755, Re = 5 x 10
6 -1.2 M ≈ 0.755, Re = 5 x 10
6

-1.0 xtr/c = 10% -1.0 xtr/c = 10%


-0.8 -0.8
-0.6 -0.6
-0.4 -0.4
cp

cp
-0.2 -0.2
0.0 0.0
0.2 0.2
0.4 0.4
0.6 Baseline, α = -0.04°, cl = 0.286, cd = 0.0083 0.6 Baseline, α = 1.96°, cl = 0.650, cd = 0.0097

0.8 GF, hGF/c = 0.50%, α = -0.06°, cl = 0.539, cd = 0.0099 0.8 GF, hGF/c = 0.50%, α = 1.94°, cl = 0.954, cd = 0.0165

1.0 1.0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
x/c x/c
Figure 5. Measured pressure distributions of the VC-Opt airfoil with and without a Gurney flap
(hGF /c = 0.50%) at Re = 5 × 106 and M ≈ 0.755 for α ≈ 0◦ (left) and α ≈ 2◦ (right)

of local supersonic flow causes differences in the interaction with the airfoil flow and in the effect on the
airfoil performance.
The MiniTED influence on the airfoil pressure distribution is shown in Fig. 5 for the baseline airfoil and
the airfoil with a Gurney flap at M = 0.755 and Re = 5 × 106 for constant angles of attack of α ≈ 0◦ and
α ≈ 2◦ . The spreading of the pressure distribution occurs for both the flow conditions with and without
supersonic flow. The MiniTED lowers the local pressures in the rear part of the suction side and raises the
pressures on the pressure side, significantly increasing the lift generation in the last 30% of the airfoil and
thus decreasing the pitching moment compared to the baseline case. For both flow conditions an additional
increase in the height of the suction peak can be observed which is caused by the increase in the effective
airfoil camber due to the turning of the flow at the trailing edge, which causes a further increase in lift in
the front part of the airfoil. For flow conditions with local supersonic flow, α ≈ 2◦ , the development of
the supersonic field is affected in addition. Due to the reduction of the plateau pressure, the streamwise
extension of the supersonic regime is increased with a downstream shift of the shock. This leads to a further
increase in lift compared to the flow condition without supersonic flow, indicating that MiniTEDs do not
produce a constant lift increase independent of the angle of attack of the airfoil. The strength of the shock
also increases and therefore drag increases.
Comparing the MiniTED effect on the pressure distribution at constant lift coefficient, the results indicate
that the lift generation is shifted from the front part of the airfoil to the rear part. Figure 6 shows the
experimental pressure distributions at cl ≈ 0.36 and cl ≈ 0.77. For the airfoil with a Gurney flap the angle
of attack is reduced to reach constant lift, leading to a reduction of the height of the suction peak compared
to the baseline airfoil. For flow conditions with supersonic flow at cl ≈ 0.77, this leads to an upstream
movement of the shock and therefore to a reduction in the size of the supersonic flow field. The shock
strength reduces and drag decreases compared to the baseline case.

-1.4 -1.4
Airfoil VC-Opt Airfoil VC-Opt
-1.2 M ≈ 0.755, Re = 5 x 10
6 -1.2 M ≈ 0.755, Re = 5 x 10
6

-1.0 xtr/c = 10% -1.0 xtr/c = 10%


-0.8 -0.8
-0.6 -0.6
-0.4 -0.4
cp

cp

-0.2 -0.2
0.0 0.0
0.2 0.2
0.4 0.4
0.6 Baseline, α = 0.46°, cl = 0.372, cd = 0.0083 0.6 Baseline, α = 2.46°, cl = 0.766, cd = 0.0131

0.8 GF, hGF/c = 0.50%, α = -1.05°, cl = 0.354, cd = 0.0097 0.8 GF, hGF/c = 0.50%, α = 1.20°, cl = 0.786, cd = 0.0109

1.0 1.0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
x/c x/c
Figure 6. Measured pressure distributions of the VC-Opt airfoil with and without Gurney flap
(hGF /c = 0.50%) at Re = 5 × 106 and M ≈ 0.755 for ca ≈ 0.36 (left) and ca ≈ 0.77 (right)

7 of 19

American Institute of Aeronautics and Astronautics


Figure 7. Numerical pressure distributions in the trailing edge region of the VC-Opt airfoil with and
without MiniTEDs at Re = 5 × 106 , M = 0.755 and α = 0◦

The investigation of the flow field around the MiniTEDs showed that the spreading of the surface pressure
distribution is caused by a strong pressure field imposed by the MiniTED, shown in Fig. 7 with numerical
pressure distributions of the trailing edge region. For the baseline airfoil the pressures around the trailing
edge are of the same order of magnitude for upper and lower airfoil sides and no strong gradient is visible
at the trailing edge itself. For the airfoil with a Gurney flap a semi-circular field of high pressure develops
upstream of the MiniTED due to the blockage of the flow in front of the device, leading to the increase in
surface pressures on the lower airfoil side downstream of x/c ≈ 0.75. A significant pressure gradient occurs
at the MiniTED itself as the airfoil flow is accelerated strongly around the lower edge of the device. This
causes much lower pressures on the back side of the MiniTED as in the baseline case, with the low pressure
region extending over the airfoil upper surface causing the reduction of the local pressures on the upper side.
The MiniTED influence on the airfoil flow was found to be very similar for the three different types of
MiniTEDs investigated. However, the spreading generated by a split flap with the same geometric height
as a Gurney flap (hGK = lSF × sin(δSF )) is always lower than the one of the Gurney flap since both the
blockage ahead of the device and the acceleration around the lower edge of the split flap are lower. No
mentionable difference was found between the divergent trailing edge and the split flap. The presence of the
gap between the flap and the airfoil does not influence the steady flow around the device. The lift increase
due to the spreading of the pressure distribution occurs in the entire angle-of-attack range and lift range
investigated. Figure 8 (left) shows the pressure difference measured at x/c = 0.90 between the upper and
lower airfoil sides as a function of the lift coefficient for the airfoil with and without MiniTEDs. The curves
have an almost linear trend for attached flow with a significant offset for the MiniTEDs, indicating a lift
increase at this streamwise station by approx. 60% for the Gurney flap and 50% for a split flap (config. no.
12) both with a height of hM iniT ED /c = 0.50%.

0.8 5.0
Airfoil VC-Opt Airfoil VC-Opt
6 6
M ≈ 0.755, Re = 5 x 10 M ≈ 0.755, Re = 5 x 10
0.7 xtr/c = 10% 4.0 xtr/c = 10%

0.6
∆cp(x/c=0.90)

3.0
αTE, deg

0.5
2.0
0.4
1.0
0.3 Baseline
Baseline
GF, hGF/c = 0.50% GF, hGF/c = 0.50%
0.2 0.0
SF, lSF/c = 2.00%, xSF/c = 98.1%, δSF = 15° SF, lSF/c = 2.0%, xSF/c = 98.1%, δSF = 15°

0.1 -1.0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 -0.4 -0.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4
cl cl
Figure 8. Measured pressure differences at x/c = 0.90 between upper and lower sides (left) and
numerical flow angles at the trailing edge (right) for the VC-Opt airfoil with and without MiniTEDs
at Re = 5 × 106 and M ≈ 0.755

8 of 19

American Institute of Aeronautics and Astronautics


1.4
Airfoil VC-Opt
6
M ≈ 0.755, Re = 5 x 10
1.2 xtr/c = 10%

1.0

MiniTED only
0.8

cl
0.6
Baseline

airfoil only

airfoil + MiniTED
GF, hGF/c = 0.50%
0.4
SF, δSF = 15°,
lSF/c = 2.0%,
0.2 xSF/c = 98.1%

0.0
20 40 60 80 100 120 140 160 180
-4
cd , 10
Figure 9. Numerical drag polars and drag components for the VC-Opt airfoil with and without
a Gurney flap (hGF /c = 0.50%) or a split flap (lSF /c = 2.0%, xSF /c = 98.1%, δSF = 15◦ ) at
Re = 5 × 106 and M ≈ 0.755

The MiniTED causes a turning of the flow at the trailing edge leading to an increase in effective camber of
the airfoil. The numerical investigations revealed that the angle αT E of the flow downstream of the trailing
edge has a linear dependence on the airfoil angle of attack or lift coefficient. Figure 8 (right) shows αT E as
a function of the lift coefficient, with a positive angle defined by a downward turning of the flow starting
from the airfoil chord line. The flow angle decreases with increasing angle of attack or lift coefficient for
all configurations shown as the development of the boundary layers affects the flow off the airfoil. With
MiniTEDs the flow angle is significantly higher than for the baseline airfoil and the slope of the curves is
increased in addition. This leads to a substantial increase in effective airfoil camber compared to the baseline
case which further increases with the airfoil angle of attack or lift coefficient. As expected, the airfoil with a
split flap has always a flow angle αT E smaller than for a Gurney flap of the same height, and therefore the
change in effective camber is also smaller. The results obtained for the divergent trailing edge (not shown)
mirror the results of the split flap, again indicating that the split flap gap has no impact on the steady
MiniTED aerodynamics.
The change in effective camber directly affects the location of the stagnation point at the leading edge.
By the analysis of the stagnation point location for the different configurations, it was found that the change
in airfoil angle of attack necessary for the compensation of the lift enhancement provided by a MiniTED is
always higher than the difference in the angle of attack caused by the altered effective camber only. This
supports the assumption that in transonic flow the MiniTED effect is not just a camber-change effect but
that the spreading of the pressure distribution has a substantial additional impact.
The performance of transonic airfoils is affected by shocks to a large extent. Since the MiniTEDs also
have an influence on the shocks, the MiniTED influence on transonic airfoils is different to their influence
at low subsonic Mach numbers. In contrast to the application at low speeds, MiniTEDs can achieve drag
reductions in transonic flows. Figure 9 illustrates this behavior with the numerically predicted drag polars
of the airfoil with and without MiniTEDs for M = 0.755 and Re = 5 × 106 . The diagram shows the total
drag of the airfoil and the drag components for the ‘airfoil only’ and the ‘MiniTED only’. The total-drag
polars exhibit the known behavior that the polars are shifted to higher lift and higher drag when MiniTEDs
are used, compare Richter et al.21 For the cases presented, the baseline airfoil has the lowest total drag up
to a lift coefficient of cl ≈ 0.59 and is outperformed by the MiniTEDs at higher lift. The configuration with
a split flap shows less minimum drag than the configuration with a Gurney flap mainly due to the different
amounts of MiniTED component drag. Both MiniTEDs produce a significant amount of component drag
that has to be compensated when total drag reductions should be achieved. The component drag of the
split flap and Gurney flap exhibit weak linear relationships to the lift coefficient of the airfoil, and the
split flap component drag measures only approx. 65-80% of the Gurney flap component drag in the case
presented. The component drag is generally dominated by pressure drag and splits into approx. 90/10%

9 of 19

American Institute of Aeronautics and Astronautics


1.4 1.1
Airfoil VC-Opt 1.0
1.2 M ≈ 0.755, Re = 5 x 10
6
0.9
1.0 xtr/c = 10%
0.8
0.8 0.7
Baseline
0.6 Baseline 0.6 SF, δSF = 0°

cl
cl

SF, δSF = 0°
0.4 0.5 SF, δSF = 7.5°
SF, δSF = 7.5°
0.4 SF, δSF = 15°
0.2 SF, δSF = 15°
SF, δSF = 22.5°
0.3 SF, δSF = 22.5°
0.0 0.2 SF, δSF = 30°
SF, δSF = 30°
-0.2 SF, δSF = 37.5°
SF, δSF = 37.5° 0.1
-0.4 0.0
-4.0 -3.0 -2.0 -1.0 0.0 1.0 2.0 3.0 4.0 5.0 6.0 70 80 90 100 110 120 130 140 150 160 170 180
-4
α, deg cd , 10
Figure 10. Measured lift curves (left) and drag polars (right) of the VC-Opt airfoil with and without
split flaps (lSF /c = 2.0%, xSF /c = 98.5%, 0◦ ≤ δSF ≤ 37.5◦ ) at Re = 5 × 106 and M ≈ 0.755

for the front/back side of the split flap and into 80/20% for the front/back side of the Gurney flap. The
compensation of the component drag can be divided into two parts: the first part is a reduction in pressure
drag which is achieved mainly at the airfoil leading edge due the change of the pressure distribution when
MiniTEDs are used. However, this part is not large enough to achieve total drag reductions on its own
and is almost independent of the MiniTED type as indicated by the ‘airfoil only’ polars. The second part
is a reduction in wave drag which is achieved when shocks are weakened. The ability of the MiniTEDs to
reduce the airfoil total drag is therefore directly related to the weakening of shocks and thus only possible
for transonic flow conditions with shocks.

III.B. Influence of geometric parameters on the MiniTED effect


The influence of the geometric parameters of MiniTEDs is important for an adaptive application of the
devices. Only if the influence exhibits consistent trends with respect to the changes in the airfoil flow and
the aerodynamic performance is an adaptive use possible. These influences were investigated for the height
of the Gurney flap and the deflection angle of the split flap in the present work, and the effect of perforation
was studied for the split flap in addition. Finally, the adaptive drag-optimal applications of the Gurney flap
and the split flap are analyzed and compared.

III.B.1. Gurney flap height and split flap deflection angle


The influences of the Gurney flap height hGF and the split flap deflection angle δSF on the transonic airfoil
performance were found to be qualitatively similar. The lift increment provided by both devices increases
with their heights both in the linear lift range and at maximum lift. Figure 10 (left) shows the lift curves
of the baseline airfoil and the airfoil with split flaps of 0◦ ≤ δSF ≤ 37.5◦ at M = 0.755 and Re = 5 × 106
as an example. The curves are shifted to higher lift and to lower angles of attack as the angle of attack at
maximum lift is reduced. A lift increase can also be seen for the separated flow conditions beyond maximum
lift.
Both the spreading of the pressure distribution and the turning of the flow at the trailing edge are seen
to increase with the device height, as shown in Fig. 11 with ∆cp (x/c = 0.90) and αT E as functions of the
MiniTED height extracted for a constant angle of attack in the linear lift range of α ≈ 0◦ . For heights in the
range of 0.25% ≤ hM iniT ED ≤ 1.00%, both ∆cp (x/c = 0.90) and αT E are smaller for the split flap than for
the Gurney flap, confirming the findings previously discussed in section III.A. The split flap exhibits very
linear trends for the deflection angles investigated, whereas the Gurney flap shows a strong increase in ∆cp
and αT E for small device heights of up to hM iniT ED = 0.50% and a reduced increase for larger heights. Thus,
the differences in the effects of the two MiniTED types increase for small heights and diminish for larger
heights. Since both the spreading of the pressure distribution and the turning of the flow at the trailing
edge determine the lift enhancement provided by the MiniTED, very similar behavior of the lift increments
were found at α = 0◦ , as shown in Fig. 12 (left). Due to the direct relation between lift and pitching
moment and the airfoil pressure distribution, the pitching moment increments show the same qualitative

10 of 19

American Institute of Aeronautics and Astronautics


0.8 5.0
Airfoil VC-Opt Airfoil VC-Opt
6 6
M ≈ 0.755, Re = 5 x 10 4.5 M ≈ 0.755, Re = 5 x 10
0.7 α ≈ 0°, xtr/c = 10% α ≈ 0°, xtr/c = 10%
∆cp(x/c=0.90)

4.0

αTE , deg
0.6
GF
SF
3.5
GF
0.5 SF
3.0

0.4 SF: lSF/c = 2.0%, xSF/c = 98.5%


2.5 SF: lSF/c = 2.0%, xSF/c = 98.5%

0.3 2.0
0.00 0.25 0.50 0.75 1.00 1.25 0.00 0.25 0.50 0.75 1.00 1.25
hMiniTED/c, % hMiniTED/c, %
Figure 11. Measured pressure differences at x/c = 0.90 (left) and numerical flow angles at the trailing
edge (right) as a function of the MiniTED height for the VC-Opt airfoil with split flaps (lSF /c = 2.0%,
xSF /c = 98.5%) and Gurney flaps at Re = 5 × 106 and M ≈ 0.755

characteristics as the lift increments, Fig. 12 (right). The influence of the MiniTED height on maximum
lift however exhibits a different behavior. The increment in maximum lift ∆cl,max of the Gurney flap has
a saturation for heights larger than hM iniT ED = 0.75% whereas the split flap shows a linear relationship in
the entire range investigated and no saturation is visible.
The influence of the MiniTED height on the drag polars is shown in Fig. 10 (right) for the split flap. With
increasing deflection angle the polars are shifted to higher lift and also to higher drag coefficients, indicating
a continuous increase in the minimum drag of the airfoil with MiniTED. The same behavior was found for
the Gurney flap height by Richter et al.21 The beginning of drag reductions achieved by the MiniTEDs
is therefore also continuously delayed to higher lift for increasing device height and deflection angle. As
total drag reductions can only be achieved with the reduction of wave drag, as discussed in section III.A,
the MiniTEDs achieve improved performance compared to the baseline case only at lift coefficients beyond
the transonic drag rise. An adaptive split flap with variable deflection angle or an adaptive Gurney flap
with variable height could realize the envelope of all drag polars and thus achieve the enhancement of the
profitable range of application of the VC-Opt airfoil to higher lift coefficients.
The shift of the polars to higher drag is primarily caused by an increase in the component drag of the
MiniTEDs. Figure 13 (left) therefore illustrates this behavior by showing the component drag for split flap
and Gurney flap as a function of the MiniTED height. Both the component drag of the Gurney flap and
the component drag of the split flap are seen to increase with the device height. The drag of the Gurney
flap is always higher than that of the split flap. The Gurney flap exhibits a roughly linear behavior above
hM iniT ED = 0.25% whereas the component drag of the split flap rather follows a quadratic function. The

0.5 -0.12
∆cl(α = 0°), GF SF: lSF/c = 2.0%, Airfoil VC-Opt
6
∆cl(α = 0°), SF xSF/c = 98.5% -0.10 M ≈ 0.755, Re = 5 x 10
0.4 α ≈ 0°, xtr/c = 10%
∆cl,max, GF
∆cl,max, SF -0.08
0.3 Airfoil VC-Opt
M ≈ 0.755
∆cm
∆cl

Re = 5 x 10
6 -0.06
GF
0.2 xtr/c = 10%
SF
-0.04

0.1 -0.02 SF: lSF/c = 2.0%, xSF/c = 98.5%

0.0 0.00
0.00 0.25 0.50 0.75 1.00 1.25 0.00 0.25 0.50 0.75 1.00 1.25
hMiniTED/c, % hMiniTED/c, %

Figure 12. Measured lift increments (left) and pitching moment increments (right) as a function of the
MiniTED height for the VC-Opt airfoil with split flaps (lSF /c = 2.0%, xSF /c = 98.5%) and Gurney
flaps at Re = 5 × 106 and M ≈ 0.755

11 of 19

American Institute of Aeronautics and Astronautics


100 70
Airfoil VC-Opt SF: lSF/c = 2.0%, xSF/c = 98.5% Airfoil VC-Opt
M ≈ 0.755, Re = 5 x 10
6 60 M ≈ 0.755, Re = 5 x 10
6

80 α ≈ 0°, xtr/c = 10% α ≈ 0°, xtr/c = 10%


50
GF front side
40

-4
-4

50 GF back side

cd , 10
cd , 10

GF SF front side
SF
30 SF back side
40 20
10
20 SF: lSF/c = 2.0%,
0 xSF/c = 98.5%
0 -10
0.00 0.25 0.50 0.75 1.00 1.25 0.00 0.25 0.50 0.75 1.00 1.25
hMiniTED/c, % hMiniTED/c, %
Figure 13. Numerical MiniTED component drag (left) and surface drag (right) as a function of the
MiniTED height for the VC-Opt airfoil with split flaps (lSF /c = 2.0%, xSF /c = 98.5%) and Gurney
flaps at Re = 5 × 106 and M ≈ 0.755

ratio of the component drag of both MiniTEDs varies between approx. 70% at hM iniT ED = 0.25% and 85%
at hM iniT ED = 1.00%. At low device heights the component drag is primarily composed of the pressure
drag of the front side of the device, as shown in Fig. 13 (right). At larger device heights the pressure drag
generated on the back side increases and also becomes important. The reduced component drag of the split
flap compared to that of the Gurney flap results from less drag on both sides of the device, but primarily on
the front side. The drag difference on the front side was found to increase with the device height, whereas
the difference remains almost constant on the back side.
The results of the present investigation showed that both the Gurney flap and the split flap exhibit
consistent trends with respect to their influence on lift, drag and pitching moment in transonic flow, and are
therefore suited for an adaptive application on a transonic airfoil.

III.B.2. Perforation
Experimental and numerical investigations at low subsonic Mach numbers had shown that a perforation
of the MiniTED can have a positive influence on its drag characteristics and can lead to drag reductions
compared to the nonpreforated device.29 This influence was also studied for MiniTEDs in transonic flow by
the perforation of a split flap. The reference was chosen as a split flap with a deflection angle of δSF = 15◦
(config. no. 8) as this device achieved the best maximum lift-to-drag ratio of all split flaps investigated at
M = 0.755. Three types of perforations (config. no. 12 - 15) were applied with horizontal slots, vertical slots,
and drill holes, as schematically shown in Fig. 14. The perforated area of the configurations was almost
constant at 9.2%, 9.4% and 11.8%, respectively.
In contrast to the findings at low Mach numbers, the influence of the perforation on the airfoil perfor-
mance at transonic speeds was found to be very small and visible primarily in the lift and pitching moment
characteristics of the VC-Opt airfoil. The perforated split flaps reached approx. 85% of the lift increment
of the nonperforated flap, with the type of perforation not being relevant. The influence on the lift-to-drag
ratio behavior is seen to be negligible for low and moderate lift coefficients. Figure 15 shows the lift-to-drag
ratio increments and the pitching moment increments as a function of the lift coefficient for M ≈ 0.755 and

75 vertical slots 11 horizontal slots 75 drill-holes

13.3
66.7
1.5

1.5
4
6.7

13.3
4

13.3 ø4

all dimensions in mm

Figure 14. Schematics of the perforated split flaps

12 of 19

American Institute of Aeronautics and Astronautics


30 -0.025
Airfoil VC-Opt SF
25 M ≈ 0.755, Re = 5 x 10
6
-0.030 SF, hor. slots
xtr/c = 10% SF, ver. slots
20
-0.035 SF, drill-holes
SF
15 SF, hor. slots
∆(cl/cd)

-0.040

∆cm
SF, vert. slots
10
SF, drill-holes
cl interpolated -0.045
5 cl interpolated

SF: lSF/c = 2.0%,


-0.050
0 Airfoil VC-Opt SF: lSF/c = 2.0%,
xSF/c = 98.5%, 6
xSF/c = 98.5%,
-0.055 M ≈ 0.755, Re = 5 x 10
-5 δSF = 15° xtr/c = 10% δSF = 15°
-10 -0.060
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2
cl cl
Figure 15. Measured lift-to-drag ratio increments (left) and pitching moment increments (right) for
the VC-Opt airfoil with split flaps (lSF /c = 2.0%, δSF = 15◦ , xSF /c = 98.5%) with and without
perforation at Re = 5 × 106 and M ≈ 0.755

Re = 5×106 . Only at lift coefficients higher than cl = 0.75, can a difference be noted with a reduced ∆(cl /cd )
for the perforated configurations compared to the original split flap. The influence on the pitching moment
increment is seen to be larger and is similar over the entire lift range investigated. The perforations here
lead to a reduction of the pitching moment change caused by the split flaps compared to the nonperforated
device. For each type of perforation an almost constant reduction can be seen, with the vertical slots and
the drill-holes both reaching the highest reductions at approx. 88% of the ∆cm of the original flap. These
changes are caused by a small reduction of the spreading of the pressure distributions which affect both lift
and moment slightly.

III.B.3. Adaptive use for drag-optimal application


Both the Gurney flap and the split flap were found to be suitable for application as an adaptive flow
control device. For the drag-optimal application, the Gurney flap height and the split flap deflection have
to be adapted to the prevailing flow condition in order to always achieve minimum drag. These optimum
characteristics can be expressed by the envelopes of the individual drag polars previously shown in Fig. 10
(right). Figure 16 shows the experimental drag polar envelopes for both types of MiniTEDs at M ≈ 0.755
and Re = 5×106 . For the purpose of a quantitative drag comparison of the two types of MiniTEDs measured

1.1 37.5°
30° (1.22%)
1.0 22.5° (1.00%)
(0.77%) 1.00%
0.9 15°
(0.52%) 0.75%
0.8 7.5°
(0.26%) 0.50%
0.7 0°
(0.07%) 0.25%
0.6
cl

GF, variable height


lit flap

0.5
flap

SF, variable deflection angle


rney
ut sp

0.4
ut Gu
witho

0.3
witho

SF: lSF/c = 2.0%, xSF/c = 98.5% Airfoil VC-Opt


6
0.2 M ≈ 0.755, Re = 5 x 10
GF: ∆cl = +0.02, ∆cd = -0.0002 xtr/c = 10%
0.1
0.0
70 80 90 100 110 120 130 140 150 160 170 180
-4
cd , 10
Figure 16. Envelopes of the measured drag polars of the VC-Opt airfoil with Gurney flaps and split
flaps at Re = 5 × 106 and M ≈ 0.755

13 of 19

American Institute of Aeronautics and Astronautics


-1.4 -1.4
Airfoil VC-Opt Airfoil VC-Opt
-1.2 SF: lSF/c = 2.0%, xSF/c = 98.5%
M ≈ 0.755, Re = 5 x 10
6 -1.2 SK: lSF/c = 2.0%, xSF/c = 98.5%
M ≈ 0.755, Re = 5 x 10
6

-1.0 xtr/c = 10% -1.0 xtr/c = 10%


-0.8 -0.8
-0.6 -0.6
-0.4 -0.4
cp

cp
-0.2 -0.2
0.0 0.0
0.2 0.2
0.4 0.4
0.6 GF, hGF/c = 0.50%, α = 1.20°, cl,corr = 0.806, cd,corr = 0.0107 0.6 GF, hGF/c = 0.50%, α = 1.45°, cl,corr = 0.861, cd,corr = 0.0116
SF, δSF = 15°, α = 1.46°, cl = 0.817, cd = 0.0105 SF, δSF = 22.5°, α = 1.20°, cl = 0.850, cd = 0.0115
0.8 0.8
1.0 1.0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
x/c x/c
Figure 17. Experimental pressure distributions of the VC-Opt airfoil with adaptive MiniTEDs at
Re = 5 × 106 and M ≈ 0.755 for cl ≈ 0.81 (left) and cl ≈ 0.86 (right)

in different wind tunnel test campaigns, the results of the Gurney flap were corrected by ∆cl = −0.02 and
∆cd = +0.0002 in order to reach the same performance of the baseline airfoil measurements for both tests.
The results show that both MiniTEDs have the same drag-optimal performance over the entire lift range
investigated. This finding could be confirmed by two-dimensional TAU simulations. The use of an adaptive
MiniTED starts at a lift coefficient of cl ≈ 0.61 and seemingly shows a performance gain for the split
flap in the range of 0.57 ≤ cl ≤ 0.72 which is caused by the absence of data for a Gurney flap lower than
hGF = 0.025%. At higher lift coefficients both MiniTEDs have identical performance within the measurement
accuracy. In Fig. 16 the MiniTED heights and deflections angles used for the drag-optimal application are
shown in addition. For the same increment in MiniTED height the application range of a constant split flap
deflection in terms of lift is always smaller than the range of constant Gurney flap height. This behavior is
caused as the split flap provides a smaller lift increment for the same change in MiniTED height than the
Gurney flap, as previously shown in Fig. 12 (left). This behavior accumulates at higher lift coefficients and
significantly larger split flap deflections are needed at high lift.
The identical drag characteristics of the airfoil with adaptive Gurney flap and adaptive split flap indicates
that due to the differences in the component drag an unequal reduction in wave drag is achieved in order to
reach the same total drag. Thus, different airfoil flows have to exist for both configurations despite the fact
that have the same aerodynamic coefficients. Figure 17 shows the airfoil pressure distributions for constant
lift coefficients of cl ≈ 0.81 and cl ≈ 0.86. At cl ≈ 0.81 the distributions show the expected behavior.
Drag-optimal performance is here reached by a split flap deflection of δSF = 15◦ and a Gurney flap height
of hGF = 0.50%. Since the lift increment provided by the split flap is smaller than that of the Gurney
flap, the split flap configuration reaches the same constant lift at an higher angle of attack. Therefore,
the flow around the airfoil with a split flap exhibits a larger supersonic flow field with a stronger shock.
The higher wave drag of the split flap configuration is then compensated by a lower MiniTED component
drag and constant total drag is reached. At cl ≈ 0.86 the distributions surprisingly show almost identical
pressure distributions although the devices have different amounts of component drag with δSF = 22.5◦ and
hGF = 0.50%, see Fig. 13 (left), and therefore different amounts of wave drag reductions would be needed
for compensation. The cause of the occurrence of both varying and identical pressure distributions within
the drag polar envelope could not be clarified within this work.
The influence of the drag-optimal application of the adaptive MiniTEDs on the pitching moment charac-
teristics of the airfoil is shown in Fig. 18. As the application of the adaptive devices starts at cl ≈ 0.61, the
pitching moment curves reflect the behavior of the baseline airfoil at lower lift coefficients. At higher lift a
linear decrease in the pitching moment can be observed with increasing lift coefficient. Due to the relatively
small number of configurations measured, the data shows a large scatter that does not allow a more detailed
investigation. However, within this scatter the drag-optimal use of an adaptive Gurney flap and an adaptive
split flap reaches similar pitching moment behaviors.

14 of 19

American Institute of Aeronautics and Astronautics


-0.04
-0.06 Airfoil VC-Opt SF: lSF/c = 2.0%, xSF/c = 98.5%
6
M ≈ 0.755, Re = 5 x 10
-0.08 xtr/c = 10% GF: ∆cl = +0.02, ∆cm = -0.001
-0.10
-0.12

cm
-0.14
-0.16
-0.18 GF, variable height
-0.20 SF, variable deflection angle

-0.22
-0.24
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3
cl
Figure 18. Measured pitching moment curves for the drag polar envelopes of the VC-Opt airfoil with
Gurney flaps and split flaps at Re = 5 × 106 and M ≈ 0.755

III.C. Influence of free stream parameters on MiniTED effect


The results in the previous sections were reported for a single transonic Mach number and Reynolds number.
Although most of the findings should be universal for the steady aerodynamics of MiniTEDs in transonic
flows, the influence of the free stream parameters Mach number and Reynolds number on the MiniTED
effect are important and were also studied in this work.

III.C.1. Mach number


The influence of the Mach number on the MiniTED effect was experimentally investigated for the Gurney
flap with various heights for transonic Mach numbers of 0.755 ≤ M ≤ 0.790 and a constant angle of attack of
α ≈ 0◦ . The MiniTED influence was found to increase with the Mach number. The lift and pitching moment
increments provided by the MiniTED increase with Mach number, and the pitching moment increment
decreases consistently for each Gurney flap height. The behavior of the lift increment is seen to depend on
the MiniTED height as a continuous increase in ∆cl is achieved only for small heights of up to hGK = 0.50%.
For these cases the influence of the Gurney flap on the airfoil flow occurs in the same manner with an increased
spreading of the pressure distribution and an increased effective airfoil camber both leading to an extension
of the supersonic flow field. For larger Gurney flaps however, the influence of the MiniTED is limited as the
the extension of the supersonic flow field reaches its limits and shock induced separation occurs.
The influence of the Mach number on the drag characteristics is larger than observed for lift and pitching
moment. Figure 19 (left) shows the drag increment as a function of the Gurney flap height for the Mach
numbers investigated for α ≈ 0◦ . It can be seen that the increase in Mach number clearly causes also an
increase in drag, with an additional progression for the Gurney flap height. Both the increase in MiniTED

80 0.80
Airfoil VC-Opt Airfoil VC-Opt
6 cl interpolated 6
Re = 5 x 10 , cl = 0.50 Re = 5 x 10 , α = 0°
xtr/c = 10% 0.75 xtr/c = 10%
60
M = 0.755
-4

cl(∆cd=0)

0.70
∆cd , 10

M = 0.775
M = 0.790
40
0.65

20 M = 0.755
0.60 M = 0.775
M = 0.790

0 0.55
0.00 0.25 0.50 0.75 1.00 0.00 0.25 0.50 0.75 1.00
hGF/c, % hGF/c, %
Figure 19. Measured drag increments (left) and lift coefficient at ∆cd = 0 (right) as a function of the
Gurney flap height for the VC-Opt airfoil at Re = 5 × 106 , M ∈ [0.755, 0.775, 0.790] and α ≈ 0◦

15 of 19

American Institute of Aeronautics and Astronautics


0.04 14

-4
GF, hGF/c = 0.50%, ∆cd(Re=5e6) = 0.0021
Airfoil VC-Opt

∆cd - ∆cd(Re = 5 x 10 ), 10
M = 0.755, α = 0°
12 GF, hGF/c = 1.00%, ∆cd(Re=5e6) = 0.0055
∆cl - ∆cl(Re = 5 x 10 )
6

0.03 xtr/c = 10%

6
10 Airfoil VC-Opt
M = 0.755, α = 0°
8 xtr/c = 10%
0.02
6

0.01
4
GF, hGF/c = 0.50%, ∆cl(Re=5e6) = 0.300
GF, hGF/c = 1.00%, ∆cl(Re=5e6) = 0.477 2

0.00 0
5 10 15 20 25 30 35 40 5 10 15 20 25 30 35 40
6 6
Re, 10 Re, 10
Figure 20. Numerical lift increments (left) and drag increments (right) as a function of the Reynolds
number for the VC-Opt airfoil with and without Gurney flaps (hGF /c ∈ [0.50%, 1.00%]) at M = 0.755
and α = 0◦

height and the increase in Mach number individually lead to an extension of the supersonic flow field and to
a strengthening of shocks. For high devices the Mach-number-induced drag-increase is therefore larger than
for low devices. This behavior directly impacts the drag reduction abilities of the airfoil with MiniTEDs.
Figure 19 (right) reflects this effect with a plot of the lift coefficient at beginning drag reductions of the
airfoil with Gurney flap compared to the baseline airfoil as a function of the Gurney flap height and the
Mach number. The efficient use of a Gurney flap is shifted to higher lift coefficients both with increasing
MiniTED height and with increasing Mach number.

III.C.2. Reynolds number


The real application of MiniTEDs on a transport aircraft wing will be conducted at very large Reynolds
numbers of up to Re = 40 × 106 . This is a significant difference to the wind tunnel Reynolds numbers that
are normally used for the basic investigation of flow control devices, as in this work, and the influence of the
Reynolds number on the MiniTED effect is important to estimate. Therefore numerical investigations were
conducted for the airfoil with and without Gurney flaps at M = 0.755 and α = 0◦ , and Re ∈ [5, 10, 20, 40] ×
106 . The results show that both the lift and the drag characteristics of the MiniTEDs are influenced by
the Reynolds number but the effect on the drag is much lager than on the lift characteristics. Figure 20
summarizes the results showing the lift and drag increments caused by a Reynolds number change starting
from Re = 5×106 . The lift increments generated with increasing Reynolds number are seen to increase similar
to a square-root function. The absolute values of ∆cl reached at Re = 40×106 seem to be disproportionately
high for a Gurney flap with a height of hGF = 0.50% compared to hGF = 1.00%. However, based on the
lift increment occurring at Re = 5 × 106 both MiniTEDs experience a constant increase of approx. 7%. No
disproportional influence was observed for high MiniTEDs although they emerge further from the airfoil
boundary layer than small MiniTEDs.
The Reynolds number change shows a larger influence on the drag increment. The drag increments
also increase with Reynolds number and the curves also approximate a square-root function. However, the
Reynolds number sensitivity of the airfoil with high MiniTEDs is larger than with low MiniTEDs. Based
on the drag increment occurring at Re = 5 × 106 , the MiniTEDs experience a maximum increase in ∆cd of
approx. 14% and 18% for hGF = 0.50% and hGF = 1.00% at Re = 40 × 106 . A detailed analysis of the drag
contributions to the total drag increment showed that the drag increase is largely affected by an increase of
the MiniTED component drag, and the increase in the component drag was found to be directly proportional
to the height of the Gurney flap. The Reynolds number influence on the MiniTED component drag is thus
independent of the MiniTED height. The drag increase caused for Re = 40 × 106 is mainly composed of
the MiniTED component drag for the small Gurney flap configuration, whereas for the high Gurney flap an
additional increase of wave drag due to the strengthening of shocks occurs, indicating an increased Reynolds
number sensitivity for high MiniTEDs.

16 of 19

American Institute of Aeronautics and Astronautics


1.1
37.5°
1.0 30°
22.5°
0.9 5°
15°

0.8 7.5° 3°
0.7 δSF = 0°


0.6

cl

= 0°
FTE

iTED
0.5 MiniTED

FTE
ut Min
0.4

ge δ
witho

ailing ed
0.3 FTE: lFTE/c = 15.0%
SF: lSF/c = 2.0%, xSF/c = 98.5% Airfoil VC-Opt
6
0.2 M ≈ 0.755, Re = 5 x 10

flexible tr
GF: ∆cl = +0.02, ∆cd = -0.0002 xtr/c = 10%
0.1 FTE: ∆cl = +0.01, ∆cd = +0.0002

0.0
0.007 0.008 0.009 0.010 0.011 0.012 0.013 0.014 0.015 0.016 0.017 0.018
cd
Figure 21. Envelopes of the measured drag polars for the VC-Opt airfoil with flexible trailing edge
(0◦ ≤ δF T E ≤ 5◦ ) and with MiniTEDs at Re = 5 × 106 and M ≈ 0.755

III.D. Comparison of MiniTEDs and the flexible trailing edge


In the present study a flexible trailing edge (FTE) was developed and experimentally investigated for com-
parisons with the MiniTEDs. The aim of this investigation was to analyze the drag-optimal application of an
adaptive FTE and to compare it to the drag-optimal application of an adaptive MiniTED. The mechanisms
whereby the FTE influences the flow are basically the same as of a conventional trailing edge flap since the
FTE changes the geometric airfoil camber only and does not cause a spreading of the pressure distribution
like the MiniTEDs do. The FTE was investigated with deflections of δF T E ∈ [0◦ , 1◦ , 2◦ , 3◦ , 4◦ , 5◦ ]. Figure 21
shows the envelope of the drag polars measured for the various FTE configurations, representing the polar
of an adaptive flexible trailing edge at M ≈ 0.755 and Re = 5 × 106 . For the purpose of a quantitative drag
comparison with an adaptive MiniTED measured in a different wind tunnel test campaign, the results of
the FTE were corrected by ∆cl = +0.01 and ∆cd = +0.0002 in order to reach the same performance of the
baseline airfoil measurements for both tests.
The comparison with the drag polar envelope of the adaptive split flap clearly indicates that the adaptive
MiniTED achieves higher performance than the adaptive FTE in transonic flow. At lift coefficients higher
than cl = 0.56 the adaptive MiniTED reaches lower drag by 0.0002 ≤ ∆cd ≤ 0.0007. In this lift range, the
airfoil with flexible trailing edge always has higher angles of attack than the airfoil with MiniTED, indicating
that the lift enhancement provided by the MiniTED is larger than by the FTE. This supports the finding
that the spreading of the pressure distribution is an important part of the MiniTED lift enhancement also
at transonic conditions. As a consequence, the pressure distributions of the FTE configuration show higher
suction peaks, larger supersonic flow fields and stronger shocks than the airfoil with split flap. The better
performance of the adaptive MiniTED can also be seen in the lift-to-drag ratio, shown in Fig. 22 (left).
Corresponding to the drag polars, the lift-to-drag ratio curves indicate an increase in cl /cd above cl = 0.56.
The maximum lift-to-drag ratio can be enhanced by approx. 6% compared to the maximum reached by
the adaptive flexible trailing edge, and the drag-optimal range of application can be increased to higher lift
coefficients.
The pitching moment of the airfoil with an adaptive MiniTED is lower than for the adaptive FTE, shown
in Fig. 22 (right). Both configurations exhibit approximately linear relationships between pitching moment
and lift, with the flexible trailing edge showing a higher slope, leading to an increasing pitching moment
difference between MiniTED and FTE at higher lift coefficients.
For the transonic conditions investigated, the adaptive MiniTED was found to be superior to the flexible
trailing edge. This result could open up the possibility for the application of an adaptive MiniTED as a
control surface on wing and empennage of transport aircraft but would need to be rechecked for the Reynolds
number influence on the performance assessment.

17 of 19

American Institute of Aeronautics and Astronautics


80 -0.04
Airfoil VC-Opt -0.06 Airfoil VC-Opt
70 6 FTE
M ≈ 0.755 M ≈ 0.755, Re = 5 x 10 MiniTED
Re = 5 x 10
6 -0.08 xtr/c = 10%
60
xtr/c = 10% -0.10
50 -0.12
cl/cd

FTE

cm
40 MiniTED -0.14
30 -0.16
FTE: lFTE/c = 15.0% -0.18 FTE: lFTE/c = 15.0%
20 SF: lSF/c = 2.0%, xSF/c = 98.5% SF: lSF/c = 2.0%, xSF/c = 98.5%
-0.20
10 FTE: ∆cl = +0.01, ∆cd = +0.0002 -0.22 FTE: ∆cl = +0.01, ∆cm = -0.003

0 -0.24
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3
cl cl
Figure 22. Measured lift-to-drag ratio curves (left) and pitching moment curves (right) for the drag
polar envelopes of the VC-Opt airfoil with MiniTEDs and with flexible trailing edge at Re = 5 × 106
and M ≈ 0.755

IV. Conclusion
The steady aerodynamics of “Miniature Trailing-Edge Devices” (MiniTEDs) were investigated on a su-
percritical airfoil in transonic flow both experimentally and numerically. The investigations were performed
for the Gurney flap, the split flap and the divergent trailing edge, using experiments in the DNW-TWG and
numerical DLR-TAU simulations.
The investigation of the effects of MiniTEDs on the steady transonic flow around the VC-Opt airfoil
has identified analogies to subsonic flow velocities and showed important differences due to the existence
of supersonic flow. Experimental and numerical results are in good agreement that all types of MiniTEDs
investigated operate by increasing the effective airfoil camber and by spreading the pressure distribution on
the rear part of the airfoil. At constant angle of attack, lift and drag are thereby increased, and pitching
moment decreases. In transonic flow, MiniTEDs show a strong influence on the supersonic flow regime,
influencing the lift and drag characteristics in particular. At constant lift coefficient a redistribution of the
lift generation from the front to the rear part of the airfoil takes place. At medium and high lift coefficients,
drag reductions can be achieved leading to an increase of the lift-to-drag ratio compared to the baseline
airfoil. The reduction of the total drag for transonic flow is dominated by the reduction of wave drag
compensating the MiniTED component drag.
The influences of the different types of MiniTEDs on the airfoil flow and the aerodynamic coefficients
were shown in detail on the basis of experimental and numerical results. The investigation of the influence
of geometric parameters showed that the variation of the height of the Gurney flap and of the deflection
angle of the split flap have similar effects, but the influence of the split flap is smaller at the same effective
MiniTED height. The divergent trailing edge was found to have identical aerodynamic characteristics as the
split flap. The maximum lift-to-drag ratio of the airfoil can be increased using small MiniTED heights or
deflection angles, and the operational range of the VC-Opt airfoil can be significantly extended to higher
lift coefficients. The perforation of the MiniTEDs showed only a small influence. The investigation of the
influence of the free stream parameters showed that increases in Mach number and Reynolds number increase
the MiniTED effect in a similar manner and cause a shift of the low-drag range of operation to higher lift
coefficients. Moreover, it was shown that both the Gurney flap and the split flap are suited for a usage as
adaptive flow control devices. The drag-optimal application is possible with the adaptive MiniTEDs and
the same effectiveness is achieved by an adaptive Gurney flap and an adaptive split flap. A comparison
with a flexible trailing edge could identify performance advantages of the MiniTEDs in the form of higher
lift-to-drag ratios, indicating that an adaptive MiniTED could also be used as a control surface.

References
1 Liebeck,
R. H., “Design of Subsonic Airfoils for High Lift,” Journal of Aircraft, Vol. 15, No. 9, 1978, pp. 547–561.
2 Zaparka,
E. F., “Aircraft and Control thereof,” US Patent Re19412, 1 January 1935, Zap Development Company, New
York (NY), 1935.

18 of 19

American Institute of Aeronautics and Astronautics


3 Gruschwitz, E. and Schrenk, O., “Über eine einfache Möglichkeit zur Auftriebserhöhung von Tragflügeln,” Zeitschrift für

Flugtechnik und Motorluftschiffahrt, Vol. 23, No. 20, 1932, pp. 597–601.
4 Boyd, J. A., “Trailing Edge Device for an Airfoil,” US Patent 4542868, 24 September 1985, Lockheed Corporation,

Burbank (CA), 1985.


5 Jeffrey, D., An Investigation Into the Aerodynamics of Gurney Flaps, Ph.D. thesis, University of Southampton, 2001.
6 Brown, J. and Filipone, A., “Aerofoil at low speeds with Gurney flaps,” The Aeronautical Journal, Vol. 107, No. 1075,

2003, pp. 539–546.


7 Singh, M. K., Dhanalakshmi, K., and Chakrabartty, S. K., “Navier-Stokes Analysis of Airfoils with Gurney Flaps,”

Journal of Aircraft, Vol. 44, No. 5, 2007, pp. 1487–1493.


8 Jeffrey, D., Zhang, X., and Hurst, D. W., “Some Aspects of the Aerodynamics of Gurney Flaps on a Double-Element

Wing,” Journal of Fluids Engineering, Vol. 123, 2001, pp. 99–104.


9 Ross, J. C., Storms, B. L., and Carrannanto, P. G., “Lift-Enhancing Tabs on Multielement Airfoils,” Journal of Aircraft,

Vol. 32, No. 3, 1995, pp. 649–655.


10 Angland, D., Zhang, X., Molin, N., and Chow, L. C., “Measurements of Flow around a Split Flap Configuration,”

AIAA-2007-0225, AIAA 45th Aerospace Sciences Meeting and Exhibit, Reno (NV), 8–11 January 2007.
11 Chandrasekhara, M. S., Martin, P. B., and Tung, C., “Compressible Dynamic Stall Performance of a Variable Droop

Leading Edge Airfoil with a Gurney Flap,” AIAA-2004-0041, AIAA 42nd Aerospace Sciences Meeting and Exhibit, Reno (NV),
5–8 January 2004.
12 Kinzel, M. P., Lesieutre, G. L., and Maughmer, M. D., “Miniature Trailing-Edge Effectors for Rotorcraft Performance

Enhancement,” Journal of the American Helicopter Society, Vol. 52, No. 2, 2007, pp. 146–158.
13 Byerley, A. R., Störmer, O., Baughn, J. W., Simon, T. W., Van Treuren, K. W., and List, J., “Using Gurney Flaps to

Control Laminar Separation on Linear Cascade Blades,” Journal of Turbomachinery, Vol. 125, 2003, pp. 114–120.
14 Donovan, L. M., Storms, B. L., and Ross, J. C., “Lift-Enhancing Tabs on Swept, Three-Dimensional High-Lift Systems,”

Journal of Aircraft, Vol. 35, No. 3, 1998, pp. 494–495.


15 Wang, J.-J., Zhan, J.-X., Zhang, W., and Wu, Z., “Application of a Gurney Flap on a Simplified Forward-Swept Aircraft

Model,” Journal of Aircraft, Vol. 43, No. 5, 2006, pp. 1561–1564.


16 Li, Y. C., Wang, J. J., and Zhang, P. F., “Effects of Gurney flaps on the lift enhancement of a cropped nonslender delta

wing,” Experiments in Fluids, Vol. 32, 2002, pp. 99–105.


17 Henne, P. A. and Gregg, R. D., “New Airfoil Design Concept,” Journal of Aircraft, Vol. 28, No. 5, 1991, pp. 300–311.
18 Köster, H., “Investigations of Airfoils with Cambered Trailing Edges,” DGLR Jahrbuch, Vol. 3, 1993, pp. 1265–1274.
19 Thompson, B. E. and Lotz, R. D., “Divergent-Trailing-Edge Airfoil Flow,” Journal of Aircraft, Vol. 33, No. 5, 1996,

pp. 950–955.
20 Bechert, D. W., Meyer, R., and Hage, W., “Drag Reduction of Airfoils with Miniflaps. Can We Learn from Dragonflies?”

AIAA-2000-2315, AIAA Fluids 2000 Conference and Exhibit, Denver (CO), 19–22 June 2000.
21 Richter, K. and Rosemann, H., “Experimental investigation of trailing-edge devices at transonic speeds,” The Aeronautical

Journal, Vol. 106, No. 1058, 2002, pp. 185–193.


22 Richter, K. and Rosemann, H., “Numerical Investigation on the Aerodynamic Effect of Mini-TEDs on the AWIATOR Air-

craft at Cruise Conditions,” 25th Congress on International Council of the Aeronautical Sciences (ICAS 2006), 8–9 September
2006.
23 Binder, B., Riethmüller, L., Tusche, S., and Wulf, R., “Modernisierung des Transsonischen Windkanals in Göttingen,”

DGLR Jahrbuch, Vol. 1, 1992, pp. 237–249.


24 Ludwieg, H., Lorenz-Meyer, W., and Schneider, W., “Der Transsonische Windkanal der Aerodynamischen Versuchsanstalt

Göttingen,” WGLR Jahrbuch, 1966, pp. 145–155.


25 Amecke, J., “Direkte Berechnung von Wandinterferenzen und Wandadaption bei zweidimensionaler Strömung in Wind-

kanälen mit geschlossenen Wänden,” Forschungsbericht DFVLR FB 1985-62, DFVLR, 1985.


26 Gerhold, T., Friedrich, O., Evans, J., and Galle, M., “Calculation of Complex Three-Dimensional Configurations Em-

ploying the DLR-TAU-Code,” AIAA-97-0167, AIAA 35th Aerospace Sciences Meeting and Exhibit, Reno (NV), 6–10 January
1997.
27 Kallinderis, Y., Khawaja, A., and McMorris, H., “Hybrid Prismatic/Tetrahedral Grid Generation for Complex Geome-

tries,” AIAA Journal, Vol. 34, No. 2, 1996, pp. 291–298.


28 Edwards, J. and Chandra, S., “Comparison of Eddy Viscosity-Transport Turbulence Models for Three-Dimensional,

Shock-Separated Flowfields,” AIAA Journal, Vol. 34, No. 4, 1996.


29 Dargel, G., “Prozesskette Hochauftrieb mit multifunktionalen Steuerflächen (Pro-HMS): Teilprojekt 2: Aerodynamik:

Schlussbereicht zum Technologievorhaben,” Interner Bericht, Airbus Deutschland, 2003.

19 of 19

American Institute of Aeronautics and Astronautics

You might also like