Present 1

Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

46th AIAA Aerospace Sciences Meeting and Exhibit AIAA 2008-381

7 - 10 January 2008, Reno, Nevada

Turbulent and Unsteady Flow Characteristics of


Delta Wing Vortex Systems

Andrej Furman∗ and Christian Breitsamter†


Institute of Aerodynamics
Technische Universität München
Boltzmannstrasse 15, D–85748, Garching, Germany

This paper presents an overview of experimental investigations on a 65 deg swept delta


wing as part of the International Vortex Flow Experiment 2 (VFE–2). Results obtained
in low–speed wind tunnel facilities include oil flow and laser light sheet flow visualization,
mean and unsteady surface pressure distributions as well as mean and turbulent velocity
components of the flow field and close to the wing surface. Details of the delta wing vortex
structure and breakdown phenomenon are discussed and analyzed. Special emphasis is on
the occurrence of an inner vortex detected for the low Reynolds number and Mach number
regime.

Nomenclature
AR aspect ratio
b wing span
bl local wing span
c mean aerodynamic chord
cp pressure coefficient
cp mean pressure coefficient
ĉp amplitude of pressure coefficient spectrum
cprms root mean square pressure coefficient
cr root chord
dd diameter of pressure probe
f frequency
fn lens focal length number
F wing area
k reduced frequency, f c/U∞
l length
M free stream Mach number
q∞ free stream dynamic pressure
rLE radius of leading edge
Rmac Reynolds number based on mean aerodynamic chord
∗ Dipl.-Ing., Member AIAA.
† PD Dr.-Ing., Associate Fellow AIAA

1 of 14

American Institute of Aeronautics and Astronautics

Copyright © 2008 by A. Furman and C. Breitsamter. Published by the American Institute of Aeronautics and Astronautics, Inc., with permission.
s wing semi span
Su0i power spectrum of velocity fluctuations
SuN0 non–dimensional power spectrum of velocity fluctuations
i

t time
tM measurement time
∆t pulse delay
T temperature
u, v, w axial, lateral and vertical velocities
ui mean velocity components
u0i velocity fluctuations
uirms root mean square velocity components
u0i u0j turbulent shear stresses
U∞ free stream velocity
x, y, z body–axis coordinates
α angle of attack
η fraction of local semi span, 2y/bl
ϕ leading–edge sweep
ρ density
ζ fraction of local height, 2z/bl

I. Introduction
Aerodynamic investigations of flow phenomena over generic and complex delta wing configurations have
been performed for many years.1–6, 13–17 The typical and well known delta wing flow physics is as follows: The
flow separates already at low angles of attack at the highly swept leading–edges. The separated shear layer
rolls up to form a large–scale vortex located over each half of the wing. Thus, two strong vortices influence
the flow field of the wing upper side. Vortex formation along the leading–edge starts from the rear part to
the apex. This primary vortex is fully developed when vorticity feeding extends over the entire leading–edge.
The vortex cross flow area reveals a rotational core with an embedded subcore, the latter dominated mainly
by viscous effects. The subcore is characterized by high axial velocities, low static pressures and enhanced
velocity fluctuations due to the steep gradient in the cross flow components. The mean velocities on the
wing upper surface are strongly increased by the leading–edge vortices resulting in high suction levels. The
corresponding suction peaks in the spanwise pressure distribution indicate the track of the vortex axis on the
wing surface. Therefore, leading–edge vortices in a fully developed, stable stage create additional lift and an
increase in maximum angle of attack improving significantly maneuver capabilities of high–agility aircraft.
Sharp leading–edge configurations are often used in delta wing research work because primary separation is
fixed and leading–edge vortex evolution is less sensitive to Reynolds number effects. Vortex aerodynamics
becomes much more complicated for rounded or blunt leading–edge configurations as the position of primary
separation varies to a certain extent depending on pressure gradient and boundary layer development. Thus,
leading–edge radius, angle of attack and Reynolds number are the main parameters influencing the onset of
vortex evolution as well as position and strength of the primary vortex whereas the angle of attack is the
main parameter for the sharp leading–edge case only. There is a strong increase in the surface pressure when
moving in spanwise direction from the station of the primary vortex suction peak to the leading–edge. This
severe lateral pressure gradient provokes boundary layer separation in that region. The separated boundary
layer rolls up by self induction and creates a small vortex, named secondary vortex, the rotation of which
is opposite to that of the leading–edge (primary) vortex. The formation of the secondary vortex depends

2 of 14

American Institute of Aeronautics and Astronautics


strongly on the presence of a laminar or turbulent boundary layer.19 Size and position of the secondary
vortex affects the primary vortex location and, thus, the associated suction level.
Further, leading–edge vortices are subject to breakdown at high angles of attack. Vortex breakdown is caused
by the stagnation of the low–energy axial core flow due to the increase of the adverse pressure gradient along
the vortex axis with increasing angle of attack. This rise in the vortex core static pressure at the wing rear part
is caused on the one hand by the diverging vortex subcore and on the other hand by the recompression in the
trailing–edge area. Vortex breakdown is indicated by the rapid expansion of the vortex core accompanied by
high velocity fluctuations. Downstream of breakdown, the fluctuation maxima are located in a limited radial
range around the burst vortex core. In addition, the breakdown flow exhibits specific instability mechanisms
leading to narrow–band unsteady aerodynamic forces.1, 13 The numerical simulation and analysis of the
breakdown flow is still a challenging problem which needs the correct representation of the turbulent flow
field and instability characteristics.
Euler methods were intensively developed in the 1980’s, also aimed to calculate leading–edge vortex flows.
Therefore, an experimental data base was needed for code validation and assessment established in frame of
the International Vortex Flow Experiment 1 (VFE–1; 1984 – 1986). The tests included force and pressure
measurements as well as flow field studies on a 65 deg swept cropped delta wing carried out in several wind
tunnel facilities.9 But Euler code results do not represent secondary vortices limiting the accuracy even for
fixed primary vortices at sharp leading edges. The vortex induced surface pressure distribution is sensitive to
viscous effects on the wing as well as in the rolled–up shear layers. Great efforts have been made in the last
decade to develop and use high fidelity computational fluid dynamics methods. Unsteady Reynolds Averaged
Navier–Stokes (URANS) Methods are available including a variety of turbulence models of algebraic type
up to Reynolds stress transport equations. Further, methods for Detached Eddy Simulations (DES) are
used as a combination of a Large Eddy Simulation (LES) to model separated flow dominated by large–scale
structures in the outer domain and a turbulence model to calculate flow quantities in the wall–bounded
domain. With such high fidelity methods available,7, 8, 22–25 there is again a strong need for an extended
experimental data base. Consequently, a second International Vortex Flow Experiment (VFE–2) has been
initiated. The VFE–2 activities are coordinated by the task group 113 of the Applied Vehicle Technology
panel (AVT) of the NATO Research Technology Organization (RTO) named ”Understanding and modeling
of vortical flows to improve the technology readiness level for military aircraft”. 18 Latest test techniques are
applied to gather high quality data. The research activities have been started in 2004 by partners of industry,
research establishments and universities from Europe and the United States. The work is still on–going.
The Institute of Aerodynamics (AER) of the Technische Universität München (TUM) is part of the RTO
AVT–113 task group focusing particularly on flow field turbulence and boundary layer quantities. 10–12 Both
sharp and rounded leading edges are investigated. The tests are conducted on a new delta wing model
which has been manufactured based on a NASA wing geometry served as reference configuration for VFE–
2.6 The TUM–AER investigations have been performed in low–speed wind tunnel facilities using laser light
sheet and oil flow visualization, steady and unsteady surface pressure measurements, stereo particle image
velocimetry and hot–wire anemometry. The results are presented as distributions of mean, turbulent and
spectral quantities, which give detailed information of the flow characteristics over the delta wing.

II. Experimental set–up


II.A. Facility
The measurements have been carried out in the low–speed wind tunnels A and B of the Institute of Aerody-
namics at the Technische Universität München.10 The wind tunnels are of closed–return type with an open
test section. The free stream turbulence intensity is less than 0.4%. The uncertainty in the temporal and
spatial mean velocity distribution is less than 0.6%. The uncertainty in free stream direction is below 0.2
deg and static pressure variations are below 0.4%.

3 of 14

American Institute of Aeronautics and Astronautics


II.B. Model
The present delta wing model10 with 65 deg leading–edge sweep has a root chord length of cr = 0.980 m, a
wing span of b = 0.914 m and an aspect ratio of AR = 1.865, Fig. 1. Leading–edge sections can be equipped
with a sharp or a rounded contour (rLE,rounded /c = 0.0015). There are 177 pressure orifices with a diameter
of dd = 0.3 mm situated on the entire wing, of which 44 are installed with unsteady pressure sensors. The
pressure orifices are positioned at five chordwise positions x/cr = 0.2, 0.4, 0.6, 0.8, and 0.95.

Figure 1. Delta wing model mounted in test section of wind tunnel facility A (left) and in test section of wind
tunnel facility B (right).

II.C. Cases
The investigations have been carried out for three angles of attack, namely at

• α = 13◦ for partly developed,


• α = 18◦ for fully developed and
• α = 23◦ for burst leading–edge vortices.

The test Mach numbers for all cases are M = 0.07 and M = 0.14 and the corresponding Reynolds numbers
based on the mean aerodynamic chord are Rmac = 1 · 106 and Rmac = 2 · 106 , respectively. Except for
the laser light sheet measurements, which have been obtained at Mach number M = 0.035 and a Reynolds
number of Rmac = 0.5 · 106 .

II.D. Measurement techniques


During the project several measurement techniques have been used:

Laser light sheet visualization11 is used in order to survey the flow field in a plane, which is illuminated
by a laser beam expanded by a cylindrical lens. The smoke particles in this plane are then recorded with a
digital photo camera. The particle size is approximately 2 µm, in order to guarantee sufficient light reflec-
tion. For this investigation an air cooled class 3B Argon–Ion–laser has been used. This laser has a maximum
power of 100 mW and the wave length of the light is between 457 ÷ 514 nm.

4 of 14

American Institute of Aeronautics and Astronautics


Oil flow visualization11 illustrates the surface stream lines on the suction side of the wing. A black foil
was stuck on the model surface to achieve good visual contrast between oil flow pigments and background
color. A mixture of yellow paint pigments, petrol and paraffin was applied on the upper surface and exposed
to the free stream flow briefly. The developed flow picture was then photographed.

Surface pressure measurements.3, 10, 11 The steady pressures are measured on the upper and lower sur-
face of the wing at five chord stations with 133 measuring tabs in total. The sampling rate of the measured
values is f = 100 Hz with an averaging time of t = 10 s. The unsteady pressure measurements were accom-
plished in four chord stations on the suction side of the wing with 12 unsteady pressure sensors per chord
station. A sampling rate of f = 2000 Hz and a sample time of t = 40 s is used. The frequency of the analog
low–pass filter was set to 256 Hz.

Stereo Particle Image Velocimetry (Stereo–PIV) 12 is performed with two cameras left and right of
a laser light sheet. A pair of 135 mm, fn = 2.8 objective lenses constitute the recording optics and are
connected to the charge coupled device (CCD) with a 1600 × 1186 pixel resolution. The light sheet was
generated by a frequency doubled, double oscillator Nd–YAG laser with a maximum energy level of 200 mJ
and a frequency of 10 Hz per pulse. The light sheet thickness was set at approximately 10 mm and the
pulse delay was set to ∆t = 21 µs.

Hot–Wire Anemometry (HWA)1, 12 is based on a dual–sensor probe of cross–wire type for measuring
the fluctuating velocities. The probes were operated by a multi–channel constant–temperature anemometer
system. By means of its signal conditioner modules, bridge output voltages were low–pass filtered at 1000 Hz
before digitization and amplified for optimal signal level. The sampling time for each channel is 6.4 s, with
the sampling rate set to 3000 Hz (Nyquist frequency of 1500 Hz), so that each sample block contains 19200
values. The use of cross–wires generally assumes some knowledge of the flow field, such as a known flow
direction to which the probe must be aligned. To determine the three velocity components, the probe has to
be rotated around its axis by 90 deg to adjust the wire plane once horizontal and once vertical against the
main flow direction. Thus, two triggered traverse sweeps are necessary to obtain the streamwise u, lateral
v and vertical w velocity components, respectively. Each digitized and temperature corrected voltage pair
of the corresponding probe positions was converted to evaluate the time–dependent velocity vector. The
numerical method used is based on look–up tables derived from the full velocity and flow angle calibration
of the probe.1

III. Analysis of results


The discussion and analysis of the measurement results address flow topology, mean and fluctuating
surface pressures, mean and fluctuating flow field and boundary layer velocities and spectral quantities.

III.A. Flow topology


III.A.1. Flow field
Laser light sheet visualization can be used to determine the flow behavior above and behind the wing as
well as the size and position of the vortices.1, 11, 20 The pictures taken from behind the wing for the stations
x/cr = 0.2, 0.4, 0.6, 0.8, 0.95, and 1.1 are shown on the right hand side of Fig. 2. They illustrate cross
sections of the vortex. Pictures from above, i.e. parallel to the wing upper surface, at z/c r = 0.01 and 0.05
are shown on the left hand side of Fig. 2. Based on such images the position of the leading–edge vortex
center (trajectories of vortex axis) and the breakdown location at high angles of attack can be determined.
Here, the partly developed leading–edge vortex at α = 13 deg is depicted where the primary separation at
the sharp leading–edge has not yet reached the apex. Also, the area of the secondary vortex can be detected

5 of 14

American Institute of Aeronautics and Astronautics


underneath the primary vortex close to the leading–edge. The comparison between sharp and rounded
leading–edge shows a delayed separation for the rounded leading–edge, which is also indicated by a smaller
vortex core and a stronger trailing–edge effect (Fig. 2).

a) sharp leading–edge b) rounded leading–edge

Figure 2. Vortex structure above and behind the wing at α = 13◦ , Rmac = 5 · 105 and M = 0.035.

III.A.2. Surface flow


Concentrating again on the flow behavior for the rounded leading–edge case at α = 13 ◦ , the corresponding
surface streamlines are shown by oil flow visualization, Fig. 3. A schematic representation depicts the flow
topology derived from the laser light sheet and oil flow pictures. This schematic highlights the separation and
attachment lines of the primary and secondary vortex as indicated by the surface streamlines. The leading–
edge (primary) vortex starts from a turbulent separation at the wing rear part, with the turbulent shear
layers of the wing upper and lower side rolling up along the leading–edge. The roll–up process does not reach
the apex at this angle of attack but progresses up to the wing front part. There, a laminar boundary layer is
present. Transition is indicated by an outboard shift of the separation line of the secondary vortex. Further,
a laminar separation exists near the apex close to the symmetry plane. This three–dimensional separation
bubble is caused by the pressure increase when the flow has turned around the leading–edge contour of the
relatively thick wing. Downstream, the inboard separated flow forms a small inner vortex favored by the
positive lateral pressure gradient between the primary vortex suction area and the local suction minimum
at the symmetry plane. The inboard vortex rotates in the same direction as the leading–edge (primary)
vortex. Separation and attachment line of the inboard vortex lie closely together and can be clearly seen in
the oil flow picture. The inboard vortex extends over the entire chord length to the wing rear edge. Further
tests have shown that the trajectory and strength of the inboard vortex depends strongly on angle of attack
(strength of primary vortex and adverse lateral pressure gradient) and Reynolds number (area of laminar

6 of 14

American Institute of Aeronautics and Astronautics


flow). Therefore, this phenomenon is only visible at certain Reynolds numbers in the medium angle of attack
range.

Figure 3. Flow topology over the wing for rounded leading–edge at α = 13◦ , M = 0.14 and Rmac = 2 · 106 .

III.B. Mean flow field


III.B.1. Pressure distribution
The quality of TUM–AER measurements have been judged by comparison with steady pressure measure-
ments obtained by NASA for the delta wing reference configuration.6, 21 Figure 4 contains the NASA results
for Rmac = 2 · 106 , M∞ = 0.2 and the results obtained by TUM–AER for Rmac = 2 · 106 , M = 0.14. The
comparison shows an excellent agreement between the steady pressure distributions, which clearly illustrates
the comparability in terms of free stream, wind tunnel and model conditions.

III.B.2. Velocity distribution


The time–averaged velocities obtained by Stereo–PIV at cross sections x/c r = 0.2, 0.4, 0.6, 0.8 and 0.95
(Rmac = 1·106 , M = 0.07) are shown for an angle of attack of α = 18 deg in Fig. 5. The velocity components
in axial direction are displayed as contour plots and in lateral and vertical direction as vector plots. For
both sharp and rounded leading–edge, fully developed leading–edge vortices are present. The structure of
primary and secondary vortex is depicted by the corresponding increased cross flow velocity vectors. The
primary vortex is associated with axial accelerated flow, with an axial peak velocity of u/U ∞ = 1.6 for both
sharp and rounded leading–edge.12 The wing inboard region is characterized by attached flow while there is

7 of 14

American Institute of Aeronautics and Astronautics


-6 -6
upper surface TUM, M = 0.14 upper surface TUM, M = 0.14
lower surface TUM, M = 0.14 lower surface TUM, M = 0.14
-5 upper surface NASA, M = 0.2 -5 upper surface NASA, M = 0.2
lower surface NASA, M = 0.2 lower surface NASA, M = 0.2

-4 -4

-3 -3
Cp

Cp
_

_
-2 -2

-1 -1

0 0

1 1
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
η η

Figure 4. Comparison between NASA– and TUM–measurements for sharp (left) and rounded (right) leading–
edge at Rmac = 2 · 106 , α = 18◦ and x/cr = 0.6.

no clear indication for a small inboard vortex by this mean velocity field. The overall flowfield pattern does
not show markable differences between the configurations of sharp and rounded leading–edge.

Figure 5. Mean velocity distribution for sharp (left) and rounded (right) leading–edge at α = 18 ◦ , Rmac = 1 · 106
and M = 0.07 .

III.C. Turbulent flow field


III.C.1. Unsteady pressures
Referring further on to an angle of attack of α = 18 deg, the unsteady surface pressures are discussed
based on root–mean–square (rms) values of the pressure coefficient and pressure amplitude spectra. The
maximum peak of pressure fluctuation intensities cprms in section x/cr = 0.6 is observed in the area of the
primary vortex outside of the suction peak near the attachment line of the secondary vortex (Fig. 6, left).

8 of 14

American Institute of Aeronautics and Astronautics


The suction peak of the mean pressure cp is broadened for the case of rounded leading–edge. It is located
closer to the leading–edge due to retarded primary separation in comparison to the sharp leading–edge. The
primary separation line, which is not geometrically fixed for the rounded leading–edge, is associated with
increased pressure fluctuation intensities in the area of the leading edge η = 0.8 ÷ 1 in comparison to the
sharp edge case. This behavior is reflected in the right part of Fig. 6, where the amplitude spectra of the
fluctuating pressure coefficient ĉp are plotted as function of reduced frequency k for each measured spanwise
station η. Raised amplitude values can be detected in direction to the leading–edge linked to fluctuations in
the primary separation and secondary separation and attachment. The spectra reveal a broadband behavior
and are typical for a fully developed leading–edge primary vortex.1, 3, 10, 11

y
-3 0.12
x/cr = 0.6
sharp Cp
sharp Cprms
-2.5 rounded Cp 0.1
x
rounded Cprms
sharp leading edge
rounded leading edge
-2 0.08

Cprms
0.02
Cp

-1.5 0.06
_

0.015

p
-1 0.04 C
^
0.01
1
0.9
0.8
0.005
-0.5 0.02 0.7
0.6
0.5

η
0
0 0.4
0.5
0 0 1 0.3
1.5
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 2 0.2
η k 2.5
3
3.5
0.1
4 0

Figure 6. Steady pressure distribution, pressure fluctuation intensity and amplitude spectra of the fluctuating
pressure coefficient for sharp and rounded leading–edge at α = 18◦ , x/cr = 0.6, Rmac = 2 · 106 and M = 0.14.

III.C.2. Reynolds stresses


The fluctuating velocity field at station x/cr = 0.6 and α = 18 deg is obtained by hot–wire measurements, the
results of which are shown for rounded leading–edge in Fig. 7. At the top left, the mean velocity distribution
is depicted, which illustrates the positions of primary and secondary vortices as well as the magnitude of
axial and cross flow velocity components. At the top right, the rms values of the axial velocity fluctuations
are plotted. The fluctuation intensities indicate high levels for the viscous core of the primary vortex and
the region of the secondary vortex and moderate levels for the separating shear layer and primary vortex
rotational core. The intensity of the lateral velocity fluctuations also displays high levels for the primary
vortex subcore as well as for the shear layer (Fig. 7, middle left). Considering the intensity of the vertical
velocity fluctuations, increased values exist mainly in the primary vortex subcore (Fig. 7, middle right). Now,
moderate levels are found for the region of the secondary vortex because the vertical velocity fluctuations are
2
the first to be damped approaching the wing surface. The shear stress u0 v 0 /U∞ shows high positive values in
the vortex core and in the surface flow under the primary vortex (Fig. 7, bottom left). High negative values
are visible in the outboard shear layer and in the shear layer over the primary vortex subcore. Negative and
positive values are determined by the direction of the lateral velocity when moving from outboard to inboard
2
along the vorticity feeding shear layer. The shear stress distribution u0 w0 /U∞ exhibit again increased levels
for the regions of vortical flow. Peak values are located in neighbored regions relative to the u 0 v 0 –maxima
and are of opposite sign. The direction of the vertical velocity determines the sign of this stress component.
A small region of slightly increased velocity fluctuations and turbulent shear stresses near the symmetry
plane of the wing is due to a weak inboard vortex, the development of which was explained above.

9 of 14

American Institute of Aeronautics and Astronautics


vrms/U∝ wrms/U∝
0.3 0.3
0.28 0.28
0.26 0.26
0.24 0.24
0.22 0.22
0.2 0.2
0.18 0.18
0.16 0.16
0.14 0.14
0.12 0.12
0.1 0.1
0.08 0.08
0.06 0.06
0.04 0.04
0.02 0.02
0 0

___ 2 ___
u’v’/U∝ u’w’/U2∝
0.005 0.005
0.003 0.003
0.001 0.001
-0.001 -0.001
-0.003 -0.003
-0.005 -0.005
-0.007 -0.007
-0.009 -0.009
-0.011 -0.011
-0.013 -0.013
-0.015 -0.015

Figure 7. Mean velocities and Reynolds stresses for rounded leading–edge at α = 18 ◦ , x/cr = 0.6, Rmac = 1 · 106
and M = 0.07.
10 of 14

American Institute of Aeronautics and Astronautics


III.C.3. Boundary layer profiles
Boundary layer profiles of mean axial velocity and velocity fluctuation components are discussed, regarding
again the case of α = 18 deg and rounded leading–edge. The measurement points closest to the wall were
placed 1 mm above the wing surface. Figure 8 top and bottom left show results for chord stations x/c r = 0.4
and x/cr = 0.6 and a lateral position of η = 0.4. The x–axis in the diagram represents the vertical distance
from wing surface based on the local wing semi span ζ = 2z/bl . The left y–axis refers to the range of the axial
velocity based on free stream velocity and the right y–axis refers to the range of the non–dimensional velocity
fluctuations. The middle picture of Fig. 8 shows the oil flow image, with the measurement positions marked
by red dots. The corresponding sections of surface pressure fluctuation intensities are also included, Fig.
8 right. Regarding the two chordwise stations, the profiles of the mean axial velocity indicate a turbulent
boundary layer. All three rms velocities show significantly increased levels in the boundary layer and a low
constant level outside. The same trend holds for the turbulent shear stresses which are multiplied by a factor
of −20 for appropriate representation in the diagram. These turbulent boundary layers are typical for the
wing inboard region close to the primary vortex attachment line. Only low surface pressure fluctuations are
evoked as the flow is accelerated in direction to the vortex axis.

a) x/cr = 0.4, η = 0.4

b) x/cr = 0.6, η = 0.4


Figure 8. Boundary layer (left), surface flow (middle) and pressure fluctuation intensity (right) for the rounded
leading edge at α = 18◦ , Rmac = 1 · 106 and M = 0.07.

III.C.4. Vortex bursting


With increasing angle of attack vortex bursting occurs over the wing. Here, vortex breakdown takes place
at x/cr = 0.75 ÷ 0.85 at α = 23 deg. The position of vortex breakdown can be determined by analyzing

11 of 14

American Institute of Aeronautics and Astronautics


the surface pressure distribution.11 If an increase in angle of attack does not cause an increase in the
primary vortex suction peak comparing stations x/cr = const., then the leading–edge vortex may experience
breakdown. This criteria fails, if the primary vortex detaches from the wing surface and thereby reduces
its influence on the pressure distribution at the same time, as the breakdown location passes the trailing–
edge upstream. The detachment of the vortex axis for strong leading–edge vortices, present at this delta
wing planform, is only observed at angles of attack beyond the ones investigated here. Therefore, the
described breakdown criteria is applicable. This criteria can also be discussed in context of the turbulence
intensity distributions, shown for the axial velocity fluctuations at x/cr = 0.8 in Fig. 9, left. The turbulence
intensity distribution exhibits an annular concentration of local rms maxima. This turbulence structure is a
characteristic feature of spiral vortex breakdown which is related to Reynolds numbers above 10 4 .1, 12
In Figure 9 right, the power spectral density of the axial velocity fluctuations S uN0 is shown for the near wall
point together with the amplitude spectra of the fluctuating pressure coefficient ĉp taken at the corresponding
surface station η = 0.775. Both spectral quantities are plotted as function of reduced frequency k. The
comparison highlights increased levels of power spectral densities in the near wall flow with some narrow–
band concentration within k = 1 ÷ 2. The amplitude spectra of pressure fluctuations reveal a narrow–band
concentration for a higher frequency range, namely k = 1.5 ÷ 2.5. That means vortex breakdown has started
to influence the surface pressure fluctuations. The narrow–band concentration of turbulent kinetic energy
at burst flow conditions reflects the helical mode instability of the vortex breakdown flow. 11, 13, 14

Figure 9. Turbulence intensity (left), power spectral density of the axial velocity fluctuations (upper right)
and amplitude spectra of fluctuating pressure coefficient (lower right) for rounded leading–edge at α = 23 ◦ ,
x/cr = 0.8 and η = 0.775.

12 of 14

American Institute of Aeronautics and Astronautics


IV. Conclusions and Outlook
A variety of experimental investigations on a delta wing model with a leading–edge sweep of 65 deg
have been performed in the low–speed wind tunnel facilities A and B of the Institute of Aerodynamics at
the Technische Universität München. The results contribute to the research work conducted within the
International Vortex Flow Experiment 2 (VFE–2). The experiments performed contain different measuring
techniques, like flow visualization using laser light sheet technique and oil flow technique, steady and unsteady
surface pressure measurements, flow field velocity measurements using Stereo Particle Image Velocimetry
and hot–wire anemometry, and boundary layer measurements based on hot–wire anemometry.
For the delta wing vortical flow structure some new and significant results are obtained, especially, when
considering a rounded leading–edge:

• For medium angles of attack, a new flow phenomenon was found for delta wings with straight leading–
edge depending strongly on Reynolds number. In addition to the classical primary vortex an inboard
vortex occur close to the wing surface. This phenomenon appears stronger for the rounded than
for the sharp leading–edge. While the primary vortex develops from the trailing–edge towards the
apex with increasing angle of attack and therewith starts here from a turbulent separation, a laminar
separation occurs at the wing surface in the region of the apex close to the symmetry plane. The flow
is attached around the leading–edge, but the pressure increases towards the symmetry plane of the
wing causing laminar separation in the inboard area. Downstream, the three–dimensional separation
bubble transforms to a spatially small and weak vortex, which is situated close to the wing surface
along the entire chord length.
• At high angle of attack, vortex breakdown dominates the wing flow associated with a characteristic
annular region of local turbulence maxima surrounding the strongly expanded core of the burst primary
vortex. Further, a narrow–band concentration of turbulent kinetic energy takes place. During the
upstream movement of the breakdown location the turbulent flow field affects more and more the wing
surface flow, thereby increasing the surface pressure fluctuations which also show coherent structures
and significant concentrations in a certain frequency domain.
• Measurements of the boundary layer allows the quantification of the time averaged velocities as well
as of the turbulent normal– and shear stresses close to the wing surface. For the Reynolds numbers
investigated here and medium angles of attack, a turbulent boundary layer starts to develop at approx-
imately 20% to 30% of the root chord for the attached flow in the inner part of the wing. Under the
primary vortex the boundary layer becomes thinner by a factor of 2 to 5 due to the strong accelerated
flow.

The comprehensive data base will be further evaluated and analyzed to improve the knowledge on the tur-
bulent and unsteady flow quantities associated with the different stages of leading–edge vortex development.

Acknowledgments
The support of this investigation by the German Research Association (DFG) is gratefully acknowledged.
The authors would also like to thank the VFE–2 partners for the fruitful and excellent co–operation.

References
1 Breitsamter, C., Turbulente Strömungsstrukturen an Flugzeugkonfigurationen mit Vorderkantenwirbeln, Dissertation,

Technische Universität München, Herbert Utz Verlag, ISBN 3-89675-201-4, 1997.


2 Breitsamter, C., and Laschka, B., Turbulent Flow Structure Associated with Vortex–Induced Fin Buffeting, Journal of

Aircraft, Vol. 31, No. 4, 1994, pp. 773–781.

13 of 14

American Institute of Aeronautics and Astronautics


3 Breitsamter, C., Experimentelle Untersuchung der instationären Feldgrös̈en und Oberfl”achendr”ucke bei wirbeldo-

minierter abgelöster Strömung an einem Deltaflügel , Jahrbuch der DGLR, Vol. I, DGLR–JT95–062, 1995, pp. 163–175.
4 Breitsamter, C., and Laschka, B., Fin Buffet Pressure Evaluation Based on Measured Flowfield Velocities, Journal of

Aircraft, Vol. 35, No. 5, 1998, pp. 806–815.


5 Breitsamter, C., Strake Effects On the Turbulent Fin Flowfield Of a High–Performance Fighter Aircraft, NNFM, Vol. 72,

Vieweg Verlag, 1999, pp. 69–76.


6 Chu, J., and Luckring, J. M., Experimental surface pressure data obtained on 65 o delta wing across Reynolds number and

Mach number ranges, NASA–TM–4645, 1996.


7 Crippa, S., and Rizzi, A., Numerical Investigation on Reynolds Number Effects on a Blunt Leading–Edge Delta Wing,

AIAA Paper 2006–3001, June 2006.


8 Crippa, S., and Rizzi, A., Initial Steady/Unsteady CFD Analysis of Vortex Flow Over the VFE-2 Delta Wing, ICAS–

2006–P2.18, 25th Congress of the International Council of the Aeronautical Sciences, Hamburg, Germany, 3–8 Sept. 2006.
9 Elsenaar, A., Hjelmberg, L., Bütefisch, K. and Bannink, W. J., The International Vortex Flow Experiment, AGARD–

CP–437, Validation of Computational Fluid Dynamics, Lisbon, Portugal, May. 2–5, 1988, pp. 9-1–15-23.
10 Furman, A. and Breitsamter, C., Delta Wing Steady Pressure Investigations for Sharp and Rounded Leading Edges, New

results in Numerical and Experimental Fluid Mechanics V, NNFM Vol. 92, 2006, pp. 77–84.
11 Furman, A. and Breitsamter, C., Investigations of Flow Phenomena on Generic Delta Wing, ICAS–2006–3.1.2, 25 th

Congress of the International Council of the Aeronautical Sciences, Hamburg, Germany, 3–8 Sept. 2006.
12 Furman, A. and Breitsamter, C., Stereo–PIV and Hot–Wire Investigations on Delta Wing with Sharp and Rounded

Leading Edge, CEAS–2007–430, 1st CEAS European Air and Space Conference, Berlin, Germany, 10–13 Sept. 2007, pp. 1749–
1761.
13 Gursul, I., Unsteady Flow Phenomena over Delta Wings at High Angle of Attack , AIAA Journal, Vol. 32, No. 2, 1994,

pp. 225–231.
14 Gursul, I., and Xie, W., Buffeting Flows Over Delta Wings, AIAA Journal, Vol. 37, No. 1, 1999, pp. 58–65.
15 Hoeijmakers, H. W. M., Modelling and numerical simulation of vortex flow in aerodynamics, AGARD–CP–494, 1991,

pp. 1-1–1-46.
16 Hummel, D., On the Vortex Formation over a Slender Wing at Large Angles of Incidence, AGARD–CP–247, High Angle

of Attack Aerodynamics, Sandefjord, Norway, Oct. 4-6, 1978, pp. 15-1–15-17.


17 Hummel, D., Documentation of Separated Flows for Computational Fluid Dynamics Validation, AGARD–CP–437, Val-

idation of Computational Fluid Dynamics, Vol. 2, Lisbon, Portugal, May 2-5, 1988, pp. 18-1–18-24.
18 Hummel, D., and Redeker, G., A new vortex flow experiment for computer code validation, Vortex Flow and High Angles

of Attack, Loen, Norway, May 7–11, 2001.


19 Hummel, D., Effects of Boundary Layer Formation on the Vortical Flow above Slender Delta Wing, RTO–MP–AVT–111,

Paper 30, Symposium on Enhancement of NATO Military Flight Vehicle Performance by Management of Interacting Boundary
Layer Transition and Separation, Prague, Czech Republic, 4–7 Oct. 2004.
20 Hummel, D., The Second International Vortex Flow Experiment (VFE–2): Objectives and Present Status, AIAA Paper

2007–4446, June 2007.


21 Luckring, J. M., Reynolds Number, Compressibility, and Leading–Edge Bluntness Effects on Delta–Wing Aerodynamics,

ICAS–2004–4.1.4, 24th Congress of the International Council of the Aeronautical Sciences, Yokohama, Japan, 29 Aug. – 3 Sept.
2004.
22 Mitchell, A. M., Molton, P., Barberis, D., and Delery, J., Characterization of vortex breakdown by flow field and surface

measurements, AIAA Paper 2000–0788, 2000.


23 Müller, J., and Hummel, D., Time–accurate CFD analysis of the unsteady flow on a fixed delta wing, AIAA Paper

2000–0138, 2000.
24 Rai, P., Finley, D. B., and Ghaffari, F., An Assessment of CFD Effectiveness for Vortex–Flow Simulation to Meet

Preliminary Design Needs, Vortex Flow and High Angles of Attack, Loen, Norway, May 7–11, 2001.
25 Schiavietta, L. A., Badcock, K., and Cummings, R., Comparison of DES and URANS for Unsteady Vortical Flows Over

Delta Wings, AIAA Paper 2007–1085, Jan. 2007.

14 of 14

American Institute of Aeronautics and Astronautics

You might also like