Sie sind auf Seite 1von 13

AIAA JOURNAL

Vol. 50, No. 1, January 2012

Validation of Unsteady Reynolds-Averaged NavierStokes


Simulations on Three-Dimensional Flapping Wings
Stephan E. Bansmer and Rolf Radespiel
Technische Universitt Braunschweig, 38106 Braunschweig, Germany

Downloaded by California Inst of Technology on October 22, 2012 | http://arc.aiaa.org | DOI: 10.2514/1.J051226

DOI: 10.2514/1.J051226
A combined experimental and computational study is presented for a wing segment undergoing a combined
pitching, plunging, and rolling motion at Reynolds number of 100,000, where transition takes place along laminar
separation bubbles. The numerical simulation approach addresses unsteady Reynolds-averaged NavierStokes
solutions and covers three-dimensional transition prediction for unsteady mean ows. The numerical simulations are
validated using high-resolution, phase-locked stereoscopic particle image velocimetry for a three-dimensional
apping case with a reduced frequency of k  0:25. The ow reveals strong unsteadiness resulting in moving laminar
separation bubbles, whose spatial extensions are varying in the spanwise direction. The experimental results are
well captured by the numerical simulations performed in this study. Because of the vortex structure in the wake
of the wing segment, the three-dimensional aerodynamics cannot be reproduced as a spanwise sequence of
two-dimensional results.



Nomenclature
A
c
cd
cd;f
cd;p
cl
cm
ct
cx , cy , cz
f
i
j
k
M
NN
Re
T
t
U1
u, v, w
u0 , v0 , w 0
v
x, y, z
x
xt
x0

=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=

z^

eff
i

=
=
=
=


eff

=
=

amplitude of wave disturbance oscillation, m


chord length, m
drag coefcient, sum of cd;f and cd;p
friction drag coefcient
pressure drag coefcient
lift coefcient
vector of x, y, and z moment coefcients
thrust coefcient
force coefcients
frequency of the apping motion, Hz
integer increment p
imaginary unit, j  1
reduced frequency
number of samples
factor
Reynolds number
period time of one apping cycle, s
time, s
freestream velocity, m=s
velocity components, m=s
turbulence velocity components, m=s
velocity vector
Cartesian coordinates, m
vector of Cartesian coordinates
Cartesian coordinate of the transition location, m
Cartesian coordinate of the point of neutral
stability, m
plunging amplitude, m
complex wave number in x direction, 1=m
effective angle of attack, deg
imaginary part of the complex wave number in x
direction, 1=m
complex wave number in y direction, 1=m
amplitude of the effective angle of attack, deg


p
L
t
%
xz

_
^
0

!

= phase difference between pitch and plunge motions,


deg
= induced angle of attack due to plunge motion, deg
= propulsive efciency
= pitch/plunge amplitude ratio
= eddy viscosity, kg=m  s
= density, kg=m3
= turbulent shear stress, kg=ms2 
= pitch angle, geometric angle of attack, deg
= vector of angular pitch, roll, and yaw velocities
= pitch amplitude, deg
= mean pitch, mean geometric angle of attack, deg
= roll angle, deg
= complex frequency of disturbance wave, Hz

I. Introduction

LIGHT physics of birds intersect with some of the most


challenging problems in modern aerospace engineering:
unsteady three-dimensional separation, moving transition with
laminar separation bubbles in boundary layers, unsteady ight
environment, aeroelasticity, and anisotropic wing structure are just a
few examples [1].
Focusing on three-dimensional aerodynamics, comparisons
between experimental and computational results of a nominally
three-dimensional apping-wing segment at low Reynolds numbers
(Re  100; 000) are presented. The apping motion is prescribed by
a combined plunging, pitching, and rolling motion with a phase
difference of 90 deg between plunging and pitching.
In the past, Neef [2] investigated the propulsive efciency of a
apping NACA 0012 wing segment using different inviscid ow
solvers. The kinematics of the wing were prescribed by combined
plunging, pitching, rolling, and torsional motions. The wingtip
vorticity was identied as a major loss of propulsive efciency. The
effect of the combined rolling and torsional motion was remarkable,
because it enables the wing to generate thrust during upstroke and
downstroke of the apping cycle.
However, the viscous effect in this low Reynolds number range of
100,000 cannot be neglected, since the laminarturbulent transition
takes place along laminar separation bubbles (LSBs). Figure 1
describes the physics of a LSB. The oncoming laminar boundary
layer separates, which is caused by a pressure increase along the
airfoil contour. According to Spalart and Strelets [3] and Rist [4], the
separated ow performs the transition process from laminar to
turbulent ow following a gradual development of the
primary instabilities from TollmienSchlichting instabilities toward

Received 15 February 2011; revision received 21 May 2011; accepted for


publication 27 May 2011. Copyright 2011 by the authors. Published by the
American Institute of Aeronautics and Astronautics, Inc., with permission.
Copies of this paper may be made for personal or internal use, on condition
that the copier pay the $10.00 per-copy fee to the Copyright Clearance Center,
Inc., 222 Rosewood Drive, Danvers, MA 01923; include the code 0001-1452/
12 and $10.00 in correspondence with the CCC.

Research Associate, Institute of Fluid Mechanics, Bienroder Weg 3;


s.bansmer@tu-braunschweig.de.

Head of the Institute, Institute of Fluid Mechanics, Bienroder Weg 3.


Senior Member AIAA.
190

191

BANSMER AND RADESPIEL

shear-stress distributions around the apping-wing segment at a


reduced frequency of k  0:25.
One feature of three-dimensional apping wings is the complex
vortex system in the wake (see Fig. 2). It is composed of a
superposition of start/stop eddies with wingtip vortices. In this
contribution, wingtip vortices due to the nite span are suppressed by
side walls at the end of the wing segment; however, a dynamically
evolving vortex sheet with spanwise variation of unsteady circulation
is present.

II. Birdlike Wing Segment


A.

Downloaded by California Inst of Technology on October 22, 2012 | http://arc.aiaa.org | DOI: 10.2514/1.J051226

Fig. 1 Sketch of a laminar separation bubble by Horton [30]


(corrected).

KelvinHelmholtz instabilities [5]. The resulting turbulent


uctuations in the ow enhance momentum transport toward the
wall, and the ow reattaches to the airfoil contour. The resulting
region of circulating ow is called the laminar separation bubble.
LSBs are usually not desired in airfoil design, because they increase
the pressure drag of the airfoil due to a higher displacement-thickness
level of the boundary layer.
In this contribution, a validated and efcient numerical tool will be
provided, which covers the unsteady ow case of a threedimensional apping-wing segment at transitional low Reynolds
numbers. This work is based upon previous two-dimensional results.
In the past, the authors developed a validated method to predict the
viscous ow around an oscillating airfoil performing a combined
pitching and plunging motion [6,7]. The ow solver of our choice
was FLOWer [8]. For the present three-dimensional investigations,
the authors migrated toward the TAU code of the DLR, German
Aerospace Center [9], which includes an implemented threedimensional transition prediction [10,11]. Therefore, the present
approach consists of three steps:
1) Generate a birdlike wing segment. The work draws on naturally
evolved airfoils. Although technical airfoils might achieve a better
aerodynamic performance, naturally evolved airfoils are assumed to
be an optimum in the parameter space of lightweight design, aerodynamic performance, and maneuverability. Therefore, the shape of
the hand pinion of a seagull was used as design paradigm, also
because the hand pinion of bird wings is known to be the thrustproducing part [12], which is favorable for future micro air vehicle
design. Seagulls were used as an inspiration because, in contrast to
storks or hummingbirds, they use the apping ight at moderate
reduced frequencies as their normal cruise mechanism. As a result,
the SG04 airfoil was developed [13], which is extruded in the
spanwise direction to a wing segment with an aspect ratio of 2.
2) Verify the TAU code simulations. Previous two-dimensional
FLOWer results have to be reproduced with the TAU code (see
Sec. V.A).
3) The computational results of the TAU code have to be validated.
Therefore, a rigid SG04 wing segment was manufactured in
lightweight design. High-resolution particle image velocimetry of
the boundary layer was used to capture velocity elds and turbulent

Fig. 2

Three-dimensional vortex system of a apping wing [32,33].

Aerodynamic Design

Compared with conventional airfoils, two major design aspects


can be found when examining the airfoil of a seagull in the vicinity of
its hand pinion: rst, a large maximum camber compared with
articial airfoils, and second, the position of maximum thickness is
located close to the leading edge (see also Fig. 3).
There are several reasons why the position of maximum thickness
is situated near the leading edge. Considering the wing anatomy, the
skeleton and muscles run in this section, whereas at the trailing-edge
region of the airfoil, only the feathers determine the airfoil shape.
From an aerodynamic point of view, there are advantages of this
design as well. On the one hand, the adverse pressure gradients along
the upper surface can be kept reasonably small. This yields thin
laminar separation bubbles with low pressure drag losses. On the
other hand, thin airfoils usually exhibit a small range of applicable
angles of attack where no stall occurs. Airfoils with their position of
maximum thickness in the vicinity of the leading edge exhibit an
increased nose radius, which results in a relatively large angle-ofattack range with attached ow. A large relative camber of 8% was
measured by Bilo [14] using narcotized birds. However, observations
in nature revealed that this value is usually smaller for gliding ight
(approximately 4%), although in wind-tunnel experiments with
living birds the maximum camber during one apping stroke was
found to vary from 8 to 12%.
Based on these design aspects, a new birdlike airfoil, the SG04,
was developed [13] (see Fig. 4). This prole represents the hand
pinion of the wing, which is known to be the thrust-producing part of
a apping wing [12], and corresponds in its shape to section 3 in
Fig. 3. Sections 4 and 5 of Fig. 3 were not used as an airfoil design
inspiration, because these sections are only determined by the shape
of feather tips. Beyond, with respect to bending stiffness, such thin
airfoils (t=c < 1%) are difcult to manufacture for validation
experiments in wind tunnels. SG04 has a maximum thickness and a
maximum camber of 4%, where the maximum camber is located at
x=c  40%. The aerodynamic design of the SG04 involved both
aerodynamic analysis and inverse design according to the bubble

Fig. 3 Comparative investigations on airfoil shapes of birds from


Oehme [34]. Shown are ve idealized bird-wing sections.

192

BANSMER AND RADESPIEL

Downloaded by California Inst of Technology on October 22, 2012 | http://arc.aiaa.org | DOI: 10.2514/1.J051226

Fig. 4

Birdlike airfoil SG04.

ramp approach introduced by Selig [15] to reduce the size of the


laminar separation bubble for a broad range of angles of attack.
Figure 5 shows the drag polar of the SG04 airfoil computed by
FLOWer, a validated unsteady Reynolds-averaged NavierStokes
(URANS) solver [7], and XFOIL. The eN method was used for
transition prediction with the critical N factor of 10. This choice
corresponds to the ow quality at the test Reynolds number of
100,000 in the wind tunnel, which was used for the validation
experiments (see also [6]). Although XFOIL is a very simple tool to
predict steady airfoil aerodynamics, its polar is in good agreement
with the results of the more complex FLOWer solver. Additionally,
the drag decompositions into friction and pressure drag are plotted,
which are approximately in the same order of magnitude in this low
Reynolds number range of 100,000. Considering the small pressure
drag values, the design goal of the SG04 airfoil with small laminar
separation bubbles was attained. To be able to compare the polar with
a SD7003 airfoil at an equivalent Reynolds number, experimental
data from Selig et al. [16] is added to the chart. Mainly due to the
higher camber, the SG04 polar is shifted to higher lift values and has
therefore a smaller drag at the same lift performance. However, the
disadvantage of thin airfoils is clearly visible with a decreased
operational range of lift.
For the three-dimensional investigations in this contribution, the
SG04 airfoil was extruded in the spanwise direction to a wing
segment with an aspect ratio of 2.
B.

Kinematics of the Oscillating-Wing Segment

In the two-dimensional case of an oscillating airfoil, the motion is


combined by a plunging zt=T motion and a pitching oscillation
t=T around the quarter-chord for one apping period
0  t=T < 1:
zt=T  z^  cos2t=T
t=T  ^  cos2t=T  =2  0

B
G

geometric angle of attack is superimposed by the induced angle of


attack  resulting from the plunging motion of the airfoil, which
nally yields the effective angle of attack eff :
eff t=T  t=T  t=T  ^  cos2t=T  =2  0
 


1 @zt=T 
 arctan 

U
@t  c
1

x4

eff  sin2t=T  0

(2)

The pitching motion affects similarly the oncoming ow vector.


However, its induced ow velocity in the z direction is a function of
the distance to the rotational axis of the pitching motion. Hence, this
inuence can be interpreted as a decambering of the airfoil.
To generate three-dimensional wing kinematics, rolling motion
^  cos2t=T is superimposed in addition (see Fig. 7).
t=T  
Consequently, the effective angle of attack varies in the spanwise
direction. However, the three-dimensional effect of wingtip vortices
due to the nite span is suppressed by side walls of the wind tunnel.
The reduced frequency k is introduced, often given by

(1)

These motions inuence the local oncoming ow conditions in the


frame of reference of the airfoil. For example, when the airfoil moves
downward, the airfoil in its frame of reference is exposed with an
additional ow component in the upward direction. Consequently,
the effective oncoming ow vector changes direction and magnitude
(see Fig. 6). This also has an impact on the local angle of attack. The

1.2

Fig. 6 Change from geometric to effective angle of attack eff t due
to the plunging motion z_t.

k

fc
U1

(3)

The inverse of k is a measure how far the undisturbed air passes the
airfoil during one apping cycle. Therefore, the reduced frequency
can be used to classify the level of aerodynamic unsteadiness. For
seagull ight conditions, k has values around 0.2. This can be derived
from cruising ight data of birds (see Pennycuick [17] and Herzog
[18]). According to this data, a kelp gull (Larus dominicanus) has a
mean chord length of 0.16 m and a apping frequency of 3.46 Hz.
Assuming a cruise speed of 8 m=s, the reduced frequency can be
determined to be k  0:22. Insects are known to have a reduced
frequency of about 1.0 [19].

0.8
V

0.4
XFOIL, N=10
SD7003 (Selig et al. [16])
FLOWer, N=10

0.000

0.010

0.020

0.030

U
Fig. 5 Drag polar of the SG04 airfoil [7]; Re  100; 000, steady
conditions, and additional SD7003 data [16].

Fig. 7 Oscillating SG04 wing segment. In the experiments, the wing


segment is limited by the wind-tunnel sidewalls. Consequently, wingtip
vortices are suppressed. Caused by the roll motion t, the gap size
between wall and wingtip is continuously changing in the experiments,
which are presented later. A rubber gasket closes this gap.

BANSMER AND RADESPIEL

III. Experimental Investigations


The objective of the experimental investigations is to capture the
oweld in the boundary layer of the oscillating-wing segment and
its turbulent quantities, such as the turbulent shear stress. The
measurements carried out serve as a three-dimensional validation
case and are not prescribing a database for parametric studies.

Downloaded by California Inst of Technology on October 22, 2012 | http://arc.aiaa.org | DOI: 10.2514/1.J051226

A.

Wind Tunnel

The experiments were carried out at the low-speed low-noise wind


tunnel (see Fig. 8). The inlet of the Eiffel-type tunnel is covered by a
eece mat 30 mm in thickness. Afterward, the air passes a
straightener made out of aluminum honeycombs, 14 mm in diameter
and 200 mm in length and then nally through a ne-mesh woven
screen. In the large settling chamber, small-scale turbulence is
dissipated, and a Boerger-type nozzle contracts the air at a 16:1 ratio.
Consequently, the air has a very low turbulence level in the 400
600 mm sized test section. The wind tunnel is driven by a 4 kW,
acoustically encapsulated, speed-controlled, three-phase asynchronous motor, which produces stable wind-tunnel speeds from 2 up to
20 m=s. The diffusor is mounted on a rail system, which allows one
to interchange modular test sections. The laboratory is lined with
open-celled acoustic foam.
B.

Motion Apparatus and Wing Segment

To create a combined plunging and pitching motion of the airfoil


as denoted by Eq. (1), a special apping-motion apparatus [13] was
used. The apparatus synchronizes pitching and plunging motion with
a mechanical gearing, depicted in Fig. 9. A broad range of motion
parameters can be adjusted: plunging amplitude from 0 to 0.1 m,
pitching amplitude from 0 to 25 deg, and apping frequency from 0
to 10 Hz. The rolling motion of the three-dimensional oscillatingwing segment is created by different plunging amplitudes at both
wingtips (again see Fig. 7). Caused by the rolling motion t, the
gap size between wall and wingtip is continuously changing. A
rubber gasket closes this gap. Its smooth shape protects the wingtip

against ow separation and avoids the generation of wingtip vortices


(see Fig. 10). Additionally, a light barrier is mounted at the rig,
enabling the connected measurement systems to be triggered
according to the operated apping frequency. The estimated motion
accuracy for the validation case is about 0.8 mm for the plunging
motion and 0.2 deg for the pitching motion. These values were
measured by capturing the airfoil position at a constant phase for 500
apping cycles and determining its standard deviation.
The SG04 wing segment used for the experimental investigations
was made of a carbon-fabric shell, which was reinforced by a closedcell rigid foam. For a high bending stiffness, a carbon-ber spar was
integrated. A top coat of polyester resin provided a smooth surface.
The mass of the wing segment is 360 g.
C.

Particle Image Velocimetry Measurements

The stereoscopic particle image velocimetry (PIV) was chosen to


capture the oweld in the boundary layer of the apping airfoil. The
advantage of this nonintrusive measurement technique is not only to
gather a complete velocity eld instead of pointwise information, but
also to obtain quantitative data with reasonable accuracy [20]. The
detailed setup is sketched in Fig. 11 as a top view of the wind-tunnel
test section. The oncoming ow from left passes the moving airfoil,
which is driven by the motion apparatus outside the test section. A
double-pulsed Nd:YAG laser (Quantel Brilliant, energy: 2
150 mJ, wavelength: 532 nm) mounted on the top of the wind
tunnel creates a thin light sheet in the chordwise direction. To capture
the three-dimensional aerodynamics around the wing segment, the
light sheet was positioned consecutively at three different spanwise
locations. Two LaVision Imager Intense cameras capture images of
the illuminated tracer particles (oil particles with a mean diameter of
about 1 m). The stereoscopic setup was necessary because the
apping-motion apparatus does not allow for direct visible access
normal to the laser light sheet, i.e., no standard PIV was possible. It
should be mentioned that the eld of view of the cameras was only
2:5 2 cm2 ; otherwise, there would not be sufcient resolution to
capture the boundary-layer ow (see also Fig. 12). These dimensions

Fig. 8 Schematic of the low-speed low-noise wind tunnel.

a) Schematic sketch
Fig. 9

193

b) Setup at wind tunnel


Flapping-motion apparatus [13,31].

194

z /c

BANSMER AND RADESPIEL

0.1
0
0

0.2

0.4

0.6

0.8

x/c

Downloaded by California Inst of Technology on October 22, 2012 | http://arc.aiaa.org | DOI: 10.2514/1.J051226

Fig. 12 Eleven measurement windows were used to capture the


boundary layer of the apping SG04 airfoil.

Fig. 10

Sketch of the rubber-gasket installation (not true to scale).

Fig. 11 Stereoscopic PIV setup as a top view.

are very small compared with the airfoil chord length of 20 cm.
Hence, the camera system had to be attached to a translation device in
order to move the system in the plunging and chordwise directions
without changing the alignment of the cameras. The light-sheet
thickness was adjusted to a value of about 2 mm. This allows to
measure the out-of-plane velocity with the stereoscopic setup. The
short distance between light sheet and camera of about 0.55 m
enabled the camera lens to focus on the light sheet with a small
bandwidth of optical depth. Therefore, it can be ensured that only
particles within the light sheet are captured in the particle images. To
capture the oweld at a constant phase angle, phase-locked imaging
was performed. For this purpose, a transistortransistor logic signal
from the apping-motion apparatus, which triggers at the beginning
of each apping cycle, was captured and shifted in time by a Stanford
delay generator DG 535. When the delayed signal was detected by
the programmable timing unit (PTU9 by LaVision), the laser ash
and the camera exposure were initiated. The PIV system was
controlled by the Davis Software of LaVision. The sampling rate was
adjusted to one particle image pair per apping cycle.

Once the particle image acquisition of at least 500 image pairs for
each of the independently captured measurement windows was
completed, the velocity vector eld of the ow around the airfoil and
its turbulent quantities were determined. First, a wobble correction
was performed as follows: When the laser light sheet touched the
airfoil surface, the so-called reection line was visible on the camera
images. The thickness of this reection line varied from 3 to
40 pixels, depending on the local curvature of the airfoil. Because of
the phase-locked imaging, this reection (which indicates the
position of the airfoil) should be always at the same location.
However, the reection line is wobbling about fractions of 1 mm in
the camera images. This wobbling had to be removed: for the later
ensemble-averaging procedure of the vector elds (to compute the
mean ow vector eld, etc.), the airfoil has to be at the same position.
The wobble correction is composed of three steps:
1) Transform the image from the camera coordinate system into
the world coordinate system. Now, the four particle images of one
capturing cycle (two cameras, each capturing at two moments t and
, P t2 , and Ptt
) are given in the same coordinate
t  t: P t1 ,P tt
1
2
system. In consequence, the correlation procedure, which is
described in the next step, has to be performed only once.
2) A distinctive area of the reection line, which can be found on
each of the 500 image pairs, has to be localized. This area is marked
on the rst of 500 images P t1 jimage1 and will be correlated with the
remaining 499 images Pt1 jimagej (see Fig. 13). The resulting 499
displacement vectors are used to move all 499 images to the origin of
, P t2 , and P tt
are treated with the same
image one. The images P tt
1
2
displacement vectors.
3) Retransform the corrected images from the world coordinate
system into the camera coordinate system.
The presented procedure can only remove translational wobbling,
angular errors of the reection line are not corrected. This could be
done by correlating two points on the reection line instead of one.
However, it is the experience of the authors that angular corrections
diminish the particle image accuracy due to unavoidable subpixel
interpolation. Having performed several image preprocessing
techniques to improve the particle image quality, the particle displacement evaluation was done in a next step using a cross
correlation scheme. The reection line was entirely masked out to
avoid correlation errors near the wing surface. A multipass
interrogation scheme with decreasing interrogation window size
(from 128 128 pixels down to 32 32 pixels), 50% overlap, and
elliptical weighting function was applied. Based upon the
measurements presented in the following sections, the measurement
uncertainty for the ow velocity in the boundary layer can be
evaluated exemplarily by a scheme of Raffel et al. [20] (see Table 1).
The scheme itself was derived from Monte Carlo simulations with
articial particle images. Consequently, each component of the
random error (for instance, the inuence of the particle diameter) can
be investigated individually. This yields a velocity uncertainty in the
boundary layer due to random effects of 0:058 m=s.
The resulting set of at least 500 vector elds for each measurement
window was postprocessed afterward. This was mandatory to lter
out nonphysical vectors, which would impair the results of the
ensemble-averaging procedure. Ensemble-averaging is the statistical
task to compute the mean velocity eld, as given by the equation
0

1
hux; ti
M
1X
v x; t
hvx; ti  @ hvx; ti A 
M i1 i
hwx; ti

195

BANSMER AND RADESPIEL

Fig. 13 Wobble correction. By correlating a distinctive area of the reection line, the particle image can moved to its correct phase position.

Moreover, the turbulent shear stress xz can be computed [21]:

Downloaded by California Inst of Technology on October 22, 2012 | http://arc.aiaa.org | DOI: 10.2514/1.J051226

xz  %  hu0  w0 i  % 

M
1X
u  huiwi  hwi
M i1 i

(4)

Finally, all measurement windows per investigated phase angle


were placed at the correct position on the airfoil to obtain the
complete distribution of the computed quantities along the airfoil
surface. This operation was done with TECPLOT 360.

IV. Numerical Simulation of the


Oscillating-Wing Segment
The numerical simulation approach is based on URANS solutions,
which are coupled with a sound transition-prediction method. Details
of this approach can be found in Radespiel et al. [6] and Krimmelbein
et al. [10,11]. This section only focuses on a brief overview of this
methodology.

A.

Simulation Approach

The time-dependent mean ow was computed by solving the


unsteady Reynolds-averaged NavierStokes equations. Linear
stability analysis, which directly investigates the velocity proles
of the Reynolds-averaged NavierStokes solution was then applied
[10,21]. Waves due to TollmienSchlichting and KelvinHelmholtz
instabilities were predicted by the stability analysis [4]. The
amplication rates were then used to predict the transition location
using an integration scheme for mode amplitude ratios that takes
unsteady ow effects into account. The largest amplitude exponents
were nally compared with a critical N factor in order to determine
the transition location.

B.

NavierStokes Code TAU

The NavierStokes solver TAU [9] requires unstructured meshes;


its discretization scheme is a nite volume approach. A secondorder-accurate central-difference scheme with scalar dissipation
was applied to evaluate convective uxes. Local-time-stepping,
preconditioning, and multigrid operations were performed to
accelerate the computation. A second-order-accurate implicit dualtime-stepping scheme was used for the time-accurate computations.
The Menter two-layer k- model [22] was chosen for the turbulence
modeling.
Table 1

C.

Linear Stability Analysis with LILO

To quantify the local amplication of disturbances, the stability


equations of the laminar boundary layer are solved. LILO [23] treats
laminar compressible boundary layers; here, its option of assuming
parallel ow is used. The harmonic-wave assumption is applied to
the velocity components u, v, w as well as for static pressure p and
temperature T, given exemplarily by the relation
~ y; z; t  q0 x; y; z; t;
q  qx;
q0 x; y; z; t  q^ expjx  y  !t

(5)

Assuming the frequency of the temporal distribution of the basic


ow state is much smaller than the frequency of the single wave
mode, the time dependence of the evolution matrices of the system of
ve stability equations vanishes. Consequently, the stability problem
can be solved at a discrete time. LILO computes the complex
eigenvalue ! of the temporal stability problem in which the user has
to specify the real wave numbers  and  for 3-D cases. LILO is also
able to detect crossow and attachment-line instabilities. However,
in the present validation case, these transition mechanisms did not
occur, neither in the experimental nor in the computational results.
D.

Transition Prediction

In many cases, the location of the nal laminar breakdown to


turbulence is dominated by the behavior of the primary instabilities
with their exponential growth. In these cases the point where the
boundary layer becomes fully turbulent correlates well with a certain
amplication factor of the most unstable primary wave that is
calculated from the point of neutral stability x0. These ndings
constitute the eN method [24]. A suitable mathematical formulation
of this method can be derived from the kinematic wave theory [25].
Disturbance waves with discrete wave frequencies are assumed to
travel downstream and to grow exponentially. Having calculated the
amplication of these waves,
 Z

x
i dx
(6)
Ax  A0  exp 
x0

one can extract the so-called N factor by taking the maximum value
of the amplitude exponent,
 Z

x
i x; ! dx
(7)
Nx  max 
!

x0

and then compare this value with a critical N factor, which yields the
transition location. While this formulation has been applied in steady

Uncertainty estimation of the PIV velocity measurement in the boundary layer [20]

Parameter
Particle diameter
Particle density
Particle displacement
Gradient of the particle displacement
Background noise
Sum of random error/maximum expected error
Root of square sum of random error/uncertainty

Value

Corresponding random error

Conversion

3 pixel
3 particles per 12 12 pixel2
10 pixels
2 pixels=100 pixels
1/28

0.03 pixels
0.03 pixels
0.02 pixels
0.04 pixels
0.02 pixels
0.14 pixels
0.065 pixels

0:126 m=s
0:058 m=s

196

BANSMER AND RADESPIEL

ow cases for many years, a new unsteady integration scheme for N


factors was introduced by Radespiel et al. [6] by using the temporal
evolution of amplication rates over time and space to compute
suitable N factors along the airfoil surface for unsteady ow
problems. Details of the formulation and its numerical implementation are given in [6,26,27]. For low reduced frequencies, the
steady transition prediction is sufciently precise [26] and will be
applied in the present contribution.

Downloaded by California Inst of Technology on October 22, 2012 | http://arc.aiaa.org | DOI: 10.2514/1.J051226

Results

Verication of Numerical Simulations

A two-dimensional motion case (without rolling motion) for the


oscillating SG04 airfoil with a reduced frequency of k  0:2 was
selected for the verication according to data from Pennycuick [17]
and Herzog [18]. Hence, the interesting aspect of aerodynamic
unsteadiness of the apping motion is included in this case. Table 2
displays the detailed parameters. The ow case is typical for cruise
ight of a bird in that a moderate amount of net thrust (cT  0:008) is
produced [7]. Net thrust is the sum of thrust and drag, considering
that these individual forces can only be separated articially (see
Windte and Radespiel [26]).
Both numerical simulations using the URANS code FLOWer, and
PIV measurements of the boundary layer for this two-dimensional
motion case were published in Bansmer et al. [7]. Grid convergence
of the numerical solution was demonstrated using a mesh sequence
of a range between 136 28 and 1088 224. To quantify the
computational uncertainty of the FLOWer results, a gridconvergence study was performed (see Fig. 14). It was found that
the FLOWer results on a 544 112 mesh were essentially grid
converged. It was further shown that the numerical results of the
FLOWer code are in good agreement with the experimental data. The
N factor sensitivity is depicted in Fig. 15. As expected, the plots with
higher critical N factors have their transition location further
downstream: the higher the critical N factor, the lower the turbulence
level, the more time disturbance waves need to grow, resulting in an
increased laminar ow length [7].
The initial objective of this verication is to reproduce the previous
numerical results of the FLOWer code with the URANS code TAU
used in this contribution. In contrast to FLOWer, TAU enables a
three-dimensional transition prediction, which is used later for the
three-dimensional test case. To exclude discretization errors due to
different mesh topologies, the original FLOWer mesh containing
544 112 cells was also used in the TAU computations (see also
Fig. 16). For the computations, the freestream Mach number was set
to 0.02. low-speed preconditioning was performed. A dual-timestepping scheme was used for the time-accurate computations. A
temporal resolution of 500 physical time steps per period was
dened. For transition prediction, the critical N factor was set to 8.
Although the ow quality at the test Reynolds number of 100,000 in
the wind tunnel, which was used for the validation experiments,
demands for a choice of Ncrit  10 [6], we used Ncrit  8 to guarantee
a robust run of the transition prediction. Later in the discussion,
Ncrit  10 is considered again.
When analyzing the rst results, the different convergence
performance between the two solvers was noticed. Figure 17 (left)
demonstrates this difference by plotting the lift coefcient over one
physical time step. While the FLOWer code quickly reaches a
converged state with 70 inner iterations, the TAU code needs about
Table 2

Motion parameters for the case of the oscillating airfoil

Parameter
Reynolds number
Reduced frequency
Chord length
Geometric mean angle of attack
Amplitude of effective angle of attack
Plunging amplitude
Pitch/plunge phase difference

1.2
0.1
1

6
0.8
0
4

0.6
0.4

-0.1

L1
L2
L3

0.2

0.25

0.5

0.75

-0.2 0

Fig. 14 Convergence of lift, drag and transition location for different


grid levels; L1 is the nest level with 1088  224 cells, apping motion
with k  0:2, Re  100; 000 [7].
8

0.8
6
0.6
4
0.4
N =6
N =8

N =10

0.2

0.25

0.5

0.75

Fig. 15 Transition prediction for different critical N factors; apping


airfoil with k  0:2 [7].

Value
Re  100; 000
k  0:2
c  0:2 m
0  4 deg
eff  4 deg
z^  0:1 m
  90 deg

0.1

z/c

V.
A.

300 inner iterations. This is mainly due to more efcient algorithms


for ux computations of the FLOWer code where the structured
meshes in FLOWer enable a direct addressability of cell neighbors.
TAU with its unstructured cell discretization compensates this
disadvantage in computational time by using its parallelization
ability.
In the right part of Fig. 17, the distribution of lift and drag
coefcients and the transition location of the upper side of the airfoil
are plotted against the nondimensional time, starting with t=T  0 at
the top dead center of the apping motion. The results of FLOWer
and TAU are in good agreement. The computational uncertainty of
the TAU results can therefore be assumed to be in the same order of
magnitude as the FLOWer results. Minor discrepancies can be
attributed to the following reasons:
1) There are different calculations of the gradients for the
reconstruction of viscous uxes. TAU uses a GreenGauss ansatz for
this purpose, FLOWer deploys a difference scheme based on a
curvilinear coordinate system due to its structured-mesh ability.
2) For differences in the numerical dissipation, see the dissipation
stencils in Fig. 18. Whereas the neighbor cells i; j 1 and i; j 2

0
0

0.2

0.4

x/c

0.6

0.8

Fig. 16 Computational mesh of the SG04 airfoil with reduced


resolution of 272  56 cells.

197

BANSMER AND RADESPIEL

0.2

1.4
0.15
1

1.2

0.8
0.1

0.05
0.8

0.8

0.6

0.6

-0.05

0.4

0.6
0.4

-0.1

0.2

100

TAU
FLOWer

101

102

TAU
FLOWer

0.4
0

0.25

0.5

0.75

0.2

-0.15
1

-0.2

Downloaded by California Inst of Technology on October 22, 2012 | http://arc.aiaa.org | DOI: 10.2514/1.J051226

Fig. 17 Comparison between FLOWer and TAU for the two-dimensional verication case, k  0:2, Re  100; 000. Left: Convergence of the lift
coefcient during one physical time step. Right: TAU and FLOWer give nearly the same result for one apping period.

contribute to the dissipation in FLOWer, the dissipation in TAU is


additionally inuenced by the diagonal neighbors.
In the next step, the three-dimensional simulation with TAU is
considered. To simulate the oscillating-wing segment with an aspect

Fig. 18 Dissipation stencils of FLOWer and TAU.

ratio of 2, the fully structured mesh was extruded in the spanwise


direction with 32 cells. According to the results in Fig. 19, this
spanwise resolution is sufcient to capture the three-dimensional
aerodynamic effects on the wing segment. The mesh size of two
million cells necessitated parallel computations using 128 CPUs. A
symmetry boundary was dened at both sides of the extruded mesh.
The kinematics of the oscillating-wing segment were identical to the
two-dimensional motion case with a reduced frequency of k  0:2.
Again, a dual-time-stepping scheme was used for the time-accurate
computations. A temporal resolution of 500 physical time steps per
period was dened. The transition prediction was based on data
extracted from three streamlines that were equidistantly distributed
in the spanwise direction. Along these streamlines, the eN method
was applied using a critical N factor of 8.
Figure 19 (left) compares the result of the two- and threedimensional TAU-simulation. The distribution of lift and drag
coefcients and the transition location of the upper side of the airfoil
are plotted over one apping period. The agreement is very good, the
full three-dimensional simulation accurately reproduces the known
two-dimensional result. Several illustrations support the twodimensional characteristics of the ow. For instance, Fig. 19 (right)
shows a plot of the distribution of the specic dissipation rate on an
isosurface of constant eddy viscosity for t=T  0:25. Similar to the
aforementioned two-dimensional results, the transition is located
near to the leading edge during the middle of the downstroke.
Beyond, there are no spanwise variations of the transition location.
Furthermore, the distribution of the shear stress on the surface of the
wing segment demonstrate the two-dimensionality of the ow (see
Fig. 20). The negative values of wall shear stress indicate the reverse
ow due to a laminar separation bubble, whose dimension does not
vary in the spanwise direction. A qualitatively overlaid pressure
distribution with its distinct pressure jump conrms the presence of a
separation bubble.

Fig. 19 Left: Comparison between two- and three-dimensional TAU computation for the oscillating-wing segment, k  0:2, Re  100; 000. Right: The
plot of the specic dissipation rate (turbulence modeling) on an isosurface of constant eddy viscosity shows a two-dimensional ow structure.

198

BANSMER AND RADESPIEL

Ncrit=8
Ncrit=10

0.2

0
0.8

Fig. 20 Distribution of the shear stress on the oscillating SG04 wing


segment at t=T  0:25, k  0:2, Re  100; 000, from TAU computation.
-0.2

Downloaded by California Inst of Technology on October 22, 2012 | http://arc.aiaa.org | DOI: 10.2514/1.J051226

B.

Oscillating-Wing Segment for k  0:25

0.25

A three-dimensional motion case (with roll motion, again see


Fig. 7) for the oscillating SG04 wing segment with an aspect ratio of
two and a reduced frequency of k  0:25 was selected for the
validation according to data from Pennycuick [17] and Herzog [18].
The detailed motion parameters are presented in Table 3. Because of
the spanwise variation of the plunging amplitude, the effective angle
of attack is not constant along the wing. One feature of this motion is
the zero amplitude of the effective angle of attack at the left side of the
wing segment. It is of interest to see, how lift and drag coefcients are
distributed at this location.
A dual-time-stepping scheme was again used for the time-accurate
computations on the mesh with 544 112 32 cells. A temporal
resolution of 500 physical time steps per period was dened. Both
side boundaries of the extruded mesh were dened as slip walls.
Hence, the effect of the wind-tunnel sidewall boundary layers was
neglected. Moreover, the entire grid moved along with the rigid-wing
segment, thereby neglecting the small effects of the upper and lower
wind-tunnel walls as well. For the computations, the freestream
Mach number was set to 0.02 and low-speed preconditioning was
employed. The transition prediction was based on data extracted
from three streamlines that were equidistantly distributed in the
spanwise direction. Along these streamlines, the eN method was
applied using a critical N factor of 8. To study the inuence of the
critical N factor, additional computations were performed with
Ncrit  10. The transition prediction also covered the detection of
crossow instabilities. However, the results of the linear stability
solver LILO demonstrated the absence of such instabilities.
The distribution of lift and drag coefcients for one apping
period is plotted in Fig. 21. The top dead center of the motion is at
t=T  0. During the downstroke, t=T varies between 0 and 0.5. The
plot highlights many similarities compared with the two-dimensional
behavior. For instance, the maximum lift (t=T  0:35) lags the
maximum effective angle of attack (t=T  0:25) with a phase
difference of 35 deg. This is due to the dynamics of start-and-stop
vortices in the wake of the wing segment and can be at least for a twodimensional consideration explained with the inviscid theory of
Theodorsen [28]. The drag coefcient plot indicates negative drag
during the downstroke of the airfoil motion, i.e., the thrust is
Table 3 Motion parameters for the case of the three-dimensional
oscillating-wing segment
Parameter

Left side

Right side

Reynolds number Re
Reduced frequency k
Chord length
c
Mean angle of attack 0
Amplitude of geometric angle of attack ^
Amplitude of effective angle of attack eff
Plunge amplitude z^a
Pitch/plunge amplitude ratio L
Phase difference plunge/pitch 

100,000
0.25
0.2 m

4 deg
7.46 deg
0 deg
0.052 m
1
90 deg

4 deg
7.46 deg
5.22 deg
0.09 m
0.59

a
The difference of the plunging amplitude between the left and right sides of the wing
causes a roll motion.

0.5

0.75

Fig. 21 Lift and drag coefcient over one apping period for the threedimensional validation case, calculated with two different critical N
factors, k  0:25, Re  100; 000.

produced in this part of the cycle mainly due to the KnollerBetz


effect: The effective angle of attack is large in the downstroke due to
plunge velocity and creates lift; however, the negative geometric
angle of attack generates a component of the lift in the upstream
direction that is acting as thrust. The inuence of the critical N factor
on lift and drag is very limited and can be merely observed during the
wing upstroke. The below mentioned transition behavior explains
this nding: the transition locations for both critical N factors of 8 and
10 are deviating from each other by 10% only.
Caused by the rolling motion of the wing segment, the effective
angle of attack is not constant along the wing. Thus, each section of
the wing segment has a different contribution to lift and drag (see
Fig. 22). The plot shows the spanwise distribution of lift and drag
coefcients over one apping period. The amplitude of the effective
angle of attack eff increases in positive y direction, the middle
section of the wing segment is at y=b  0:5. On the side with large
amplitudes of the effective angle of attack (y=b > 0:5), the largest lift
variations occur. In particular, the phase difference between effective
angle of attack and lift obtains its smallest value in this region. This
phase difference increases for decreasing eff because the pitch/
plunge amplitude ratio L is increasing from 0.59 to 1 in this special
validation case. The mean angle of attack with its value of 4 deg does
not change in the spanwise direction. Hence, the mean lift coefcient
remains almost constant along the wing segment, c z
0:85. The
drag coefcient also has its largest temporal variation for large eff .
However, the phase difference between effective angle of attack and
drag coefcient is nearly constant in the spanwise direction.
The rolling motion of the wing segment also affects the transition
characteristics over one apping period (see Fig. 23). The distribution of the wall shear stress cf for t=T  0:375 and the overlaid
pressure distribution indicate a laminar separation bubble, whose
spatial extension is varying in the spanwise direction. However,
nonexisting crossow instabilities and the fact that the tips of the
wing segment are modeled as slip walls yield a nearly linear
transition distribution in the spanwise direction. Turbulence and
transition are further investigated and compared with experimental
data for three spanwise sections (for eff  f4:2 ; 2:7 ; 1:2 g) in
Fig. 24. In each of the three sections, the transition moves toward the
leading edge during the downstroke and back toward the trailing
edge during the upstroke motion. The smaller the amplitude of the
effective angle of attack eff , the less the transition moves along the
airfoil surface. The transition locations computed by the TAU code
for the two different critical N factors show only small differences of
about 10% and match well with the experimental determined
transition locations from the PIV measurements.
The approach to extract transition locations from PIV data shall be
discussed briey (see Fig. 25). Shown are the turbulent shear-stress
distributions xz from the PIV experiments and the numerical
computations
for
three
spanwise
sections
(eff 
f4:2 deg; 2:7 deg; 1:2 degg) at four different phase angles starting

199

BANSMER AND RADESPIEL

Downloaded by California Inst of Technology on October 22, 2012 | http://arc.aiaa.org | DOI: 10.2514/1.J051226

Fig. 22 Spanwise distribution of lift and drag coefcients over one apping period, Ncrit  10, k  0:25, Re  100; 000.

was used to recover the turbulent shear stress from the URANS
solution. The results demonstrate good agreement between
numerical and experimental data. However, a closer look at the
results of the numerical simulation reveals strong local overshoots of
u0 w0 at the location of laminar turbulent transition, whereas the
experimental u0 w0 distribution is smooth along the surface. This is
due to the two-layer k-" turbulence model, whose dominant property
in turbulence production is well known [29]. For future
computations, it will be interesting to see the performance of
Reynolds-stress turbulence models, particularly because they are
able to reproduce the anisotropy of turbulence for this strongly
nonequilibrium ow case. Previous investigations of a twodimensional apping airfoil [7] revealed the relaminarization process
for 0:5  t=T  0:75 to be the most difcult to simulate.
Interestingly, the relaminarization process in the present threedimensional simulation is well captured.
Can the three-dimensional aerodynamics of the apping-wing
segment simply be reproduced as a spanwise sequence of twodimensional airfoil aerodynamics at equivalent effective angle of
attack? Figure 26 answers this question. Shown are lift and drag
coefcients and transition location over one apping period for two
spanwise sections of the wing segment. The equivalent twodimensional aerodynamics of the SG04 airfoil, which were
computed with TAU, are overlaid, assuming the airfoil moves with
the local motion of each wing section. Hence, the same distribution of
the effective angle of attack over one apping period is ensured for
each two-dimensional apping case with its corresponding wingsection counterpart.
Compared with the two-dimensional case, there is less lift
produced for the wing section with a large amplitude of the effective
angle of attack (left plot) during the apping downstroke in the threedimensional case. This is due to the vortex dynamics in the wake of

0.009
0.007
0.004
0.002
0.000
-0.002
-0.004
-0.006
-0.008
-0.010

Fig. 23 Shear-stress distribution on the wing segment for t=T  0:375,


k  0:25, Re  100; 000, from TAU computation. The negative values of
shear stress and the overlaid pressure distribution indicate the presence
of a laminar separation bubble, whose spatial extension varies in the
spanwise direction.

from t=T  0 at the top dead center. The turbulent shear stress
characterizes the viscous momentum transport across the boundary
layer and can be used to detect regions of turbulent ow. Thus, the
transition location in the measurements can be dened as the
beginning of the turbulent wedge that starts from the shear layer of
the LSB. To avoid errors in its localization due to insufcient
resolution of apping of the laminar part of the LSB, the point where
2
reaches 0.1% and
the normalized Reynolds shear stress u0 w0 =U1
demonstrates a clearly visible rise is dened as the transition
position. Derived from the Boussinesq approximation, the relation


 @u @w

(8)
u0 w 0   t
% @z @x

TAU, Ncrit=8
TAU, Ncrit=10
experiment
eff

0.8

1
8

0.8
5

0.8
6

5
0.6

0.6

0.6

4
0.4

4
0.4

4
0.4

3
3

0.2

0.2

2
0.2
0

0.25

0.5

0.75

2
1

0.25

0.5

0.75

1
1

0.25

0.5

0.75

Fig. 24 Comparison of the transition locations over one apping period with experimental PIV data for three selected spanwise sections.

Downloaded by California Inst of Technology on October 22, 2012 | http://arc.aiaa.org | DOI: 10.2514/1.J051226

200
BANSMER AND RADESPIEL

Fig. 25

PIV data and numerical simulation.

201

BANSMER AND RADESPIEL

3D simulation
2D simulation
1.2

0.2

1
1

0.1

0.8
0.1

0.8

0.8

0.8
0.6

0.4

-0.1

0.2

Downloaded by California Inst of Technology on October 22, 2012 | http://arc.aiaa.org | DOI: 10.2514/1.J051226

0.4

0.6

0.6

0.6

0.25

0.5

0.75

0.2

0.4

0.4
0.2
0

-0.1

0.25

0.5

0.75

0.2

Fig. 26 Comparison of lift and drag coefcients and transition location over one apping period for two selected spanwise sections with their equivalent
two-dimensional aerodynamics.

vortices in the three-dimensional case creates an efciency loss of


about 10%.

100%

80%

VI. Conclusions

60%

40%

20%

3D simulation
2D simulation

0.2

0.4

0.6

0.8

Fig. 27 Comparison of the propulsive efciency for the two- and threedimensional case.

the wing segment. The unsteady part of the 3-D wake is composed of
start-and-stop vortices with a strength that varies in the spanwise
direction, and in between these spanwise structures the formation of a
streamwise vortex street takes place, so that conservation of vorticity
is fullled. The induced velocities of the streamwise vortex street
change lift and drag coefcients of the wing segment as shown in
Fig. 26. In particular, they explain the induced drag of the wing
segment that lowers the propulsive efciency p of the apping
motion. According to Windte and Radespiel [26], the propulsive
efciency can be determined by
ct  cx  cx;stat 0 

(9)

RT



ct U1 dt

_ dt y=bconst
_  cy y_  cz z_  cm  c
0 cx x

p  R T

A birdlike wing segment is considered, oscillating with a


nominally three-dimensional combined pitchplungeroll motion at
a reduced frequency of k  0:25 in the transitional Reynolds number
range of 100,000. A sophisticated numerical computation scheme
based on the unsteady Reynolds-averaged NavierStokes solver
TAU was set up. To validate the numerical simulations, the boundary
layer of the oscillating-wing segment was investigated with highresolution stereoscopic particle image velocimetry. Good agreement
in the ow prediction has been demonstrated. The transition
locations of experimental and numerical result are especially in good
agreement. The two-layer k-" turbulence model delivers generally
adequate results; however, systematic discrepancies between
simulation and experiment were found regarding the turbulence
intensity. The results demonstrate that the numerical simulation
scheme provided in this contribution may be deemed as acceptable
tools for conceptualization and parametric studies to understand the
apping-wing aerodynamics of natural yers. An analysis of the
three-dimensional simulation also highlighted its signicance: the
three-dimensional aerodynamics of the apping-wing segment
cannot be reproduced as a spanwise sequence of two-dimensional
results due to the complex vortex structure in the wake.

Acknowledgments
As a part of the priority program SPP 1207, this project is funded
by the German Research Foundation (DFG). The majority of
computations were performed at the facilities of the North-German
Supercomputing Alliance (HLRN). The authors thank A. Probst, S.
Mahmood, and N. Krimmelbein for valuable discussions and their
support.

(10)

This equation also contains the thrust coefcient ct . In contrast to


the thrust determination of conventional aircraft, thrust and drag can
only be separated articially for apping yers. In Eq. (9), this is
done by subtracting the drag of the SG04 airfoil at the mean angle of
attack from the force coefcient cx . Figure 27 shows the spanwise
distribution of the propulsive efciency for the three-dimensional
TAU computation. The efciency is not constant along the span. As
expected, on the side with large amplitudes of the effective angle of
attack eff and low pitch/plunge amplitude ratios L (y=b > 0:5)
the propulsive efciency is larger. The efciencies of the twodimensional reference cases of Fig. 26 are indicated as diamonds and
prove the above mentioned hypothesis: the formation of streamwise

References
[1] Abate, G., Ol, M., and Shyy, W., Introduction: Biologically Inspired
Aerodynamics, AIAA Journal, Vol. 46, No. 9, 2008, pp. 21132114.
doi:10.2514/1.35949
[2] Neef, M., Analyse des Schlagugs Durch Numerische Strmungsberechnung, Ph.D. Thesis, Inst. of Fluid Mechanics, Technische
Universitt Braunschweig, Braunschweig, Germany, 2002.
[3] Spalart, P., and Strelets, M., Mechanisms of Transition and Heat
Transfer in a Separation Bubble, Journal of Fluid Mechanics, Vol. 403,
2000, pp. 329349.
doi:10.1017/S0022112099007077
[4] Rist, U., Instability and Transition Mechanisms in Laminar Separation
Bubbles, RTO-AVT-VKI Lecture Series 104, von Karman Inst. of
Fluid Dynamics, Rhode -Saint-Gense, Belgium, 2003.
[5] Oertel, H., and Delfs, J., Strmungsmechanische Instabilitten,
Springer, Berlin, 1996,

Downloaded by California Inst of Technology on October 22, 2012 | http://arc.aiaa.org | DOI: 10.2514/1.J051226

202

BANSMER AND RADESPIEL

[6] Radespiel, R., Windte, J., and Scholz, U., Numerical and Experimental
Flow Analysis of Moving Airfoils with Laminar Separation Bubbles,
AIAA Journal, Vol. 45, No. 6, 2007, pp. 13461356.
doi:10.2514/1.25913
[7] Bansmer, S., Radespiel, R., Unger, R., Haupt, M., and Horst, P.,
Experimental and Numerical Fluid-Structure Analysis of Rigid and
Flexible Flapping Airfoils, AIAA Journal, Vol. 48, No. 9, 2010,
pp. 19591974.
doi:10.2514/1.J050158
[8] Kroll, N., Rossow, C. C., Schwamborn, D., Becker, K., and Heller, G.,
MEGAFLOW: A Numerical Flow Simulation Tool for Transport
Aircraft Design, International Council of the Aeronautical Sciences,
Paper 1105, Toronto, 2002.
[9] Kroll, N., and Fassbender, J., MEGAFLOWNumerical Flow
Simulation for Aircraft Design, Notes on Numerical Fluid Mechanics
and Multidisciplinary Design, Vol. 89, Springer, New York, 2005,
pp. 8192.
[10] Krimmelbein, N., and Radespiel, R., Transition Prediction for ThreeDimensional Flows Using Parallel Computation, Computers and
Fluids, Vol. 38, 2009, pp. 121136.
doi:10.1016/j.compuid.2008.01.004
[11] Krimmelbein, N., Radespiel, R., and Nebel, C., Numerical Aspects Of
Transition Prediction for Three-Dimensional Congurations, 35th
AIAA Fluid Dynamics Conference & Exhibit, AIAA Paper 2005-4764,
Toronto, 69 June 2005.
[12] Bilo, D., Flugbiophysik von Kleinvgeln; II. Kinematik und
Aerodynamik des Flgelaufschlages beim Haussperrling (Passer
domesticus L.), Ph.D. Thesis, Department of Zoology, University of
Munich, Munich, 1970.
[13] Bansmer, S., Scholz, U., Windte, J., Khler, C., and Radespiel, R.,
Flow Field Measurements on an Oscillating Airfoil for Flapping Wing
Propulsion, AIAA Paper 2008-581, 2008.
[14] Bilo, D., Flugbiophysik von Kleinvgeln: I. Kinematik und
Aerodynamik des Flgelabschlages beim Haussperrling (Passer
domesticus L.), Ph.D. Thesis, Dept. of Zoology, Univ. of Munich,
Munich, 1970.
[15] Selig, M., Low Reynolds Number Airfoil Design Lecture Notes, VKI
Lecture Series, von Karman Inst. of Fluid Dynamics, Rhode -SaintGense, Belgium, 2428 Nov. 2003.
[16] Selig, M., Guglielmo, J., Broeren, A., and Gigure, P., Summary of LowSpeed Airfoil Data, Vol. 1, SoarTech, Virginia Beach, VA, 1995.
[17] Pennycuick, C., Wingbeat Frequency of Birds in Steady Cruising
Flight: New Data and Improved Predictions, Journal of Experimental
Biology, Vol. 199, 1996, pp. 16131618.
[18] Herzog, K., Anatomie und Flugbiologie der Vgel, Gustav Fischer
Verlag, Stuttgart, 1968.
[19] Byrne, D., Buchmann, S., and Spangler, H., Relationship Between
Wing Loading, Wingbeat Frequency and Body Mass in Homopterous
Insects, Journal of Experimental Biology, Vol. 135, 1988, pp. 923.
[20] Raffel, M., Willert, C. E., Wereley, S. T., and Kompenhans, J., Particle

[21]
[22]

[23]
[24]

[25]
[26]

[27]

[28]
[29]
[30]

[31]

[32]

[33]

[34]

Image VelocimetryA Practical Guide, Springer, New York,


2007.
Hinze, J. O., TurbulenceAn Introduction to Its Mechanism and
Theory, McGrawHill, New York, 1959.
Menter, F., Two-Equation Eddy-Viscosity Transport Turbulence
Model for Engineering Applications, AIAA Journal, Vol. 32, No. 8,
1994, pp. 15981605.
doi:10.2514/3.12149
Schrauf, G., LILO 2.1 Users Guide and Tutorial, Ver. 2.1, GSSC
TR 6, 2006.
Van Ingen, J. L., Suggested Semi-Empirical Method for the
Calculation of the Boundary Layer Transition Region, Department of
Aerospace Engineering, Delft University of Technology, Rept. VTH74, Delft, The Netherlands, 1956.
Schrauf, G., Transition Prediction Using Different Linear Stability
Analysis Strategies, AIAA Paper 94-1848-CP, 1994.
Windte, J., and Radespiel, R., Propulsive Efciency of a Moving
Airfoil at Transitional Low Reynolds Numbers, AIAA Journal, Vol. 46,
No. 9, 2008, pp. 21652177.
doi:10.2514/1.30569
Windte, J., Scholz, U., and Radespiel, R., Validation of the RANSSimulation of Laminar Separation Bubbles on Airfoils, Aerospace
Science and Technology, Vol. 10, No. 6, 2006, pp. 484494.
doi:10.1016/j.ast.2006.03.008
Theodorsen, T., General Theory of Aerodynamic Instability and the
Mechanism of Flutter, NACA TM 496, 1935.
Murakami, S., Overview of Turbulence Models Applied in CWE1997, Second European and African Conference on Wind
Engineering, Genova, Italy, June 2226 1997.
Horton, H., Laminar Separation Bubbles in Two and Three
Dimensional Incompressible Flow, Department of Aeronautical
Engineering, Queen Mary College, University of London, London,
1968.
Bansmer, S., Strmungsanalyse an Einem Zwei- und Dreidimensional
Schlagenden FLGELSEGMENT, Ph.D. Thesis, Inst. of Fluid
Mechanics, Technische Universitt Braunschweig, Braunschweig,
Germany, 2002.
Ellenrieder, K., Parker, L., and Soria, J., Flow Structures Behind a
Heaving and Pitching Finite-Span Wing, Journal of Fluid Mechanics,
Vol. 490, 2003, pp. 129138.
doi:10.1017/S0022112003005408
Parker, L., Soria, J., and Ellenrieder, K., Thrust Measurements from a
Finite-Span Flapping Wing, AIAA Journal, Vol. 45, No. 1, 2007,
pp. 5870.
doi:10.2514/1.18217
Oehme, H., Vergleichende Proluntersuchungen an Vogelgeln,
Beitrge zur Vogelkunde, Vol. 16, Nos. 16, 1970.

F. Ladeinde
Associate Editor

Das könnte Ihnen auch gefallen